You are on page 1of 328

THE ANTI-DYNAMIC TUBE IN

MASS FLOW SILOS

A THESIS SUBMITTED TO THE UNIVERSITY OF


WESTERN AUSTRALIA FOR THE DEGREE OF
DOCTOR OF PHILOSOPHY

DEPARTMENT OF CIVIL AND


RESOURCE ENGINEERING

DESIREE NORTJE

2002
ABSTRACT

Storage of granular solids in silos has been the practice for many years. Engineers
have been faced with the problem of making the silos empty more efficiently and
minimising the forces acting on the walls of the silo during material discharge. To
this end the anti-dynamic tube was invented. The tube has a smaller diameter than
the silo and consists of several portholes along its height and around its
circumference. When the discharge gate of the silo is opened the granular material
enters the tube through the portholes, flows down the inside of the tube and exits
the silo through the discharge gate. Most tubes have been installed such that there
was sufficient space between the base of the tube and silo bottom for the granular
material to flow simultaneously through the discharge gate.

The flowing material causes a down drag on the tube from the friction of the
granular material on the walls of the tube. Previous research has underestimated
the magnitude of these frictional forces resulting in catastrophic buckling failure of
the tubes, blocking the discharge gate of the silo. A blockage of the discharge gate
requires top emptying of the silo resulting in financial losses and down time of
equipment. A steel model silo with an anti-dynamic tube was set up in the
laboratory to measure the friction on the tube during material flow. From the
results of these experiments, an equation has been derived to estimate the
magnitude of the down-drag force. Furthermore, an empirical expression was
developed for the effects of the speed of the flowing material on the magnitude of
the down-drag force.

To keep construction costs down, it is necessary to optimise the wall thickness of


the tube. There is currently no theory for the buckling capacity of a thin walled
cylindrical shell with multiple perforations around its height and circumference.
Therefore additional experiments were undertaken on a cylindrical shell with
multiple perforations subjected to a combination of an axial as well as an external
lateral pressure. Following on from the experiments, finite element analyses were
undertaken to compare with the experimental results. For each finite element
analysis an out-of-roundness was introduced as an initial wall imperfection. From
these analyses and the cylinder experiments, a method of producing interaction
curves for tubes with varying ratios of open area has been developed.
DISCLAIMER

No portion of the work presented in this thesis has been submitted in support of an
application for another degree or qualification from this, or any other, university or
institute of learning.

Désirée Nortje, August 2002.


ACKNOWLEDGEMENTS

My sincere thanks are given to my supervisor, Dr Ken Kavanagh, for his


encouragement, insight, informality, approachability, understanding and extreme
patience. Without his motivation and kindness I would have given up in the early
stages of the project. I would also like to thank Dr Kavanagh for supporting the
application for my research scholarship provided by the Department of Civil and
Resource Engineering.

Thanks are also due to the workshop staff who assisted me in setting up both my
model silo and cylinder experiments. Special thanks goes to Jim Carrol, Jim Waters,
and Neil McIntosh for their assistance with the laboratory machinery. Without their
help and guidance the experiments would have been cumbersome, difficult to
execute and probably never would have happened. Thanks are also due to
Wladyslaw Bzdyl and Sun Nichersen for maintaining the electrical equipment and
their endless patience in explaining my queries.

ACI Glass, Penrith Plant, kindly allowed me to print and bind my thesis at their
offices. I am very grateful for their generosity in making space available for me and
providing invaluable assistance for binding.

Finally, there are not enough words which can describe my gratitude to my
husband, Richard, and my two daughters, Stephanie and Jennifer, for their patience
while I undertook this PhD research. Many of our decisions revolved around this
project and it’s eventual completion date. We can now finally start making our
plans become a reality.

Thank you to all.

Desiree Nortje, August 2002


LIST OF SYMBOLS

hopper half angle


a constant
stress ratio

c
strain in the circumferential direction

r
strain in the r direction

strain in the q direction

angle in the circumferential direction in the hopper

m
internal angle of friction of the granular material

t
anti-dynamic tube wall friction angle

w
wall friction angle

material bulk density

angle in the material. w


< < m

angle subtended by the hopper axis and centered at the hopper vertex

c
vertical stress on the silo axis

h
horizontal stress

J
Janssen static vertical pressure

JH
Janssen static horizontal pressure

v
vertical pressure

vw
vertical stress adjacent to the hopper wall

shearing stress

v
shear stress in the vertical direction at an arbitrary distance from the silo

axis

vw
shear stress in the vertical direction adjacent to the hopper wall

angle of inclination of the major principal stress

A silo cross sectional area


surface area of a cylindrical shell
Ac plan area of a cut out in the wall of a cylindrical shell

AP projected plan area of an object


B factor derived from the Mohr circle in the circumferential section of the silo
Cc Reimberts characteristic constant of the silo

constant
dt diameter of the anti-dynamic tube

dg diameter of the silo discharge gate

D silo diameter
flexural rigidity of a thin walled shell
E Youngs modulus
factor derived from the Mohr circle in the hopper
a constant
F vertical pressure distribution factor in the cylindrical section of the silo
a constant
FH vertical pressure distribution factor in the hopper

G a constant
h height of cone of material above the top of the silo
h step size used in the Runge-Kutta method of numerical integration
H total height of the silo
Hh height of hopper

ht height of the anti-dynamic tube

iN ratio of zN to hopper height =zN/ Hh

K stress ratio
K0 at rest stress ratio

Ka active stress ratio

Knn factor used in the Runge-Kutta method of numerical integration

Kp passive stress ratio

L length of a cylindrical shell


m coefficient of friction
M bending moment
N axial force
Nx axial load at buckling of a cylindrical shell with cut outs

N0 axial load at buckling of a cylindrical shell with no cut outs

P silo perimeter
Q stress ratio
q shear stress
R hydraulic radius = A/P
r radius of a thin walled cylinder
rc radius of a cut out in the wall of a cylindrical shell
r0 radius of the hopper at the level of the transition

rav average radius

rb radius of the bottom of an element in the hopper

rt radius of the top of an element in the hopper

sN pressure normal to the hopper wall

Sv dimensionless pressure ratio = v


/ D

t wall thickness of a cylinder


u,v,w displacements in the cartesian plane
Uc strain energy in the circumferential direction

UB bending energy

x,y,z cartesian co-ordinates


z depth co-ordinate
Z dimensionless depth ratio = z/H
z element thickness
zN depth of the maximum pressure in the hopper
CONTENTS
ABSTRACT

DISCLAIMER

ACKNOWLEDGEMENTS

LIST OF SYMBOLS

CHAPTERS

1 INTRODUCTION
1.1 Introduction 1.1
1.2 Silo Types and Flow Patterns 1.6
1.3 Silo Inserts 1.7
1.4 Introduction to Wall Pressures 1.10

2 CLASSIC WALL PRESSURE THEORIES


2.1 STATIC WALL PRESSURES
2.1.1 CYLINDRICAL SECTION
2.1.1.1 Janssen 2.1
2.1.1.2 Reimbert 2.4
2.1.1.3 Janssen vs Reimbert’s theory 2.7
2.1.2 HOPPER SECTION
2.1.2.1 Walker 2.8
2.1.2.2 Jenike Radial Pressures 2.11
2.1.2.2.1 Linear Normal Wall Pressure 2.11
2.1.2.2.2 Radial Pressure Field in the Solid 2.16
2.1.2.2.3 Position of the Maximum Pressure in the Hopper 2.21
2.1.2.3 Walters Static Hopper Pressures 2.23

2.2 DYNAMIC WALL PRESSURES


2.2.1 CYLINDRICAL SECTION
2.2.1.1 Walters 2.25
2.2.2 HOPPER SECTION
2.2.2.1 Flow/Slip in the Hopper by Equilibrium of a Slice 2.33
2.2.2.2 Walters Pressures in Converging Channels 2.39
2.2.2.3 Jenike Radial Stress Field 2.49

2.3 SWITCH PRESSURES


2.3.1 CYLINDRICAL SECTION
2.3.1.1 Jenike Upper Bound Pressures 2.62
2.3.1.2 Walters Switch Pressure in the Cylinder 2.68
2.3.2 HOPPER SECTION
2.3.2.1 Jenike Switch Pressure in the Hopper 2.74
2.3.2.2 Walters Switch Pressure in the Hopper 2.78

3 WALL PRESSURE MEASUREMENTS


3.1 LITERATURE SURVEY
3.1.1 STATIC PRESSURES
3.1.1.1 Cylindrical Section 3.1
3.1.1.2 Hopper Section 3.3
3.1.2 DYNAMIC PRESSURES
3.1.2.1 Cylindrical Section 3.5
3.1.3 STRESS RATIOS 3.8

3.2 EXPERIMENTAL SET-UP


3.2.1 Steel Model 3.11
3.2.2 Bulk Solid material 3.12
3.2.3 Data Acquisition 3.12
3.2.4 Strain Gauge Bridges 3.14
3.2.5 Floating Pressure Cells 3.16
3.2.5.1 Ball type pressure cell 3.17
3.2.5.2 Tube type pressure cell 3.17
3.2.5.3 Plate type pressure cell 3.18
3.2.6 Pressure Cell Calibration 3.19
3.2.7 Multi-turn potential meters 3.21
3.2.8 Gate Switches 3.22

3.3 EXPERIMENTAL RESULTS


3.3.1 Description 3.23
3.3.2 Static Tests 3.24
3.3.3 Dynamic Tests 3.27
3.3.3.1 Switch Pressure 3.30
3.3.4 Stress Ratios 3.32

4 ANTI-DYNAMIC TUBE THEORY


4.1 LITERATURE SURVEY
4.1.1 Pieper 4.1
4.1.2 Reimbert 4.2
4.1.3 Ravenet 4.2
4.1.4 McLean 4.3
4.1.5 Ooms and Roberts 4.4
4.1.6 Kaminski and Zubrzycki 4.7
4.1.7 Schwedes and Schulze 4.10

4.2 EXPERIMENTAL SET-UP


4.2.1 Anti-Dynamic Tube Model 4.12

4.3 EXPERIMENTAL RESULTS 4.15

4.4 MATHEMATICAL MODEL 4.23


4.4.1 Tube Parameters 4.25
4.4.2 Variable Vertical Pressure across a Slice 4.26

5 BUCKLING OF THIN CYLINDRICAL SHELLS


5.1 ELASTIC SHELL BUCKLING THEORY
5.1.1 Cylinder subjected to uniform external lateral pressure 5.1
5.1.2 Cylinder subjected to axial pressure
5.1.2.1 Special Case 5.10
5.1.2.2 General Case 5.19
5.1.3 Cylinder subjected to combined axial and lateral pressure 5.24

6 PERFORATED CYLINDRICAL SHELLS


6.1 LITERATURE SURVEY
6.1.1 Tennyson 6.1
6.1.2 Almroth and Holmes 6.3
6.1.3 Starnes Jr 6.5
6.1.4 Scutella 6.7
6.2 DISCUSSION 6.10

6.3 PERFORATED CYLINDER EXPERIMENTS


6.3.1 Experimental set-up 6.14
6.3.2 Experimental results 6.19

6.4 FINITE ELEMENT ANALYSIS


6.4.1 Description 6.26
6.4.2 Cylinder with 16.5% open area 6.30
6.4.3 Cylinder with 36.6% open area 6.35
6.4.4 Solid Cylinder 6.39

6.5 COMPARISON WITH LABORATORY TESTS


6.5.1 Cylinder with 16.5% Open Area 6.43
6.5.2 Cylinder with 36.6% Open Area 6.44
6.5.3 Interaction Plots for Cylinders with Multiple Perforations 6.45

7 CONCLUSIONS

7.1 SILO WALL PRESSURES

7.1.1 Static pressures 7.1

7.1.2 Dynamic pressures 7.2

7.1.3 Switch pressures 7.3

7.1.4 Stress Ratios 7.4

7.2 ANTI-DYNAMIC TUBE FRICTIONAL DRAG 7.5

7.3 PERFORATED CYLINDERS 7.8

7.4 INTERACTION CURVES 7.8

8 APPENDICES

8.1 APPENDIX A: IMPLEMENTATION OF THE RUNGE-KUTTA


METHOD
8.1.1 The Runge-Kutta Equations A.1
8.1.2 Equilibrium Slice Method A.1

8.2 APPENDIX B: CALIBRATION CONSTANTS


8.2.1 Pressure Cell Calibration B.1
8.2.2 Anti-Dynamic Tube Support Calibration B.3
8.3 APPENDIX C: CHECK LISTS
8.3.1 Pre-Static Test Check List C.1
8.3.2 Pre-Dynamic Test Check List C.2

8.4 APPENDIX D: MODEL SILO WALL PRESSURES TESTS


8.4.1 Static Test Results D.1
8.4.2 Dynamic Test Results D.5

8.5 APPENDIX E: ANTI-DYNAMIC TUBE TESTS


Frictional Drag Test Results E.1

8.6 APPENDIX F: SHELL THEORY


8.6.1 Uniformly Compressed Circular Ring F.1
8.6.2 Flexural Rigidity of a Shell F.1

8.7 APPENDIX G: PERFORATED CYLINDER TEST RESULTS


8.7.1 Cylinders with 16.5% Open Area G.2
8.7.2 Cylinders with 36.6% Open Area G.4
8.7.3 Lateral Pressure Tests G.9

8.8 APPENDIX H: EIGENVALUE BUCKLING MODE SHAPES


8.8.1 Cylinders with 16.5% Open Area H.1
8.8.2 Cylinders with 36.6% Open Area H.4
8.8.3 Solid Shell H.7

9 REFERENCES
INTRODUCTION 1.1

CHAPTER 1
INTRODUCTION

1.1 INTRODUCTION

A silo is a structure well known to most people. It’s use is for the storage of any
bulk material which is of a granular nature such as grain, wheat, lupins, salt, sugar,
cement, coal, etc… While there are many silos in existence, this does not imply that
all knowledge about silos has been determined and that very little is left to still be
discovered. Silos in all sizes are being constructed all around the world, some of
which operate very successfully, and others which do not.

Silo discharge is classified into two main groups, either concentric or eccentric
discharge. In eccentric discharge the gate is off-centre with respect to the centre
line of the silo. Due to this eccentricity, flowing material causes large bending
stresses on the walls of the silo. These bending stresses are erratic in nature and
difficult to predict due to the erratic nature of eccentric flow. In the case of
concentric discharge, the centre of the gate aligns with the centre line of the silo, or
the centre of the group of gates aligns with the centre line of the silo. The research
covered in this thesis focuses only on concentrically discharging silos.

Generally speaking a silo consists of two main sections, namely a hopper and an
upper bin. The joint between the upper bin and the hopper is referred to as the
transition. A flat bottom silo has an effective transition which is formed within the
stored material which does not exit the silo during discharge. This remaining
material is referred to as the “dead material” . Figure 1.2b shows a flat bottom silo
with dead material forming an effective transition with the walls.

Silos vary in shape from circular to square and rectangular. Depending on the
shape of the silo, there may be one discharge gate in the hopper as for the circular
case, or several discharge gates as for the rectangular silo. Figure 1.1 a,b,c shows
some of the typical silos in use. In figure 1.1d, a bank of silos has been shown
where the interstitial areas between the silos have also been used for the storage of
material, shown by the shaded area. Some silos also have their cones inverted as
shown in figure 1.2. This type of silo is used mainly for storing and blending bulk
materials which are in powder form, such as raw meal, cement and lime. This type
of silo has not been considered in this thesis.
1.2 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

(b) (c) (d)


square or rectangular silo a bank of
doubly with a group of circular silos
symmetrical outlet gates
silo

Figure 1.1: Typical silo geometries.

Effective
transition

Dead material

Figure 2.1a: Figure 2.1b: Flat bottom


Inverted cone silo.
silo

Some of the problems associated with the storage of bulk materials are the
segregation of particles of varying sizes, excessive wear on the walls due to the
flowing material and cracked walls from the flowing material. The wall pressures
are generally referred to as the static and dynamic pressures, ie the filling and
emptying pressures, respectively.

One of the methods proposed to alleviate these problems is a silo insert called the
anti-dynamic tube. Other names for the anti-dynamic tube are the tremmie tube,
decompression tube, discharge tube and static-flow pipe. This method consists of
placing a tube centrally inside the silo, which has a smaller diameter than the silo.
The tube may extend the full height of the silo with multiple perforations around its
circumference and along its length. The material then flows into the tube through
INTRODUCTION 1.3

the holes, and down the tube to the discharge gate. This type of tube causes the
silo to empty in successive layers resulting in a first-in-last-out situation. For
materials which degrade with time (either biologically or mechanically) this is highly
undesirable. An alternative arrangement is a shorter tube which extends only a
portion of the silo height. The optimum length of the tube is determined from the
internal friction angle of the material. In this arrangement the silo empties in two or
three stages only. In some instances, port holes are accommodated at the base of
the tube allowing material discharge to occur simultaneously from the bottom of
the silo as well as through the tube at higher levels.

This PhD thesis considers anti-dynamic tubes placed in mass flow silos storing free
flowing granular materials, such as sand, grain, lupins etc. This introductory
chapter gives a background to the various silo types and flow patterns. A brief
background of the different types of silo inserts used to overcome material flow and
wall pressure problems has been given, as well as a general description of the wall
pressures acting on the silo during filling and material discharge.

Chapter two consists of a study of the classic theories for static, dynamic and
switch pressures acting on the wall of the silos. This includes the Janssen, Jenike,
Walker and Walters theories for the pressures in a silo. From these pressure
theories it has been established that one of the main factors affecting the
determination of the horizontal wall pressures is the assumption of a suitable stress
ratio. Thus a section of the literature survey has been dedicated to the stress ratio
as recommended by other researchers.

Chapter three gives an overview of the wall pressures as measured by researchers


world wide in either model silos or full scale silos. This has been categorised into
the static, and dynamic pressures for the cylindrical and hopper section of the silo.
There is not much data available for the measurement of the stress ratio in silos
and consequently the literature survey covering stress ratios is relatively short.

Following the literature survey is a description of the steel model silo set up in the
structures laboratory. To enable the measurement of the pressures during material
flow, three novel types of floating pressure cells were developed. These pressure
cells were inexpensive, easy to construct, easy to calibrate and were found to be
very responsive to the instantaneous pressures found in the flowing material. The
results of the test are discussed in section three of this chapter, while a full set of
the data has been given in Appendix D in the form of graphs. Throughout this
1.4 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

thesis, reference is made to the appendices for further detailed results.

Chapter four is dedicated to the anti-dynamic tube. The first section consists of the
limited research work by others that has been done on tubes. In many cases the
reports are merely descriptive, rather than relating to the actual frictional drag on
the tube. There has been much verbal discussion with Professor Roberts, from the
Centre for Bulk Solids and Particulate Technologies in Newcastle, Australia, about
the difficulty of determining the friction force acting on the tube during material
flow. In one such a discussion, recount was given of a tube installed in a full scale
silo which started punching through the base of a flat bottomed silo when the
discharge gate was opened. It was mentioned that several attempts were made to
support the tube from the walls and the roof of the silo to prevent excessive loading
on the silo bottom from the tube. This discussion gives a good appreciation of the
extent of the drag forces on the tube during material flow.

The second section of chapter four gives an in-depth description of the tube
experiments undertaken for this PhD research, and the method of measuring the
force on the tube during material flow. The experimental results of the drag force
measurements with a full set of graphs have been given in Appendix E. From the
literature survey on the pressures in the material, a mathematical model of the
pressures exerted on the anti-dynamic tube has been presented in the last section
of chapter four.

Since the tube is a shell structure, chapter five of this thesis has been dedicated to
shell theory, in particular, cylindrical thin shells. The topic of a thin shell is
appropriate to the anti-dynamic tube, as the wall thickness of the tube would need
to be a optimised for financial reasons. Furthermore, during material flow there is
wear on the walls of the tube which results in thinning of the walls over time. There
is no theory for the structural stability and strength capacity of a thin shell with
multiple perforations around it’s circumference. Consequently the classic theories of
shells subject to axial, lateral and a combination of both pressures have been
studied.

There has been limited work from previous researchers who have conducted tests
on thin cylindrical shells with either one or two cut outs, placed at the mid height of
the shell. The shells considered had a varying ratio of radius to wall thickness as
well as the diameter of the cut out in the shell wall. These shells were subjected to
an axial load only and the results from this work has been presented in the
INTRODUCTION 1.5

literature survey of chapter six. Included in this survey is the work from an honours
thesis, L Scutella, University of Western Australia, 1998, which described tests on
thin shells with multiple perforations subjected to an axial load only. In Scutella’s
work, four different percentages of open area were considered. The open area is
defined as the ratio of the area of the cut outs to the surface area of the shell.

Scutella’s research presented a good basis for comparison with the experiments
conducted in this PhD research on similar shells subjected to both axial and
external lateral pressures. However only two percentages of open area were
considered, 16.5% (a cut out radius of 51mm) and 36.6% (a cut out radius of
76mm). These tests consisted of subjecting the shells to a combination of a varying
axial load and a constant external lateral pressure.

The final section of chapter six describes the large deflection, non-linear finite
element analyses undertaken on a solid shell and shells with the same open areas
as the laboratory experiments. In this thesis the term “solid shell” has been
employed to describe a thin walled cylindrical shell with no holes in the wall, hence
solid. As there is no shell which has perfectly curved walls, the finite element
analyses included initial geometric imperfections which were imposed on the walls.
As there are an infinite variety of geometric imperfections which could be imposed
on the walls in such analyses, it was considered reasonable to use the mode shapes
from elastic eigenvalue buckling analyses. However, there is no way of predicting
which mode shape as an imposed wall imperfection will result in the lowest failure
load in the large deflection, non-linear buckling analyses. Consequently, the first
thirty mode shapes from the eigenvalue buckling analyses were expanded and
imposed as imperfections in an axial load analysis and an external lateral pressure
analysis for each shell. From these analyses the final mode shapes were chosen as
the required wall imperfection in the non-linear buckling analyses of the shells
subjected to simultaneous axial loads and lateral pressures. The degree of wall
imperfection imposed on each shell was varied till the finite element results
reasonably matched the results from the shell experiments in the laboratory. Finally
from these analyses, interaction curves have been presented as a design tool for
shells with multiple perforations of varying open area ratios.

Finite element analysis is a topic undergoing much research on a continuous basis.


It is therefore necessary to stress that in this PhD research, finite element analysis
has been used as a means to an end, similar to the use of pressure cells, strain
gauges and other equipment in the laboratory tests. Consequently no attempt has
1.6 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

been made to give an in depth theory of finite element analysis.

The final chapter in this thesis presents the overall conclusion of this PhD research
including the work on silo wall pressures, the resulting frictional drag on an anti-
dynamic tube during material flow and the recommended interaction curves for thin
cylindrical shells with multiple perforations.

1.2 SILO FLOW PATTERNS

Silos are defined according to the resulting flow pattern within the silo. A W Jenike
(1964) defined types of flow pattern, mass flow and funnel flow, as shown in figure
1.3a and b respectively. Figure 1.3c shows a combination of the two types of silos.

Funnel
flow
cylinder

Dcr
Mass flow
hopper

(a) (b) (c)

Figure 1.3: Basic silo types: a) Mass flow ; b) Funnel flow ; c) Expanded flow

1.2.1 MASS FLOW SILOS


Mass flow silos are characterised by steep hopper half angles and smooth wall
surfaces enabling all the material to flow when the discharge gate is opened. This is
most desirable for materials which degrade or consolidate with time, as mass flow
guarantees complete discharge of the material. If the material becomes segregated
during filling, re-mixing during discharge can be ensured. Mass flow silos are
classified according to the hopper shape; axi-symmetric silos have conical hoppers
and plane flow silos have wedge shaped hoppers with long slotted openings or a
group of discharge gates.

1.2.2 FUNNEL FLOW SILOS


In funnel flow silos the material forms a funnel within itself above the hopper
outlet, causing a last-in-first-out situation. The silo does not completely empty
when flow has stopped and an area of dead material remains inside the silo. This is
INTRODUCTION 1.7

undesirable for most materials, but has the advantage of protecting the walls
against excessive wear. Funnel flow silos have shallow half angles or are flat
bottomed and cause segregation problems and erratic discharge.

1.2.3 EXPANDED FLOW SILOS


A third type of silo is the expanded flow bin which combines the two types of flow.
The critical pipe diameter Dcr in the funnel flow cylindrical section determines the
minimum dimension for the mass flow hopper below.

1.3 SILO INSERTS


Since the anti-dynamic tube is a silo insert a brief discussion of types of inserts has
been given.
Most silo inserts are discharge devices used as a solution for a poorly flowing
hopper. Ideally, the silo should first be designed with a gravity flow hopper that
would discharge the material satisfactorily, as gravity flow is the cheapest and most
reliable method of discharge. If a gravity flow hopper cannot be installed then an
appropriate discharge aid should be considered. Reed and Duffell (1983) have
categorised discharge aids into three types:
1) Pneumatic: those which rely on the application of air to the material to
initiate flow
2) Vibrational: those which rely on the application of high frequency low
amplitude vibrations to the hopper wall
3) Mechanical: those which rely on mechanical means to discharge/extract the
material from the hopper.

1.3.1 PNEUMATIC DEVICES


These devices rely on controlled quantities of air at low pressure being applied to
the material thereby reducing its strength and improving the flow characteristics.
By introducing the air at the wall of the hopper, the wall friction is reduced and the
material in the region of the wall becomes “liquid”. Reed and Duffell (1983) state
that this method works best with dry (or very low moisture content) materials less
than 300microns in size.

1.3.2 VIBRATIONAL DEVICES


Vibrational devices rely on the ability of the material to transmit vibrations thereby
reducing the strength of the bulk solid. This method should not be operated when
1.8 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

the discharge gate is closed as the vibrations cause densification of the material
and a higher strength material than initially loaded results. Their effectiveness in
handling sticky, flaky and fibrous materials is doubtful as they rely on the ability of
the material to transmit the vibrations.

1.3.3 MECHANICAL DEVICES


The simplest mechanical device consists of chains suspended in the silo. If the
material arches, an upward pull of the chain breaks the arch and the material starts
flowing. Alternatively, paddles within the material which rotate about a horizontal
or vertical axis maintain the material in a state of continuous motion. For hard
coarse materials, wear on the chain and paddles can be significant. Screw feeders
are a commonly used mechanical method of discharging and controlling flow rate of
the bulk solid, but are also subject to high wear rates.

1.3.3.1 Binsert
The binsert approximates the concept of the anti-dynamic tube except that it has
sloping sides, not parallel like the anti-dynamic tube.

The binsert is a patented device invented by JR


Johanson (1982) consisting of a cone-in-cone insert
as shown in figure 1.4. The smaller hopper inside the
silo hopper must be designed in accordance with
mass flow principles. He states that the location of
the insert relative to the outer hopper is such that the
included angle, satisfies mass flow criteria.
Figure 1.4: Binsert Johanson states that the outer hopper half angle,
must be twice the angle required for mass flow.

1.3.3.2 Anti-dynamic tube

The Reimberts (1976) claim to be the inventors of anti-dynamic tube in France in


the 1950’s. Figure 1.5c is a photograph taken from the Reimbert’s book (1976)
illustrating the use of their tubes. The main purpose of the tube was to alleviate the
pressures on the walls during flow of the material. The Reimbert tube is a small
diameter tube with multiple perforations along it’s length, which is placed centrally
in the silo, as shown in figure 1.5a. This design caused the silo to operate on a last-
in-first-out basis. A modification to the Reimbert tube is reported in Ooms and
Roberts (1985) and was installed in full scale silos in Port Adelaide. The
INTRODUCTION 1.9

modification consisted of reducing the overall height of the tube and omitting all the
holes such that the silo empties in two stages only, as shown in figure 1.5b.

Material
Flow

(a) (b)

Figure 1.5: a) Reimbert tube; b) Roberts tube

Figure 1.5c: Photograph taken from Reimbert(1976) showing the


use of anti-dynamic tubes.
1.10 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

1.4 INTRODUCTION TO WALL PRESSURES

The two pressures which are of interest to researchers are the vertical pressures
acting on the silo bottom, and the horizontal (or normal) pressures acting on the
silo walls. Depending on which type of pressure exists in the silo, the vertical
pressure is either larger than, or smaller than, the horizontal pressure.
The pressures exerted by the stored material in the silo can be divided into three
main types. The first type of pressures are those which develop during loading of
the solid into the empty silo, and are generally referred to as the static pressures.
Since the silo consists of two main sections, ie the cylindrical and the hopper
sections, the static pressure distribution in each section is different. When the
material is loaded into the silo, an active pressure field develops and the lines of
major principal stress are nearly vertical. The associated minimum stress ratio
possible is the active stress ratio, Ka which is less than one. The static vertical

pressures are greater than the pressures acting normal to the silo walls. The static
normal pressures are the minimum pressures which can be expected in a silo.

The second type of pressures are those which develop during discharge of the
material from the silo, and are generally referred to as the dynamic pressures.
Again, the distribution of the dynamic pressures in the cylindrical section are
different to those found in the hopper. When the discharge gate is opened the
relative motion of the solid with respect to the wall is the same as the wall moving
in towards the solids, and hence the pressure field changes from an active to a
passive state. The stress ratio now has a value greater than one, and the lines of
major principal stress are approximately horizontal. The dynamic pressures acting
normal to the silo walls are greater than the dynamic vertical pressures, as well as
the static normal pressures, but are not necessarily the largest that can be
expected.

Based on the above discussion, a passive stress state cannot theoretically develop
in the cylindrical section of the silo, since the walls are parallel and there is no
relative motion in towards the material. However, it is doubtful whether a perfectly
parallel wall can be achieved in practice, and consequently many researchers allow
for the possibility of a small convergence in the walls of the cylindrical section. Thus
in the theories which have been presented in this chapter, a passive stress state
has been assumed in the cylindrical section, to determine the possible distribution
of flow pressures.
INTRODUCTION 1.11

The third pressure on the walls of the silo also occurs during discharge of the
material. This is the switch pressure, which is a transient pressure exerted over a
small area of the walls for a small period of time. The switch pressure is caused by
the change-over from a static to a dynamic stress state and travels quickly up the
hopper the instant the gate is opened. The switch pressure may extend the full
height of the silo, or it may become trapped at the transition from the cylindrical to
hopper section. In both the hopper and the cylindrical section, the switch pressure
is the largest expected pressure acting on the walls of the silo. Most researchers ,
and silo design codes, give the switch pressure as a multiple of the static pressure,
the actual ratio being dependant on the theory developed.

The direction of the major principal stresses have been shown by the lines in figure
1.6 for the three types of pressures found in the silo. Figure 1.6a shows the near
vertical lines of the major principal stress during filling of the silo. Figure 1.6b
shows the lines of the major principal stress as near horizontal during emptying of
the silo, with the assumption that the passive stress state develops in the cylinder.
Figure 1.6c shows the location of the switch at an instant in time when the switch
has travelled up in to the cylinder and the stress changes from an active to a
passive state.

(a) (b) (c)

Figure 1.6: Schematic representation of the lines of major principal stress.


CLASSIC WALL PRESSURE THEORIES: LITERATURE SURVEY STATIC PRESSURES 2.1

CHAPTER 2

CLASSIC WALL PRESSURE THEORIES

2.1 STATIC WALL PRESSURES

2.1.1 CYLINDRICAL SECTION

2.1.1.1 The Janssen Theory


Prior to Janssen’s experiments and his paper of 1895, it was assumed that the
material in a silo exerted a triangular hydrostatic pressure distribution down the
height of the wall. Now the Janssen formula is used in engineering standards world
wide for computing the initial filling pressures in the vertical section of the silo.

Roberts(1995) reports on Janssen’s experiments which were performed on wooden


model silos filled with wheat. The silos being of square cross-section with sides
measuring approximately 200, 300, 400 and 600mm. The models were mounted on
adjustable screws, while the bottoms were formed by close fitting movable boards
connected to a weigh bridge. In this way the pressure on the bottom of the silo
could be measured.

The Janssen theory was developed by considering the equilibrium of the vertical
forces acting on a horizontal elemental slice in the cylindrical section of the silo, as
shown in figure 2.1.
s = length of the side
s
A = silo cross sectional area
P = silo perimeter
z = depth from the top

z dz = element thickness
v v = average vertical pressure

h d v /dz = change in vertical pressure


dz
through the element thickness.
= shearing stress between the material

v+ d v dz and the silo wall


dz h = horizontal pressure

= material density
Figure 2.1: Forces acting on = wall friction angle
a horizontal elemental slice
2.2 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Equilibrium of forces on the elemental slice is given by :


d v
v dz A P dz v A A dz (2.1)
dz

Simplifying the above equation gives :


d v P
(2.2)
dz A

In equation (2.2) P/A is the perimeter divided by the cross-sectional area and for a

circular silo P/A = D/( D2/4) = 4/D (2.3)

The shearing stress between the material and the wall is given by :

h Tan h (2.4)

Janssen assumed a constant stress ratio of K= h / v =Ka , through the depth of


the silo.

Substituting equations (2.3) and (2.4), and substituting for h = K/ v into


equation (2.2) gives the following expression for the vertical pressure:

z
4 h D
v dz 1 e z 4K /D (2.5)
D 4K
0

The derivation of Janssen’s equation was based on the following assumptions:


1) the vertical pressure is constant across a horizontal slice

2) the stress ratio K = h / v is assumed constant at all depths in the silo

3) the bulk density, does not vary with depth


4) the wall friction is fully mobilised and the material is on the point of slip

Roberts(1995) reports that Janssen showed from his experimental results that the
vertical pressure distribution across the cross-section was not uniform. He showed
that the pressure was higher in the centre, at 1.15 times the average value, and
lower in the corners of his models, at 0.8 times the average value. He concluded
that the wall pressures could be estimated with sufficient accuracy by assuming a
constant pressure distribution across a horizontal slice.

In this thesis the horizontal pressure on the silo wall is of interest, which is given by
D z 4K /D
h
K V 1 e (2.6)
4
CLASSIC WALL PRESSURE THEORIES: LITERATURE SURVEY STATIC PRESSURES 2.3

The graph in figure 2.2 shows the distribution of the horizontal pressure as a
function of the height to diameter ratio of the silo. The values used in the graph

are: =16.8kN/m 3 , D=0.96m, =Tan22 0.404. The stress ratio used for the

purposes of this graph is the “at rest” ratio given by: K=0.29=1-Sin45 material

friction angle of 45 ).

From equation 2.6, as the depth z, tends to infinity, the exponential term tends to
D
zero, and the pressure tends towards a maximum given by the asymptote: .
4

Therefore, in a very tall silo, there is no increase in the vertical pressure at the silo
bottom when more material is loaded on top. The additional material weight is
carried by the walls of the silo.

This asymptote is inversely


Horizontal Pressure (kPa)
0 3 6 9 proportional to the wall friction
0.0 angle and therefore as the wall
friction angle increases the
0.4
horizontal pressure acting on the

0.8
wall decreases, as shown by the

curve labelled 2 in figure 2.2.


1.2
H/D Ratio

The horizontal pressure is directly


related to the stress ratio since
1.6
2K the horizontal pressure is
2.0 calculated from the vertical
2
pressure by multiplying by the
2.4
stress ratio. As can be seen in
figure 2.2, increasing the stress
2.8
ratio by a factor of two has the
3.2 greatest influence in the upper

Figure 2.2: Graph of Janssen regions of the silo.


horizontal pressure distribution
2.4 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

2.1.1.2 The Reimbert Theory

Marcel and Andre Reimbert (1976) developed a theory for the static pressures in
the cylindrical section of the silo. They approached the problem by taking a
horizontal cut through the silo and considering the vertical forces acting on the free
body diagram of the upper portion, as shown in figure 2.3a. The depth of the silo,
z, is measured from the top of the silo, while the cone of material on top has a
height of, h. The Reimberts state that if there were no friction on the walls of the
silo, the graph of vertical pressure would be a straight line as shown in figure 2.3b.

The cone of material above the silo is shown by the offset, h/3 in figure2.3b.

h/3 Stress
h

z
w

Depth
(z)
z + h /3
v
Figure 2.3a: Free
Body Diagram Figure 2.3b: Graph
of vertical stress

However, since there is friction on the walls of the silo, the equation of vertical
equilibrium, can be written as:
h
vA 1 AP 2A Pz 0 (2.7)
3
where the cross-sectional area of the silo is given by A, the circumference is given
by P, and the volume of the cone of material above the silo is given by Ah/3. In

equation 2.7, v is the vertical reaction from the material below acting on the upper

portion, is the shear stress acting on the sides of the walls and is the bulk

density of the material.

From their experimental results, the Reimberts state that the shape of the curve of
the shear stress on the walls of the silo is as shown schematically in figure 2.4.
CLASSIC WALL PRESSURE THEORIES: LITERATURE SURVEY STATIC PRESSURES 2.5

Stress

Line of hydraulic
stress in the silo
for zero friction

(z) v max
Depth
(z) z+ h/3

Figure 2.4: Graph of shear


stress

The Reimberts state that with increasing depth the curve of the shear stress
approaches an asymptote, which is parallel to the line of the hydraulic stress, and
has been shown dotted in figure 2.4. Therefore, as the depth of the silo increases
the equation of vertical equilibrium becomes:
h P
v max 1P 2 z 0 (2.8)
3 A

The shear stress has the following expression, which was derived from their
experiments:
2
z
(z) (2.9)
z Cc

The Reimberts define the characteristic constant, Cc, as follows:

v max h/3
Cc (2.10)

Now the unknown term, vmax , in the expression for the friction function is
contained in the expression for the characteristic constant. As the silo is being
filled, v reaches a maximum limit and any additional elemental slice of material of
thickness dz, loaded into the silo is carried by friction on the walls. This can be
expressed as:

P dz = We = A dz (2.11)

But the friction term can also be expressed as : = hmax (2.12)


Equating equation 2.11 and 2.12 gives:

hmax = ( A)/( P) = ( / ) (2.13)

This expression is a constant since the vertical pressure in the silo becomes
constant as the depth increases and therefore the horizontal stress also becomes
constant.
2.6 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

The maximum vertical stress can be found from the maximum horizontal stress by
using an appropriate stress ratio. The stress ratio they propose is the active stress

ratio. Therefore, substituting for vmax into the expression for the characteristic

constant gives:

Cc = ( /( Ka) - h/3) (2.14)

And hence, the friction function is fully described as:

= ( z2) / (z+ /( Ka)-h/3) (2.15)

The Reimbert theory then considers the general expression for friction as given by
equation 2.12, and state that since the horizontal stress varies with depth, the
expression for the friction can be written as:

(z) = h dz (2.16)

Therefore, the derivative of the friction function as given in equation 2.15 would
give an expression for the horizontal pressure. The derivative of equation 2.15 is:

d z C c (2 z) ( z2 ) Cc2
1 (2.17)
dz z Cc 2 z Cc 2

Therefore, the final expression for the horizontal stress can be written as:

Cc2
h 1 (2.18)
z Cc 2

From figure 2.4, the total weight of the material in the silo, z+ h/3, equals the

shear stress on the walls of the silo plus maximum vertical stress acting on the
bottom. This is expressed as:

vmax + = z+ h/3 (2.19)

The final expression for the vertical stress on the bottom of the silo, can be found
since the expression for the friction function is given in equation 2.15. Therefore,
the expression for the vertical stress is:

1
z h
v z 1 (2.20)
Cc 3
CLASSIC WALL PRESSURE THEORIES: LITERATURE SURVEY STATIC PRESSURES 2.7

2.1.1.3 Discussion of Janssen vs Reimbert Theory

Briassoulis(1991) states that both the Janssen and Reimbert’s theories are
“unconditionally applicable for any silo geometry and stored material”.

For the silo model and material used in the laboratory, a comparison between the
two theories is shown in figure (2.5). The Janssen theory gives a greater vertical
stress as the depth increases, while the Reimbert theory gives a greater horizontal
stress in the upper parts of the silo. Both theories tend to the same value of
horizontal pressure as the depth of the silo increases.

Vertic al Pressure Horizontal Pressure


0 5 10 15 20 0 2 4 6 8
0 0

0.4 0.4
Reimbert
Janssen
0.8 0.8
Depth below surface

1.2
Depth below surface

1.2

1.6 1.6
Janssen
2 2
Reimbert
2.4 2.4

2.8 2.8

3.2 3.2

Figure 2.5: Comparison between Janssen and Reimbert theories

Re-writing Janssen’s expression for the horizontal pressure, equation (2.1.6) as:
4K z
D 1 D
h 1 e (2.21)
4 K

Noting that for a cylindrical silo A/P = D/4; Reimbert’s expression for the horizontal
pressure, equation (2.20) can be re-written as:
2
D z
h 1 1 (2.22)
4 Cc

Therefore, only the terms in square brackets need be considered when comparing
equation (2.21) and (2.22).

For a silo with no cone of material on top, the expression for the characteristic

abscissa in Reimbert’s theory can be re-written as Cc = D / (Ka 4 ) and substituting


2.8 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

this in equation (2.22) gives the term in square brackets as: 1- ( 4 K z / D + 1) -2

The graphs in figure 6 show the influence of the wall friction angle and the stress
ratio, K, on the terms in square brackets for both equations of horizontal pressure.
The Janssen term has been shown with a blue line and the Reimbert term in a red
line. The graphs show that as the wall friction angle increases, the Janssen and
Reimbert expressions tend to coincide, while a smaller stress ratio causes them to
diverge.
Horizontal pressure (kPa) Horizontal pressure (kPa)
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.4 0.8 1.2 1.6
0 0
Janssen
Depth below surface (m) Reimbert
Depth below surface (m)

1 1
10o
50o

2 2 1

3 3
Janssen
Reimbert 0.5
4 4

5 5
(a) Wall friction angle varies (b): Stress ratio varies
from 10 to 50 degrees from 0.5 to 1

Figure 2.6: Comparison of terms in Janssen and Reimbert equations: (a):


Wall Friction ; (b) Stress ratio

2.1.2 HOPPER SECTION


The theory for the wall pressure in a convergent hopper was presented by Walker in
1960. The theory is included in this section for completeness only, since the results
appear to be at a variance with physical intuition and with the results of Janssen
theory in the cylindrical section

2.1.2.1 Walker Theory

Walker (1960) considered a smooth walled hopper with a hopper half angle ( ,
filled with a nearly incompressible material. His initial premise was that the
principal stress planes in the hopper were vertical and horizontal, so that the shear
stress on a vertical plane is zero. Under the assumption of zero shear, the normal
stresses must increase hydrostatically with depth.
CLASSIC WALL PRESSURE THEORIES: LITERATURE SURVEY STATIC PRESSURES 2.9

Therefore: V gz zh (2.23)

Vertical pressure (kPa)


0 10 20 30 40
0.0
Depth below top of silo (m)

0.5 Therefore, plotting a graph of the


Cylindrical section

Janssen static pressure in the silo using


1.0
the Janssen theory for the

1.5 cylindrical section, and the Walker


theory for the hopper section,
2.0
gives the diagram shown in Figure

2.5 2.7 for the model silo in the


Hopper

Walker laboratory. This graph illustrates


3.0 the vertical pressure distribution
in the silo.
Figure 2.7: Graph of static vertical
pressure in a model silo

If the material in the hopper lies within the yield surface, and only the wall is
assumed to slip, Walker gives the Mohr circle for an element on the wall as shown
figure 2.8. Point P on the circle represents the stresses at the wall which is
tangential to the wall yield locus. The wall friction angle is given by w and the

material friction angle is given by m. Walker assumes the horizontal stress to be

equal to the minor principal stress, 3


, on the circle.

material effective
yield locus hopper wall
yield locus

P
m

W 2
0 3 N C 1

Figure 2.8: Mohr’s circle for stresses at the hopper


2.10 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

The angle which the normal stress, N


, acting on the hopper wall makes with the

horizontal stress, is given by 2 in Figure 2.8. The value of N


is obtained from the

geometry of the Mohr circle.

Sin 2 Cos w
N 1 C 1 (2.24)
Sin w
2 Sin w

From equation 2.24, it can be seen that Walker gives the normal stress at the wall
as a function of the hopper half angle and the wall friction angle. The material
friction angle does not influence the value of N. Figure 2.9 shows the variation of

the ratio given by C in equation 2.24 as a function of the wall friction angle. The
four curves shown in figure 2.9 are for a hopper angle of 5 increasing to 20 in 5
increments.

0.8

0.7

0.6

0.5

0.4

0.3
Constant C

0.2
=20
0.1
=5
0
0 10 20 30 40 50 60 70
Wall friction angle ( w

Figure 2.9: Variation of C as a function of the wall friction angle

As can be seen, the shallower the hopper slope, the higher will be the normal wall
stress according for the same wall friction. The lowest normal wall stress will be
achieved using a shallow hopper half angle, ie 5 , and a higher wall friction angle.
However too high a value for the wall friction angle will inhibit mass flow and result
in funnel flow.
CLASSIC WALL PRESSURE THEORIES: STATIC PRESSURES: HOPPER 2.11

2.1.2.2 Jenike Theory


Jenike has written several papers on the pressures exerted on the hopper walls as
well as the pressures within the solid stored in the hopper. The theories developed
were covered in several papers. To fully understand Jenike’s derivations, a brief
summary, in chronological order, describing the outline of each theory is given
below. A similar outline has been given in chapter 2.2 for Jenike’s theories of
pressures due to the flowing material in the silo.

1) Radial pressure field in the solid stored in a hopper.


In this theory Jenike (1961) considers an element of material, at an arbitrary
location in the stored solid, and evaluates the forces acting on this element by
considering equilibrium of all forces. The equations which result after extensive
algebraic manipulation, are lengthy differential equations which are solved using
numerical methods. Jenike shows from the results of these equations that the
pressures in the solid increase linearly from zero at the vertex of the hopper. Jenike
defines this as a radial stress field.

2) Linear normal pressure exerted on the hopper wall .


In this analysis, Jenike (1968) states that the weight of the material in a hopper
without a surcharge, is carried by the vertical components of the shear and normal
wall stresses. By simple equilibrium of forces, Jenike derives an expression for the
pressure normal to the hopper wall. This analysis was extended by Jenike (1973) to
include normal pressures acting on the walls of hoppers with a surcharge.

3) Position of the maximum pressure in the hopper.


Having found the expressions for the normal wall pressure and the radial pressure
field, Jenike(1968) equates the two expressions to determine the location, iN, of the

maximum normal wall pressure in the hopper.

Jenike’s theory for the linear normal pressures on the hopper walls has been dealt
with first in this thesis, followed by the radial pressure field and the location of the
maximum static pressure.

2.1.2.2.1) Linear normal pressure exerted on the hopper wall

Jenike(1968) states that the pressures in a hopper have been shown


experimentally to decrease to zero at the vertex of the hopper. Therefore, Jenike
assumes a linear pressure distribution acting normal to the wall during filling of a
2.12 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

hopper with no surcharge as shown in figure 2.10(b) below. This pressure field
cannot extend all the way to the free surface in a hopper without a surcharge, but
will be topped off by a compatible pressure field, with pressures decreasing towards
zero at the free surface. The maximum pressure in the hopper occurs at the
interface of the two pressure fields. The distance of the maximum pressure from
the vertex has been defined by Jenike using the ratio iN, as shown in figure 2.10(c).

r
i=1
Hh
N max

zN
N
Depth

z iN=zN/Hh

Pressure
(a) (b) (c)

Figure 2.10: (a) Hopper; (b) Pressure distribution; (c) Ratio i

Jenike states that the total weight of the material must be supported by the walls
and ignores any support gained from the material below. Therefore, the sum of the

vertical components of the wall shear and normal pressure, and N, acting on an

element of thickness dz, equals the weight of the element in the hopper.

The vertical components of the wall support, acting over an area of 2 r dz / Cos .
are given by:

( N Sin N Tan w Cos ) 2 r dz / Cos (2.25)

The radius of the hopper is given by: r = z Tan (2.26)

The volume of the element is approximately given by:


2 2
V= z T an dz (2.27)

Therefore, the weight of the element is given by:

W Tan2 z 2 dz (2.28)

By equating equation 2.25 and 2.28, re-arranging and integrating over the hopper
wall from 0 to Hh gives the following, (where Hh is the height of the hopper):
CLASSIC WALL PRESSURE THEORIES: STATIC PRESSURES: HOPPER 2.13

Hh
Tan Hh3 6 ( Tan Tan w) nz dz (2.29)
0

In Figure 2.10(c), for 0< z <z N : N


=z N max
/z N

And for z N < z <Hh : N


= (Hh-zN) N max / (Hh-z )

Substituting for N
in equation 2.29, and integrating gives the following expression:

3
zN 1 Hh3 2
Hh z N Hh3 3
zN
Tan Hh3 6 ( Tan Tan w) N (2.30)
3 zN Hh zN 2 2 3 3

Solving for N gives the expression for the normal pressure on the hopper wall as:

N
= D/ [ 2(Tan + Tan ) ( 1+z N / Hh)]
w
(2.31)

1 Figure 2.11 shows the


zN/Hh = 0.25
Ratio zN/Hh

zN/Hh = 0.5 maximum normal wall


zN/Hh = 0.75 pressure in the hopper as
0.8
Locus of Maxima given by equation 2.31,
for various values of zN/Hh
0.6
The graph was drawn for
3
values of =18 kN/m ,
0.4
hopper diameter D=1m,
hopper half angle =15 ,
0.2
and a wall friction angle

w = 22 . From equation
0
2.31, the material bulk
0 2 4 6 8 10 12 14
Normal wall pressure (kPa) density and diameter of
the hopper have the
Figure 2.11: Normal Pressures on the Hopper Wall
for various values of the Ratio iN greatest effect on the wall
pressure. As can be seen,
from figure 2.11 the maximum pressure decreases, as the position of the maximum
increases within the hopper. The locus of the maxima has been shown in figure
2.11. Since the maximum cannot be at the free surface, the locus has been
2.14 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

arbitrarily terminated at a value of iN=0.875

The shape of the locus can be closely approximated by a cubic polynomial of the
3 2
general form: iN = A N
+ B N
+C N
+D. The constants in the equation are different

for each hopper being analysed.

Having derived an expression for the normal wall pressure in a hopper with no
surcharge, Jenike(1973) adopts the same approach in deriving an expression for a
hopper with a surcharge. This surcharge, , calculated from the Janssen equation
J

at the level of the transition, increases the pressure at the top of the hopper, from
zero to some value t
at the transition. Jenike states that this surcharge is

supported by the walls above the location of the maximum normal pressure, as
shown by the shaded area in Figure 2.12 (b).

J
t

Additional pressure
r p1 p2 due to material in the
Hh
cylindrical section
N max

N
zN Pressure distribution
z in hopper with no
surcharge

Pressure
(a) (b)

Figure 2.12: (a) Hopper ; (b) Pressure distribution with surcharge

By equilibrium, the additional pressure is given by:


Hh
2 Tan Tan w
J
p2 p 1 z dz (2.32)
Tan
zN

where p1 is the pressure at some level, z, for no surcharge and p2 is the additional

pressure at the same level, z, for a surcharge acting on the hopper

and (p2 - p1) = t (z – z N) / (Hh-zN) (2.33)

The ratio i is defined by Jenike as: i = z /Hh

Substituting equation 2.33 into equation 2.32, and substituting for iN=zN/Hh gives:
CLASSIC WALL PRESSURE THEORIES: STATIC PRESSURES: HOPPER 2.15

Hh
2 Tan Tan w t
J
(i iN ) i di (2.34)
Tan 1 iN
zN

Integrating and solving for t gives:

3 Tan
t J (2.35)
1 iN 2 iN Tan Tan w

Figure 2.13 shows the normal wall pressure acting on a hopper with a surcharge.
The Janssen pressure, J
, for the hopper surcharge was calculated for a cylinder

height of six times the hopper diameter, since the vertical pressure in a cylinder
does not increase substantially beyond this height to diameter ratio. The graph has
3
been drawn for values of =18 kN/m , a hopper diameter D=1m, a hopper half
angle =15 , and a wall friction angle w = 22 . As the position of the maximum

pressure on the hopper walls increases, so the pressure at the transition also
increases.

1
zN/Hh=0
zN/Hh=0.25
0.8 zN/Hh=0.5
Ratio iN

zN/Hh=0.75

0.6

0.4

0.2

0
0 20 40 60 80 100
Normal wall pressure (kPa)

Figure 2.13: Normal Pressures on the Hopper Wall for


various values of Ratio iN for a hopper with a surcharge

The dotted lines in figure 2.13 indicate the lines of pressure for no surcharge in the
hopper, as given in figure 2.11.
2.16 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

2.1.2.2.2) Radial pressure field in the solid stored in a hopper.

Jenike [1961] defines a radial stress field as a field in which all the stresses
increase linearly with the radius, r, for the initial loading of material in the hopper.
Youngs modulus for the material is eliminated during the process of derivation and
Poisson’s ratio is assumed constant. The material is assumed to be non-linearly
elastic and is assumed to slip at the walls. In his derivation Jenike considers an
element in the hopper with a set of spherical co-ordinates (r, , ) with origin at the
vertex of the hopper. Jenike does not give a complete diagram of the element in his
papers. The element shown in figure 2.14 has been drawn to fully understand the
shape of the element and the directions of the forces acting on it. The areas of each
face of the element have been shown in figure 2.15, with the expressions for the
length of each side and the area of each face.

Considering equilibrium in the r-direction (refer figure 2.14): +

r d
r dr A bott 2r dr d d Sin r A bott r A right Cos ………………...
r 2

r d d
+ r d A right r dr d d Cos Cos A right …………..…….
2 2

d d
d A right r dr d d Cos 2 c Sin r dr d …..…….
2 2

d
+ Cos r 2 dr d d Sin 0 (2.36)
2

Divide throughout by:( r2 dr d d Sin )

1 1 r 1 1 d
2 r r
r Cot c Cos 0 (2.37)
r r r r r r 2

Re-arranging gives the following final equation for equilibrium in the r-direction:

r 1 r 1 d
2 r c r Cot Cos 0 (2.38)
r r r 2

This expression differs from Jenike’s expression by the term Cos( +d /2) instead of
Cos( ). By small angle approximations, this difference in the equations has little
effect on the final solution and the term d /2 has been ignored.

The expression for equilibrium in the direction is given after figures 2.14 and 2.15
CLASSIC WALL PRESSURE THEORIES: STATIC PRESSURES: HOPPER 2.17

d /2
c
Plan View Section B-B

r
r dr
r dr r
r
r r
d /2

r
r d

r
d
d

Section A-A

A A

Element in the
stored solid
B

Plan on hopper
Figure 2.14: Element in the stored
material in a conical hopper
2.18 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

The position of the element


within the hopper is defined as
Area back follows:
r = radius from hopper vertex
dr = small change in r
= angle of rotation of r

S4 d = small change in
S3 S6 S5
d
d = small change in the
horizontal angle

The lengths of the element


sides are defined as follows:
Area front
S1 = r d
S2 = (r+dr) d
S3 = r Sin( +d ) d

Area top S4 = (r+dr) Sin( +d ) d


(r+dr) Sin S5 = r Sin d
S2 Area right S6 = (r+dr) Sin d

The areas of the element faces


(r+dr) Sin( +d )
can be estimated as follows:
Area bottom = S1 (S3+S5)/2
r Sin
S1 =r2 d d Sin
dr
r Sin( +d ) Area top = S2 (S4+S3)/2
Area left r2 d d Sin +

d 2r dr d d Sin

= Abott + 2r dr d d Sin
r
Area right = dr (S5+S6)/2
Area bottom
r dr d Sin
Area left = dr (S3+S4)/2
r dr d Sin +

Figure 2.15: Element in the Stored r dr d d Cos


Material in a Conical Hopper =Aright + r dr d d Cos
Area front = Area back
= dr (S1+S2)/2 r dr d

The volume of the element is given by: V = dr (Atop + Abott)/2 r2 dr d d Sin


CLASSIC WALL PRESSURE THEORIES: STATIC PRESSURES: HOPPER 2.19

Considering equilibrium in the -direction : +

d d
Cos A right d Cos A right r dr d d Cos …………………….….…
2 2

d r d
- r A right r d A right r dr d d Cos ………………………………..
2 2

d
+ Sin r 2 dr d d Sin r A bott ………………..………………………………………………
2

r dr A d d
- r bott
2 r dr d d Sin c r dr d 2 Cos 0 (2.39)
r 2 2

Dividing throughout by:( r2 dr d d Sin ) , collecting terms and re-arranging gives :

r 1 d
3 r c Cot Sin 0 (2.40)
r r 2

This expression differs from Jenike’s expression by the term Sin( +d /2) instead of
Sin( ). By small angle approximations, this differences has little effect on the final
solution, and has therefore been ignored.

To complete the solutions of equations 2.38 and 2.40, Jenike(1961) first finds
expressions for , ,
r c
and r
. As the material is loaded in the hopper of the silo,

it contracts both vertically and horizontally. Therefore, the material does not reach
the limiting state of stress but is in an elastic-active state of pressure.

The elastic stress strain relations in the hopper are therefore given by:

1 r c
r E
1 1 2

1 r c
E (2.41 a, b, c)
1 1 2

1 c r
c E
1 1 2

The strains given in equations 2.41 a,b,c can be written as functions of the radial
displacement as follows:

r = - u/ r
(2.42 a,b)
= c = -u/r
2.20 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

From equation 2.42b it follows that = c

Jenike defines the stress ratio in the hopper as: k = / r (2.43)

This is similar to the definition used by Janssen for the stress ratio in the cylindrical
section of the silo.
In a Mohr circle for the stresses on the element, the average stress is given by:
= ( r+ )/2 (2.44)

The general equation for the radial stress is given by Jenike as: = rS (2.45)

where S is the radial stress field and is a function of the co-ordinates r and
Therefore, from equations 2.43, 2.44 and 2.45:

r =k /(1+k) = k r S /(1+k) (2.46)

= c = 2k /(1+k) = 2k r S /(1+k) (2.47)

The shear stress on the wall is given by:

r = Tan w= 2k r S Tan w /(1+k) (2.48)

The derivatives of r , and r are as follows:

r r 2 dS
S r (2.49)
r r 1 k dr

2k dS
r (2.50)
r 1 k d

r r d r w 2k dS 2k rS d w
Tan w r (2.51)
d 1 k d 2
w 1 k Cos w d

r r d r w 2k dS 2k rS d w
Tan w S r (2.52)
r dr r 1 k dr 2
w 1 k Cos w dr

Now the derivatives of r , and r given by equations 2.49, 2.50, 2.51 and 2.52

can be substituted into the equations of equilibrium given by equations 2.38 and
2.40. In his analysis, Jenike has assumed dS/dr and d w/dr to be zero. Therefore,

collecting terms in dS/d and d w/d and dividing throughout by (1+k)/(2k ) gives

the following:

dS S d w 2 1 k
Tan w S 2 Tan w Cot Cos 0 (2.53)
d Cos 2 w d k 2k

dS 1 k Sin
4STan w 0 (2.54)
d k
CLASSIC WALL PRESSURE THEORIES: STATIC PRESSURES: HOPPER 2.21

Substitution of ds/d from equation 2.54 into equation 2.53 gives the following for
d w/d :

d w 3 2k 1 k 1 k
4 Cos 2 w Sin Cot Sin w Cos w Cos Cos 2 w
d k 2kS 2kS

(2.55)

Assigning a constant value to k, the stress ratio in the hopper, equations 2.53 and
2.55 can be solved by numerical integration. The value of S, for the given value of
k, is then substituted into equation 2.51 to determine the value of the stress
normal to the wall, , which is a linear function of the hopper radius, r. From

equation 2.51, the stress field, S, must be unitless since the density, , has units of
3
kN/m and radius, r, has units of m. This results in the units of kPa for the stress,
, normal to the wall.

2.1.2.2.3) Position of the Maximum Pressure in the Hopper

The locus of maximum pressures is unique for a given hopper and stored solid, as
shown in figure 2.11. This locus has been drawn by assuming values for the ratio of
zN/Hh. Jenike states that the position, zN, of the peak pressure is determined from

the intersection of the locus of maximum pressures with the linear (radial) pressure
field shown in figure 2.10. This has been shown schematically in figure 2.16.

D
Locus of maxima
Hopper

Radial pressure field


zN

Peak
pressure

Figure 2.16: Position of Peak Pressure in a Hopper

The normal wall pressure acting on the hopper due to the radial pressure field, has
been given by equation 2.47, and is repeated as follows : = 2 k r S /(1+k).

Referring to figure 2.15, the radius, r, in equation 2.47 is measured from the vertex
of the hopper, and does not denote the radius of the hopper. Therefore, at the level
of the maximum pressure, let r=rN, which can be written in terms of iN as follows:
2.22 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

zN iN Hh
rN
Cos Cos

2 k S iN Hh
Then equation 2.47 becomes: (2.56)
Cos 1 k

The equation for the maximum pressure has been given by equation 2.31 and is

repeated as follows: N
= D / [ 2( Tan + Tan w
) ( 1+z N / Hh)]

In equation 2.31, D is the diameter of the hopper as shown in figure 2.16, and can
be re-written as :
D = 2 r = 2 HhTan

And by definition iN=zN/Hh

2 Hh Tan
Therefore, equation 2.31 becomes: N (2.57)
2 Tan Tan w 1 iN

By letting equation 2.56 equal equation 2.57, and re-arranging, an equation for the
position of the peak pressure, ratio iN, is given as follows:

1 k Sin
iN 1 iN (2.58)
2 k S Tan Tan w

where S is the radial pressure field in the solid and is determined from equations
2.54 and 2.55.

The position of the maximum pressure is directly related to the stress ratio, k, and
the hopper half angle, , and is indirectly related to the stress field, S, and the

hopper wall friction angle, w.

Re-arranging equation 2.58 in terms of w and gives the following:

1 k Sin
Tan w Tan (2.59)
iN 1 iN 2kS

A graph of equation 2.59 for k=0.8, S=0.4 and varying the hopper half angle has
been plotted in figure 2.17 below. Unlike the graph given in Jenike (1968), the
contours of iN in figure 2.17, decrease from iN=0.8 closest to the horizontal axis to

iN=0.1 closest to the vertical axis. This is the reverse of the graph given by Jenike.

Increasing the values of both the stress ratio, k, and the stress field, S, from 0.3 to
0.9 has the effect of decreasing the required wall friction angle for a given hopper
half angle. Since these graphs represent filling conditions in the hopper, the stress
CLASSIC WALL PRESSURE THEORIES: STATIC PRESSURES: HOPPER 2.23

ratio, k, cannot have a value greater than one. Increasing the stress field, S,
beyond a value of 1 gives a negative value for the wall friction angle, which is not
possible.

50

w
iN=0.1
Wall friction angle
iN=0.2
40 iN=0.4
iN=0.6
30 iN=0.8

20

10

0
0 5 10 15 20 25
Hopper half angle

Figure 2.17: Contours of iN for a conical hopper


for k=0.8 and S=0.4

2.1.2.3 Walters Static Hopper Pressures.

In his paper Walters(1972) gives the same analysis for the static vertical pressure
in the hopper as for the dynamic vertical pressure in the hopper. The full analysis of
the equilibrium on a horizontal elemental slice in the hopper section, by Walters,
has been given in chapter 2.2.1.1 of this thesis. The difference between the static
and the dynamic case, is that in the static case, the shear stress at the wall is the
minimum value. This minimum value is determined from the intersection of the wall
yield locus and the Mohr circle representing the stresses at the wall, as shown
graphically in figure 2.18.

Walters analysis of a horizontal elemental slice in a hopper results in the same


differential equation (equation 2.132) for both the static and dynamic conditions,
except that the constants E and FH are as follows:

Sin mSin 2 2 D
E (2.105 repeated)
1 Sin m Cos 2 2 D
2.24 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

2 2 2
1 Sin m
Cos 2 Sin m
Sin
FH (2.128 repeated)
2
Cos 1 Sin m
2ySin m

where the +ve sign refers to static conditions and the –ve sign refers to dynamic
conditions.

Material yield locus

Wall
yield
locus Dynamic shear value

Static shear value

Figure 2.18 General Mohr circle for stresses in the material adjacent to the wall.
CLASSIC WALL PRESSURE THEORIES: DYNAMIC PRESSURES: CYLINDRICAL SECTION 2.25

2.2 DYNAMIC WALL PRESSURES

2.2.1 CYLINDRICAL SECTION

2.2.1.1 The Walters Theory

Walters (1973a) follows the same assumption made by Jenike, that the lines of
major principal stress are approximately vertical during initial filling of the solids
into the silo, as shown in figure 2.19(a). When the discharge gate is opened and
the material starts flowing, the lines of major principal stress become nearly
horizontal, as shown in figure 2.19(b). In both cases, the angle which the major

principal stress makes with the wall is S


and D
, for the static and dynamic cases.

v
vw

S D dz
H

v
v dz
z

(a) (b) (c)

Figure 2.19: Lines of major principal stress (a) Static,


(b) Dynamic; (c) Force balance on an elemental slice

In the same manner as Janssen, Walters solves for the equilibrium of vertical forces
acting on an elemental slice of thickness, dz, in the cylindrical section of the silo.
The result is stated again as follows:
d v 4
vw
(2.60)
dz D
where v is the average vertical stress acting across the elemental slice and vw is

the shear stress on the silo wall. To solve equation 2.60, the shear stress at the
wall, vw must be related to the average vertical stress, v acting across the slice.

First vw is related to v and then vw is related to vw .

Since the vertical stress acting across a horizontal elemental slice is not constant,
Walters assumes the average vertical stress, v, is related to the vertical stress

adjacent to the wall, vw, by a distribution factor, as follows:

vw =F v (2.61)
2.26 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

The distribution factor has been determined by Walters from the mohr circle given
in figure 2.20, which has been greatly enlarged in this thesis for clarity. The
distribution factor has been given the symbol F and not D (as in Walters), to avoid
confusion with D used for the silo diameter. The calculations have been continued
after figure 2.20
vw

v
Stresses at the centre

w
m

H
C
Average stresses

v
Stresses at the wall

vw
x
Material yield locus

Wall yield locus

Figure 2.20: Mohr Circle for Stresses at the Wall


and in the Material for Flow Conditions
CLASSIC WALL PRESSURE THEORIES: DYNAMIC PRESSURES: CYLINDRICAL SECTION 2.27

The green circle in figure 2.20, represents the stresses at the centre of the silo, the
blue circle represents the average stresses, and the red circle represents the
stresses at the wall. All three circles are tangential to the material effective yield

locus. The vertical stress, vw, and the shear stress, vw, at the wall, are determied

from the intersection of the wall yield locus with the circle for wall stresses. Walters
assumes that the horizontal stress, H
, remains constant across the slice as shown

in figure 2.20.
The circle for average stresses has a radius, x, and the centre of the circle is given
by ( +
H v
)/2

The centre of the circle is related to the internal angle of friction of the material as
follows:

H v x 2x
which can be re-written as: v H (2.62)
2 Sin m Sin m

The radius of the Mohr circle, x, for average stresses is given by:
2
2 v H
x2 v H (2.63)
2

Substituting equation 2.62 into equation 2.63, solving for x and re-arranging gives
the following expression for the radius of the circle:
Tan m 2 2
x H H Sin2 m vw
Cos 2 m (2.64)
Cos m

The shear stress at the centre of the silo is


Silo Silo
centre wall zero, while the shear stress is a maximum at
line
=max the silo wall. Walters assumes that the shear
vw
v
stress varies linearly as shown in figure 2.21.
The shear stress at an arbitrary distance, r,
0 from the centre of the silo is given by v, and
r
is related to the shear stress at the wall, vw,
D/2
as follows:

Figure 2.21: Shear stress = r / (D/2) (2.65)


v vw
variation across the silo

Substituting for x from equation 2.64, and from equation 2.65 into equation 2.63
gives the following expression for the distribution factor, F given in equation 2.61,
2.28 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

as follows: vw =F v where F is given by:

Cos w 1 Sin2 m 2 Sin2 m Sin2 w


F (2.66)
3 /2
2 2
4 Tan m Sin m Tan
Cos w 1 Sin2 m 1 1 w
3 Tan2 w Tan 2
m

The graph of the distribution factor, F, as a function of the material friction angle,

m, and the wall friction angle, w, has been shown in figure 2.22 below.

3.0 w=89

2.8

2.6

2.4
Distribution Factor (F)

2.2 w=40

2.0

1.8

1.6

1.4

1.2

1.0
w=10
0.8
0 10 20 30 40 50 60 70 80 90
Material Friction Angle ( m)

Figure 2.22: Distribution factor as a function of both


material friction angle ( m) and wall friction angle ( w)

The curves of the wall friction angle in figure 2.22 have been given in increments of
10 . From equation 2.66 for F, a wall friction angle of 90 has no meaning as this
would require a division by zero, therefore a maximum value of w= 89 has been

shown in figure 2.22. From the figure it can be seen that for all values of material
friction angle and wall friction angle, the vertical stress at the wall of the cylinder,

vw, is greater than the vertical stress, v, at the centre of the cylinder.

Now that vw has been defined as a function of v the next step in the solution of

equation 2.60, is to define the shear stress at the wall, vw as a function of vw.
CLASSIC WALL PRESSURE THEORIES: DYNAMIC PRESSURES: CYLINDRICAL SECTION 2.29

To find an expression of vw as a function of vw the Mohr stress circle for the

stresses at the wall, as given in figure 2.20, has been repeated in figure 2.23
below. From the Mohr circle Walters derives an expression for the ratio of vw/ vw.

Material yield Stresses at


locus the wall
Wall
yield P N
locus vw

2 D m
w

H W vw O

Figure 2.23: Mohr circle for stresses at the wall

Let the radius of the circle be denoted by ‘a’ = PW = NW

Let B = vw/ vw =P H / (OW – W vw )

And P H = a Sin2 D ; OW = a / Sin m ;W vw= W H = a Cos2 D (2.67a,b,c)

The subscript D in D refers to the angle shown in figure 2.19b for the case of

material discharge ie the dynamic case.

Therefore, B = Sin2 D Sin m / (1 - Sin m Cos2 D ) (2.68)

From triangle MWO: +2 D = /2 + w (2.69)

Therefore, 2 D = /2 + w - (2.70)

From triangle PMW: = Arc Cos (MW/a) (2.71)

And MW = OW Sin w (2.72)

Substituting equation 2.67b for OW into 2.72 gives:


MW = a Sin w / Sin m (2.73)

Substituting equation 2.73 into equation 2.71 and re-arranging gives :

2 D /2 + w - Arc Cos( Sin w / Sin m ) (2.74)


2.30 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Substituting equation 2.74 into equation 2.68 and re-arranging gives the following
expression for B:

Sin w Cos 2 m
B (2.75)
2 2 2
Cos w 1 Sin m 2 Sin m Sin w

Therefore, vw =B vw (2.76)

The graph of B as a function of the material friction angle has been given in figure
2.24 below. Walters gives the plot of the factor B as a function of the wall friction
angle, therefore the shape of the graph in figure 2.24 does not correspond to that
given by Walters. The curves of varying wall friction angle have been given in
increments of 10 . The first two curves for wall friction angles of 5 and 10 have
been plotted in dotted lines for clarity only.

3 w=10 w=5

w=20

2.5

1.5
Ratio (B)

w=40
1

0.5

w=80
0
0 10 20 30 40 50 60 70 80 90
Material Friction Angle ( m)

Figure 2.24: Ratio B as a function of material friction angle.

Substituting equation 2.61 into equation 2.76, and then substituting this into the
equation of equilibrium of forces (equation 2.60) gives the following:

d v 4BF
v
(2.77)
dz D

where F is given by equation 2.66 and B is given in equation 2.75.

Plotting the product of the distribution factor, F, and the ratio, B, gives a set of
curves similar in shape to the curves in figure 2.24. This product has been shown in
CLASSIC WALL PRESSURE THEORIES: DYNAMIC PRESSURES: CYLINDRICAL SECTION 2.31

figure 2.25 below. The curves for wall friction angles of 5 , 10 and 20 have been
plotted in dotted lines for clarity only. For most silo problems the wall and material
friction angles lie within the range shown by the red dotted line in figure 2.25.

w=10 w=20
3

2.5
w=40

1.5
Product BF

0.5
w=80

0
0 10 20 30 40 50 60 70 80 90
Material Friction Angle ( m)

Figure 2.25: Product of distribution factor, F, and ratio, B.

Before Walters gives the solution to equation 2.77, he first puts the variables in
dimensionless form by dividing throughout by D and letting SV equal v
/ d and

Z=z/D as follows:
dS V
4BFS V 1 (2.78)
dZ

The solution to equation 2.78 for the vertical pressure in a silo has the form :
1 4BFZ
SV 1 e (2.79)
4BF

Walters gives the horizontal pressures acting normal to the silo wall from the
relationship given below:

SH = SV*B*F/m (2.80)

As Walters points out these equations are of similar form to the classical Janssen
equation, with the factor BF in place of Janssen’s Ka , where Tan w and Ka is the

active stress ratio.


2.32 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

The vertical pressures acting across a slice have been given in figure 2.26a, and the
horizontal pressures acting normal to the silo wall have been shown in figure 2.26b.
The static pressures acting in the silo have been calculated using the Janssen
equations as given by equation 2.5 for the vertical pressures, and equation 2.6 for
the horizontal pressures. The vertical dynamic pressures have been calculated
using Walters equation 2.79, and the horizontal pressures have been calculated
using equation 2.80. To keep the equations consistent, the Janssen equation has
been divided throughout by D to present it in dimensionless form.

Vertical pressure Horizontal pressure


0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6
0

1 Janssen
Janssen
static
static
2
Walters
dynamic
Depth within silo Z=z/D

3 Walters
dynamic

8
(a) (b)

Figure 2.26: Static and Dynamic Pressures: (a) Vertical pressure


(b) Horizontal pressures

From figure 2.26 it can be seen that the vertical dynamic pressures are
approximately 3 times less than the static pressures for values of Z greater than 5.
However, the dynamic horizontal pressures are greater than the static pressures by
a factor of approximately 2.6. This excludes the effect of a switch pressure which
has been discussed in chapter 2.3. The dynamic pressures approach a constant
value at a height to diameter ratio of approximately 2, whereas in the case of static
pressures, the asymptote is only reached at height to depth ratios of approximately
5.
CLASSIC WALL PRESSURE THEORIES: STATIC PRESSURES: HOPPER 2.33

2.2.2 HOPPER SECTION

2.2.2.1 Flow/Slip in the Hopper by Equilibrium of a Slice

In order to examine the variation of stress within the hopper, a slice equilibrium
model similar to the Janssen model has been investigated, as shown in figure 2.27.
The model assumes that the wall slip and the material flow occur simultaneously,
so that the Mohr circle is tangential to the yield surface and the wall stress is
defined by the friction angle . A circular hopper with a half angle (
w
, a radius of

r0 at the transition has been assumed. The depth (z) is measured from the
transition, and the stress at the transition is given by Janssen theory. The vertical
stress v is assumed to act uniformly over the slice, and the pressures normal and

tangential to the wall are given by N and respectively. The material in the

hopper has a bulk density of .

It will be shown that the slice model leads to a first order differential equation with
non-constant coefficients, and that a solution can be obtained numerically.
Importantly, the first order differential equation allows only one boundary condition
for stress at the top of the hopper. The variation of stress with depth and the stress
at the hopper bottom are dependent on the material properties and the hopper half
angle, and are obtained from the numerical equation solution.

The radius of the top of the element is


given by:

V
rt r0 z Tan (2.81)
r0
The radius of the bottom of the element
is given by:
rt
dz rb r0 (z dz) Tan (2.82)
rb
z And the average radius of the element is
N
given by:
rav r0 (z dz / 2) Tan (2.83)

The shear stress acting on the wall of


Figure 2.27: Elemental slice
the hopper is given by:
of material in the hopper
v
N Tan w Tan w (2.84)
Q

The pressure normal to the hopper wall N, has been substituted by v / Q, where
the expression for Q will be derived later.
2.34 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

By taking the sum of the vertical forces equal to zero, and considering the
downwards direction as positive, gives the following equation for equilibrium:

dz dz
v rt 2 v
v
dz rb 2 N Sin 2 rav Cos 2 rav rav 2 dz 0
z Cos Cos

(2.85)

Substituting equations 2.81, 2.82, 2.83 and 2.84 for rt, rb, rav and N in equation

2.85, and re-arranging, collecting and canceling common terms, gives the equation
below.

v 2 1 1 Tan w
Tan r0 1 zTan 1 r0 zTan v
2
z r0 zTan Q Q Q

(2.86)
The boundary value problem given in equation 2.86 is a first order differential
equation and can be re-written as:

v
c v (2.87)
z
where c is the term given in the square brackets in equation 2.86.
2
The term (r0-zTan ) in the denominator of c, was found to make the solution of
equation 2.87 unstable. Therefore, the differential equation was re-written as:

2 v 2
r0 zTan C v r0 zTan (2.88)
z

1 1 Tan w
where C 2 Tan r0 1 zTan 1 r0 zTan (2.89)
Q Q Q

V
To find the expression for Q in equations 2.84
D and 2.86, the stresses acting on an element of
-
material, adjacent and parallel to the hopper
+
- wall, has been shown in Figure 2.28. The
-
- corresponding Mohr circle has been shown in
N
-
Figure 2.29. The sign convention used is that of
+
Gere and Timoshenko (1996), where
compressive stresses and shear stresses acting
in a clockwise direction are negative, and shear
Hopper wall
stresses acting in a counter clockwise direction
are positive. A positive or negative symbol has
Figure 2.28: Stresses
acting on an element been placed next to the arrowhead of each
CLASSIC WALL PRESSURE THEORIES: STATIC PRESSURES: HOPPER 2.35

stress in Figure 2.28. The hopper wall has a half angle inclined to the vertical,

and a wall friction angle of w. The stress acting normal to the wall is N
and the

stress acting at right angles to N is D. The vertical stress is V and acts on a


horizontal plane as shown in Figure 2.28.

In Figure 2.29 below, the arrowheads on the axes indicate the positive directions of
normal and shear stresses. The circle has been drawn tangential to the material
yield locus since the material has not yet yielded. The maximum shear stress which
can occur between the element and the wall, is determined by the point of
intersection of the wall yield locus and the circle. This point locates the normal
stress N on the circle. Using the method of origin of planes, the plane of the

hopper wall has been plotted at an angle, to the vertical and has been shown by

the dotted line through N


in Figure 2.29. The intersection of this line with the circle

locates the origin of planes labeled OP. Since the vertical stress acts on a horizontal
plane, a line has been drawn through OP to locate the point of the vertical stress on

the circle, labeled V


. Thus the conjugate pairs of stress D
and H
are located,

since the other points have been established.

Material yield locus

H
Wall
yield N
locus

2 M

W
+
C O

D
OP
V +

Figure 2.29: Mohr circle for an element adjacent to the hopper wall.

The radius of the circle can be given by:

2
N D 2 2
r N W where W=Tan W . (2.90)
2
2.36 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Let = D / N which is less than 1 since D < N on the Mohr circle.

Then the radius of the circle can be re-written as:

N 2 2
r 1 4 w (2.91)
2

The vertical stress can be given by:

N D
v r Cos 2 (2.92)
2

N
Tan w
2 Tan w
where ATN ATN (2.93)
N D 1
2

Substituting equation 2.91 and 2.93 into 2.92, and re-arranging gives the following
expression for the vertical pressure acting in the hopper under filling conditions:

v Q N (2.94)

1 2 2 2 Tan w
where Q 1 1 4 w Cos ATN 2 (2.95)
2 1

The only unknown in the expression for Q is the stress ratio given by D/ N .

To solve the boundary value problem given by equation 2.88, the Runge-Kutta
method of numerical mathematics has been used. This is stated as follows:

h
K n1 2 K n2 2 K n3 K n4 (2.96)
n 1 n 6

where: Kn1= (zn; Nn)

Kn2= (z n +h/2 ; Nn+K n1 h/2)


(2.97 a,b,c & d)
Kn3= (z n +h/2 ; Nn+K n2 h/2)

Kn4= (z n +h ; Nn+K n3 h)

The solution of equation 2.88 using equation 2.96 and 2.97a,b,c and d, has been
done using an excel spreadsheet. The full set of calculations showing the
implementation of the Runge-Kutta method on the differential equation 2.88, has
been given appendix A.
CLASSIC WALL PRESSURE THEORIES: STATIC PRESSURES: HOPPER 2.37

Vertical Pressure The solution of equation 2.88 gives a


curve as shown in Figure 2.30. The
shape of this curve is in contrast to
that given by Walker, refer Figure 2.7.
According to the theory given in this
thesis, the maximum vertical pressure
Depth of hopper

is at the transition from cylinder to


hopper, and not at the base of the
hopper, as suggested by Walker.
Furthermore, the pressure decreases
rapidly with depth in the hopper.

Figure 2.30: Typical vertical


pressure curve in a hopper

To determine the effect of changing the variables on the solution of equation 2.88,
a sample silo was analysed. In this example, the silo has a height of 7m, a
diameter of 2m, a hopper depth of 2.6m and hopper half angle of 15 , a wall
3
friction angle of 20 and a bulk density of 17kN/m . The value of the stress ratio, K,
used to determine the vertical pressure at the level of the transition was 0.4. These
results have been shown respectively in Figures 2.31 a,b,c,d and e, for varying
hopper radius, material bulk density, wall friction angle, hopper half angle and the
stress ratio, .

Vertical Pressure (kPa) Vertical Pressure (kPa)


0 20 40 60 0 20 40 60 80
0 0

1
2
2

4 3
Depth (m)
Depth (m)

6 4

r=1 5 =10
8 r=2 =15
r=3 6
=20
10 7

(a): Changing silo radius (b): Changing material bulk density

Figure 2.31 (a & b): Effect of variables on vertical pressure in the hopper.
2.38 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

In Figure 2.31(a), the silo radius has been varied from 1m to 3m. It can be seen
that at depths equal to one silo diameter, there is an increase in the vertical
pressure. For a radius of 1m, the vertical pressure at a depth of 2m is 3.34kPa,
while for a radius of 3m the vertical pressure at a depth of 6m is4.76kPa. This is an
increase of 42.5%. In Figure 2.31(b), the material bulk density has been increased
3 3
from 10kN/m to 20kN/m . At a depth of 1m, there is an increase in the vertical
3
pressure of 9.27kPa to 18.54kPa for an increase in bulk density from 10kN/m to
3
20kN/m . Thus when the bulk density is doubled, so too is the vertical pressure.
Since the bulk density of a material can vary with time, moisture content and
source, a range of densities needs to be specified when designing a silo.

In Figures 2.31 c & d below, the effects of changing the stress ratio and the wall
friction angle have been shown. From Figure 2.31 (c) it can be seen that at a depth
of 1m, the vertical pressure is 15.76kPa for a stress ratio of 0.3, while for a stress
ratio of 0.9 the vertical pressure increases to 54.44 kPa. This is an increase of
approximately 245%. Therefore, using the correct stress ratio has a significant
effect on the value of the calculated vertical pressure. From Figure 2.31(d) at a
depth of 1m below the transition, it can be seen that changing the wall friction
angle from 10 to 20 has the effect of decreasing the vertical pressure from
22.53kPa to 15.11kPa. This is a decrease of approximately 33%.

Vertical pressure (kPa) Vertical pressure (kPa)


0 20 40 60 80 0 20 40 60 80
0 0

1 1

2 2

3 3
Depth (m)

Depth (m)

4 4

5 =0.3 5 =10
=0.6 =15
6 6
=0.9 =20
7 7

(c): Changing hopper stress ratio (d): Changing wall friction angle

Figure 2.31 (c & d): Effect of variables on vertical pressure in the hopper.
CLASSIC WALL PRESSURE THEORIES: STATIC PRESSURES: HOPPER 2.39

Vertical Pressure (kPa)


From Figure 2.31 (e) it can be
0 20 40 60 80
0 seen that increasing the hopper
half angle from 10 to 30 has
1
the effect of increasing the
2 vertical pressure, up to a depth

3 of 0.75r below the transition.


Beyond this point, the vertical
Depth (m)

4
pressure decreases with
5 =10 increasing hopper half angle.
=20
6
=30
7

(e): Changing the hopper half angle

Figure 2.31 (e): Effect of variables


on the vertical pressure in the hopper

2.2.2.2 Walters Pressures in Converging Channels

Walters (1972b) developed his theory for pressures acting on the walls of axially
symmetric hoppers in the same manner as his theory for pressures in vertically
sided silos. Figure 2.32a below shows the lines of the major principle stress in the
hopper when the material starts flowing. The stresses acting on a horizontal
elemental slice have been shown in figure 2.32b.
D

v
D
w
z Area A

dz
w
Area (A+dA)

v +d v

(a) (b)

Figure 2.32: (a) Lines of major principle stresses (b) Elemental slice
2.40 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Walters gives the results of the vertical force balance on the elemental slice in the
hopper as follows:

d v 1 dA P
v w w Tan g (2.99)
dz A dz A

where A is the cross sectional area of the element at a depth z from the top of the
hopper, P is the perimeter at that depth, v is the uniform average vertical stress

acting across the slice, w is the shear stress acting along the wall, w is the stress

normal to the hopper wall and is the hopper half angle.

To solve equation 2.99, Walters finds a relationship between the third term in

equation 2.99, ( w+ wTan ), and the vertical stress at the hopper wall, vw, from

the geometry of the Mohr circle given in figure 2.33. As before, the green circle
represents the stresses along the centreline of the hopper and the red circle
represents the stresses at the hopper wall. The stresses at a distance r from the
centre of the hopper have been shown in figure 2.33 by the blue circle. Since
Walters makes the assumption that the horizontal stress, H
, remains constant

across the slice, the points P and N can be located on the Mohr circles.

Walters further defines the angle as that angle which the line through point P
makes with the horizontal axis of the graph in figure 2.33.

Let the radius of the circle for the stresses at the wall be denoted by a, then:

w = a Sin2 D
(2.100)

Similarly w = OC + C w = a/Sin m + aCos2 D (2.101)

Therefore, substituting equations 2.101 and 2.100 into the third term of equation
2.99 and rearranging gives the following:

w + w Tan = a ( Sin m Sin2 D + ( 1+Sin m Cos2 D )Tan ) / Sin m (2.102)

To eliminate the radius, a, from the expression in equation 2.102, Walters finds a
relationship between the vertical shear stress at the wall and the vertical stress at
the wall, as follows:

Let vw =E vw (2.103)
CLASSIC WALL PRESSURE THEORIES: DYNAMIC PRESSURES: HOPPER SECTION 2.41

vw

O
w

vr

from hopper centreline


Stresses at distance r
Stresses at
the centre
w

Stresses at the wall


vc
m

v
x

vw
Material yield locus

C
2
P

H
N

D
2
a

w
Wall yield locus

Figure 2.33: Mohr circle for stresses in the hopper during discharge

where both vw and vw are found from the Mohr circle to be:

vw= a Sin (2 +2 D) and vw= a /Sin m – a Cos(2 +2 D) (2.104a,b)


2.42 DES NORTJE: PhD THESIS: THE ANTI-DYANMIC TUBE IN MASS FLOW SILOS

Substituting 2.104a,b into 2.103 and rearranging gives the expression for E as
follows:

Sin mSin 2 2 D
E (2.105)
1 Sin m Cos 2 2 D

The expression for angle 2 D


in equation 2.105 is found from the Mohr circle shown

in figure 2.34 as follows:

In triangle MCO, MC = OC Sin (2.106)

In triangle NCO, OC = a / Sin m (2.107)

Therefore, MC = a Sin / Sin m (2.108)

In triangle PCM, = ArcCos(MC/a) (2.109)

Substituting for MC from equation 2.108: = ArcCos (Sin / Sin m) (2.110)

From triangle MCO, +2 +2 =


D
/2 + (2.111)

Therefore, 2 +2 D
= /2 + - /2 + - ArcCos (Sin / Sin m) (2.112)

The expression for 2 + 2 D


in equation 2.112 is similar to the expression derived

for 2 D
in equation 2.92, except w is replaced by . It can be seen from figure 2.34

that as the hopper half angle tends to zero, ie a vertically walled silo, the angle

tends to the wall friction angle w. The method for determining the angles w and

has been given at the end of this section.

Material yield
locus

P
Wall yield vw
locus N
w

M
a 2 D

m
2 w

C vw O
H

Figure 2.34: Mohr circle for


stresses at the wall
CLASSIC WALL PRESSURE THEORIES: DYNAMIC PRESSURES: HOPPER SECTION 2.43

Now the radius, a, can be eliminated from the expression in equation 2.102 by

dividing ( w + w Tan ) by vw to give the following:

w w Tan Sin mSin2 D 1 Sin mCos2 D Tan


(2.113)
vw 1 Sin m Cos 2 2 D

To simplify equation 2.113, Walters adds and subtracts Tan S in mCos(2 +2 D


)

from the numerator on the RHS, which after rearranging gives the following:

w w Tan Sin mSin 2 2 D


Tan E Tan (2.114)
vw 1 Sin m Cos 2 2 D

Substituting equation 2.114 into equation 2.99 gives:

d v 1 dA P
v E Tan vw g (2.115)
dz A dz A

Equation 2.115 still cannot be solved due to the unknown term of vw, therefore

Walters defines the relationship between the average vertical stress, v, and the

vertical stress at the wall, vw, as before (refer chapter 2.2.1.1):

vw = Fh v (2.79 repeated)

where Fh is the distribution factor associated with the hopper stresses. To

determine the expression for the distribution factor Fh, Walters defines the

relationship between the average vertical stress, v , and the vertical stress at a

distance, r, from the centreline of the hopper, vr , as shown in figure 2.33.

Line of average vertical stress, v


Figure 2.33 shows a typical
CL horizontal elemental slice in a
Vertical stress, vr
hopper, with the line of average
at distance r
vertical stress acting on the
element. The line of the real
vertical stress has been shown
Horizontal elemental slice arbitrarily by the curved line in
figure 2.33. The vertical stress at
r
a distance, r, has been shown by
Figure 2.33. Average and general stress
the dotted vertical line.
distribution across an elemental slice in the
hopper.

In Walters paper, the relationship between v and vr has the general form:
2.44 DES NORTJE: PhD THESIS: THE ANTI-DYANMIC TUBE IN MASS FLOW SILOS

rw
1
v 2 r vr dz (2.116)
r2 0

Walters assumes that the vertical shear stress, vr at a distance r from the centre of

the hopper is a linear function of the vertical shear stress at the wall of the hopper,

vw , as before, so that:

vr/ vw = r/rw vr = vw r / rw (2.117)

where r is an arbitrary distance from the centre, and rw is the radius of the hopper

at the wall. Let the radius of the circle through point N be x, then:

2 2 2
x = vr + (x / Sin m- r ) (2.118)

And hence:

Sin m 2 2
x H H Sin2 m vr Cos 2 m (2.119)
2
Cos m

In the Mohr circle for stresses at the wall, 2 +2 D


< /2 + m. If the hopper half

angle is greater than this limit, mass flow in the hopper can not occur. From the
geometry of the Mohr circle:

v = 2x / Sin m- H (2.120)

Substituting for x from equation 2.119, and for vr from equation 2.117 into

equation 2.120, and re-arranging, gives the following:

v
H
1 Sin2 m 2 Sin m 1 cr 2 / rw 2 (2.121)
2
Cos m

2
where c=(Tan / Tan m)

Substituting equation 2.121 for v into equation 2.1116, the average stress across

the slice can be integrated to give the following:

2
v
H
1 sin2 m 2 Sin m 1 1 c 3 /2
2 3c
Cos m

H
1 sin2 m 2 y Sin m (2.122)
2
Cos m
CLASSIC WALL PRESSURE THEORIES: DYNAMIC PRESSURES: HOPPER SECTION 2.45

3 /2
2 Tan2 m Tan2
where y = 2[1- ( 1 - c ) ]/3c =
3/2
1 1 (2.123)
3 Tan2 Tan2 m

From the Mohr circle in figure 2.32, H= vw/Tan (2.124)

Substituting for vw from equation 2.103 into equation 2.126 gives:

H =E vw/Tan where E is given in equation 2.105 (2.125)

Substituting for H from equation 2.125 into equation 2.122 gives the following:

1 Sin2 m 2 y Sin m
v E vw (2.126)
Tan Cos 2 m

Now substituting for E from equation 2.105 into equation 2.126, and re-arranging
gives the following:

Cos 1 Sin2 m 2 ySin m vw


v vw (2.127)
2 2 2 Fh
Cos 1 Sin m 2 Sin m Sin

Cos 1 Sin2 m 2 Sin2 m Sin2


where FH (2.128)
Cos 1 Sin2 m 2 y Sin m

Substituting for vw from equation 2.127 into equation 2.115 gives the following:

d v 1 dA P
v E Tan FH v g (2.129)
dz A dz A
1 dA 4 Tan
For a conical hopper: (2.130)
A dz D 2 zTan

P 4
and (2.131)
A D 2 zTan

Substituting equation 2.130 and 2.131 into equation 2.129, and re-arranging gives
the following:
d v 4 v
EF Tan F 1 g (2.132)
dz D 2zTan H H

The equation for the vertical force balance on the elemental slice given in equation
2.132 can now be solved. As before Walters puts this equation into dimensionless
form;
Let Sv= v/( gD) , Z=z/D , Z0= z0/D (z0 = 0 at the top of the hopper)

2 E FH
and let M FH 1 (2.133)
Tan
2.46 DES NORTJE: PhD THESIS: THE ANTI-DYANMIC TUBE IN MASS FLOW SILOS

Then equation 2.132 can be re-written as:


dS v 2 MTan Sv
1 (2.134)
dZ 1 2 MTan

The solution to the linear first order differential equation, 2.134 is:
M 1
1 2ZTan 1 2ZTan
Sv 1 (2.135)
2 Tan M 1 1 2Z 0 Tan

To find the vertical pressure in the hopper, the unknowns in the expressions for E
and M need to be determined.

From equation 2.112 for 2 + 2 D


, the graph in figure 2.34 has been plotted for

determining the value of 2 D


and . The curves of material friction angle varying

from 10 to 80 has been shown in increments of 10 . To demonstrate the use of


figure 2.34, a curve has been plotted for a silo with a wall friction angle of 22 a
material friction angle of 45 and a hopper half angle of 25 . The value of 22 for
the wall friction is entered along the x-axis. The intersection of this vertical line with

the curved line gives a value of 2 D


,for this example, on the y-axis as 54. To this

value of 2 D
a value of 2 is added 2 = 50 which gives 2 +2 D
as 104.

180
70
160

140
40
120
+2 )

m=10
100
D

80
or (2

60
D

40
2

20

0
0 10 20 30 40 50 60 70 80
( w) or ( )

Figure 2.34: 2 D
+2 as a function of w and
CLASSIC WALL PRESSURE THEORIES: DYNAMIC PRESSURES: HOPPER SECTION 2.47

140
m=45

120

100
2 D
+2 =104

80
+2

60
+2 )

2 D
=54
40
D
or (2

20

w=22 =39
D

0
2

0 4 8 12 16 20 24 28 32 36 40 44

( w) or ( )

Figure 2.35: 2 D
+2 as a function of w and
for a material friction angle, m =45

The intersection of this horizontal line with the curve gives the value of as 39

along the x axis of the graph.

Having thus found the value of , this can be used in equation 2.128 to evaluate FH

Substituting equation 2.123 for y into equation 2.128 shows that equation 2.128 is

the same as equation 2.84 for F, except that w is replaced by . The graph for FH

is therefore, the same as figure 2.26 for various curves of angle .

All the unknowns in the expression for the average vertical pressure in the hopper
given by equation 2.135 have been defined and the stress distribution in a hopper
without a surcharge can be plotted, as shown by the graphs in figure 2.36. For all
three graphs the scales of the vertical pressure, on the horizontal axis, and the
depth, on the vertical axis, are the same. Figure 2.36a was calculated for a material
friction angle of 50 , wall friction angle of 22 and the hopper half angle varied from
5 to 15 in increments of 5 .
2.48 DES NORTJE: PhD THESIS: THE ANTI-DYANMIC TUBE IN MASS FLOW SILOS

Dimensionless Average vertical pressure Sv


0 0.1 0.2 0.3
0 0.1 0.2 0
0.3 0.1 0.2 0.3
0
=5 m=25 w=15
=10 m=30 w=20
0.2 =15
m=35 w=25

0.4
Z=z/D

0.6

0.8

1
(a) (b) (c)

Figure 2.36: Average vertical pressure in conical hopper : (a) Effect of


changing hopper half angle , (b) Effect of changing material friction angle
(c) Effect of changing wall friction angle

Similarly figure 2.36b was calculated for a wall friction angle of 22 , a hopper half
angle of 10 and the material friction angle varied from 25 to 35 in increments of
5 . In figure 2.36c the material friction angle used was 50 , the hopper half angle
was 10 and the wall friction angle varied from 15 to 25 in increments of 5 .
From figures 2.36a and b it can be seen that a larger hopper half angle and higher
material friction angle both have the effect of lowering the vertical pressure acting
in the hopper, while from figure 2.36c it can be seen that the wall friction angle has
a negligible effect on the vertical pressure in the hopper.

For a hopper with a surcharge pressure due to the cylindrical section above the
solution to equation 2.134 would be as follows:

M 1 M
1 2ZTan 1 2ZTan 1 2ZTan
Sv 1 S v0 (2.136)
2 Tan M 1 1 2Z 0 Tan 1 2Z 0 Tan

where all variables are as previously, Z0 is the dimensionless depth ratio at the

transition, Sv0 is the vertical pressure acting at the transition due to the material in
CLASSIC WALL PRESSURE THEORIES: DYNAMIC PRESSURES: HOPPER SECTION 2.49

the cylindrical section of the silo. The value of Sv0 can be calculated from equation

2.97 (the Walters expression for the vertical pressure in a cylinder).

Dimensionless average
vertical pressure, Sv

0 0.05 0.1 0.15 0.2


0
Figure 2.37 shows the
dynamic average vertical
0.5
stress acting in a silo filled

Cylinder
Height to diameter ratio, Z

with material with a


1 friction angle of 50 , a wall
friction angle of 22 and a
1.5 hopper half angle of 10 .
The dotted line in figure

2 2.37, shows the pressure


distribution in the hopper
Hopper

with no surcharge acting in


2.5
the cylindrical section.

Figure 2.37: Dynamic average vertical stress


acting in a silo

2.2.2.3 The Jenike Radial Stress Field

Jenike (1968) defines the condition of flow as a particular case of failure, which
occurs when pressures are such that shear occurs without destroying the isotropy
of the material. During flow, the bulk density of the material is a function of the
pressures. When the pressures are constant, the solid shears at a constant density.
When the pressures increase, the solid compacts and the density increases, and
when the pressures decrease, the solid expands and the density decreases. Thus,
flow can proceed indefinitely. In the hopper the mass of solid contracts laterally and
expands vertically which implies horizontal, or nearly horizontal, major principal
stresses and a plastic-passive state of pressure exists in the hopper. This state may
extend to the top of the bin. Because the solid slides along the walls as it flows, the

vertical pressure at the wall, vw is accompanied by a frictional stress at the wall w


.

There is a change in the wall pressure at the transition from the cylindrical section
2.50 DES NORTJE: PhD THESIS: THE ANTI-DYANMIC TUBE IN MASS FLOW SILOS

to the hopper section, with the pressures decreasing to zero at the vertex of the
hopper. The speed of the flowing material is generally sufficiently slow and close to
the steady state condition for the inertia forces to be negligible. Therefore, the
conditions of equilibrium are satisfied and the vertical force supported by the walls
is equal to the weight of the stored material. The element in the hopper is the same
as that shown in figure 2.18 and figure 2.19. The symbols used in this chapter are
the same as those used in chapter 2.1.2.3.2 resulting in the same equations of
equilibrium as previously. These equations have been repeated here for continuity.

Considering equilibrium in the r-direction +

r 1 r 1 d
2 r c r Cot Cos 0 (2.56 repeated)
r r r 2

Considering equilibrium in the -direction +

r 1 d
3 r c Cot Sin 0 (2.58 repeated)
r r 2

In Bulletin 108, Jenike (1961) gives the following relationships from the Mohr circle
as follows:
1 Sin mCos2 (2.137)

1 Sin m Cos2 (2.138)


r

c 1 Sin m (2.139)

from the assumption that in axial symmetry the circumferential stress is equal to
either the major or minor principal stress of the median plane.

r = Sin mSin2 (2.140)

In Bulletin 108, Jenike(1961) gives the following relationship for the average
vertical stress in the hopper:

v=r s (2.141)

where is the density and is assumed a function of r and as follows: = (r, )

s is the radial stress field and is also function of r and as follows: s=s(r, )
The angle between the major principal stress and the ‘r’ co-ordinate in figure 2.18
is given as ; = (r, )
CLASSIC WALL PRESSURE THEORIES: DYNAMIC PRESSURES: HOPPER SECTION 2.51

The derivative of the shear stress given in equation 2.140 therefore becomes:

r r r
r r r

s
Sin mSin2 r rs s 2 Sin m Cos2 (2.142)
r r r

The derivative of the average vertical stress with respect to is as follows:

s
1 Sin mCos2 r rs 2 Sin mSin2 (2.143)

The derivative of the average vertical stress with respect to r is as follows:

r r r
r r r

s
1 Sin mCos2 r s 2 Sin mSin2 (2.144)
r r

Substituting equations 2.142 and 2.143 into the equation of equilibrium in the
direction given by equation 2.58 gives the following:

s
Sin mSin2 r rs s 2 Sin mCos2 .............
r r r

1 s
...... 1 Sin mCos2 r rs 2 Sin mSin2 3 Sin mSin2 .......
r

1 d
...... Cot Sin mCos2 Sin m Sin 0 (2.145)
r 2

Now substituting for =r s into equation 2.145 and collecting terms in s/ r, s/ ,

s and the remaining terms gives the following:

s s
A B Cs D 0 (2.146)
r

where A = r Sin mSin2 (2.147)

B= (1-Sin mCos2 (2.148)

C Sin mSin2 r 1 Sin mCos2 2 Sin m rCos2 .......


r r
2.52 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

...... 2 Sin m Sin2 3 Sin mSin2 Sin mCot 1 Cos2 (2.149)

D = - Sin (2.150)

Solving for s/ r in equation 2.146 gives :


s D C B s
s (2.151)
r A A A

Similarly in the r direction, substituting equations 2.142 and 2.144 into the
equation of equilibrium given by equation 2.56, and collecting terms in s/ r, s/ ,
s and the remaining terms gives the following:
s s
E F Gs H 0 (2.152)
r

where: E r 1 Sin Cos2 (2.153)


m

F Sin Sin2 (2.154)


m

G 1 Sin mCos2 r Sin mSin2 2r 2 Sin mCos2 .......


r r

..... 2 1 Sin mCos2 1 Sin mCos2 1 Sin m Sin m Sin2 Cot

(2.155)

H Cos (2.156)

Substituting for s/ r in equation 2.151 into equation 2.152 and collecting terms in
s/ , s and the remaining terms, gives the following:

EB s EC ED
F G s H 0 (2.157)
A A A

Due to the lengthy size of equation 2.157, each term has been evaluated
separately. To maintain an overview of the solution, equation 2.157 has been
repeated several times during the following calculations.

The first term of equation 2.157 is as follows:

EB 1 Sin mCos2
F Sin mSin2 r 1 Sin mCos2
A r Sin m Sin2

Sin2 m 1 Cos 2 m
............. (2.158)
Sin m Sin2 Sin mSin2
CLASSIC WALL PRESSURE THEORIES: DYNAMIC PRESSURES: HOPPER SECTION 2.53

EB s EC ED
Equation 2.157 repeated: F G s H 0
A A A

EC
G 1 Sin mCos2 r Sin m Sin2 2r ..........
A r r

...... 2 Sin mCos2 2 1 Sin mCos2 1 Sin mCos2 1 Sin m ......

1 Sin2 mCos 2 2
...... Sin mSin2 Cot 1 Sin mCos2 r .......
r Sin mSin2

2r 1 Sin mCos2
.... Cos2 2 1 Sin m Cos2 3 1 Sin mCos2 ....
Sin2 r

Cot
.... 1 Sin mCos2 Cos2 1 (2.159)
Sin2

2 2
Now multiplying out all the terms , replacing Cos 2 with(1-Sin 2 ), cancelling out,
and after many lengthy algebraic manipulations, the second term of equation 2.157
can be simplified to:

EC Cos 2 m 2r
G Sin m 3 2 Cos2 Sin m ......
A Sin mSin2 Sin2 r

Cot
...... Cos2 1 Sin m 1 (2.160)
Sin2

EB s EC ED
Equation 2.157 repeated: F G s H 0
A A A

ED 1 Sin mCos2 Sin


H Cos (2.161)
A Sin m Sin2

Solving for s/ in equation 2.157 is as follows:

EC ED
G H
s A A
s 0 (2.162)
EB EB
F F
A A

As described previously, solving the terms in equation 2.162 is a lengthy process,


equation 2.162 has again been repeated throughout the calculations to maintain an
overview of the solution progress.

To evaluate the first term in equation 2.162, substitute equations 2.160 and 2.158
to give the following:
2.54 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

EC
G
A Sin m 3 1 2Sin mSin2
Sin mSin2 ..........
2
F
EB Cos m Cos 2 m
A

2rSin m Cot Sin m


...... Cos2 Sin m Cos2 1 Sin m 1
2
Cos m
r Cos 2 m

This can be simplified to give the following:

EC
G
A 1 2 Sinm Sin2 2 r Cos2 Sin m Sin m
........
EB Cos 2 m Cos 2
r
F m
A

Sin m
..... Sin2 Sin m 3 Cot 1 Sin m 1 Cos2 Cos2 (2.163)
2
Cos m

This is the same as the f(r, ) term given in Jenike’s Bulletin 108 (1961).

EC ED
G H
s A A
Equation 2.162 repeated: s 0
EB EB
F F
A A

To solve the second term in equation 2.162, substitute equations 2.161 and 2.158
to give the following:

ED 1 Sin mCos2 Sin


H Cos
A Sin m Sin2
EB 1 Sin2 mCos 2 2
F Sin mSin2
A Sin mSin2

which can be simplified to give:

ED
H
A Sin m Sin
Sin 2 (2.164)
EB Cos 2 m Cos 2 m
F
A

This is the same as the g(r, ) term given in Jenike’s Bulletin 108 (1961)
CLASSIC WALL PRESSURE THEORIES: DYNAMIC PRESSURES: HOPPER SECTION 2.55

Thus, equation 2.162 has the form:

s
f (r, ) s g(r, ) 0

Therefore, ( s/ f(r, )s - g(r, ) (2.165)

where f(r, ) is given by equation 2.163 and g(r, ) is given by equation 2.164

Having solved for s/ , equation 2.165 can be substituted into equation 2.151 to
solve for s/ r as follows:

s B GA EC B HA ED
C s D (2.166)
r A FA EB A FA EB

As previously, the evaluation of both terms in equation 2.166 is a lengthy process


and equation 2.166 has been repeated throughout the following calculations to
maintain an overview of the solution progress.

Evaluating the first term in equation 2.166 :

B GA EC 1 Cos2 1 2 SinmSin2
C .........
A FA EB r Sin m Sin2 rSin2 Cos 2 m

2 r Cos2 Sin m Sin m Sin m


........ Sin2 Sin m 3 ......
2 2
Cos m r Cos m

Sin m
....... Cot 1 Sin m 1 Cos2 Cos2 Sin m Sin2 r ......
2
Cos m r

........ 1 Sin mCos2 2 Sin m rCos2 3 Sin mSin2 ......


r

...... 2 Sin m Sin2 Sin mCot 1 Cos2

This can be simplified by multiplying out all the terms and then collecting terms in
/ r, / , / r, / and the remaining terms. After considerable mathematical
manipulation the result is as follows:
2.56 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

B GA EC 1 2 Sin m Sin m
C 1 Cos2 Sin m 2 Sin2 ......
A FA EB r r Cos 2 m Cos 2 m r

1 Sin m
..... 1 Sin m Cot Sin2 Cos2 1 (2.167)
r r Cos 2 m

s B GA EC B HA ED
Equation 2.166 repeated: C s D
r A FA EB A FA EB

To evaluate the second term in equation 2.166, substitute equations 2.147, 2.148,
2.150, and equation 2.163 as follows:

B HA ED 1 Sin mCos2 Sin m Sin


D Sin 2 Sin
A FA EB r Sin mSin2 Cos 2 m Cos 2 m

Multiplying out, cancelling and collecting terms, the above equation can be further
simplified to give:

B HA ED 1 Sin m Cos
D Cos 2 (2.168)
A FA EB r Cos 2 m Cos 2 m

Both equations 2.167 and 2.168, can be multiplied throughout by the radius, r, of
the hopper. This then gives the final expressions as found in Bulletin 108.

Finally the two differential equations have now been defined as follows:
s
f (r, ) s g (r, ) 0 (2.169)

s
r h(r, ) s j(r, ) 0 (2.170)
r
where f(r, ) is given by equation 2.163, g(r, ) is given by equation 2.164.
CLASSIC WALL PRESSURE THEORIES: DYNAMIC PRESSURES: HOPPER 2.57

Sin m Sin m
h(r, ) 1 2 1 Cos2 Sin m 2r Sin2 ......
Cos 2 m Cos 2 m r

r Sin m
..... 1 Sin m Cot Sin2 Cos2 1 (2.171)
r Cos 2 m

Sin m Cos
j(r, ) Cos 2 (2.172)
Cos 2 m Cos 2 m

Jenike further simplifies his calculations by making the assumption that is only a

function of and that the density g is a constant. Then the terms / r, / / r


become zero and equations 2.169 and 2.170 become:

s
f ( )s g( ) 0 (2.173)

s
r h( ) s j( ) 0 (2.174)
r

Both equations 2.173 and 2.174 are first order linear partial differential equations.

2 Sin mSin2 Sin m


where f ( ) Sin2 Sin m 3 .........
Cos 2 m Cos 2
m

Sin m
....... Cot 1 Sin m 1 Cos2 Cos2 (2.175)
2
Cos m

Sin m Sin
g( ) Sin 2 ( as before ) (2.164 repeated)
2
Cos m Cos 2 m

Sin m
h( ) 1 2 1 Cos2 Sin m 1 Sin m Cot Sin2 Cos2 1
2
Cos m

(2.176)
Sin m Cos
j( ) Cos 2 ( as before ) (2.172 repeated)
2
Cos m Cos 2 m

In Bulleting 108, Jenike(1961) gives the solutions to the differential equations given
in equations 2.173 and 2.174. Then Jenike assumes s/ r to be equal to zero. This
simplifies the solution process considerably to give only two unknowns in equations
2.174 and 2.175. These unknowns are / and s/ .
2.58 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

The unknowns / and s/ can easily be solved for. Equation 2.174 now
becomes:

h( )s = -j( ) (2.177)

2
Multiplying equation 2.177 throughout by Cos m/Sin m and dividing throughout by

2s, re-arranging and collecting terms ,gives the following expression for / as
follows:

2
Cos Sin m Cos 2 s Cos m
.............
2s Sin m Cos2 Sin m

sSin m 1 Sin m Cot Sin2 cos 2 1


........... 1 (2.178)
2sSin m Cos2 Sin m

Equation 2.178 represents the


1 1
variation of with respect to along a
given ray of radius, r as shown in
2
2
figure 2.38.

On the axis of the hopper, equals


r zero and equals 90 . At some
arbitrary distance in the hopper, when
varies from 1 to 2, varies from 1

to 2 on the same ray. Thus a family of

solutions can be plotted for equation


Figure 2.38: Variation of
with respect to within the 2.178 for a range of material friction
hopper.
angles, m, and stress field values of s.

Substituting for / from equation 2.178 into equation 2.173 and solving for s/
gives the following:

s s Sin2 Sin 2 s Sin m Cot 1 Cos2 Sin2


(2.179)
Cos2 Sin m

Equations 2.178 and 2.179 can be solved using numerical methods and applying
the physical boundary conditions of = 0 on the axis of the hopper, and = at
the wall.
CLASSIC WALL PRESSURE THEORIES: DYNAMIC PRESSURES: HOPPER 2.59

The solution to equation 2.179 is substituted into the expression for the average
vertical stress given by equation 2.141, to give a radial stress field as follows

v =r s( ) (2.180)

In Bulletin 108, Jenike(1961) considers the


general case of a hopper as shown in figure
w2
w1 2.39, where the walls of the hopper have
different friction angles, the slopes of the

1
2 hopper walls are not the same on either side of
the axis and the wall friction angle approaches
the material friction angle, ie rough walls. In
this thesis only the solution of an axially
symmetric hopper (ie one hopper half angle)

Figure 2.39: General and a wall friction angle less than the material
Hopper geometry friction angle has been considered.

Since the density in equation 2.180 does not equal zero, the radius does not equal
zero and the stress field s( ) does not equal zero, the radial stress field given by
equation 2.180 cannot extend upwards to a free surface. Therefore, the stress field
in a hopper without a surcharge deviates significantly from a radial stress field in
the upper part of the hopper.

The solution of equation 2.178 for can be plotted on a graph with as the vertical
axis and as the horizontal axis. Jenike (1961) states that the boundary conditions
of equation 2.178 are not mathematically uniquely defined, but can be determined
from the physical boundary
conditions. There are two
(0, )
physical boundary conditions
( , ) imposed on the equation for
/ which are symmetrically
(0, /2)
located as shown in figure 2.40.
On the axis of the hopper, =
(- , )
/2 and = 0; and at the wall
= . Since there is no direct
way of finding a solution which
(0,0)
connects two boundary points,
Figure 2.40: General shape of the function Jenike has computed several
/ in ( , ) co-ordinates
solutions and interpolated the
required functions.
2.60 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Having thus solved for the radial stress field, Jenike derives the normal wall
pressure acting on the hopper during flow from equation 2.137, which has been
repeated below:

1 Sin mCos2 (2.137 repeated)

In the derivation of , the normal pressure


D N

on the hopper wall, Jenike uses the (x,y) co-


ordinates as shown in figure 2.41 below.

B
The hopper has a diameter, D, at the top, a

N
-x Hh width of B=2y at some distance –x above
the vertex, and a hopper half angle of .

The polar co-ordinates used in the derivation

y of the radial stress can be expressed in


terms of (x,y) as follows:
x

Figure 2.41: Co-ordinate r = Sin (B/2) at the wall. (2.181)


system in the hopper

Substituting equation 2.181 into equation 2.180 and in turn substituting this into
equation 2.137 gives the following expression for the normal stress acting on the
wall:
B
N s( ) Sin 1 Sin mCos2
2
which Jenike has written in the following form:

N Sin
s( ) 1 Sin mCos2 (2.182)
B 2

Putting equation 2.182 as a function of the depth of the hopper gives the following
equation for the normal stress acting on the hopper wall:
x Sin
N s( ) D 1 Sin mCos2 (2.183)
Hh 2

where B has been replaced with Dx/Hh . In equation 2.183 above, the term x/Hh is

the depth within the hopper as a ratio of the hopper height. From equation 2.183 it
can be seen that N
is a linear function of the radial stress s( ), the bulk density of

the material, , the diameter of the hopper at the transition, D, and the depth ratio,
x/Hh. The graph in figure 2.42 shows the hopper normal wall stress, and has been

plotted for s( ) = D = =1.


CLASSIC WALL PRESSURE THEORIES: DYNAMIC PRESSURES: HOPPER 2.61

1 1
=25 =25 m=25
0.9 0.9
0.8 0.8

0.7 =15 0.7


Depth ratio (x/Hh)

0.6 0.6

0.5 0.5
0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0
0.0 0.1 0.2 0.3 0.0 0.1 0.2 0.3
Normal wall pressure N
(kPa) Normal wall pressure N
(kPa)
(a) (b)

Figure 2.42: Normal wall stress N


, acting in the hopper during flow

Figure 2.42(a) shows the effect on the value of N


, of varying the hopper half angle

from 25 to 15 , for a material friction angle of 45 and an angle of 60 . Using the


same scale, figure 2.42(b) shows the effect of changing the material friction angle
from 45 to 25 ( =60 as in figure (a) and =15 ). The third curve in figure
2.42(b) shows the effect of changing the angle from 60 to 45 ( m=45 as in

figure (a) and =15 ).

From equation 2.183, the greatest effect on the hopper normal wall pressure is due
to the radial stress field s( ), the material bulk density, , and the diameter of the
hopper at the transition, D. The graphs in figure 2.42(a,b) show that the hopper
half angle has a greater effect on the normal wall pressure compared to the
material friction angle and the angle of the major principal stress, .
2.62 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

2.3 SWITCH PRESSURES

2.3.1 CYLINDRICAL SECTION

2.3.1.1 Jenike Upper Bound Pressures

The pressures exerted by a solid stored in a silo are affected by the wall
imperfections of the silo (ie deviations from cylindrical) and the boundary layers of
material which form at the walls. Therefore, Jenike only considers the bounds on
wall pressures, the minimum being described by Janssen (equation 2.5) and the
critical upper bound which Jenike has based on the considerations of strain energy
of the stored material..

During flow of a mass solid, Jenike(1973b) states that the energy is lost at the
maximum rate possible. Therefore, the recoverable strain energy tends towards a
minimum, which is approached as closely as the wall imperfections will allow. Using
Janssen’s assumption of the stresses being independent of the horizontal co-
ordinates, Jenike solves for the strain energy in the cylinder in one dimension. He
neglects the strain energy due to shear stresses and assumes the cylinder walls to
be rigid. Hence in his analysis, the vertical co-ordinate, z, is the independent
variable. Figure 2.43 shows a typical horizontal element in a cylinder, of cross
sectional area A.

℄ In his derivation, Jenike assumes the vertical


pressure, V , to be the major principal stress,

the horizontal pressure H, is the minor principle


z
v
stress and the circumferential pressure, C, is

the intermediate stress. Therefore:


dz

1 = V = RS (2.184)

2 = 3 = C = H = K RS (2.185)
Figure 2.43: Horizontal
Element in the cylinder where K= H/ V is the stress ratio.

Since Jenike(1973b) does not give a definition of the symbol S, it has been
assumed in this thesis that S has been used to denote a stress field. Furthermore,
while equation 2.184 appears to be similar to equation 2.63, the symbol R denotes
the hydraulic radius of the cylindrical section and not the radial co-ordinate as
previously. There is no explanation for this change of symbols in either Bulletin 108
(Jenike 1961) or in the paper relating to the switch pressure, Jenike (1973b)
CLASSIC WALL PRESSURE THEORIES: SWITCH PRESSURES: CYLINDRICAL SECTION 2.63

The recoverable strain energy in an element of thickness dz is given by:

0 0

U A dz Vd V 2 Hd H (2.186)

V H

The recoverable parts of the energy expression given in equation 2.185 are as
follows:
d v d H d H d V
d V 2 , d H 1 (2.187a,b)
E E E E

In equation 2.187, Young’s Modulus of the stored solid, E, and Poison’s ratio, , are
assumed constant.

Substituting equations 2.187a,b into 2.186 gives the following:


z z
1 1
U A RS RS 2 K RS dz 2 K RS 1 K RS RS dz
E E
0 0

Collecting terms gives:


z
2
A R 2 2
U 1 4 K 21 K S dz (2.188)
E
0

As the switch proceeds up into the cylindrical section, it is located at some arbitrary
level, z0, as shown in figure 2.44. Above this level the stress ratio K is equal to the

Janssen stress ratio, and below this level the stress ratio varies and therefore is a
function of S.

Vertical Pressure

Static
Janssen pressures:
static K=KJanssen
pressure
z0

Switch
Cylindrical silo

pressure
Dynamic
pressures:
Depth

K varies

Figure 2.44: Location of the switch pressure during flow


2.64 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Jenike (1973b) assumes that the switch propagates slow enough so that
acceleration terms are negligible, and therefore the equation of equilibrium of a
horizontal elemental slice as shown in figure 2.43 is as follows:

1d v v
K 1 0 (2.189)
dz R

The derivative of equation 2.184 is as follows:

d v dS
R (2.190)
dz dz

And re-arranging equation 2.184 in terms of S gives: S = v/R (2.191)

Substituting equations 2.190 and 2.191 into equation 2.189 gives the following
differential equation in terms of the stress field S:

dS K 1
S (2.192)
dz R R

Kz / R
1 e
Equation 2.192 integrates to: S (2.193)
K

Substituting the expression for S given by equation 2.193, into the equation for the
strain energy, given by equation 2.188, gives the following;

1 2
U 21 4 K 21 K ...... X ......
3
K

z0 Kz0 / R 1 2 Kz0 / R
........ K 2 1 e 1 e (2.194)
R 2

Re-arranging equation 2.192 gives K as a function of the stress field S, as follows :

dS
1 R
K dz (2.195)
S

By applying variational calculus to equation 2.188 for strain energy, the minimum
energy can be obtained by letting U = 0.

h
A R2
U {2S S 4 SK S S SK 41 SK SK }dz (2.196)
E
o

Both S, (S+ S), and SK, (SK+ SK) must satisfy equation 2.195, which is the
equilibrium equation re-arranged. Substituting into equation 2.195 gives:
CLASSIC WALL PRESSURE THEORIES: SWITCH PRESSURES: CYLINDRICAL SECTION 2.65

RdS/dz + SK = Rd(S+ S)/dz + (SK+ SK) (2.197)

Cancelling out SK and RdS/dz and re-arranging equation 2.197 gives SK as follows:

R d( S)
SK (2.198)
dz

Substituting equation 2.198 into 2.196 and integrating by parts gives the following:

2 h
A R R
U 4 S 41 KS S .............
E z
0

h
R d
.......... S 2S 1 2 K 4 S 41 KS dz (2.199)
dz
z
0

There is a boundary condition on S such that at z=z0, S=SJ (the Janssen static

pressure) and therefore no variation in S is admissible.


At z=h, any value of S is possible to force ΔU to be zero for any value of S.

Therefore: 4 S 41 KS at z = h must always hold. (2.200)

Thus Jenike(1973) has set the second term in the integral sign of equation 2.199
equal to zero. Jenike re-arranges equation 2.200 to give an expression for the
stress ratio in the material as follows:

K = /(1- ) (2.201)

Jenike then states that the integrand in equation 2.199 must vanish to zero for any
admissible value of S , and since S is arbitrary, this requires that:

R d
2S 4 KS 4 S 41 KS in the range z0< z < h (2.202)
dz

Substituting for K from equation 2.195 into equation 2.202 gives the following:

1 RS R d 1 RS
2S 4 4 S 41
dz

Cancelling terms and dividing throughout by 2 gives:


2
2 R
S 21 S (2.203)
2

Jenike now introduces a new co-ordinate system such that:


z z
0
x (2.204)
R 21
2.66 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Then a general solution to equation 2.203 is as follows:


x -x
S= Ae + Be +2 / (2.205)

where A and B are constants determined from boundary conditions.

From the boundary condition at z=z0 ; S=S0 it follows that x =0 and substituting

this into equation 2.205 gives:

A + B = S0 – 2 / (2.206)

At the bottom of the silo: z=h and x=X; and from equation 2.201: K= /(1- ) gives
the following solution for A:

x 1
KM 1 S N e M KN
0
A (2.207)
x x
KM 1e KM 1e

where M = (2(1- )) and N=2 / (2.208a,b)

Substituting equation 2.207 for A into equation 2.206 gives the expression for the
constant B.

Three curves of the switch pressure envelope, as given by equation 2.205, have
been shown in dotted lines in figure 2.45 for Poisson’s ratio, 0.3. The solid
curves are the Janssen static pressures in the silo for the same H/D ratios and wall
friction angles. The first two curves were calculated for a H/D ratio of 10(D=1) and
5(D=2), both with a wall friction angle of 20 . The third curve was calculated for a
H/D ratio of 10, but varying the wall friction angle to 15 . Also shown in figure 2.45
is the Janssen horizontal static pressure distribution in dimensionless form, ie

H
/ D, for a static stress ratio of K=0.4. The values of the switch pressure at three

different levels in the silo (ie 1.5, 3.5, 5, and 7m) have been given as a multiple of
the static horizontal pressure next to the graph for all three curves.

The graph of Jenike’s upper bound of switch pressures shows that the envelope of
switch pressures tends towards an asymptote, and the greatest changes in
pressures occur in the top half of the cylindrical section. It can also be seen that the
switch pressure varies from approximately two to four times the static pressure
value. The greatest horizontal pressures acting on the wall of a silo during flow
would therefore occur on a very tall silo with a low wall friction angle .
CLASSIC WALL PRESSURE THEORIES: SWITCH PRESSURES: CYLINDRICAL SECTION 2.67

Dimensionless horizontal pressure


0 0.5 1 1.5 2
0

1
S=3.10SJH S=4.01SJH S=3.53SJH
2

3
S=2.20SJH S=2.89SJH S=2.45SJH
4

5 S=2.02SJH S=2.48SJH S=2.17SJH

7 S=1.95SJH S=2.20SJH S=2.01SJH

8
H/D=10; w=20
Depth

9
H/D=5; w=20

10 H/D=10; w=15

Figure 2.45: Jenike switch pressure envelope

As the switch reaches the top of the cylindrical section the stress field, S, tends to a
minimum. From equation 2.195 this implies that the stress ratio, K, approaches
infinity, which makes Jenike’s analysis less reliable for positions close to the top of
the silo. Jenike states that the switch pressure stops at a height approximately one
silo diameter below the top, which Jenike states has been observed experimentally.
Although the expression for the switch pressure does not give an area of influence,
Jenike states that the maximum pressure acts over an area of approximately one
third the diameter of the silo, (D/3). Jenike emphasises that the curves given in
figure 2.45 are only an upper bound to the wall pressures which can be expected
during flow of the material. Small deviations in the shape of the cylinder as well as
thin boundary layers of material on the wall, will reduce the maximum pressure.
2.68 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

2.3.1.2 Walters Switch Pressure

Walters(1973) considers an instant in time during the discharge of the silo when
the switch pressure is at a height z below the top of the silo as shown in figure
2.46.

Horizontal wall pressure

Janssen
static
pressure
1
Switch
pressure
h

2
Walters
dynamic
pressure
Depth

Figure 2.46: Location of the Switch Pressure according to Walker

Walters assumes that above the switch, at depth h, the stress field is undisturbed
and therefore the static pressure field applies. Below the switch, dynamic pressures
exist with a surcharge pressure equal to the static pressure at that point. Below the
switch the dynamic pressures are given by the differential equation as follows:

d v 4BF
v
(2.95 repeated)
dz D

where F is given by equation 2.84 and B is given by equation 2.93.

With a uniform surcharge pressure acting above the level of the switch, the limits of
integration for equation 2.95 become = J
(the Janssen static pressure) at z=z0.

The solution to equation 2.95 then becomes:

D 4BFz / D 4BFz / D
1 e e (2.209)
V 4BF J

1 4BFZ 4BFZ
In dimensionless form: S 1 e S e (2.210)
V 4BF J

where S= / D and Z=z/D as before


CLASSIC WALL PRESSURE THEORIES: SWITCH PRESSURES: CYLINDRICAL SECTION 2.69

Therefore, below the switch, at depth Z h, the pressures are as follows:

1 4BF Z h 4BF Z h
S 1 e S e (2.211)
V 4BF J

The normal wall pressures at the level of the switch are determined as follows:

SH = BF SV/Tan w = BF SV/ for dynamic pressures (2.212)

-4 K z / D
and SJH=K SJ= (1-e )/(4 ) for static pressures (2.213)

Multiplying equation 2.211 by equation 2.212 and substituting equation 2.213 for SJ

gives the following expression for the horizontal wall pressures:

1 4BF Z h BF 4BF Z h
S 1 e S e (2.214)
H 4 K JH

At the level of the switch, Z=h, and the first term in equation 2.214 becomes zero
Therefore, at the level of the switch, Walters gives the horizontal pressures as:

BF
S S (2.215)
H K JH

which is simply the ratio (BF/ K) multiplied by the static horizontal pressures.

Figure 2.47 shows the ratio (BF)/( K) for a Janssen stress ratio of 0.4, for material
friction angles varying in increments of 20 .

60
m=80

50

40
Ratio (BF)/( K)

m=60

30

20

10

0
0 10 20 30 40 50 60
Wall friction angle w)

Figure 2.47: Walters Switch Ratio for horizontal wall pressures:


for material friction angle varying from 20 to 80
2.70 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Equation 2.214 gives the height over which the switch pressure acts from the term
Z-h in the exponent. By changing the co-ordinate system as shown in figure 2.48
below, the distance over which the switch acts can be determined. The horizontal
axis is at the level of the switch and the vertical axis lies on the dynamic pressure
graph.

Horizontal wall pressure

Janssen
static
pressure

0
Switch
x pressure

Walters
dynamic
pressure
Depth

Figure 2.48: New co-ordinate system to determine height of switch

For the new co-ordinate system, as SH tends to zero the value of x can be

determined as follows:

1 4BF x BF 4BF x
0 1 e S e (2.216)
4 K JH

Multiplying equation 2.216 by 4 and re-arranging gives the following:

-4BFx
A=e (2.217)

1
BF
where A 1 S (2.218)
K JH

The solution to equation 2.217 is n (A) = -4BFx and therefore x is as follows:

x = n (A) / (-4BF) (2.219)

Equation 2.219 depends on the factors B and F, which in turn, are functions of the
material friction angle and the wall friction angle. Therefore, the depth over which
the switch pressure acts varies for each silo.
CLASSIC WALL PRESSURE THEORIES: SWITCH PRESSURES: CYLINDRICAL SECTION 2.71

Figure 2.49 shows the switch pressure at various levels in a tall silo with a wall
friction angle of 20 and a material friction angle of 40 . The silo height to diameter
ratio is 10:1. The pressures have been plotted in dimensionless form according to
equation 2.97 and 2.98 for the dynamic pressures and equation 2.6 divided by D
for the Janssen static pressure. The switch pressures have been plotted using
equations 2.214 and 2.215. The calculations were done for the switch at level 1.5,
3.5, 5 and 7m below the top of the cylinder. As can be seen from figure 2.49, the
maximum value of the switch pressure decreases from 5.2SJH to 3.2SJH as the

switch moves up the cylinder. However, the area over which it acts remains
constant, in this example the area of influence of the switch is approximately 1.0
silo diameter.

Dimensionless horizontal pressures


0 1 2 3 4 5 6
0

1
SH = 3.1 SJH
2

3
SH = 4.6 SJH
4

5 SH = 5 SJH

7 SH = 5.2 SJH
Depth

8
Static pressures (SJH)
9
Dynamic pressures
10 Switch pressures (SH)

Locus of switch

Figure 2.49: Walters Switch pressure at various levels in a silo of H/D=10

Figure 2.50 shows the switch pressure in a silo with a height to diameter ratio of 5.
The wall friction angle, the material friction angle and the levels of the switch are
the same as in the previous example . From figure 2.50 it can be seen that the area
of influence of the switch is still approximately 1 silo diameter. However, the
2.72 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

maximum values of the switch pressure have decreased by 11.5%=(5.2-4.6)/5.2 at


the lowest level to 39%=(3.08-1.87)/3.08 at the highest level.

Dimensionless horizontal pressures


0 1 2 3 4 5
0

1
SH = 1.87 SJH
2

3
SH = 3.38 SJH
4

5 SH = 4.1 SJH

7 SH = 4.6 SJH

8
Depth

Static pressures (SJH)


9
Dynamic pressures
10 Switch pressures (SH)

Locus of switch

Figure 2.50: Walters Switch pressure at various levels in a silo of H/D=5

Figures 2.51 and 2.52 show the switch in a silo with a H/D ratio of 10, as in figure
2.49. However, the material friction angle has been varied from 40 in figure 2.49,
to 25 in figure 2.51, while the wall friction angle has been varied from 20 in figure
2.49 to 15 in figure 2.52.

By reducing the material friction angle by 37.5%, the magnitude of the switch
pressures is decreased by approximately 44%. A reduction in the wall friction angle
of 25% has the effect of increasing the switch pressure by 179% at the lowest level
to 136% at the highest level in the silo.

Therefore, using the Walters equation for determining the switch pressures during
flow, the worst case pressures would occur in a silo with a high H/D (height to
diameter) ratio combined with a high material friction angle and a low wall friction
angle. A “short” silo with a rough wall would experience lower switch pressures, but
consequently may not undergo mass flow of the material.
CLASSIC WALL PRESSURE THEORIES: SWITCH PRESSURES: CYLINDRICAL SECTION 2.73

Dimensionless horizontal pressures


0 1 2 3
0

1
SH = 1.71 SJH
2

3
SH = 2.56 SJH
4
SH = 2.78 SJH
5

6
SH = 2.89 SJH
7

8 Static pressures (SJH)


Depth

Dynamic pressures
9 Switch pressures (SH)
Locus of switch
10

Figure 2.51: Walters Switch pressures in a silo of H/D=10, =30 ,


m w
=20

Dimensionless horizontal pressures


0 1 2 3 4 5 6 7 8
0

1 SH = 4.03 SJH

3 SH = 6.61 SJH

4
SH = 7.51 SJH
5

6
SH = 8.08 SJH
7

8 Static pressures (SJH)


Depth

Dynamic pressures
9
Switch pressures (SH)
10 Locus of switch

Figure 2.52: Walters Switch pressures in a silo of H/D=10, m


=40 , w
=15
2.74 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

2.3.2 HOPPER SECTION

2.3.2.1 Jenike Switch Pressures in the Hopper

The material starts to flow at the instant the discharge gate of the silo is opened.
vertical support of the solids has been removed and the material above the outlet
starts to expand vertically downwards. This reduces the vertical pressure within
the material and causes a change from a static to a dynamic stress field. The major
principal stresses now arch across the outlet of the silo. As more material expands
the region of flow extends upwards into the hopper and the switch travels upwards.
Jenike(1969) states that this change in stress fields results in a deficiency in the
wall support during flow. Figure 2.53 shows an instant when the switch is at level z
above the vertex of the hopper. Above the switch the material is still in the static
state while below the switch the material is in the passive state of stress. The
shaded volume of solid between the two stress fields does not belong to either, but
is in transition from an active to a passive state. The area under the dynamic
pressure curve represents the total weight of the solid which has not changed
significantly compared to the curve for the static pressures. Therefore, there is a
deficiency in the wall support as shown by the shaded area between the static and
dynamic pressure curves. Equilibrium is maintained by a switch pressure which is
exerted on the walls of the hopper and travels upwards from the discharge gate to
the transition, where it can become locked in position, or move up into the
cylindrical section of the silo. The force of the switch pressure is equal to the
shaded area under the pressure curve.

N t Pressure

J
Static pressures
Dynamic pressures
Switch pressure

Deficiency in
Depth

z wall support

Figure 2.53: Deficient wall support in the hopper during flow


CLASSIC WALL PRESSURE THEORIES: SWITCH PRESSURES: HOPPER SECTION 2.75

Since both the static and dynamic curves are known, Jenike (1968) states that the
magnitude of the concentrated force needs to be superimposed over the dynamic
pressures to obtain the envelope of the design pressures for flow conditions.

Jenike(1969) assumes that the switch pressure has a triangular distribution acting
normal to the hopper wall over a depth of 0.3D parallel to the hopper wall as shown
in figure 2.54. During flow, the mass of the material in the hopper remains constant
for both static and dynamic conditions. When the switch is located at the transition,
the deficiency in wall support is due to the difference between the static hopper
pressures, t
, and the dynamic hopper pressures, N
. This difference in pressure

( t- N
) is also equal to the difference between the initial surcharge acting on the

hopper given by the Janssen pressure, J


, and the radial flow pressure, , as

shown in figure 2.53.

J
t S
Pressure

SZ
dz

SZ
0.3D

zs h
zb
Depth

Figure 2.54: Switch pressures acting on the hopper wall.

The vertical components of the switch pressure, SZ


and the shear stress, SZ
at

level zs from the vertex of the hopper, acting on an elemental slice of thickness dz,

given in figure 2.54, are as follows:

Sin and SZ
Cos (2.220)
SZ

In figure 2.54 the change in pressure during flow varies from 0 at level zb to a

value of S
, at level h. Therefore, SZ
at a level, zs, can be given by:

zs zb
SZ S
(2.221)
h zb
2.76 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Therefore, the additional force due to the switch acting normal to the wall can be
given by:
h h
dz dz
J
Ah SZ
Sin P SZ
Cos P (2.222)
Cos Cos
zb zb

where Ah is the area of the hopper at the level of the transition and is given by:
2 2 2
Ah = r = h Tan (2.223)

and P is the perimeter of the elemental slice given by:


P = 2 r = 2 z Tan (2.224)

Note that in the integration of the area under the curve, Jenike has ignored the
error as shown by the red shaded area in figure 2.5, due to the approximation of
the area as a right angle triangle.

The shear stress along the wall is given by SZ


= SZ
Tan w
(2.225)

Substituting for SZ
from equation 2.221, SZ
from equation 2.225, and for Ah and

P given by equations 2.223 and 2.224, into equation 2.222 gives the following:

h
2 2
J
h Tan 2 Tan Tan Tan w s
z s2 z s z b dz (2.226)
h zb
zb

In equation 2.226, h-zb= 0.3D Cos . Therefore, integrating equation 2.226 and

cancelling out Tan , gives:

2 h
1 1 3 1 2
J
h Tan 2 Tan Tan w S
z z z
0.3 D Cos 3 s 2 s b z
b

s 1 2 zb 2
2 Tan Tan w h 2h zb 3 (2.227)
0.3 D Cos 6 2
h

Equation 2.227 can be simplified to give:

s
zb 2
J
Tan Tan Tan w 2h zb 3 (2.228)
0.9 D Cos 2
h

Solving for the additional pressure at the transition due to the switch, S
, gives:
CLASSIC WALL PRESSURE THEORIES: SWITCH PRESSURES: HOPPER SECTION 2.77

0.9D Cos J
Tan
s
(2.229)
zb 2
Tan Tan w 2h zb 3
2
h

where h is the height of the hopper from the vertex (not the gate), and zb is the

bottom of the area of influence of the switch, refer figure 2.54, and zb =h-0.3DCos

Therefore, all the variables in equation 2.229 are known and S


can be calculated.

Jenike then adds this additional pressure at the transition to the dynamic static
pressure in the hopper at the level of the transition as follows:

= t+ S
(2.230)

where t is given by equation 2.53 in chapter 2 2.1.2.3.1 and S


is given by

equation 2.229 above.

The shape of the curve for the switch pressure acting at the transition between the
cylindrical and hopper sections, is as is shown in figure 2.54.
2.78 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

2.3.2.2 Walters Switch Pressures in the Hopper.

In his derivation of the switch pressures in the hopper, Walters(1972) adopts the
same approach as for the cylindrical section given in chapter 2.3.1.2. Figure 2.55
shows the lines of the major principle stresses above the switch and below the
switch during flow of the material when the discharge gate of the silo is opened.

Static
pressures
zs

Switch

Dynamic
pressures

Figure 2.55: Switch pressure at depth z


from the transition.

Therefore, below the switch, the dynamic pressures are given by equation 2.136 for
a hopper with a cylindrical section above, resulting in a surcharge pressure acting
at the level of the transition.

M 1 M
1 2ZTan 1 2ZTan 1 2ZTan
Sv 1 S0 (2.136 repeated)
2 Tan M 1 1 2Z 0 Tan 1 2Z 0 Tan

2 E FH
where M FH 1 (2.133 repeated)
Tan

Cos 1 Sin2 m 2 Sin2 m Sin2


FH (2.128 repeated)
Cos 1 Sin2 m 2 y Sin m

Sin mSin 2 2 D
E (2.105 repeated)
1 Sin m Cos 2 2 D

In equation 2.136, Z0=Zs the level of the switch, and the variable Z now starts from

below the switch. (Note Z=z/D). S0 in equation 2.136 is the Janssen vertical

pressure in dimensionless form, ie SJ= v/ D. However, to determine the dynamic

vertical pressure below the switch, at depth Zs, the value of S0 should be replaced

with the static value of the vertical pressure in the hopper, Sv.
CLASSIC WALL PRESSURE THEORIES: SWITCH PRESSURES: HOPPER SECTION 2.79

Walters(1972) gives the expression for the dimensionless normal wall stress as a
function of the dimensionless average wall stress as follows:

SN=EFHSv/Tan w
(2.231)

This was determined from the general expression (in dimensionless form):

T=Tan w
SN , where T is the dimensionless wall shear stress = / D
w

Therefore, SN= / D = T/Tan


vw w
(2.232)

In chapter 2.2.2.1 it was shown that Walters(1972b) relates the wall shear stress

vw
to the vertical pressure at the wall, , by equation 2.103. Substituting
vw

equation 2.103, in dimensionless form, into equation 2.232 gives:

SN= ESvw/Tan w
(2.233)

where Svw is the dimensionless form of the vertical wall pressure acting at the wall.

Walters then relates vw


to the average vertical pressure acting across the hopper

slice by equation 2.127. Substituting the dimensionless form of equation 2.127 into
equation 2.233 results in equation 2.231

Substituting equation 2.231 into equation 2.136 gives the dimensionless normal
pressure during flow of the material as follows:

M 1 M
EFH (1 2ZTan ) 1 2ZTan EFH 1 2ZTan
SN Sv 1 Sv
2 Tan Tan w M 1 1 2Z 0 Tan Tan w 1 2Z 0 Tan

(2.234)
The envelope of the switch pressure is found by letting Z=Z0 in equation 2.234.

Since the term (1-2ZTan )/(1-2Z0Tan ) becomes equal to 1, the first term equals

zero, and equation 2.234 becomes:

SN=EFHSv/Tan w
(2.235)

where Sv is the dimensionless form of the static vertical pressure in the hopper.

Substituting the dimensionless form of the Walters equation for the static normal
wall pressure, into equation 2.235 results in the following expression:

(EFH )D
SN SNS (2.236)
(EFH )S
2.80 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

where (EFH)D is the dynamic value of the constants E and FH ; and (EFH)S is the

static value of the constants E and FH. This ratio is as follows:

2 2 2
Sin Sin 2 2 1 Sin m
Cos 2 Sin m
Sin
m D
1 Sin m
Cos 2 2 D Cos 1 Sin
2
2ySin
m m

(2.237)
2 2 2
Sin Sin 2 2 1 Sin m
Cos 2 Sin m
Sin
m s
1 Sin m
Cos 2 2 s Cos 1 Sin
2
2ySin
m m

Since the variables in equation 2.237 are all constant for a given hopper geometry,
the ratio (EFH)D/(EFH)S for the envelope of the switch pressures is also a constant.

The shape of the switch pressure envelope is therefore dependant on the shape of
the static pressure curve.
In his paper, Walters substitutes the following expression for 2 +2 in equation
2.237:

2 +2 = /2 + ArcCos(Sin /Sin m) (2.112 repeated)

where the +ve sign refers to static conditions and the –ve sign refers to static
conditions.
The value of can be determined as shown in figure 2.35 of chapter 2.2.1.1.
Walters states that the ratio given in equation 2.237 gives a value of 3.23 for a
material friction angle, m=50 , a wall friction angle, w=25 and a hopper half

angle, =4 . However, this was checked on a spreadsheet, where the value of


was solved by trial and error, and found to be 28.279 for dynamic conditions and
1.582 for static conditions. These values of in equation 2.237 give the following
numerical values:
(FH)D=1.620 ; (E)D=1.012 and (FH)S=1.000 ; (E)S=0.004

Substituting these values in equation 2.237 gives a switch pressure ratio of 447.8,
which is excessively large. The spreadsheet was checked several times by
comparison with hand calculations and found to be correct. It is therefore
suggested in this thesis that the switch pressure ratio as given by Walters in
equation 2.237 is not reasonable.
CLASSIC WALL PRESSURE THEORIES: SWITCH PRESSURES: HOPPER SECTION 2.81
WALL PRESSURE MEASUREMENTS: LITERATURE SURVEY: STATIC PRESSURES 3.1

CHAPTER 3
WALL PRESSURE MEASUREMENTS

3.1 LITERATURE SURVEY

3.1.1 STATIC PRESSURES

3.1.1.1 Cylindrical Section

Bishara et al (1981) undertook finite element (FE) analyses of the pressures in the
cylindrical section of a silo of 7.3m internal diameter and 24.4m tall (H/D=3.3). The
material used in their simulations was a granular cohesionless sand. They state that
the FE horizontal pressures were shown to be 10% larger than the calculated
Janssen horizontal pressure. However, on closer inspection of their graphs for the
horizontal pressure, the amount by which the FE solution is larger than the Janssen
formula, varies from 28% at 1.2 diameters from the top, to 11% at 2.6 diameters
from the top. This is higher than they have reported.
From the graphs of their results for the vertical pressures, the finite element
solution gives values higher than Janssen for a depth of 0 to 1.4 diameters, while
below this level the Janssen formula gives higher values. At full depth, the Janssen
formula gives a 26% higher vertical pressure than the finite element solution while
at a depth of 0.9 diameters, the Janssen formula gives a 14.6% lower vertical
pressure than the FE solution. They reported that the distribution of the vertical
pressure across a horizontal plane was about 50% higher in the centre of the silo
than in the vicinity of the walls.
The stress ratio Bishara et al (1981) used in their calculations was not given, so it
has been interpolated from their graphs for the purposes of this thesis. The stress
ratio used in their calculation of the Janssen horizontal pressure, was found to be
approximately 0,4 taken from three points: (9m depth: 47/120=0.39, 14m depth:
43/106=0.41, 19m depth: 33/84=0.39). The graphs from their finite element
results suggest a stress ratio of approximately 0.51 taken from the same three
points. (53/103=0.51, 50/97=0.51, 46/87=0.53)

Suzuki et al (1985) conducted tests on small and medium sized model silos. The
dimensions were 0.3m internal diameter by 1.7m tall for the small model
(H/D=5.7), and 1.4m internal diameter by 6.4m tall for the medium sized model
(H/D=4.6). The test material used in the small model was Milo(™) and in the
medium model tests were done using Milo(™), maize, soybean meal and alfalfa
3.2 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

meal pellets were used. They report that their test results from both models
compare favourably with the Janssen horizontal pressure, particularly at depths of
less than half the model diameter (ie in the upper part of the silo). From their
graphs of the measured horizontal pressure in the smaller model, their plotted test
results are 6% , 25% and 37% less than the Janssen equation.
As part of their experiments, Suzuki et al also measured the vertical pressure at six
points across the bottom of the model for various levels of fill. Their graph has been
reproduced in this thesis by scaling the points off their report and replotting them,
as shown in figure 3.1. The average values of the vertical pressure are shown in
dotted lines in figure 3.1.
5
As can be seen from this graph,
Suzuki et al show that the central
Vertical pressure on model bottom

4 80kg fill
3.8 vertical pressure is the highest, and
3.4 60kg fill
rapidly diminishes to a minimum
3 within one third of the radius from
2.9 40kg fill
the centre. The central pressure is
2.1
2 20kg fill approximately 26% greater than
the average, while the minimum is
1 approximately 15% less than the
average.

0 Their measured value of the stress


0 0.25 0.5 0.75 1
ratio which they calculated from the
Distance from centre to wall: r/ro
measured average vertical stress
Figure 3.1: Radial Distribution of
and the measured horizontal stress
Vertical pressure taken from Suzuki et
was K= 0.51.

Blight et al (1989) conducted a set of tests on two identical full-scale silos


containing cement. The silos were 20m internal diameter and 65m overall height
and were strain gauged across the height with six temperature compensated
gauges. For the calculation of emptying pressures, a stress ratio of 1 was used to
calculate the horizontal pressures. Their internal angle of friction was measured at
42 , (which is higher than the design values of 28 ). They measured the stress ratio
as being 0.35 to 0.37 while the calculated value of the at-rest stress ratio (given by
Ko=1-Sin ) was 0.29. They showed that the measured pressures were within the

envelope of Janssen’s pressures using a Ko ratio of 0,35. However, the test results
also showed the measured pressures in areas of low overburden were larger than
the calculated values. This is equivalent to areas close to the top of the silo.
WALL PRESSURE MEASUREMENTS: LITERATURE SURVEY: STATIC PRESSURES 3.3

Molenda et al (1993) have conducted experiments to determine the effects of filling


method on the wall loads. They tested concentric, eccentric and uniform sprinkle
filling methods using soft and hard wheats. From their experiments they found that
the grains aligned themselves parallel to the free surface during filling. For the
concentric filling methods, the bulk density of the material after filling was
approximately 6% lower than for the cases where the grains were uniformly
sprinkled in the silo.

It is well established that the Janssen equation for the static vertical pressure in the
silo, gives a good initial estimate of the minimum loads to be expected in the
cylindrical section of the silo. The vertical pressure in the centre of the cylindrical
section has been shown to be greater than the average value by approximately
26% to 50%, while Janssen only found it to be 15% greater. For the purposes of
this thesis, a value of 30% greater than average will be used for the vertical
pressure in the centre of the silo (where 30% is the average of the three values
quoted). The shape of the individual particles affects the bulk density of the
material in the silo due to the filling method employed. This effect does not show
up in the finite element analyses, as can be seen by the average vertical pressures
being higher than those calculated from the Janssen equation.

3.1.1.2 Hopper section

Blair-Fish and Bransby (1973) conducted tests on a model silo 0.15m square and
0.45m in height, with a hopper half angle which can vary between 30 and 20 ,
filled with dry sand. From their report, the normal pressure on the hopper wall is
approximately constant throughout the depth of the hopper.

Van Zanten and Mooij (1977) conducted tests on a model silo, 1.5m in diameter
and 6m tall, fitted with a hopper half angle of 15 . Two types of fill materials were
used, viz PVC powder and sand. The graph of their results shows a large scatter of
data, which has the average minimum value in the lower portion of the hopper and
the maximum occuring at approximately half the hopper height. At the transition,
the pressure in the hopper is shown to be approximately three times greater than
the pressure in the cylinder. In their graph, Van Zanten and Mooij also show the
calculated line of pressure in the cone according to Jenike. Of their nearly 50 data
points for normal wall pressure in the hopper, only the four maximum points fall
outside the limit given by Jenike.
3.4 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Suzuki et al (1985) measured the wall pressures in the hoppers of two mass flow
silos filled with milo. Their results show the pressures in the hopper to be the
greatest at the base decreasing towards the transition between hopper and

Dimensionless parameter P/Pmax


0 0.25 0.5 0.75 1
parameter z/zh
Dimensionless

0
b
0.25
a
0.5
0.75 position of
1 pressure cells

Figure 3.2: Distribution of normal pressure on hopper wall


(a) no surcharge, (b) silo fully loaded

cylindrical section. Their results show the static pressure in the hopper at the
transition is approximately 2.3 times greater than in the cylindrical section. A
diagrammatic representation of the pressure distribution in the hopper has been
taken from the graph of their experimental results and shown in figure 3.2. In this
thesis, the vertical axis is the ratio below the transition, z, to the total depth of the
hopper,zh. The horizontal axis is given as the ratio of the pressure, P, to the
maximum pressure, Pmax , at the base of the hopper when the silo is fully loaded.
From figure 3.2 it can be seen that the pressure distribution varies from nearly
linear for no surcharge in the hopper, to a curved distribution for a fully loaded silo,
with a minimum at approximately one third the depth.

Kmita (1991) gives a very different pressure distribution for the normal pressure on
the hopper wall. The tests were conducted on a plane flow silo 0.8m wide and
3.6m overall height, filled with rinsed grit of particle diameter ranging between
3mm and 5mm. The silo has a hopper half angle of 15 and hopper height of 1.2m.
Kmita conducted tests for the case of the silo being filled from empty as well as
partially emptied and then re-filled to the same height. The pattern of the pressure
distribution is shown to be the opposite of that given by Suzuki et al in figure 3.2,
with a maximum occurring at approximately one fifth the depth of the hopper below
the transition. There is no difference in the pressure distribution for the partial
emptying and re-filling case compared to filling from completely empty. Kmita
shows the maximum to be five times greater than the maximum wall pressure in
the vertical section of the model.

It is clear from the varied results for the pressures in the hopper, that a large
variety of factors influence these pressures and the exact pressure distribution
WALL PRESSURE MEASUREMENTS: LITERATURE SURVEY: STATIC PRESSURES 3.5

cannot be predicted. However, all the research reports studied indicate the static
pressures generally do not exceed those given by Jenike, which is therefore a good
initial assessment of the maximum static pressures to be expected in the hopper.

3.1.2 DYNAMIC PRESSURES

3.1.2.1 Cylindrical Section

Pieper (1969) conducted tests on three model silos filled with a quartz sand. The
two cylindrical models were 0.6m in diameter and 3m tall, and 0.8m in diameter
and 6m tall. The third model was a square silo of 0.7m cross section and 5m tall.
This gives H/D ratios of 5.1, 7.5 and 7.1 respectively. In the graph of flow results,
Pieper shows the flow pressures in all three models to be approximately 1.3 times
the static value.

Blair-Fish and Bransby (1973) conducted tests on a sand filled mass flow silo,
150mm square cross section and 375mm tall (H/D=2.5). with a 30 hopper half
angle. They presented their results for the measured pressures in the form of a bar
graph at the point measured and the flow results were presented for each
increment of emptying

Their flow results have been


Ratio of flow/static pressure
0 1 2 3 scaled and given as a ratio of
the static pressure at the point,
as shown in figure 3.3. This
graph shows their flow results
only reached a maximum of 3
times the static pressures in the
cylinder, while in the hopper,
the maximum did not exced 2.5
times the measured static
value.

Figure 3.3: Test results scaled from


Blair-Fish and Bransby

Richards ( 1977) conducted experiments on a model silo 0.6m in diameter and 1m


in height with a 15 hopper half angle, filled with sand (wet sand to study minimum
opening dimensions and dry sand to study flow pressures). Richards reports that as
soon as the gate was opened only slightly, the normal wall pressures just below the
3.6 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

transition increased substantially above the static pressure, although the maximum
pressure did not always occur at the start of flow when the silo was full. The flow
rate was varied by a factor of 6 in these experiments and had no effect on the
measured wall pressures. When the flow was stopped, the overpressures remained
at the points where they were recorded.

In the tests conducted by Van Zanten and Mooij (1977) on model silos filled with
PVC powder and sand, they considered flow in a “perfect” silo as well as flow in a
silo with irregularities on the walls. The following discussion does not include the
measured pressures at the points of irregularities in their model. The geometric
aspect ratio of their model silo is shown in figure 3.4 below. First they measured
the circumferential distribution of the overpressure at four points, 90 apart, at the
transition during flow. They report that the distribution was highly assymmetrical
and “simultaneous peak pressures at two points occurred only occasionally”.

For the discussion in this thesis, their test results of


the vertical distribution of the overpressures for sand,
have been divided into two sections. Those between
4D

the transition and 1.5D above the transition, and


those at levels greater than 1.5D above the transition.
1.5 D

In the lower section of the cylinder, the flow pressures


are greater than the pressure envelope given by
Jenike Strain Energy. Their test results show these
1.9 D

pressures to be 5.8 times the maximum Janssen static


value and twice the maximum Jenike Strain Energy
value. In the upper portion of the cylinder, their test
Figure 3.4:
Geometric aspect of results are 4.3 times the Janssen static pressure, but
the model silo are within the envelope given by Jenike Strain Energy.

At the transition, the maximum pressures are 15% less than the value given by
hydrostatic pressure at a depth of 4D, and 39% less than the Jenike Peak Pressure.
At a depth of 0.7D below the transition, the maximum test results are 1.4 times
greater than the Jenike static value in the hopper. In the hopper, nearly all the test
results lie outside the line of the Jenike peak Pressure. (Note: In their report, the
Jenike peak pressures are smaller than the Jenike static value in the hopper)

For their tests with PVC powder, the same division in their results has been made in
this thesis. In the lower section of the cylinder, the pressures are approximately 3.7
times the maximum Janssen static value and twice the value given by Jenike Strain
WALL PRESSURE MEASUREMENTS: LITERATURE SURVEY: STATIC PRESSURES 3.7

Energy. In the upper portion, the results are twice the Janssen static value, and are
also within the envelope given by Jenike strain energy. At the level of the
transition, nearly half the test results lie beyond the maximum value given by
hydrostatic pressure. In the hopper, the results again lie beyond the Jenike peak
pressure envelope. At a depth of 0.7D below the transition the test results are 1.2
times the Jenike static value in the hopper.

Nielsen and Andersen (1982) conducted tests on full scale silos, 7m in diameter and
46m tall (H/D=6.6), filled with barley. They conducted several tests for various
arrangements of filling and emptying but only their tests for central emptying have
been considered in this thesis. The
Flow/Static
pressure cells were placed at four points,
0 1 2 3
7m
90 apart, around the circumference, and
at seven different vertical levels, giving a
Depth below surface

11 total of 28 pressure cells. From the

18 results of their filling and emptying tests,


46m

the ratios of flow vs static pressures have


25
been plotted in figure 3.5. These results
32
show that in the upper half of the silo, the
39 flow pressures are only 1.5 times the
static pressures whereas in lower portion,
Figure 3.5: Ratio of flow pressures
the ratio is approximately 1,5 to 3.
to static pressures taken from
Nielsen and Andersen’s test results

Rombach and Eibl (1995), conducted finite element tests on material flow in the
hopper. The hopper half angle in their model was 20 , and the wall friction angle
was 21.8 . Their results show the dynamic pressure, 0.2 seconds after flow was
initiated, was 1.4 times the static pressure.
3.8 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

3.1.3 STRESS RATIOS

In his derivation of the vertical pressure in a silo, Janssen assumes a constant value
of 0.4 for the stress ratio. Jenike and Johanson(1969) recommend a constant value,
assuming Poisson’s ratio for the material does not vary and the material remains
isotropic. They recommend a minimum value of 0.4 in the cylinder and 0.8 in the
hopper.

In their finite element analysis of the storage and flow pressures in silos, Bishara et
al (1981) give the following expressions for the horizontal and vertical pressures on
the walls, immediately after filling:

0.33
D 0.75 H 0.55 0.62
h 0.46 and v 1.82 D H
0.25 0.25

Since the stress ratio is defined as : K= h/ v : an equation for K can be found by

substituting in their expressions for h and v to give the following:

D 0.2 H0.38 0.08


K 0.253 (3.1)
0.58

Equation (3.1) implies that the stress ratio is independent of the internal friction
angle of the material. It would be reasonable to expect the stress ratio to have a
term for the internal friction angle in the expression.

Furthermore, substituting for (D, H, and ) the values 1m, 3m, 0.4, and 16kN/m3
respectively in equation (3.1), gives a value of K = 0.071. Irrespective of the
material used (whether cohesive or free flowing), this value is too low to give
reasonable results.

1m: K=0.4
Ravenet (1983) reports of tests done by the
5m: K=0.67 Reimbert brothers in 1943 on full scale, flat
bottomed silos in France which were exhibiting
10m: K=0.81
signs of being overstressed. According to Ravenet,
15m: K=0.89 the Reimbert brother’s strain gauged these silos to
measure the horizontal and vertical pressures on
Figure 3.6: Stress the walls. From the graph of Reimbert’s results,
ratios in silo from
Reimbert’s tests given in Ravenet’s report, the stress ratio at various
WALL PRESSURE MEASUREMENTS: LITERATURE SURVEY: STATIC PRESSURES 3.9

points along the height of the silo has been determined by scaling off the graph.
These values are shown in figure 3.6. These results show the stress ratio increases
with increasing depth in the silo.

Briassoulis (1991) derives an expression for the stress ratio in the Reimbert
formula from the ratio of the horizontal stress to the vertical stress, and gives this
as follows:

2 y /p
K(y) Ka (3.2)
1 y /p

In the above expression, Ka is the active stress ratio, p = r/ (2 K a ), and y is the


depth below the surface of fill. Briassoulis states that the Reimberts assume the
stress ratio, K, to decrease with depth, which is shown in equation 3.2. However,
the results given in figure 3.6 from Ravenet’s (1983) report of the Reimberts’ tests,
show the stress ratio to increase with depth.

Referring to the Mohr circle for stresses at the hopper wall, as shown in figure 3.7,
an expression for the stress ratio at the hopper wall under static conditions can be
derived as follows:

Wall yield
locus

2
W H V

O 3 C 1

Op

plane on which the


plane on which the vertical stress acts
horizontal stress
acts
Figure 3.7: Mohr circle for stresses at the hopper wall under static conditions.

OC=O H+ HC (3.3)
3.10 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

But OC= r / Sin W and H C= r Sin( W+2 ) ( 3.4a&b)

Substituting these values into equation (3.3) gives the expression for the horizontal
stress as follows:

1 Sin W
2 Sin W
H r (3.5)
Sin W

Similarly, the expression for the vertical stress is given as follows:

1 Sin W
2 Sin W
V r (3.6)
Sin W

Therefor the expression for the stress ratio in the hopper becomes:

1 Sin W
2 Sin W
Kh (3.7)
1 Sin W
2 Sin W

As the hopper half angle varies to a minimum, in the limit, this would give the
expression for the stress ratio at the wall in the cylindrical section of the silo as:

K=(1-Sin2 w)/(1+Sin
2
w) (3.8)

These expressions (3.7 and 3.8) imply the stress ratio is a constant value and not
dependant on the level of overburden material in the silo.
WALL PRESSURE MEASUREMENTS: EXPERIMENTAL SET-UP 3.11

3.2 EXPERIMENTAL SET-UP

3.2.1 STEEL MODEL SILO

Nielsen and Askegaard (1977) studied the effects of model scale on the results of
pressure measurements. Their experiments were done in a centrifuge, using a
sample of 40mm in diameter and 150mm in height (H/D ratio of 3.75), filled with a
cohesive material (silica gel) and dry sand. They concluded that test results on
models filled with a cohesionless material can be transferred to a geometrically
similar full scale silo, provided the model is “not too small”.They do not give a
definition of what is considered not too small. However, they stated that it was not
necessary to test cohesive materials in a centrifuge, if the model diameter was at
least 12 times larger than their model. This value of 480mm (=12x40mm) was
therefore used as an indication of their description “not too small”, for the purposes
of this this research.

The available sections in the laboratory used to set up the model, were just more
than twice the minimum requirement, defining it as a large model. Therefore, it has
been assumed in this thesis that there should be no scaling errors applicable to the
test results.

A steel model silo of height 3.21m,


0.98m in diameter (H/D=3.28) and
outlet opening of 0.18m in diameter, was
set up in the laboratory as shown in
figure 3.8. The cylindrical section of the
BUCKET ELEVATOR

980 model was made up of four equal semi-


2400

circular sections, 1.2m in length. The


sections were bolted together through
the outside flanges, so that there were
no obstructions to the material flow. The

25 hopper was made up in two halves, and


810

fitted to the cylindrical section through


matching flanges at the transition. A
180 hopper half angle of 25 was chosen from
Jenike (1967) as the maximum angle
which would still cause mass flow of the

Figure 3.8 : Steel Model Silo material. The silo was filled by means of
a bucket elevator.
3.12 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

3.2.2 BULK SOLID MATERIAL

The pressures exerted on the silo walls during flow are affected by obvious criteria
such as wall friction angle, hopper slope, shape and the number of outlets in the
hopper. The less obvious influences are the material properties such as
cohesiveness, the ability to segregate and the degree of segregation due to the
filling technique. Arnold (1991) states that particle segregation influences the flow
pattern in the silo and hence the wall pressures during flow as well as during
storage. Coarser materials have a better degree of flowability than finer material,
as well as a lower coefficient of wall friction. Based on this, a uniformly graded, dry
cohesionless sand of particle diameter between 0.8mm and 1.6mm was chosen as
the fill material in the experiments. The internal friction angle of the material was
determined in a triaxial cell and was found to be 45 . The wall friction angle
between the material and the silo (also determined in a triaxial cell), was found to
be 22 . The material density was determined in the laboratory by allowing a sample
of the material to fall from a height of 1m into a container of known volume and
self weight. Two different methods of filling the container were used to simulate the
filling operations of the model. The material was allowed to rain into the container
and also to fall in a constant stream. There was very little difference between the
density as determined from the two different filling methods and an average
density of 16 kN/m3 was used.

3.2.3 DATA ACQUISITION

Data acquisition was obtained from a 16 channel AD card (Metrobyte Dash16)


driven by a fortran computer code (K.Kavanagh 1986). The original code was
written for a 2 cycle internal timing clock, which had to be hand calibrated in these
experiments for the a/d card with a single cycle clock. The program was structured
around background filling and emptying of 512 word buffers. Calculations were
post-processed from data transferred to hard disc. Maximum data rates of 25kHz
were obtainable without data interuption due to disc transfers.

The input file has the four entries: number of channels, number of combinations,
number of buffers and the input frequency. Therefore, if eight different types of
pressure cells were used, the number of channels would be eight. As there were no
combinations of pressure cells, this entry was entered as zero. The number of
buffers and the number of channels entered in the input file, affected the time
WALL PRESSURE MEASUREMENTS: EXPERIMENTAL SET-UP 3.13

taken for data acquisition and size of the output data file. The output data file had
the form of columns and rows, where the columns represented the individual
channels and the rows represented the next data point for each channel.

As the number of data points in the output file was used as a time reference, the
exact sampling frequency of the single cycle clock in the computer, had to be
determined. This hand calibration was done by timing a series of blank runs for an
input file of 1 channel, 0 combinations, 5
buffers and input frequency. The input
800
frequency was varied from 500 to 8000 Hz

600 in steps of 500. The number of 5 buffers was


Actual frequency

chosen so that at higher input frequencies,


400 the time could be measured with a stop
watch. If too few buffers were used, the test
200 duration was too short to enable timing on a
stop watch. For each input frequency, ten
0 tests were run and the average was taken as
2000

4000

6000

8000
0

the sampling time for that input frequency.


Input frequency The results of these tests are shown in figure
3.9. The slope of this graph gives a factor of
Figure 3.9: Actual frequency of 0.095369 applied to the sampling frequency
computer data acquisition
in the input file.

The total time, in seconds, taken for the computer to record data, was required to
ensure data acquisition was not cut short before the end of a test, resulting in a
loss of results. The computer’s total sampling time was found to be given by:

Sampling Duration=( 512 NB)/ (F i 0.095369) (3.9)

The time between rows of data, tr , was determined from the following equation:

tr = N C /( F i 0.095369) (3.10)

where NB is the number of buffers, NC is the number of channels and Fi is the input
frequency.

The size of the output data file was determined from:

Number of data rows = 512 NB/NC (3.11)


3.14 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

3.2.4 STRAIN GAUGE BRIDGES

Figure 3.10: Location of Strain


Gauge Bridges

From the literature on wall pressures in silos it was recognised that the pressures at
the transiton during flow, were likely to be the greatest. Therefore, the first attempt
at measuring the pressures was concentrated in the area just above and below the
transition, in three groups of four points around the circumference of the model, as
shown in figure 3.10. The concept was to make small holes in the silo walls and to
place an instrumented bridge across each opening. The strain in each bridge was
recorded as a measurement of the applied pressure during flow of the material.

The holes were made in the silo walls by cutting 50mm disks at each point shown in
figure 3.10. These disks were then fixed to the middle leg of the bridge as shown in
figure 3.11. Due to the fact that the disks were curved in one direction only, the
bridges were placed on the silo wall so that the curvatures of the disks and silo
were aligned. The holes were covered with a clear plastic sheeting on the inside of
the silo wall. The dimensions of the aluminium bridge is shown in figure 3.11, with
two strain gauges glued to the top and bottom of the bridge. The strain gauges
used were Kyowa type KFG-5-120-C1-11, with a temperature compenstion for steel
and a gauge length of 5mm.

After the gauges were connected in series to give an average reading of deflection,
they were tested for specified resistance to ensure no gauges had become damaged
in the soldering process. Tape was placed over the gauges to reduce the risk of
damage during handling.
WALL PRESSURE MEASUREMENTS: EXPERIMENTAL SET-UP 3.15

10mm
50mm disk
cut out of
silo wall 40mm

10mm
Strain gauges

40mm
20mm 2mm

Aluminium
bridge 10mm

Figure 3.11: Aluminium strain gauge bridge

Each bridge was loaded and unloaded three times to ensure the gauges had
adhered to the bridge before they were calibrated. The gauges were then connected
to the amplifier, and the bridges loaded up to 5 kilograms in increments of 1
kilogram. Readings were taken after each increment. The bridges were then
unloaded to 3 kilograms, a reading taken, and allowed to stand for 30 minutes after
which another reading was taken. This was done to check for drift in the gauges.
The results for each bridge were plotted on a graph of load versus voltage output
and the slope of the graph gave the calibration factor for each bridge.

Static readings were taken after the silo had been completely filled. The silo was
emptied in stages into drums which were placed under the hopper. Emptying
stopped when the drum was full and reloaded into the silo. Readings were taken for
3.16 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

each stage of emptying. The results from the bridges did not record any strain for
the static pressures or for the first stage of emptying, the output being attributed to
electrical noise. The likely cause was that the bridges were too flexible relative to
the model wall so that the material arched over the bridges. One bridge became
dislodged during testing, and the plastic sheeting managed to hold the material in
place without rupturing. This lockup of material is a strong indication that arching
will occur over small diameters, where the opening is softer than the bin wall. This
phenomenon may be more critical, the smaller the diameter of the opening.

Due to the failure of these experiments, a radically different pressure measuring


cell was developed. Rather than being mounted to the silo wall, the cells were
placed in the material during filling, and the cells travel with the material during
flow.

3.2.5 PRESSURE CELLS

Richards (1977) reports of vibrations being felt and heard during discharge from a
mass flow model. The frequencies of the individual fluctuations were reported to be
in the range of 15Hz to 85Hz, the frequency increasing with flow rate and being
independent of particle size. Therefore, a responsive measurement and fast
recording system was needed to investigate the material pressures during flow.

Three types of material pressure cells, as shown in figure 3.12, were developed in
the laboratory to allow continuous measurement of the pressures during filling and
emptying of the silo contents. The concept of the cells was to place a standard
pressure sensor in a small hollow object with a flexible wall, which could then be
filled with an incompressible, low viscosity oil. The most important criteria was to
ensure that there were no air bubbles retained in the oil or pressure sensors to
affect the incompressibilty.

(a) (c)

Y (b)

Z X
Z

Figure 3.12: Floating Pressure Cells: a) Ball type, b) Tube type, c) Plate type
WALL PRESSURE MEASUREMENTS: EXPERIMENTAL SET-UP 3.17

3.2.5.1 The Ball Type Pressure Cell

The first type of cell was the ball cell, where a pressure sensor (psi absolute) of
capacity 15psi, was modified to fit inside a hollow thin walled rubber sphere, as
shown in figure 3.13. The front ports and the side connection holes were clipped
off, leaving only the shell of the sensor which was then fitted in the sphere. A large
container filled with silicon oil was placed inside a vaccuum chamber to remove any
air bubbles in the oil. The sphere and sensor were then placed in the vaccuum
chamber and the air bubbles removed. The opening in the sphere, through which
the sensor was inserted, was sealed off with a glue suitable for elastomeric
materials. Removing the air bubbles from the inside of the sensor proved to be very
difficult, and in some ball cells the prescence of an air bubble showed up during
calibration.

Electrical
Connection
pins and
holes
cables
Pressure sensor shell

Absolute port

Pressure port

Thin walled
rubber sphere

Figure 3.13: Pressure sensor modified to fit inside rubber sphere.

3.2.5.2 The Tube Type Pressure Cell

The second type of pressure cell was the tube type cell which consisted of a 14mm
diameter hollow Tygon(™) tube, 100mm long, as shown in figure 3.14. The
pressure sensor used in these cells was a standard 5psi differential pressure sensor,
which was chosen above the 15psi sensor, as the sensor’s capacity produced a
larger signal-to-noise ratio. The smaller capacity sensors were not available in the
absolute form, so that one port was blocked off from the atmosphere with a plug,
leaving the remaining port as the active port. A solid end cap to close off the end of
the tygon tube, and a plug with a central hole, were fabricated in the laboratory.
The separate components for this cell were filled with oil by placing them in the
container of oil in the vacuum chamber. Once the air bubbles were removed, the
cell was then assembled while submerged in the oil, taking care not to introduce
new air bubbles while handling the tube and components.
3.18 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

flexible tubing pushed over port


and plug, and fastened with a tie

Pressure 14mm diameter tygon


sensor
plug with
End cap
centre hole
open port
blocked off

Electric pins
and cables
Figure 3.14: Tube type pressure cell

The sensor and the plug were connected using a flexible plastic tubing which was
pushed over both the port and the plug and fixed on with a plastic tie. The plug and
end cap were pushed into the tube and also fixed in place with a plastic tie. This
system proved very successful as the air bubbles from the active port of the sensor
were easily removed. Since the Tygon(™) tube was translucent, a visual check of
the existence of air bubbles was easily made. The tube cell was left to stand
overnight in a vertical position resting on the sensor, and any trapped air bubbles
would then float up into the tygon tubing.

3.2.5.3 The Plate Type Pressure Cell

As shown in figure 3.13, the ball cell measured the average pressure acting in three
dimensions, where as the tube cell measured the average pressure acting in a
plane. The third type of pressure cell was therefore developed to measure the
average pressure acting in one direction only. This was the plate type cell which
consisted of an aluminium plate with a slot cut out the middle connected to the
pressure sensor as shown in figure 3.15.

port blocked threaded


standard pipe connection
off end cap
pressure
sensor

10mm wide slot

Electric pins
rigid plastic 100mmx30mm aluminium plate
and cables
tubing covered with rubber membrane
on both faces

Figure 3.15: Plate type pressure cell


WALL PRESSURE MEASUREMENTS: EXPERIMENTAL SET-UP 3.19

Again, a standard 5psi differential pressure sensor was used which had one port
plugged off. Two threaded holes were made at either end of the aluminium plate
extending into the slot. Both faces of the plate were sealed by a rubber membrane
which was glued on to the plate surface. One hole was closed off with a threaded
end cap, and a standard pipe fitting was used to screw into the other hole. A short
length of stiff plastic tubing was pushed into the pipe fitting and then heated up in
boiling water to fit over the active port of the sensor. The individual components of
the pressure cell were placed in the container of low viscosity oil and a vacuum
applied to remove all the air bubbles. Once the air bubbles were removed the cell
was assembled while submerged in the oil.

3.2.6 PRESSURE CELL CALIBRATION

The cells were placed in an air tight vessel with an absolute pressure sensor, first to
determine their responses to an instantaneous pressure and then to calibrate each
cell. The inlet and outlet to the vessel were sealed off with silicon sealant and
allowed to stand for five hours before applying air. Compressed air was supplied to
the vessel by means of a valve and regulator, which gave a digital read out of the
pressure supplied to the vessel. The set up is shown in figure 3.16.

compressed inlet and outlet


air supply sealed with silicon

valve and
digital
readout cables from
unit pressure
ii
iii cells

air tight container


i

iv amplifier

To computer

Figure 3.16a: Calibration of pressure cells


(i) Ball cell, (ii) Tube cell, (iii) Plate cell, (iv) Plain sensor
3.20 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Figure 3.16b: Photograph of pressure Cell Calibration

Before calibration, the cells were tested to ensure there were no air bubbles
trapped inside and also to determine their responsiveness to an instantaneous
pressure. The instantaneous pressure was applied to the cells by first regulating the
pressure to a known value with the valve closed and once the pressure had been
reached the valve was suddenly opened. The cells were then subjected to a series
of instantaneous pressures and the output recorded on the computer. When the
cells contained trapped air bubbles, the output of the cell compared to the plain
pressure sensor exhibited a lag in the peak reponse, as well as a damped response
in the decay curve of the graph. As shown in figure 3.17, the dotted shows there is
no lag in the peak response of the cells.

No lag in Cells very responsive


peak to instantaneous
response pressures

Ball cell

Plain sensor
Voltage
output

Voltage
output

Data point Data point

Figure 3.17: Typical response of a ball type cell


WALL PRESSURE MEASUREMENTS: EXPERIMENTAL SET-UP 3.21

Once the responsiveness of each cell had been determined, they were calibrated by
applying a series of pressures, in increments of one, up to thirty two kilopascals.
Before the cell output was recorded, the air pressure in the vessel was first set at a
constant for each increment, thereby ensuring good data readings. For each
increment of pressure, 170 data points were recorded. The average voltage output
for each pressure increment was plotted as a point on the graph, and a linear least
squares regression analysis applied to the data. The graph for each cell has been
shown in Appendix B, with the calibration constant.

The pressure cells were connected to the data logger by means of 6m long cables
which enabled the cells to flow freely during material discharge. The cells were
placed at various depths in the silo during the filling operation, and their positions
measured.

3.2.7 MULTI-TURN POTENTIAL METERS

To determine the vertical position of the

multi-turn pressure cells during material flow, two


potential multi-turn potentiometers were each
meter
connected to a pulley, supported at the

pulley top of the silo. Two flat plates, 100mm in


diameter each, were fixed to an

inextensible inextensible wire which connected the


wire plates to the pulleys, as shown in figures
3.18a and 3.18b.
100mm
flat plate The multi-turn potentiometers were
connected to the data logger and
calibrated by recording the voltage
output for each quarter, or half turn of
the pulley. The pulleys each had a
circumference of 685mm, giving a
reading for every 171mm travelled. The
Figure 3.18a: Multi-turn
potential meters graph of results and the calibration
constants for each multi-turn potential
meter has been shown in Appendix B.
3.22 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

The plates were placed in the material during filling for each test, and their
positions from the top of the silo measured. When the material started flowing, the
position of the pressure cells relative to the multi-turn potential meters was
therefore easily determined.

Figure 3.18b: Photograph of Multi-Turn Potential Meter


connected to the Pulleys on top of the Model Silo

3.2.8 GATE SWITCHES

To determine the exact time or data point when the gate was opened, two on/off
switches were positioned on the gate. One switch was triggered by the closed gate
and the second switch was triggered by a leverarm which had been placed on the
sliding gate. When the gate was opened the voltage output changed from zero volts
to five volts, and when the gate was fully opened the leverarm triggered the second
switch, which changed the voltage output from five volts back to zero volts. This
also gave the time taken and hence the number of data rows, to open the gate.
WALL PRESSURE MEASUREMENTS: EXPERIMENTAL RESULTS 3.23

3.3 EXPERIMENTAL RESULTS

3.3.1 DESCRIPTION

A total of twenty four tests were performed on the silo, of which four tests gave no
useable results due to problems with the data logger. The first five tests were
performed on the silo without an anti-dyanmic tube installed, to measure the static
and dynamic pressures in the material. The remaining fifteen tests were undertaken
to determine the frictional drag on tubes of varying lengths and diameters. In the
drag tests, the pressure cells were placed in the silo to gain additional data about
the speed with which the swich pressure travels up the silo, as well as to measure
the effect of the tube on the wall pressures. The discussion of the drag tests on the
anti-dynamic tube is given in chapter four.

Owing to the large amount of effort involved in filling the silo for each test, a check
list was made to ensure correct preparation was carried out. An example of the
check lists for both the static and dynamic tests has been given in Appendix C. The
static test check list included checking the electrical signals from all the cells, the
multi-turn potential meters and the switches on the gate. The positions of the
pressure cells from the top of the silo were measured and recorded, as well as the
cell orientation relative to the silo wall. The check list also noted the calibration
constants, the input data file and the method of filling the silo for later reference.
During the filling stage a stopper was placed in the pulleys of the multi-turn
potential meters to avoid the falling material causing the pulleys to turn and reach
their full rotation before the dynamic tests were performed.

The dynamic test check list included the input data file, calibration constants and
checking electrical signals from the data loggers. At the end of each test, the cone
of material at the top was levelled off and the silo was filled to it’s capacity. The
sampling frequency for each test was set approximately seven seconds longer to
ensure the start and finish of each test would be recorded. Each test was timed
with a stop watch to compare the test duration with the output from the multi-turn
potential meters. The cables from the pressure cells were laid out individually next
to the silo to ensure they would not become entangled. It was essential to leave
enough spare cable for the pressure cells to flow freely down the silo. The stoppers
from the pulleys were removed before the start of each dynamic test. However, due
to the slip-stick nature of the material flow, the pulleys overshot their turning
giving exaggerated flow rates. The friction of the pulleys was increased to avoid this
problem.
3.24 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

A complete list of the tests performed is given in table 3.1.

TABLE OF SILO TESTS


Static and Dynamic pressures
Test Pressure cells with
Mtpm
Number good test results

-1 1 Ball
1 1 Ball 1
2 4 tubes 1
3 3 tubes 1
4 3 tubes 1
5 2 tubes 3 plates 1
6 No useable results:
to Elec tric al problems with
9 the data loggers
Anti-Dynamic Tube Drag Force
Test Tube Tube Hopper
Number Diameter Length Gate
diameter
(mm) (mm) (mm)
10 140 2610 180 T ube support broke
11 140 2610 180
12 140 2610 180
13 140 2610 180
14 140 3 sec tions 180 Middle sec tion broke off ten
895,840,845 sec onds after flow started.
15 140 3 sec tions 180
895,840,845
16 140 3 sec tions 180
895,840,845
17 140 3 sec tions 140
895,840,845 Smaller gate opening
18 140 3 sec tions 140 resulted in a funnel
895,840,845 flow pattern.
19 140 2 sec tions 140
895,840
20 120 1800 180
21 120 1800 180 Flow down the inside
22 120 1800 180 and outside of the tube
23 120 1800 180

Table 3.1: Static and Dynamic test list

3.3.2 STATIC TESTS

The silo was filled by means of a bucket elevator without any attachments at the
outlet, for the first five tests. From figure 3.19 it can be seen that the position of
the cone of material varied with the filling process, making it a combination of
eccentric and central filling. To determine the effect of this on the density of the
material several samples were taken from various positions in the silo by placing a
WALL PRESSURE MEASUREMENTS: EXPERIMENTAL RESULTS 3.25

Cone of material small bucket of known weight and volume in


changes position
the material during filling. Each sample was
during filling
then weighed to determine the density at that
level. The bucket elevator was also equipped
with a flexible hose which was positioned in the
silo to cause a central filling situation. Again
Bucket elevator

samples were taken from various depths in the


silo and the density calculated. Table 3.2
shows the densities at various levels for both
filling methods. The difference in the material
density between the two methods was 2.7%
which was not considered large enough to have
a noticeable effect on the results. Therefore, an
average value of 16kN/m3 as determined in the
Figure 3.19: Position of the
laboratory has been used throughout this
cone of material
thesis.

(1) (1)
Test No Depth Mass Density Test No Depth Mass Density
3
(m) (g) kN / m (m) (g) kN / m3
Free fall from bucket elevator Flexible hosing fixed to bucket elevator
7 2.62 2299.7 16.13 6 2.62 2270 15.92
1.89 2353.7 16.53 1.66 2306.8 16.18
0.98 2320.7 16.29 0.67 2196.8 15.38
8 2.8 2335.5 16.39 14 2.84 2325.9 16.32
- - - 2.2 2338.3 16.41
1.53 2244.6 15.73 1.43 2263.5 15.87
0.47 2286.9 16.04 0.93 2072.5 14.47
9 2.8 2334.1 16.38 15 2.71 2340.4 16.43
2.02 2327.2 16.33 1.87 2247.2 15.75
1.35 2306.9 16.18 1.13 2313.1 16.23
- - - 0.74 2098.5 14.66
Average 16.22 Average 15.78

OVERALL AVERAGE = 16,0 kN/m3


(1) Depth of container from top of silo
(2) Container self weight: 87.92g and volume: 0.001345m3

Table 3.2: Density measurements during filling of the silo

The results from the first five static tests have been shown in table 3.3 and plotted
in figure 3.20, with the equivalent Janssen horizontal pressure. The Janssen
pressure was calculated using a stress ratio of 0.4. The pressure cells used in these
3.26 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

tests were tube type cells, which were all placed vertically in the material
approximately 100mm from the wall. The tube cells measured the average of the
circumferential and horizontal pressures acting in the plane, and is given by:

C H
tube
2

From Jenike (1964), the horizontal and circumferential pressure acting in the
horizontal plane of the axially symmetric cylindrical section are equal in magnitude.
Therefore, the horizontal pressure acting normal to the wall is given by:

H tube

T est Depth Pressure Janssen 0


below c ell (kPa) horizontal
surfac e ( H + C )/2 pressure
1 1.1 5.57 5.25
2 0.6 3.15 3.35
Depth below surface (m)

0.93 4.73 4.67 1


1.61 6.65 6.61
2.11 8.06 7.55
2.58 3.1
3 0.69 4.04 3.74
Hopper Cylinder

1.11 6.5 5.28 2


1.39 6.64 6.08
2.28 7.93
2.54 3.88
4 0.28 2.57 1.74
1.04 4.53 5.05 3
2.2 5.33 7.69 8
0 2 4 6
5 1.26 5.66 5.73
Horizontal Pressure (kPa)
2.17 8.29 7.65

Table 3.3: Static Test Results Figure 3.20: Static test results

The pressure cells were placed in the silo at various depths. The exact orientation
and depths of the cells from the top of the silo were recorded on a drawing which
formed part of the check list, as shown in Appendix C. The filling process was
therefore interrupted to place the pressure cells in the material. The total time
taken to fill the silo, place the cells and note the depth and orientation, was
approximately three to four hours. Thus, there was not enough time between
successive layers for the material to be affected by the process, and filling has been
considered as a continuous operation.
WALL PRESSURE MEASUREMENTS: EXPERIMENTAL RESULTS 3.27

Although most of the static test results are slightly greater than those shown by the
line of the Janssen horizontal pressure, these results are deemed to be in good
agreement with the Janssen theory. Therefore, the static results will be used to
compare the ratio of the dynamic to the static pressures. This ratio gives an
indication of the factors applied to wall pressures, and hence to the pressures
exerted on the anti-dynamic tube in the material. From the literature survey, the
vertical pressure in the centre of the silo has been shown to be approximately 30%
greater than the average value. Therefore, the vertical pressure exerted on the
anti-dynamic tube can be estimated from the ratio of dynamic to static results by
multiplying the ratio by a factor of 1.3.

3.3.3 DYNAMIC TESTS

The dynamic tests all followed directly after the static tests with no waiting time
between tests in which the material had time to settle or de-aerate.

At the end of each static test, the computer program was shut down and a new
input data file was entered for the dynamic test. Zero offsets were recorded with
the static pressure acting on the cells. Hence, the dynamic tests show a negative
pressure when the cells passed through the gate of the silo with the moving
material. This negative value is equivalent to the static pressure at the depth the
cell was placed during filling. Figure 3.21 is a typical sample output showing the
general trend of the dynamic tests.

12
Second peak
First
8 peak
Pressure (kPa)

0
zero initial static pressure
gate opened

offsets
-4
20 30 40 50 60 70
Time (sec)

Figure 3.21: Sample output of a typical dynamic test


3.28 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Nearly all the tests showed a first peak pressure at the instant the discharge gate
was opened and the material started moving. The pressure then reduced to
approximately the static value (zero on the graph), although in some cases the
pressure was considerably higher. The second pressure peak shown on the graph
occurred when the cell passed through the transition and entered the hopper. As
the cell left the silo the pressure reduced to a negative value. When the cell was
placed at, or just above, the transition, the first and second peak curves in the
output graph, merged to form a single curve. However, when the cells where
placed below the transition only a single peak appeared in the dynamic output
curve, followed by a rapid decrease in pressure as the cell moved down the hopper.

A list of all the tests performed on the silo, as well as a complete set of the output
results for each test, is given in appendix D.

A table of the results for the first five dynamic tests is given in table 3.4, showing
the values of the first and second measured peak pressures. The total pressures
acting on the silo wall during flow are equal to the dynamic value plus the static
value, which have also been given in table 3.4. The values plotted in figure 3.22 are
the total pressures acting on the silo walls.

T est Depth Static Dynamic pressures T otal pressure


No from top pressure First Sec ond First Sec ond
of silo (m) peak peak peak peak
-1 1.1 3.64 4.83 8.89 8.47 12.53
1 1.1 5.57 3.73 5.89 9.3 11.46
2 0.6 3.15 4.13 6.67 7.28 9.82
0.93 4.73 2.13 11.49 6.86 16.22
1.61 6.65 - 12.26 18.91
2.11 8.06 - 11.93 19.99
3 0.69 4.04 5.18 14.67 9.22 18.71
1.11 6.5 1.66 22.18 8.16 28.68
1.39 6.64 2.35 20.13 8.99 26.77
2.28 7.93 - 12.36 20.29
2.54 3.88 - 3.88
4 0.28 2.57 0.53 3.26 3.1 5.83
1.04 4.53 2.46 4.24 6.99 8.77
2.2 5.33 10.4 21.41 15.73 26.74
5 1.26 5.66 7.89 12.86 13.55 18.52
2.17 8.29 5.24 11.25 13.53 19.54

Table 3.4: Dynamic Test Results


WALL PRESSURE MEASUREMENTS: EXPERIMENTAL RESULTS 3.29

The ratios of the measured dynamic to static pressures have been shown in figure
3.22(b). This ratio was calculated by dividing the value of the first peak pressure by
the measured static pressure at the depth the cell was placed. To calculate the ratio
of the dynamic to static pressure at the transition, the value of the second peak in
the output curve was used for the dynamic pressure, and an average value of 7.10
kPa was used for the static pressure at the transition. This average value has been
calculated from the results of six static tests.

Shown in figure 3.22(c) is the ratio of the measured dynamic pressure to the
calculated Janssen static pressure. For test numbers three to five, tube type
pressure cells were used. These tubes were placed horizontally parallel to the silo
wall and thus measured the average of the pressures acting in the meridian plane.
Therefore an equivalent average static pressure has been calculated as follows:

Average pressure at the point: AV =( H+ V )/2 , and stress ratio: K= H/ V

Therefore: AV= V (K+1)/2 where V is the Janssen vertical pressure given in

equation 2.1.5, using =16.8kN/m3, D=0.98m, =0.404 and K varies.

A varying stress ratio has been used in the above calculation of the Janssen
pressure. For a depth of fill of zero to one diameter K=0.5; from one to two
diameters fill level, K=0.3; and for two to three diameters fill, K = 0.2. These
values have been determined from the static tests and have been explained in
chapter 3.3.4.1.

0
Depth below surface (m)

3
0 10 20 30 0 1 2 3 4 0 1 2
(a) (b) (c)

Figure 3.22: Dynamic test results: (a) Measured horizontal pressures (kPa) ;
(b) Ratio of measured dynamic to measured static pressure ; (c) Ratio of
measured dynamic pressures to Janssen static.
In all three graphs: * First peak ; Second peak
3.30 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

The experimental results show the dynamic pressures to be between one and two
times greater than the Janssen static pressures, and one to four times greater than
the experimental static pressures. This is smaller than the values taken from the
tests by Blair-Fish and Bransby (1973), as well as Van Zanten and Mooij (1977).

3.3.3.1 Speed of the Switch Pressure

The height of the pressure cells above the gate and the sampling rate of the
computer, give an estimate of the speed with which the switch pressure travels up
the silo. The calculation of the time between columns of data and the time between
rows of data has been given in equations 3.9 and 3.10, and has been used to
determine the time when the cells registered a change in pressure. A sample output
data file has been given in table 3.5.

Sec onds T ube 2 T ube 4 T ube 5 Plate 6 Plate 7 Plate 8 Switc hes

- 0.0948 - 0.012 - 0.018 0.184 - 0.024 - 0.009 0.032 - 0.002


- 0.0711 - 0.021 0.001 0.157 - 0.014 - 0.029 0.042 - 0.002
- 0.0474 - 0.104 0.011 0.157 - 0.097 - 0.1 0.022 4.972
- 0.0237 - 0.067 - 0.018 0.157 - 0.014 0.001 0.052 4.982
0.0000 - 0.058 0.001 0.221 - 0.033 - 0.07 - 0.018 4.977
0.0237 - 0.095 0.001 0.231 - 0.088 - 0.019 0.012 4.977
0.0474 - 0.169 - 0.018 0.231 0.004 - 0.019 0.032 4.979
0.0711 0.007 - 0.027 0.24 - 0.005 0.001 0.042 4.975
0.0948 0.888 0.02 0.231 1.028 - 0.321 0.192 0.001
0.1185 4.281 0.523 0.695 2.816 0.624 0.9 0.001
0.1422 5.236 1.755 1.456 2.881 0.995 0.69 0.008
0.1659 4.986 1.3 1.196 2.816 0.794 0.7 0.003
0.1896 4.921 1.167 1.112 2.844 0.804 0.71 - 0.002
0.2133 5.217 1.044 1.103 3.037 0.754 0.631 0.003
0.2370 5.625 1.082 1.149 3.406 0.784 0.631 - 0.004

Table 3.5: Sample Output Data File

For this test the sampling frequency was 5300 Hz, therefore the time between
columns and rows of data was 0.00198 and 0.0237 seconds, respectively. It is
reasonable to assume that the pressure wave may have passed over a pressure cell
before the computer was able to record the data change.

In the spreadsheet the cell number represents both the channel number of the data
logger and the column number in the data output. Tube cell 2 and plate cell 6,
which were placed together at a height of 0.87m above the discharge gate,
registered a pressure change at 0.0810 seconds after the gate was opened. This
gives the pressure wave speed was 10.74m/s for this sample output. By the time
the pressure wave reached tube cell 4 and plate cell 7, which were both placed at
WALL PRESSURE MEASUREMENTS: EXPERIMENTAL RESULTS 3.31

the same level higher in the silo, the computer had already recorded channel 4,
with no pressure change. However since plate cell 7 registered a pressure change,
the time taken for the wave to reach these cells has been taken from channel 5.
This is reasonable, since tube cell 5 was placed together near the top of the silo and
therefore registered no pressure change at the instant in time when the pressure
wave reached plate cell 7.

Tube cell 5 and plate cell 8 were placed together at a height of 2.51m above the
discharge gate. The pressure wave passed over these cells at a time of 0.1106
seconds after the gate was opened. This calculation gives the speed of the switch
as 22.69m/s.

Due to the possibility of large errors being introduced when trying to determine
when the pressure wave passed over the cells, only those cells which were placed
in the upper sections of the silo have been used. This gives the average speed of
the switch pressure travelling up the silo. Table 3.6 gives the average speed from
13 tests using only those cells which were placed in the upper third of the silo.

T est Sampling Height of Switc h


number frequenc y c ell above pressure
(Hz) gate (m) speed (m/s)
5 401 2.88 23.05
6 477 2.54 16.82
10 477 2.78 21.80
11 429 1.99 23.24
12 496 2.18 18.65
13 496 2.61 19.47
14 505 2.51 22.69
15 668 2.47 22.49
16 906 2.04 17.84
17 906 2.33 12.9
18 291 2.36 13.25
19 305 2.29 17.09
20 944 2.64 52.28
21 954 2.63 58.02
Average speed: 20.31m/s

Table 3.6: Switch Pressure Speed

The average speed from all the tests has been shown in table 3.6 and is found to
be 20.31m/s. The results from test numbers 17, 18, 20 and 21 have been excluded
from the calculation of the average due to the large scatter of their results
3.32 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

3.3.4 STRESS RATIO

By placing two types of cells at the same level


V in close proximity to each other, the stress ratio
C can be determined for filling as well as during
flow as shown in figure 3.22. When a tube cell
H
is placed horizontally and parallel to the silo
wall the average of V and H is measured.
Figure 3.22: Directions of Therefore by placing a plate cell to measure V
the stresses in the cylinder
or H, the stress ratios can be determined.

3.3.4.1 Static Stress Ratio

The results of the static stress ratio from test number 11 to 14 is given in table 3.7.

T est Depth Average Measured Stress


No. from Pressure pressure Ratio
top(m) (kPa) V H H V
1.47 2.86 0
4.48 1.24 0.28
1.47 1.24

0 to 1D
10
0.71 0.14 K
Depth from top (m)

0.2 0.08 0.40


0.71 0.08
2.8 2.58 1
4.33 0.83 0.19
2.8 0.83

1 to 2D
1.77 2.54 Ko
11 4.18 0.90 0.22
1.77 0.90
1.22 2.28 2
3.19 1.37 0.43
1.22 1.37
Hopper

1.43 1.36 Ka
12 0.10 2.62 26.47
1.43 0.10
1.24 1.42 3
13 4.39 - 1.6 - 0.35
1.24 4.39
0.0 0.4 0.8
2.34 3.72 Stress Ratio
6.26 1.18 0.19
2.34 1.18
1.17 3.84 Figure 3.23: Graph of
14 6.72 0.96 0.14
1.17 0.96 static stress ratios
0.7 1.67
1.92 1.42 0.74
0.7 1.42

Table 3.7: Static Stress Ratios

The data from test numbers 12 and 13 did not give reasonable results and have
therefore been ignored. Thus, the overall average of the measured stress ratio is
0.32. This is approximately equal to the stress ratio for the material at rest namely:

Ko = 1 – Sin = 1 – Sin 45 = 0.293


WALL PRESSURE MEASUREMENTS: EXPERIMENTAL RESULTS 3.33

The active stress ratio given by : Ka = (1 – Sin )/(1 + Sin ) = 0.17, is


considerably less than the overall average. However from the graph in figure 3.23,
it can be seen that the stress ratio decreases with the depth of material in the silo.

It is proposed in this thesis, to divide the silo into three sections to calculate the
stress ratio. At a depth of zero to one diameter below the top of the silo the ratio
can be approximated from: K = 1–Sin 2 = 0.5.

From one diameter to two diameters below the top, the ratio approaches the value
for the material at rest: Ko = 0.29.

At depths greater than two diameters and in the hopper, the stress ratio is
approximately equal to the active stress ratio, Ka = 0.17.

From Briassoulis (1991) the Reimberts give the stress ratio as decreasing with
depth, which is in agreement with these test results.

3.3.4.2 Dynamic Stress Ratio

For each pair of pressure cells placed at the same level, two ratios have been
determined; one at the start of flow and one when the cells passed through the
transition. These results are shown in table 3.8 and plotted in figure 3.24. There is
a large scatter in the measured stress ratios.

T est Depth Stress 0


No. below Ratio
surfac e H V
(m)
11 1.22 1.89
2.42 1.95
Depth below surface (m)

1
1.77 0.46
2.3 1.73
12 1.43 3.2
2.4 0.58
13 2.4 1.52
Hopper Cylinder

14 2.34 0.4 2
1.17 1.77
2.3 1.69
0.7 1.71
2.45 1.22
15 1.41 0.51
3
2.4 0.53
0.74 0.91 0 1 2 3
2.41 0.31 Stress Ratio
16 1.8 2.8
Figure 3.24: Graph of
Table 3.8: Dynamic Stress Ratios Dynamic stress ratios
ANTI-DYNAMIC TUBE THEORY: LITERATURE SURVEY 4.1

CHAPTER FOUR
THE ANTI-DYNAMIC TUBE
4.1 LITERATURE SURVEY

Other names for anti-dynamic tubes are decompression tubes, discharge tubes,
static flow pipes and tremmie tubes.

4.1.1 PIEPER
Pieper (1969) conducted tests which measured the force exerted on a horizontal
bar during flow. A 10mm diameter bar was placed in the material, 1.5m above the
outlet of a 0.8m square silo filled with quartz sand of 15kN/m3 bulk density. The
bar was supported by a frame hanging from the ceiling, as shown in figure 4.1. The
graph shown in figure 4.2, has been taken from the results by Pieper, as the
average of the two values for the left hand and right hand support. In the paper,
Pieper does not say if the deflection of the bar itself was taken into account in the
output. Furthermore, no description is given of the method to take into account any
downward movement of the bar supports, which may have caused it to derive
additional support from the walls of the model. It must therefore be assumed that
these results may have an experimental error in the value of the force recorded.
Pieper states that the output of the vertical force on the bar during filling is

0.8m 70
65.1
60

50
Pressure (kPa)

40
3.35

30
Pb
20

10
1.5m

0
10mm 0 5 10 15 20 25 30
bar Time (min)
filling emptying

Figure 4.1: Horizontal tie Figure 4.2: Average values


placed in silo during flow. taken from Pieper’s test results
(Tests by Pieper)
4.2 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

approximately the same as the Janssen static value for that depth.
As can be seen from figure 4.1, the force suddenly increases fourfold when the gate
is opened, decreasing slowly with the decrease in the height of the material. The
maximum pressure on the bar recorded by Pieper was 65.1 kPa approximately one
and a half minutes after the gate was opened.

4.1.2 REIMBERT
The Reimbert brothers (1976) claim to be the inventors of the anti-dynamic tube as
a means to produce a homogeneous outflow from a silo. Initially, this consisted of
perforated tubes fitted to the side walls of the silo, connected to inclined tubes in
the hopper, as shown in figure 4.3(a), or alternatively placed centrally as shown in
figure 4.3(b). The Reimbert’s state that the centrally placed tube caused the
material to flow down the tube in successive layers, resulting in a first-in-last-out
situation. In both cases, material flow along the walls is eliminated, thereby
reducing the wall pressures as well as wear on the walls. The Reimberts claim to
have successfully installed the tubes as a retro-fit in many existing concrete silos
which were cracked due to excessive wall pressures.

(a)
(b)

Figure 4.3: The Reimbert’s tube: (a) Section and Top View of
Tube down the side walls of the silo; (b) Centrally placed tube.

However, according to Ravenet’s report (1983), the Swedish specialist Bergau,


records that anti-dynamic tubes were first used at the beginning of the 1900’s by
Miersch in the Frankfurt/Main Silos; Duhle used them in the Alexander Dock Silo, in
Liverpool; and Huart and Kvapil also made use of them in the early 1900’s.

4.1.3 RAVENET
Ravenet (1983) conducted tests with an anti-dynamic tube in a transparent model
to emulate flow in layers as reported by the Reimbert brothers. Ravenet states that
ANTI-DYNAMIC TUBE THEORY: LITERATURE SURVEY 4.3

the tube failed to discharge the material as the holes frequently became blocked.
After several attempts, layered flow was achieved. He states that when the tube
operated succesfully, the dynamic to static pressure ratio was approximately 1.35.

4.1.4 M C LEAN
McLean (1985) reports on general arrangements of the anti-dynamic tube being
successfully installed in silos to allow safe side discharge. These arrangements
alleviate the bending stresses in the wall associated with an eccentric outlet. The
diagrams shown in figure 4.4 have been taken from McLean’s report by scaling his
drawings. While it is uncertain if McLeans’ diagrams were drawn to scale, the ratio
of the tube to silo diameter in his diagrams is 0.2. This is approximately twice the
ratio given by other researchers.

The material is drawn


from two or three levels
in the case of the silo
with multiple outlets.
McLean reports on the
importance of adequate
support given to the
side discharge chutes
Single Multiple
outlet outlets which protrude into the
path of the flowing
(a) (b) material.
Figure 4.4: Side Discharge Outlets: (a) Single Outlet
(b) Multiple Outlets

McLean (1985) gives the following equations to determine the vertical forces acting
on objects placed in the flowing material:
D
For a tall silo: Fv 2.5A p 1 eJ (4.1)
4 K

K
where J 4z Hs (4.2)
D

and Ap is the projected plan area of the object and Hs is the height of surcharge
above the silo. If there is no cone of material above the top of the silo, equation 4.1
reduces to a modification factor of 2.5 times the Janssen equation multiplied by the
projected area of the object.

Both Pieper and McLean report that the pressure on an object submerged in the
4.4 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

material is approximately equal to the Janssen static pressure during filling of the
silo. However McLean’s formula (equation 4.1) gives the force during flow of the
material as 2.5 times the static value, whereas Pieper’s test results show the flow
force to be nearly four times the static value. Thus Pieper’s results give values that
are approximately 1.6 times larger than predicted by McLean.

4.1.5 OOMS AND ROBERTS


Ooms and Roberts (1985) conducted tests on a flat bottomed acrylic model silo,
3.8m tall and 1.2m in diameter, fitted with an anti-dynamic tube and filled with
wheat. Their model and tube arrangement has been shown in figure 4.5. Unlike the
Reimbert’s tube, theirs did not extend the full height of the silo and was open at the
top, with port holes only at the base of the tube. The purpose of their tests was to
determine the effectiveness of this tube arrangement in controlling flow patterns
and wall pressures, before installing tubes in full scale silos. After installation of the
tube, the model silo emptied in two stages as shown in figure 4.5, thus operating
as two short silos in series. They report that with no tube installed, the flow
pressures at the effective transition were over three times the static measured
value. However, the flow pressures were nearly equal to the measured static
pressures after the tube was installed.

Ooms and Roberts state that the minimum height of the tube is determined by the
angle , such that no effective transition is formed between the material and the
silo wall during the first stage of discharge. Therefore, from the geometry of the
model silo and tube:

Stage 1 Effective
transition

Stage 2
hmax

hmin
Dead
material

(a) (b)

Figure 4.5: Ooms and Roberts Tube: (a) Discharge


Sequence; (b) Tube geometry
ANTI-DYNAMIC TUBE THEORY: LITERATURE SURVEY 4.5

hmin = H–D/(2Tan ) (4.3)

Similarly, the maximum height of the tube is determined such that no effective
transition intersects the wall during the second stage of discharge. Therefor

hmax =D/(2Tan ) – (D/2) Tan (4.4)

The angle ( ) of the effective transition to the vertical is given by Jenike et al


(1973c) and Arnold et al (1989) to be a function of the internal friction angle of the
material, whereas Hasra and Bazur (1980) give as a function of both the wall and
internal friction angles.

Ooms and Roberts derived an expression for the static vertical pressure in a silo
with a tube installed, by considering the equilibrium of a horizontal element.

The forces acting on a horizontal slice


of thickness dz in the silo at the level of
zo
z the anti-dynamic tube have been
shown in figure 4.6. The silo and tube
v
diameters are given by D and d
t w dz respectively. At the level of z=zo the
vertical pressure, zo is equal to the
static pressure given by the Janssen
v+( v/ z)dz
equation (2.1)

Figure 4.6: Forces acting on a


horizontal slice

v zo / B eBz o e Bz /B (4.5)

2
and B=4(DKo o+dKi i)/(D -d2) (4.6)

In their derivation, Ooms and Roberts assumed the stress ratios on the inside and
outside of the tube to be different. In equation 4.5, Ko is the stress ratio on the
outside of the tube and Ki is the stress ratio on the inside of the tube.
Assuming fully mobilised flow along the outside of the tube wall, Ooms and Roberts
have given the vertical drag as a direct function of the internal friction angle and
stress ratio, as shown in equation 4.7.

Ht BH t 1
Fv Ki i d zo e 1 (4.7)
B B B

However, as there is no direct relationship between the vertical drag force on the
4.6 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

outside of the tube and the internal properties of the tube, equation 4.7 has been
re-written as shown in equation 4.8. Therefore, Ko and o have been substituted for
Ki and i, and the drag down force becomes:

Ht BH t 1
Fv Ko o d zo e 1 (4.8)
B B B

The effect of the tube diameter on the vertical pressure and the vertical drag down
force as given in equation 4.5, has been shown in figure 4.7 (a) and (b)
respectively. These graphs were calculated for the material propeties and silo
geometry of the model used in the laboratory. The tube height used in the
calculations is 2.6m at a depth of 0.6m below the top of the silo. By increasing the
tube to silo diameter ratio from one quarter to a half, the vertical pressure as
calculated by equation 4.5 decreases from 84% of the Janssen pressure to 61% of
the Janssen pressure. Thus, the vertical static pressure in a silo, and hence the
horizontal static pressures on the walls, can be reduced by introducing a larger tube
to silo diameter ratio.

0 0.0

0.4 0.4
Depth below surface (m)

0.8
Janssen 0.8
1.2
Height of tube (m)

1.2
1.6
0.5
2 1.6 0.5

2.4 2.0
0.25 0.25
2.8
2.4
3.2
0 10 20 30 0 2 4 6 8
Vertical pressure (kPa) Vertical drag (kN)
(a) (b)
Figure 4.7: Effect of tube to silo diameter ratio of 0.25 and 0.5 on:
(a) Vertical pressure ; and (b) Vertical drag on outside of tube

The function B in the vertical drag down force given by equation 4.6, was calculated
assuming equal internal and external stress ratios, and equal internal and external
wall friction angles. The vertical drag force increases with increasing height of the
tube as expected, as this is directly related to the tube surface area. The graph in
figure 4.7(b) shows the relation to be approximately linear at depths of 1.4m and
greater.
ANTI-DYNAMIC TUBE THEORY: LITERATURE SURVEY 4.7

The vertical drag force on the inside of the tube during flow has been given in the
report by Ooms and Roberts, as:

d2 d 4 i Ki Ht / d d2Ht
Fvi zo 1 e (4.9)
4 4 i Ki 4

This expression contains only the properties relating to the inside of the tube, as
would be expected. The same material and silo properties as used in the above
calculations, was used in the calculation of the vertical drag inside the tube. For a
tube of length 2.6m the vertical drag force is 2.07kN, which is 45% of the vertical
drag on the outside of the tube. This implies that the drag should be increased by a
factor of 1.5 when the material flows inside and outside the tube at the same time.

4.1.6 KAMINSKI AND ZUBRZYCKI


Kaminski and Zubrzycki (1985) conducted experiments on a concrete model silo
1.25m in diameter by 3.78m tall, fitted with anti-dynamic tubes and filled with
wheat. They report that the wheat had a bulk density 8 t/m3 , which is too heavy
for wheat and should probably be 8kN/m3. In their report, they do not give the
diameters of the tubes nor do they give an indication of the number of tubes
installed. A sketch of their model arrangement is shown in figure 4.8.
The purpose of their tests was to determine
the vertical forces acting on perforated and
support
non-perforated tubes during material flow.
spring The tubes were suspended by steel rods
dynamometer
from a supporting structure above the
steel
model silo. Spring dynamometers, with a
rods
maximum capacity of 2kN, were fixed to
tube each steel rod to measure the loads on the
tubes as shown in the diagram. The results
dS
of their tests showed that the drag on the
non-perforated tubes increased by a factor
of 6 times the filling vertical force, whereas
Figure 4.8: Kaminski and
for a perforated tube there was no increase
Zubrzycki model silo
in the force during discharge of the material.

Kaminski and Zubrzykci state that the flexibility of the tube supports has a large
influence on the measured value of the vertical force. They conducted experiments
by varying the flexibility of the supports and then measuring the deflection of the
bottom of the tube, dS, during material flow.The results of their tests have been
4.8 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

reproduced in figure 4.9. The vertical axis in their graph shows the ratio of the

measured vertical deflection, dS , to the hydraulic radius of the tube, rh. Thus, they

define the flexibility parameter as the ratio of the deflection, ds, to the hydraulic
radius of the tube, rh. This seems inappropiate as the flexibility can not be a factor
of the tube geometry. The horizontal axis is the ratio of the measured vertical
force, FV , to the calculated Reimbert value, FR. They state that an increase in the
flexibility parameter beyond 25.6, does not give a further decrease in the measured
vertical force on the tube. The lowest measured vertical force was 4% of the
-3
calculated Reimbert value. For an inflexible support (a value of 3.3x10 ), the
measured vertical force was 70% of the calculated Reimbert value. They propose
correction factors to be applied to the calculated Reimbert value, as shown in the
graph by the stepped solid line, which represents an envelope of their test results.

25.6

22.4

19.2
ds / rh (x 10 )
-3

16

12.8

9.6

6.4

3.2

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 FV / FR

Figure 4.9: Kaminski and Zubrzycki test results for varying support flexibility

They state that their results were compared with measurements on full scale silos.
However, there is no reference or description of the extent of the full scale tests.
In their report they state that the optimal geometric parameters for an anti-
dynamic tube were previously derived in a dissertation in 1977. Unfortunately this
reference is not available in English, and hence the recommendations from their
report have been included in this thesis for completeness only. The parameters are
as follows:

1) The diameter of the tube should fall within the following range:

0.32 dt / H 0.64 (4.10)

where H is the hydraulic radius of the silo and dt is the tube diameter.
ANTI-DYNAMIC TUBE THEORY: LITERATURE SURVEY 4.9

In this relationship, the tube diameter is a linear function of the silo diameter.

2) The total area of the tube perforations should fall within the following limits:

0.08 A 0.2 (4.11)

and A = Ah/At = Ah/ dt Ht (4.12)

where Ah is the total area of the perforations in the tube wall, and At is the surface
area of a solid tube. This parameter gives the tube perforations as a linear function
of the tube surface area.
3) Flow characteristics of the tube perforations

0.25 0.8 (4.13)


3
and = A * dt /(dG * D) (4.14)

where dG is the silo discharge gate diameter, and D is the silo diameter.

For a cylindrical or square silo, H =D/4, equation (4.10) can be re-written as:
0.32 D/4 dt 0.64 D/4 (4.15)

Re-writing the second tube parameter given in equation (4.11), in terms of the
tube diameter gives:

0.08 dt Lt Ah 0.2 dt Lt (4.16)

Re-writing the third tube parameter gives the discharge gate diameter as a function
of the area of the tube perforations:

A h dt 2 A h dt 2
dg (4.17)
0.25 D 0.8 D

The maximum and minimum limits as given by equations (4.15), (4.16) and (4.17)
have been plotted in figure 4.10.

Therefore, for a tube diameter of 2m, the acceptable range for the total area of the
tube perforations must fall within 0.5m2 upto 1.25m2 per meter length of tube,
while the silo discharge gate must fall within 0.2m up to 1.25m in diameter.
4.10 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Tube
2

1
Tube perforation area
2
20 15 10 5 1 2
Silo diameter
0.1

0.2
Discharge

0.3
gate

0.4

Figure 4.10: Maximum and minimum limits for tube parameters

4.1.7 SCHWEDES AND SCHULZE


Schwedes and Schulze (1991) report on a discharge tube being succesfully installed
in two cement clinker silos of 16m diameter and 7m tall, with a funnel flow hopper.
The tube diameter was 1.8m, with a wall thickness of 25mm. The ratio of the tube
diameter to the silo diameter was 0.11. The vertical hole spacing was set at 3.5m
c
/c with four rectangular holes at each level. The holes were orientated vertically,
with an open area of 0.6m2 (0.6x1.0m). In their report the maximum vertical stress
inside the tube of diameter dt was calculated from:

v= dt/(4K ) (4.18)

This is the Janssen equation with the exponential term approximately equal to one.
Using the above formula given by Schwedes and Schulze, the pressure on the tube
during flow of the material, for the silo and tube model used in this research, would
be: v = ( 16 0.14) / (4 0.4 Tan19 ) = 4.07 kPa

For the tube length of 2.61m and diameter of 140mm, this gives a total drag down
force of:

Fv=4.07 * 2.61* *0.14 = 4.66 kN = 475.5 kg

This is 2.1 times larger than the value determined using the equation
recommended by Ooms and Roberts.
THE ANTI-DYNAMIC TUBE: EXPERIMENTAL SET-UP 4.11

4.2 EXPERIMENTAL SET-UP

4.2.1 ANTI-DYNAMIC TUBE MODEL

The model silo set up in the laboratory has been described in chapter 3.2.1 of this
thesis. The same cohesionless material as described in chapter 3.2.2, was used for
the experiments of the frictional drag on the anti-dynamic tube. Anti-dynamic tubes
of 0.14m and 0.12m in diameter, and 0.4mm wall thickness, were placed centrally
in the model silo and suspended from a support at the top of the model silo, as

shown in figure 4.11. The angle of wall friction ( t) between the tube and the

material was measured in a standard shear box test and was found to be 19 .

strain gauged
support beam

inextensible
BUCKET ELEVATOR

980 chain
2400

anti-dynamic
tube: 2.61m
total length

locating rods
25
810

180

Figure 4.11 : Model Silo and Tube

The support frame was made up from rectangular hollow sections (RHS) of 65mm x
35mm, welded to two short struts with base plates which were clamped to the top
flange of the model silo, as shown in figure 4.12. The deflection of the support
frame is directly related to the frictional drag acting on the side of the tube during
flow of the material. The beam in the support frame was deliberately orientated
across its weaker axis to ensure adequate deflection, as well as provide sufficient
space for fixing of the strain gauges. The chain connecting the tube to the support
4.12 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

beam was an inextensible chain, thus avoiding excessive vertical movement of the
tube during material flow. The chain was located across the centre of the support
beam.

average strain readings


65 x 35mm RHS from top and bottom of
beam

Base plates clamped


on to top flange of
model silo

inextensible chain
threaded rod
and eye

cover over
bent plate
top of tube

nut, top and


bottom pop-rivets

anti-dynamic
tube

Figure 4.12: Support frame and tube connection

The chain was connected to the top of the tube by a strip of plate which was bent
to fit across the tube, as shown in figure 4.12. Two holes were drilled through the
side of the plate, which was then connected to the top of the tube. A hole was
drilled through the centre of the strip to accommodate a threaded bolt and eye with
two nuts, top and bottom of the bent strip. The chain was passed through the eye
of the threaded bolt and fixed over the top of the support beam. The cover to the
top of the tube was fixed around the threaded bolt to prevent material flowing
down the inside of the tube. For three tests this cover was removed giving the drag
down force for material flow on the inside as well as the outside of the tube.

The model tube lengths varied from 1.8m to 2.6m and were made up from three
sections of galvanized pipes which were riveted together. To ensure the tubes
THE ANTI-DYNAMIC TUBE: EXPERIMENTAL SET-UP 4.13

inextensible Anti-dynamic
chain tube

cover over
top of pipe Disks bolted to
side of tube
skirt around
gap (not Strain gauged
shown) 5mm diameter
threaded rods
disk inside
the pipe
Skirt around the
strain gap to prevent
gauged rods entry of material
(not shown)

Figure 4.13a: Anti-Dynamic Tube Figure 4.13b: Detail of Disks and


Sections Rods

remained in a central position during filling of the silo, six locating rods, three near
the base and three at mid height of the tube, were used. The locating rods were
removed when the fill level reached the level of the rods. The three sections of the
tube were connected by means of disks and 5mm threaded rods, placed inside the
tubes, as shown in figures 4.13a and4.13b. These 5mm rods were strain gauged
at three points around their circumference to give the average strain reading.

The support beam at the top of the silo, and the 5mm rods were calibrated by
hanging a series of weights at the bottom of the tube and recording the voltage
output. The weights were added in 10kg increments up to a maximum of 160kg.
The calibration constants of the strain gauges, with units of kg/V, were determined
from the slopes of the data output which plotted as a straight lines as shown in the
Appendix B.2.

To determine the rate of flow of the material during discharge, the multi-turn
potential meters (mtpm) and 100mm plates, as described in chapter 3.2.7, were
used. The filling process was interrupted to place the plates in the material and
record their positions. The total time taken to empty the silo was also recorded and
compared to the flow rate given by the multi turn potential meters.

A photograph of the inside of the silo with the tube installed has been shown in
figure 4.14. The photograph shows the group of experiments with the tube divided
into three sections.
4.14 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Figure 4.14: View of the Anti-Dynamic tube


installed in the steel model silo.

4.3 EXPERIMENTAL RESULTS


A total of 14 tests were conducted with the anti-dynamic tube installed in the model
silo. The tube lengths, diameters and the diameter of the silo discharge gate were
varied. A matrix of tube tests performed is shown in table 4.1. The first group of
tests, Group A, were done with a silo discharge gate diameter of 160mm, while the
second group of tests, Group B, had a discharge gate diameter of 140mm.

Number of tests T ube diameter T ube diameter


performed (140mm) (120mm)
Silo Solid T ube Solid Flow down
disc harge tube split in 3 tube inside &
gate 2610 mm sec tions 1800 mm outside of
diameter long long solid tube
Group A: 160 mm 3 3 2 3
Group B: 140 mm 3

Table 4.1: Matrix of tube tests

Three tests were performed with a tube of 2610 mm long, followed by another
three tests with the tube separated into three sections as shown in figure 4.13a.
These six tests had a silo discharge gate diameter of 160mm and a tube diameter
of 140mm.
THE ANTI-DYNAMIC TUBE: EXPERIMENTAL RESULTS 4.15

The opening size of the silo discharge gate was then varied from 160mm to 140mm
by placing a wooden disk in the base of the hopper, and another three tests
performed with the tube split into three sections. The tube diameter was then
changed to 120 mm and the length reduced to 1800 mm. Two tests were then
performed with a silo discharge gate diameter of 160mm. Finally, the last three
tests were performed with material flowing down the inside and outside of the tube,
with a silo discharge gate of 160mm in diameter. These variables on the tube were
considered to be of greatest influence on the magnitude of the drag down force.

The complete set of tube test results has been given in Appendix E. In nearly all the
tests there was an initial peak force which only lasted a matter of seconds, followed
by a second peak force of a longer duration. The maximum value of the initial peak
varied considerably between tests, as can be seen in the two typical test results
shown in figure 4.15 and 4.16 below. In figure 4.15, the initial peak value was
lower than the second peak, where as in figure 4.16, the initial peak value is
greater than the second peak value. The second peak represents established flow in
the cylindrical section of the silo. Once the level of material drops below the top of
the tube, there is a non-linear decrease in the drag force, which can be seen in
figures 4.15 and 4.16. In figure 4.16, the drag force on the top and middle portion
of the tube is decreasing, while on the bottom portion of the tube there is a
constant drag force acting, until the material level drops below the bottom portion
of the tube.

240 245.6
Drag Down Force (kg)

180
157.1
120

60

10 20 30 40 50 60 70 80
Time (seconds)
Figure 4.15: Test number 12:Drag down force on solid tube 2.61m long
4.16 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

358.0
320

Drag Down Force (kg) 280

240 245.5
212.5 211.6
200

160

120

80
66.5
40 52.6

0 10 20 30 40 50 60 70
Time (seconds)

Figure 4.16: Test number 16: Drag Down Force on tube


split into three sections

The first test performed to determine the drag down force, was test number 10. As
can be seen from the output of this test in Appendix E, the pop rivets fixing the
plate to the tube broke approximately 16 seconds after the discharge gate was
opened. The tube moved down the silo with the flowing material and blocked the
silo outlet. This test had to be abandoned and the silo top emptied before testing
could resume. The plate was re-fixed to the top of the tube using larger diameter
pop rivets, 5mm, to ensure this did not occur again.

Silo Disc harge Gate Diameter (160 mm) The results from Group A
T est T ube Initial Shear Constant Shear Tests for the solid tubes,
No. length Drag Stress drag Stress
(kg) (kPa) (kg) (kPa) have been given in table 4.2
T ube Diameter (140 mm)
11 2610 183.1 1.56 252.7 2.16
The shear stress on the tube
12 2610 157.1 1.34 245.6 2.10
13 2610 193.8 1.66 278.4 2.38 has been determined from
T ube Diameter (120 mm) the drag down force divided
20 1800 74.8 1.08 - -
by the surface area of the
21 1800 177.5 2.57 134.3 1.94
T ube Diameter (120 mm) tube. The column of
Flow down inside and outside of tube constant drag in the table
22 1800 118.3 0.86 127.9 0.92
23 1800 110.9 0.80 104.8 0.76 refers to established flow in
24 1800 - 103.4 0.75 the silo, which occurs after
the switch pressure has
Table 4.2: Results from Group A Tests
THE ANTI-DYNAMIC TUBE: EXPERIMENTAL RESULTS 4.17

travelled up the silo. In some graphs of the output, the established flow is clearly
visible, as in test numbers 13, 14, 16, 17 and 18. Except for test number 13, these
tests correspond to the tube being split into three sections. In test number 13, the
section of constant flow is very short due to the level of material going below the
top of the tube, as is the case for the remaining tests using a solid tube.

The results for the Group A tests with the tube split into three sections have been
given in table 4.3 below. For this group of tests, there was an initial peak drag
force followed by a second peak representing established flow in the cylindrical
section of the silo.

Silo Discharge gate Diameter (160mm)


Test Tube Initial Drag Force (kg) Shear Established Flow Shear
No. length Top Middle Bottom Stress Drag Force (kg) Stress
(kPa) Top Middle Bottom (kPa)
Tube Diameter (140 mm)
840 254.4 2.20 296.6 2.57
14 840 254.4 2.31 414.3 2.79
895 74.6 1.86 197.4 4.92
840 287.6 2.28 287.6 2.28
15 840 201.6 3.00 201.6 3.00
895 88.7 2.21 88.7 2.21
840 358 3.86 245.5 0.90
16 840 212.5 4.25 211.6 3.85
895 52.6 1.31 66.5 1.66

Table 4.3: Group A Test Results for the tube split into three sections

Gate = 160 mm: Initial peak


3 The test results for Group A tests have

2.5 been plotted in graphs as shown in


figures 4.17 and 4.18. These test
Drag Force (kN)

2
results have been divided into two sub-
1.5
groups showing the initial peak drag
1
value, shown in figure 4.17, and the
0.5 peak drag value for established flow,
0 shown in figure 4.18.
0.0 0.5 1.0 1.5 2.0 2.5
2
Tube Surface Area (m )

Figure 4.17: Group A Tests: Initial Peak


Drag Force
4.18 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

The drag force on the vertical axis, has been given in units of kiloNewtons, while
the tube surface area on the
Peak Drag Values for Established Flow horizontal axis has been given
3
in units of meters squared.
2.5
From figure 4.17 and 4.18 it
2 can be seen that there was a
Drag Force (kN)

1.5 wide scatter of results, with

1 the maximum initial drag on


the tube reaching
0.5
approximately 2.6 kN while
0
the maximum drag value for
0.0 0.5 1.0 1.5 2.0 2.5
2 established flow reached
Tube Surface Area (m )
approximately 2.8 kN.
Figure 4.18: Group A Tests: Established
Flow Drag Force

The results from the Group B tests have been given in table 4.4. For these tests the
tube was split into three sections and the discharge gate diameter was changed to
140mm diameter. The data output showed no initial peak value in the drag force on
the tube when the discharge gate was opened, as can be seen in the graphs given
in Appendix E.

The values given in table 4.4 represent the maximum values of the drag force on
the tube. For each section of the tube, the shear stress has been calculated by
dividing the drag force by the surface area of the tube. The drag for each section
has been determined by
Silo Discharge Gate Diameter (140mm)
subtracting the value of the
Tube Diameter (140 mm)
drag force for the lower
Test Tube Drag Force (kg) Shear
No. length Top Middle Bottom Stress sections. Thus for test number
(kPa) 17, the drag force acting on
840 91.9 0.16
17 840 85.9 1.05 this section is equivalent to
895 46.3 1.15 85.9kg subtract 46.3kg, which
840 108.3 0.26
18 840 98.5 1.21
gives 39.6kg. This value was
895 52.8 1.32 then divided by the surface
840 96.2 0.86
area over the length of 0.84m.
19 840 -
895 64 1.59

Table 4.4: Group B Test results


THE ANTI-DYNAMIC TUBE: EXPERIMENTAL RESULTS 4.19

The results for Group B


Peak Drag Values for 140mm Gate
0.7 tests have been shown in
0.6 figure 4.19 for established

0.5 flow only, as there was no


initial peak drag for these
Drag Force (kN)

0.4
0.3 tests. From figure 4.19 it

0.2 can be seen that the

0.1 maximum drag force

0 reached a value of
0.36 0.38 0.40 approximately 0.64 kN.
2
Tube Surface Area (m )

Figure 4.19: Group B Tests: Established


Flow Drag Force

The time taken to completely empty the silo was determined using a stop watch.
This time gives the average flow rate of the material, compared to the results from
the multi turn potential meters, which give the variation of the flow rate during
material flow. These average flow rates are given in units of meters per second,
which represents the rate of travel of the top surface of material. Therefore, the
total height of the silo was divided by the time measured on the stop watch. While
it is accepted that this is not an accurate measurement of the flow rate of a silo, it
gives an indication of the variability of the flow rate for varying sizes of discharge
gate diameter. The results are given in table 4.5 and 4.6 below.

Silo Gate Diameter 160 mm


Silo gate diameter 140 mm
T est T ime from Flow
Test Time from Flow
Number stop watc h Rate
number stop watch Rate
(sec onds) (m/s)
(seconds) (m/s)
3 88 0.0365
17 355.09 0.0090
5 76.72 0.0418
18 378.87 0.0085
6 87.2 0.0368
19 364.96 0.0088
8 80.72 0.0398
Average 366.31 0.00877
9 80.28 0.0400
11 81 0.0396
Table 4.6: Average Flow Rates of
12 77 0.0417
silo with gate diameter of 140mm
13 94 0.0341
15 74.84 0.0429
16 75.25 0.0427
20 82.07 0.0391
22 79.34 0.0405
23 81 0.0396
Average 81.34 0.0396

Table 4.5: Average Flow Rates of


silo with gate diameter of 160mm
4.20 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

From table 4.5, it can be seen that the average time taken to empty the silo was
one minute and thirty four seconds. For the total silo height of 3.21 meters, this
gives an average flow rate of 0.0396 m/s. This is faster than the rate given by the
multi turn potential meters, since the material undergoes free fall in the hopper
section of the silo. The average flow rates for the silo with a discharge gate
diameter of 140mm are given in table 4.6. The discharge gate of 160mm has a plan
area of 0.0201 m2 while the gate of diameter 140mm has a plan area of 0.01539
m2. The gate of diameter 140mm is approximately 23.4 percent smaller than for
the gate of diameter 160mm. The comparison between the flow rates shows that
the 140mm gate gives an average flow rate of 4.5 times slower than the 160mm
gate.

The average value of the initial shear stress from table 4.2 and 4.3 for Group A
tests is 2.25 kPa. This frictional stress corresponds to a silo with a gate diameter of
160mm, which has a plan area of 0.0201m2. The average shear stress for
established flow is 2.52 kPa for the same silo discharge gate . These values exclude
the results for material flow down the inside as well as the outside of the tube. The
average value of the shear stress for established flow for Group B tests, is 0.95
kPa, taken from table 4.4. This corresponds to a silo gate diameter of 140mm
which has a plan area of 0.01539m2 . These results have been shown in graphical
form in figure 4.20.

20
Exponential
Trend Line
16
Power
Shear Stress (kPa)

12 Trend
Line

Linear Trend Line


0
0 0.01 0.02 0.03 0.04
2
Silo Gate Plan Area (m )

Figure 4.20: Shear Stress vs


Discharge Gate Area
THE ANTI-DYNAMIC TUBE: EXPERIMENTAL RESULTS 4.21

Only three trend lines could be fitted through the two data points. These are an
exponential curve, a power curve and a linear trend line. The equation for each of
the trend lines are:

Exponential Curve:
207.12x
= 0.0392 e

Power Curve:
6 3.6537
= 4x10 A

Linear Trend Line:

= 101.85 A

where is the shear stress acting on the walls of the tube, and A is the plan area of

the gate.

To determine which of these lines would be most appropriate, the drag force on a
full scale silo which is five times larger than the model silo has been calculated. The
full scale silo would have a diameter of 4.8m and a height of 16m, while the
discharge gate would be 0.8m in diameter. The corresponding tube size would be
0.7m in diameter and 13.05m tall. For a discharge gate of 0.8m in diameter, the
plan area is 0.503m2 . This would give the following shear stresses:

43
Exponential Curve: = 6.897x10 kPa.

6
Power Curve: = 3.248x10 kPa.

Linear trend Line: = 51,23 kPa.

The full scale anti-dynamic tube would have a surface area of approximately
2
28.7m . Thus the corresponding total friction force acting on the full scale tube, as
45 6
determined from each of the trend lines, would give 1.979x10 kN, 93.22x10 kN,
and 1470 kN, respectively. Clearly the calculated value from the exponential curve
is far too large and therefore not practical. The value determined from the power
6
curve is 9.5x10 tonnes of force, which is equally impossible for the scope of the
assumed model. When the silo discharge gate is opened the material suddenly
changes from a stationary state to a state of constant motion under the
4.22 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

acceleration of gravity. Therefore, the impact force as the material starts moving
can be calculated as follows:

2 2
F = mg = 16 (4.8 16/4 – 0.7 13.05/4) = 4552.1 kN (4.19)

If the value of the friction force as determined from the linear trend line was
greater than the calculated impact force, it would imply that the material had an
acceleration greater than gravity. Since the value of the friction force calculated
from the linear trend line, 1470 kN, is less than the force calculated for impact, as
given in equation 4.19, the linear trend line can be accepted as a reasonable curve
applied to the two data points, as shown in figure 4.20.

Therefore, it is proposed in this thesis that an equation for estimating the effect of
the speed of the flowing material on the magnitude of the friction stress acting on
the sides of a tube is as follows:

= 101.85 A (4.20)

where is the shear on the walls of the tube and A is the plan area of the discharge
gate.

This equation is an empirical equation and it is not proposed that this is an


acceptable expression for all silo and tube arrangements. This expression is limited
to the scope of this thesis only.
THE ANTI-DYNAMIC TUBE: MATHEMATICAL MODEL 4.23

4.4 MATHEMATICAL MODEL

A mathematical model of the friction force on an insert tube has been developed by
considering the equilibrium of vertical forces acting on an elemental slice of
material as shown in figure 4.11. The figure represents an element of material of
thickness dz, from a silo fitted with an anti-dynamic tube. The lengths of the
element sides are S1, S2, and S3 where:
S1 = R d S2 = r d S 3 = (R- r) (4.21)
The top and bottom surface areas of the element are equal, and are given by:
2 2
A (S 1 +S 2 ) S3/2 d (R -r )/2 (4.22)
2 2
The volume of the element is given by: V = A dz = d (R -r )dz/2 (4.23)

A top

R
A silo
S1
S3

dz r
S2

A tube
A bottom
t

d
v
v dz
z
Figure 4.11: Forces acting on an elemental slice of
material in a silo fitted with an anti-dynamic tube

The areas of the element in contact with the silo wall and tube wall are given
respectively by:
A silo = S1 dz = R d dz and A tube = S2 dz = r d dz (4.24)

By considering downward forces on the element as positive, equilibrium of the


vertical forces acting on the elemental slice gives:

v
v A V s A silo t A tube v dz A 0 (4.25)
z

Noting that = K v Tan w , and substituting equations 4.22, 4.23, and 4.24 in

equation 4.25 gives the following differential equation for the vertical pressure
acting on the element :
4.24 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

d v
G v (4.26)
dz
2 2
where : G = [2 (Ks Tan sw R + Kt Tan tw r)]/[ (R -r )] (4.27)

This is the same expression as derived by Ooms and Roberts (1985).

To solve equation 4.26, multiply throughout by a function (z) and add and subtract

(z) v from the left hand side of the equation. Re-arranging the terms gives:

(z) v (z) v (z) G (z) v (z) (4.28)

Noting that the first term of equation (4.28) is the derivative of the product of

(z) and v, and if the second term equals zero then:

z v z (4.29)

From the second term in equation (4.28), v 0 , therefor let: G = (z)/ (z) , and
Ln (z) G
z
Therefore, z e Gz C Ae Gz (4.30)

Substitute equation 4.30 into 4.29 and solving gives:

1
v Ce Gz (4.31)
G

To find the constant of integration: at z = 0, v = 0, and C= -1/G , Then equation


4.31 becomes:

v 1 e Gz (4.32)
G

O’Neil (1995) gives this general method for solving linear first order differential
equations.

Substitute equation (4.32) into the expression for t to give:

t Kt v Tan t K t Tan t 1 e Gz (4.33)


G
The total friction force acting on the tube can be found by multiplying equation
4.33 by the circumference of the tube and integrating over the length of the tube:

h
1 Gh
FF t 2 r dz Kt Tan t 2 r h e C (4.34)
G G
0

From the boundary conditions at h=0, FF = 0 , and therefore the constant of


integration can be found:
THE ANTI-DYNAMIC TUBE: MATHEMATICAL MODEL 4.25

1
C (4.35)
G
Substituting equation 4.35 in equation 4.34 gives the expression for the total
friction force acting on the tube.

Gh
FF Kt Tan t 2 r e 1 (4.36)
2
G
where G is given in equation 4.27

4.4.1 TUBE PARAMETERS

From the first geometric parameter defined by Kaminski and Zubrzycki 1985, the
diameter of the tube divided by the hydraulic radius of the model silo used in this
research gives a value of 0.5833 which lies within their recommended optimum
range. Therefore, this model tube and silo is considered a good geometric
arrangement. As mentioned previously in this paper, the second and third
parameters have no meaning for a tube with zero perforations and have not been
considered.

As can be seen from equation 4.36 the friction force on the tube is directly related
to the stress ratio within the material at the tube wall, the material density and

the tube radius. Substituting for =16kN/m3, t=19 , s=22 ,r=140mm,


R=960mm, h=2.61m and using various values of the stress ratio, the variation of
the friction force has been shown in figure 4.12. The stress ratios used in figure
4.12 have been discussed on the following page. From this figure it can be seen
that the correct value of stress ratio must be used to determine the friction force.

Friction Force FF
0 50 100 150 200 250 300 350
0
1: Active Stress ratio
2: At Rest Stress ratio
0.5 3: Minimum given in AS 3774
Length of tube (m)

4: Blight’s Upper limit


1
Experimental results

1.5

2.5
Figure 4.12: Graph of Friction
force for various stress ratios
3 1 2 3 4
4.26 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

The active stress ratio is given by: K a (1 Sin ) /(1 Sin ) = 0.172, for the tube

and silo wall. Using this value in equation 4.36 gives the first curve in figure 4.12.

The stress ratio for the “at rest” condition is given by K o (1 Sin ) =0.2929 for

the tube and silo wall, which gives the second curve in figure 4.12

The stress ratio given in the Australian Code [10] is :

1 Sin2 2 Sin2 2
Cos 2
K (4.37)
2
4 Cos 2

where is the coefficient of wall friction and is the material friction angle
For the silo and tube model this gives: Kt = 0.1768 and Ks = 0.1790.
The minimum value recommended by AS 3774 is 0.35, which is given in figure 4.12
by the third curve.

The fourth curve of the friction force in figure 4.12 has been calculated using the
upper limit recommendation of the stress ratio for granular materials from Blight

(1993). This upper limit is given by: K = 1+Sin = 1.7071

Substituting the values for the model into equation 4.36 and using the Blight upper
limit stress ratio gives a total friction force of 293.1 kg. Substituting a stress ratio
of 1.7071 into the expression given by Ooms and Roberts, equation 4.7, the total
friction force acting on the tube is:

1.7071 Tan19 0.14 16 3.3012 2.61 16 2.61


FF z0 e 1
3.3012 3.3012 3.3012

FF 2.8904 kN 294.64 kg

In the above expression, the values of the wall friction and stress ratio inside the
tube have been taken equal to the outside of the tube, and the v0 has been
assumed zero. In both equations 4.9 and 4.36 for the friction force, the variables G
and B have the same numerical value because of the assumption that the internal
tube wall has the same surface as the external tube wall.

4.4.2 VARIABLE VERTICAL PRESSURE ACROSS A SLICE

From the literature survey of the pressure measurements in the silo, the vertical
pressure in the centre of the silo is approximately 15% to 50% greater than the
average vertical pressure. In this thesis an average value of 30% is recommended.
Suzuki et al (1985) give an approximation of the vertical pressure distribution
THE ANTI-DYNAMIC TUBE: MATHEMATICAL MODEL 4.27

across a horizontal element in the silo. This distribution has been given in chapter
3.1.1.1 and has been shown schematically in figure 4.13
This pressure distribution
Silo
center has been adapted to a silo
Silo
line wall fitted with an anti-dynamic
1.26
tube. Since the vertical

1.1 pressure near the silo wall


Average
vertical has been shown to be less
stress than average, this has

0.8 been assumed at the wall


1/3 R 3/4 R of the tube. Similarly,
Figure 4.13: Vertical Pressure distribution since the vertical pressure
according to Suzuki et al.
at the centre is greater
than average, it has been
assumed in this thesis that the vertical pressure at the midspan between the tube
and silo wall is greater than average. This results in the pressure distribution shown
in figure 4.14.

Silo
Tube Silo
center
wall wall
line 1.1 to 1.30

F2 Average
vertical
F1 F3 stress

0.90 to 0.90 to
0.70 0.70
1/4 1/2 3/4

Figure 4.14: Assumed vertical pressure distribution


in a silo fitted with an anti-dynamic tube

The shape was chosen such that the area of the curve below the average pressure
line equals the area above the average pressure line. Two distributions were
considered where the maximum and the minimum varied from 10% above and
below the average line to 30% above and below the average pressure line.

The equations for the two distributions are :

v 1 0.1 Cos 2 x for the 10% variation (4.38)


4.28 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

and

v 1 0.3 Cos 2 x for the 30% variation (4.39)

Thus the element has been divided into four segments as shown in figure 4.15. The
vertical forces acting over the element segments, F1, F2 and F3, have been shown in
figure 4.14. These forces were determined by integrating the expression for the
vertical pressure distribution across the element and multiplying by the average
length of the segment.

R
S1
S6
S3
S5

S4
dz r
S2

3
/4 1
/2 1
/4
d

Figure 4.15: Segments of the elemental slice

The lengths of the segments S4, S5 and S6 are given by the following:

S4 = (r+(R-r)/4) d = (3r+R) d /4

S6 = (r+(R-r)3/4) d = (r+3R) d /4 (4.40a,b)

For the distribution of a 10% variation above and below the average pressure line,
the vertical force, F1 is given by:

S2 S4
F1 v 0.1 v Cos 2 x dx (4.41)
2

where (S2+S4)/2 is the average length of the segment and is given by:

(S2+S4)/2 = (7r+R) d /4 (4.42)

Integrating equation 4.41 from 0 to 0.25(R-r) and substituting equation 4.42 in


equation 4.41 gives:

v 0.1 v 7r R 7r v 0.7r v R v 0.1R v


F1 d
4 2 4 16 8 16 8
THE ANTI-DYNAMIC TUBE: MATHEMATICAL MODEL 4.29

which can be simplified to give:

7r v R v 1 0.1
F1 d (4.43)
8 8 2

Similarly the expression for the vertical force acting over the middle segment is
given by:

F2 v 0.1 v Cos 2 x dx S5 (4.44)

and the length of the segment S5 is given by S5 = (r+R) d /2 (4.45)

Integrating equation 4.44 and substituting equation 4.45 gives the expression for
F2, after simplification, as follows:

r v R v 1 0.1
F2 d (4.46)
2 2 2

And the expression for the vertical force, F3, acting over the third segment of the
element is given by:

S1 S6
F3 v 0.1 v Cos 2 x dx (4.47)
2

where (S1+S6)/2 is the average length of the segment and is given by:

(S1+S6)/2 = (r+7R) d /4 (4.48)

Again integrating equation 4.47 and substituting equation 4.48 into 4.47, after
simplification, gives the following expression for the force F3:

r v 7R v 1 0.1
F3 d (4.49)
8 8 2

Therefore, there are three forces acting upwards on the element from the material
below. These three forces are:
( F1/ z)dz , ( F2/ z)dz , and ( F1/ z)dz (4.50a,b,c)

These three forces have an expression which can be found in a similar manner to
the definition of the vertical forces acting downwards. The final expression for the
vertical forces acting upwards on the element are:

7r R v 1 0.1
F1 v dz d (4.51)
8 8 z 2
4.30 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

r R v 1 0.1
F2 v dz d (4.52)
2 2 z 2

r 7R v 1 0.1
F3 v dz d (4.53)
8 8 z 2

The terms containing v in the expressions of the vertical forces acting downwards

cancel out with the v terms in the expressions of the vertical forces acting upwards
on the element. Therefore, only terms containing ( v/ z)dz will appear in the
expression for vertical equilibrium of the forces acting on the element, which is as
follows:

2 2
R r d dz Tan R d dz Tan r d dz 7r R v
dz d . . . . . . . . . .
2 H s H t 16 z

0.7r 0.1R v r R v 0.1 r R v


dz d dz d dz d .............
8 z 4 z 2 z

r 7R v 0.1 r 0.7 R v
dz d dz d 0 (4.54)
16 z 8 z

Dividing equation 4.54 throughout by dzd and cancelling terms gives the following
expression:

3
0.1 r R v
2 Ks sR Kt tr v R2 r2 (4.55)
2 z

which can be written as:

v
C v J (4.56)
z

2 Ks sR Kt t r
where C (4.57)
3
0.1 r R
2

R2 r2 R r
and J (4.58)
3 3
0.1 r R 0.1
2 2

The expression for C in equation 4.57 differs from the expression for G given in

equation 4.27 by the term: /((3 /2-0.1)(R-r)), which is inversely proportional to


the difference between the silo radius and the tube radius.
THE ANTI-DYNAMIC TUBE: MATHEMATICAL MODEL 4.31

The expression for J in equation 4.58 differs from the right hand term in equation

4.26 by the term: (R-r)/(3 /2-0.1), which is linearly related to the difference
between the silo and tube radius.

Solving equation 4.56 by the same method as given for equation 4.29 gives the
solution for v as:

J
v 1 e Cz (4.59)
C

where J and C have been defined in equations 4.57 and 4.58.

If the silo wall and the tube wall have the same coefficient of friction and assuming
Ks=Kt=K, then the expression for J/C becomes:

J ( R2 r2 ) (R r)
(4.61)
C K (R r) K

The total drag force on the tube can be determined from:

h
FF t 2 r dz (4.62)
0

The shear stress on the walls of the tube is determined from:

t= h Tan t = Kt v t (4.63)
h
Therefore, the friction force is given by: FF 2 rKt t v dz (4.64)
0

Substituting the expression for v from equation 4.59 into the equation 4.64, and
integrating over the height of the tube gives the following expression for the friction
force acting on the tube walls:

h
FF 2 rKt tJ e Ch 1 (4.65)
C

where h is the total height of the tube, Kt is the stress ratio at the wall of the tube,

t
is the coefficient of friction of the tube wall, r is the radius of the tube and C and

J are constants defined in equations 4.57 and 4.58.


4.32 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Thus the expression for the friction force assuming a vertical pressure distribution
across the element, as given by equation 4.38, has been determined. Substituting

in the values for the model, where =16kN/m3, t=Tan19 , s=Tan22 , and

assuming the Blight (1993) upper limit on the stress ratio of K=1+sin , is
applicable gives the total friction force on the tube of 230.43kg.

The second expression for the vertical pressure distribution, as given by equation
4.39, can be solved in a similar manner. For this pressure distribution the final
expression for the friction force is as given in equation 4.65. However the constants
J and C are as follows:

R2 r2
J (4.66)
3 1
R r
2 4

2 Ks sR Kt tr
C (4.67)
3 1
R r
2 4

Substituting the model constants for the silo and tube walls into equations 4.66 and
4.67, and solving equation 4.65 gives a total friction force acting on the tube of
327.60kg.

These equations give values of the friction force in accordance with the values
measured in the laboratory, refer table 4.3.
BUCKLING OF THIN CYLINDRICAL SHELLS: CLASSICAL THEORY 5.1

CHAPTER 5

INTRODUCTION TO CYLINDRICAL THIN SHELL


BUCKLING THEORY
Shell structures are defined as thin by their radius to thickness ratio, r/t. The
definition of small and large in terms of r/t is very ambiguous in most texts. The
minimum limit for the radius to thickness ratio is approximately, 150 to 200
whereas the maximum limit is solely dependent on the practical aspects of
constructing such a thin shell. As the r/t ratio gets larger, so the walls of the shell
deviate from their initial curved shape. Thus there are “imperfections” in the shape
of the shell. These imperfections greatly influence the maximum load carrying
capacity of the shell. The early theories dealing with shell analysis assume a
perfectly curved wall and therefore the load capacities are much higher than the
results obtained from experimental research. These imperfections are most
pronounced in the axial load capacity of the shell.
The middle surface of an element of the cylindrical shell wall, shown in figure 5.1,
is synonymous to the neutral axis of a beam. The shell element has a radius r, a
wall thickness t and sides of length dx and rd . The x, y, and z-axes are as shown
in figure 5.1, where the x-axis is parallel to the length of the shell, the y axis is
tangential to the shell wall and the z-axis is positive towards the centre of the shell.
The corresponding displacements along the x, y and z axes are u, v, and w
respectively.

dx

rd
t/2

dz
middle
surface -t/2

Figure 5.1: Element of a cylindrical shell wall


5.2 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

5.1 UNIFORM EXTERNAL LATERAL PRESSURE

In the consideration of a thin cylindrical shell subjected to a uniform external lateral


pressure, the simple equation, as given in appendix F, for a uniformly compressed
circular ring can be used as the theoretical buckling load, if the ratio of length to
shell diameter is larger than fifty. For shorter cylinders, Timoshenko and Gere
(1963) use the general equations for the deformation of a cylindrical shell, to
determine the critical buckling load due to a uniform external pressure. This allows
the effects of end restraints to be taken into account, which introduces local
bending for shells with rigid end supports. The analysis of a thin shell requires the
determination of the equilibrium of forces acting on a small element of the shell
wall. Figure 5.2 shows a cylinder subjected to a uniform external lateral pressure,
q, around it’s circumference and along it’s length. The internal membrane forces
acting on the deformed shell element have been shown in Figure 5.3.

z
x

y d

Figure 5.2: External lateral pressure


on a Cylindrical shell z

Ny

dx
Nyx r
x
Qx
Qx dx
x Nxy

Nx

Nx
Nx dx
x
Qx rd
N xy
N xy dx
x
Nyx
Nyx d
y
Ny
Ny d

Figure 5.3: Enlarged Deformed Element showing the internal forces


BUCKLING OF THIN CYLINDRICAL SHELLS: CLASSICAL THEORY 5.3

The general equations of equilibrium of membrane forces for deformation of a


cylinder due to uniform external lateral pressure are given below:

In the x-direction:

N yx 2
Nx v w
r Ny 0 (5.1)
x x x

In the y-direction:
Ny N xy
r Qy 0 (5.2)
x

In the z-direction:

Qy 2
Qx v w
r Ny 1 qr 0 (5.3)
2
x r r

In equations 5.1, 5.2 and 5.3, Timoshenko and Gere assumed that the resultant
forces, except Ny , are small. They neglected all terms containing products of these
resultants with the derivatives of the displacements u, v and w.

Figure 5.4 shows the internal moments acting on the sides of the deformed
element.

z
My
y
dx
Myx
Mxy r
x
Mxy
Mx Mx

rd
Myx

My

Figure 5.4: Deformed Element showing the internal moments


5.4 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Equilibrium of moments with respect to the x, y, and z axes given in figure 5.4 are
as given below. Again Timoshenko and Gere assumed that the bending and twisting
moments are small and therefore neglected the products of these moments with
the derivatives of the displacements u, v, and w.

M xy My My M xy
1
r rQ y 0 Qy (5.4)
x r x

Myx Mx Mxy Mx
1
r r Qx 0 Qx (5.5)
x r x
where Qx and Qy are the shear forces acting perpendicular to the element sides as

shown in figure 5.3.

Substituting for Qy and Qx from equations 5.4 and 5.5 into equations 5.1, 5.2, and

5.3 gives the following:

In the x-direction:

Nx N yx 2
v w
r Ny 0 (5.6)
x x x

In the y-direction:

Ny N xy My M xy
1
r 0 (5.7)
x r x

In the z-direction:

2 2
Mx My 2
1 1 v 1 w
r Ny 1 qr 0 (5.8)
2 r 2 r r 2
x

Timoshenko and Gere give the following expressions for Nx, Nxy and Ny:

t /2

Nx dz Et u v w (5.9)
x 2 x r r
1
t /2

t /2

N xy dz Et 1 u v (5.10)
xy 21 r x
t /2

Ny=-qr (5.11)
BUCKLING OF THIN CYLINDRICAL SHELLS: CLASSICAL THEORY 5.5

Substituting equation 5.11 and the derivatives of equations 5.9 and 5.10 and into
equation 5.6 for the forces in the axial direction, and multiplying this result
2
throughout by r(1- )/Et gives the following:

In the x-direction:

2 2 2 2
2 u w 1 u 1 r v v w
r r r 0 (5.12)
2 x 2 2 2 x x x
x

2
where =qr(1- )/Et (5.13)

The equation for equilibrium in the x-direction, given by Timoshenko and Gere
(1963) has been repeated below:

2 2 2 2
2 u w 1 u 1 r v v w
r r r 0 (5.14)
2 x 2 2 2 x x x
x

The difference between the two equations (5.12 & 5.14) lies in the second term
in brackets, which comes from the substitution for Nxy. Thus there is a printing error

in equation 5.14 given in Timoshenko and Gere. Working back from this equation,
and integrating this second term in brackets with respect to , and multiplying by
2
Et/r(1- ) gives the following substitution for Nxy :

Et 1 u 1 v
N xy (5.15)
21 r 1 x

Unless =0 , which for steel is not true, equation 5.15 does not equal equation
5.10, and therefore equation 5.14 is incorrect.

Only small deflections from the uniformly compressed form of equilibrium have
been considered, and therefore Ny differs by a small amount from the value of –qr:

Ny = -qr+Ny’ (5.16)

where Ny’ is the small change in the value of Ny due to the deformation of the shell.

Also taking into account stretching of the middle surface during buckling gives the
following expressions for Ny and the external pressure q:

Ny=Ny(1+ 1) and q=q(1+ 1)(1+ 2) (5.17a,b)


5.6 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

where the axial and circumferential strains are given by:

1= u/ x and 2= ( v/ -w)/r (5.18a,b)

The bending moments in the x and y directions are given by:


2 2
w v w
Mx D (5.19)
x2 2 2 2
r r x

2 2
1 v 1 w w
My D (5.20)
2 2 2 2
r r x x

The twisting moment is given by:


2
D v w
M xy 1 (5.21)
r x

Substituting the derivatives of equations 5.10, 5.16, 5.20 and 5.21 into equation
5.7 for the forces in the y-direction gives the following:

In the y-direction:
2 2 2
1 u 2 1 v v w
r r .............................
2 x 2 2 2
x
2 3 3 2
v w 2 w 2 v
................ r r 1 0 (5.22)
3 2 2
2 x x

2 2
where =t /(12r ) (5.23)

Similarly substituting the derivatives of equations 5.16, 5.19, and 5.20 into
equation 5.8 for the forces in the z-direction gives the following:

In the z-direction:

3 3 4 4
u v v 2 v 4 w w
r w 2 r r ..........
x 3 2 4 4
x x

4 2
2 w w
.......... 2r w (5.24)
2 2 2
x

The displacements in the x, y, and z-directions, for the buckled shape, are given in
Timoshenko and Gere(1963) by the following equations, respectively :

u = Asin(n )sin( x/L) , v = Bcos(n )cos( x/L) , w = Csin(n )cos( x/L) (5.25a,b,c)
BUCKLING OF THIN CYLINDRICAL SHELLS: CLASSICAL THEORY 5.7

where 2n is the number of half waves around the circumference and A, B and C are
arbitrary constants.

The derivatives of these displacement equations with respect to x is:


2 2
u A x u
Sin n Cos and u (5.26a,b)
x L L 2 2
x L

2 2
v B x v
Cos n Sin and v (5.27a,b)
x L L 2 2
x L

2 2 4 4
w C x w w
Sin n Sin and w and w (5.28a,b,c)
x L L 2 2 4 4
x L x L
Similarly the derivatives of equations 5.25a,b,c with respect to is:
2
u x u 2
An Cos(n )Sin and n u (5.29a,b)
L 2

2
v x v 2
BnSin n Cos and n v (5.30a,b)
L 2

2 4
w x w 2 w 4
CnCos(n )Cos and n w and n w (5.31,a,b,c)
L 2 4

Substituting the derivatives given in equations 5.26 to 5.31 into equations 5.12,
5.22 and 5.24, after considerable manipulation, gives the following three
simultaneous equations with three unknown constants A, B, and C, in Timoshenko
and Gere as follows:

In the x-direction:

2 1 1
A n2 B n n C 0 (5.32)
2 2

In the y-direction:
1 1 2
A n B n2 n2 2
1 C n n3 n 2 0 (5.33)
2 2

In the z-direction:

A Bn n3 2 n 2 C1 4
n4 2 n2 2 1 n2 0 (5.34)

where
2 2
qr 1 1 t r
and and (5.35a,b,c)
Et 12 r L
5.8 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

These three equations can be written in matrix form as follows:

0
2 1 n
2
n 1 A
2 2

0 1 n
2
1 1 n
2
1 n
2 2
n 1 B
2 2

2 2 2 2 2
0 n 2 n 1 n n 1 1 C

(5.36)

Buckling of the cylinder is only possible if the constants A, B, and C in equations


5.32, 5.33 and 5.34, are not equal to zero. Therefore, the determinant of the above
matrix, equation 5.36, must be equal to zero.

The equation for the critical lateral pressure can be simplified by further neglecting
small terms in the determinant to give the following final expression for the critical
lateral pressure acting on the cylinder:

2 2 2
qcr Et t n
2
1 2n 1 1 (5.37)
2 2 2 2 2
r 1 12 r n L 2 2
1 2 n L
2 2 n 1 1
r 2 2
r

The expression as given by equation 5.37 has been plotted in graphical form in
3
figure 5.5, for a steel shell with a Young’s Modulus of 200x10 Mpa, a Poisson’s ratio
of 0.3 and a length to diameter ratio of 4. A group of curves has been plotted for
n=2 to n=8 and the ratio of thickness to radius was varied from 0.00025 to 0.005.

For a thickness to radius ratio of 0.0035, the shell buckles at a pressure of 0.0166
MPa with 8 half waves around the circumference, as shown by point A in figure5.4.
However, a similar shell will buckle at a higher pressure with only 6 half waves, as
shown by point B, and with 4 half waves as shown by point C in figure 5.5.
Therefore, as the lateral pressure acting on a shell increases, so the number of half
waves into which the shell buckles decreases.
BUCKLING OF THIN CYLINDRICAL SHELLS: CLASSICAL THEORY 5.9

0.08
n=2 n=8

0.07 n=6
External lateral pressure, qcr (MPa)

0.06

0.05 C

0.04
n=4
0.03
B

0.02
A
0.01

0
0 0.1 0.2 0.3 0.4 0.5
100t/r
Figure 5.5: Critical External Lateral Pressure acting on a Thin Cylindrical Shell

Figure 5.6 shows the pressure on a cylinder as given by equation 5.37 for a steel
cylinder with the number of half waves around the circumference kept constant at
n=3. A group of curves has been plotted for length to diameter ratios of 5, 6 and
50. As can be seen from figure 5.6, the critical pressure on the shell decreases as
the length to diameter ratio increases, for a given thickness to radius ratio.

Figure 5.7 shows the critical lateral pressure on the same cylinder, however the
number of half waves around the circumference has been increased to 8. As can be
seen for any given thickness to radius ratio of the shell, there is very little change
in the value of the critical pressure for a cylinder of length 5 times the diameter
compared to a cylinder of length 50 times the diameter.

Therefore, the statement made by Timoshenko and Gere that the equation for a
uniformly compressed circular ring can be used to calculate the critical pressure for
a cylinder if the length to diameter ratio is large, is only partially true. Figures 5.6
and 5.7 show that as the number of half waves around the circumference increases,
so the critical pressure tends towards the same value, regardless of the cylinder’s
length.
5.10 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

0.04 1.0

L=5D L=5D
0.8

Critical Lateral Pressure, qcr, (Mpa)


Critical Lateral Pressure, qcr, (Mpa)

0.03

0.6
L=6D
0.02

0.4
L=50D
0.01
0.2
L=50D

0.00 0.0
0 0.1 0.2 0.3 0.4 0.5 0 0.2 0.4 0.6 0.8 1
100t/r 100t/r
Figure 5.6: Critical Lateral Figure 5.7: Critical Lateral
pressure for n=3 pressure for n=8

5.2 SYMMETRICAL BUCKLING DUE TO A UNIFORM


AXIAL PRESSURE
The stiffness of a shell around it’s circumference (the membrane stiffness), is much
larger than the bending stiffness of the shell, ie along its axis. Therefore, a great
deal of membrane energy can be absorbed without deforming too much. However
to absorb the same amount of strain energy in bending, the shell has to deform
much more. When the shell is loaded such that most of it’s strain energy is in the
form of membrane compression, the shell will fail dramatically in buckling if this
stored up membrane energy is converted to bending energy.

Two types of buckling exist, non-linear collapse and bifurcation buckling. Non-linear
collapse is predicted by means of a non-linear stress analysis, while the onset of
bifurcation buckling is predicted by means of an eigenvalue analysis. An idealised
load-deflection curve has been given in figure 5.8 for an axially compressed
cylinder, showing the general shape of the various types of buckling paths. For all
paths of buckling, as the load approaches the maximum load, the load-deflection
curve has a nearly zero slope. If the load is maintained, snap through buckling
occurs resulting in dramatic and instantaneous failure.
BUCKLING OF THIN CYLINDRICAL SHELLS: LITERATURE SURVEY 5.11

Limit load of a perfect


shell
A
PL

B Bifurcation
PC C

Post-Buckling
PS Curve
E
D
F
Load “P”

Limit of an imperfect
shell
Deflection “ ”

Figure 5.8: General P- curve for a


non-linear analysis

In figure 5.8, the cylinder deforms approximately axi-symmetrically along the line
OA, until the limit load PL is reached at point “A”. At point A, snap through buckling

occurs and the load-deflection curve follows the line AC. The path OAC is called
fundamental or primary buckling and is associated with axi-symmetric buckling.

The bifurcation buckling point, B, lies between O and A, as shown in figure 5.8. At
this buckling load, the deformations begin to grow in a pattern which is different
from the pre-buckled pattern. Failure of this new deflection mode occurs if the post-
bifurcation load-deflection curve has a negative slope and the applied load is
independent of the deformation amplitude. The post bifurcation buckling path BD is
called the secondary or post buckling path and along this path the cylinder buckles
non-symmetrically.

In both load paths of collapse and bifurcation buckling, the maximum limit occurs at
loads for which some or all of the material is stressed beyond its elastic limit point.

In the case of real structures which contain imperfections there is no such thing as
true bifurcation buckling. The loading path of a real shell follows the curve OEF,
with the failure corresponding to snap through at point E. The collapse load ,Ps, at
point E involves significant non-symmetric displacements.
5.12 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Although true bifurcation buckling is fictitious, a bifurcation buckling analysis is


valid as it gives a good approximation of the actual failure load and mode. The
difference between the bifurcation load PC of the perfect shell, and the actual
collapse load PS of the imperfect shell, depends on the magnitude of the initial
imperfections. A plot of PC/PS versus deflection characterises the sensitivity of the
maximum load, PS, to initial geometric imperfections. This corresponds to the term
of an “imperfection sensitive” shell.

Figure 5.9 shows a plot of test data from Brush and Almroth of the normalised
buckling stress, Pc/Ps, for various r/t ratios of cylindrical shells. The curve shows
that the greater the r/t ratio (ie smaller wall thicknesses) the lower is the Pc/Ps
ratio. Therefore, shells with very thin walls are highly sensitive to initial
imperfections in the walls.
Normalised buckling stress

Theory

1.0

0.8
A design recommendation

0.6

0.4

0.2

0
1500

2000

2500

3000

3500
500

1000

Radius/Thickness
Practical range

Figure 5.9: Test data for cylinders subjected to axial


compression. Taken from Brush and Almroth
BUCKLING OF THIN CYLINDRICAL SHELLS: LITERATURE SURVEY 5.13

5.2.1 BUCKLING DUE TO A UNIFORM AXIAL PRESSURE


( SPECIAL CASE )

Ncr/unit circumference Timoshenko and Gere (1963) give a theoretical


method for determining the critical axial buckling
load acting on a thin shell by considering the
potential energy of the shell during deformation.
The shell in figure 5.10 shows a typical thin shell
subjected to a uniform axial pressure, Ncr per unit

x circumference length, which Timoshenko and


Gere assume remains constant during buckling.
The shell has a radius r, a wall thickness t, and a
length of L. They define the middle surface of the
z
shell as the surface which is located in the centre
y
of the wall thickness. The associated
Figure 5.10: Axial buckling
displacements for the axes system, x,y,z , shown
of a thin cylindrical shell
in figure 5.10 is u,v and w, respectively.

In this theory they have used energy methods to determine the critical axial force
by equating the strain energy to the work done by the external forces. The strain
energy of the shell during buckling consists of the strain due to axial compression
as well as strain of the middle surface in the circumferential direction and bending
of the middle surface.

The strain energy term is made up of bending energy and the strain energy due to
stretching of the middle surface.

U = UC+ UB (5.38)

UC is the circumferential strain energy and is given by:


L 2
Eh 2
UC a c
2 1 a c
dy dx (5.39)
2
2 (1 )
0 0

where a is the axial strain and c is the circumferential strain.

The circumferential strain energy of the middle surface has been given in
Timoshenko and Gere, using the following expressions for the axial and
circumferential strains. These strains have been shown schematically in figure 5.11.
5.14 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

o is the axial strain before buckling and is given by:


Ncr
o = Ncr/Et (5.40)
o
a a is the axial strain and is given by:

a = o- w/r (5.41)
c

c is the circumferential strain and is given by

c =w/r - o (5.42)

The function for the axial displacements has been


Figure 5.11: Schematic
diagram of strains due given by Timoshenko and Gere as:
to axial pressure
w = A Sin(m x/L) (5.43)

Substituting equation 5.40, 5.41, 5.42 and 5.43 into equation 5.39 gives the
following expression for the change in circumferential strain energy:

L 2
2
Uc Et A Sin m x A Sin m x r d dx
2 o r L o r L
2 1
0 0
(5.44)
L 2
Et 21 A Sin m x A Sin m x r d dx
2 o r L o r L
2 1
0 0

In equation 5.38, UB is the bending energy term and is given in Timoshenko and

Gere as:
L 2

UB D 2
dy dx (5.45)
2 x
0 0

where D is the flexural rigidity, (the derivation of D is given in Appendix F):


3
Et
D (5.46)
2
12 1

In the expression for the change in bending strain energy given by equation 5.45

x is the curvature of the shell along its length and is given by:
2
x = w/ x 2 (5.47)
BUCKLING OF THIN CYLINDRICAL SHELLS: LITERATURE SURVEY 5.15

Substituting the second derivative of the displacement function given by equation


5.43 into equation 5.47, gives the curvature of the shell along it’s axis as:
2
x = A(m x/L) Sin(m x/L) (5.48)

Substituting equation 5.48 into 5.45 gives the following expression for the change
in bending strain energy:
4 l
D 2 m m x
UB A 2 r Sin2 dx (5.49)
2 l l
0

Substituting equation 5.49 and 546 into equation 5.38 gives the final expression for
the increase in strain energy in the shell during buckling by Timoshenko and Gere
as:
L
2 2 4 4
m x A EtL A m
U 2 tE o
A Sin dx rLD (5.50)
L 2r 4
2L
0

The work done by the compressive forces, W, is equal to force times distance. The
distance through which the axial force travels is given by:

X = ( a- 0)+ L (5.51)

where L is the shortening due to bending effects. In general terms L is


determined as given in figure 5.12, which shows the wall of the shell buckling in the
vertical plane.
The distance through which the axial force travels is
given by L=L0-L 2 , where L0 is the original length,
L2 is the length after buckling, z is the radial axis,
and dz is the change in radius due to buckling.
dx

L0 By small angles :
dz dz/dx
L2
L L
dz
L2 Cos dx 1 ....... dx
2 dx
0 0
(5.52)
L
1 2
L0 z dx
Shell wall 2
0

L
1 2
L L0 L2 z dx
Figure 5.12: Shell wall 2
during buckling 0
5.16 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Therefore, the work done is given by: W = Ncr ( a - c) + Ncr ( L) (5.53)

The radial displacements in the z-direction are given in equation 5.42. Substituting
equation 5.40, 5.41 and 5.42 into 5.53, and multiplying around the circumference
of the shell gives:

L L
2
A m x 1 Am m x
W 2 r Ncr Sin dx Cos dx (5.54)
r L 2 L L
0 0

Timoshenko and Gere equate the strain energy in the shell to the work done by the
axial compressive force, to give the following expression for the critical axial stress
acting on the shell:

Ncr B2 E D B2 E
cr D (5.55)
2 2
tS t Dr B t r B2
2

where B= m /L, S is the shell circumference and m is the number of half waves
along the length of the shell. The shell in figure 5.10 has been shown with 5 half
waves along it’s length. Since D, E, S, r and t are all constants, the minimum axial
force required to cause the shell to buckle is given by:

d cr 2 E tS
tS Ncr min 2B D S (5.56)
dB 2 3
r B

The expression for the critical load given by equation 5.55, has been plotted in
figure 5.13 for a steel shell with a Youngs modulus of 200x103 MPa, a Poisons ratio
of 0.3 and a wall thickness of 0.8mm. The graphs in figure 5.13 have been plotted
as a function of the number of half waves along the length of the shell. Four curves
have been shown for a cylinder with a length to diameter (2r) ratio varying from 2
to 8, in increments of 2.

Figure 5.13 shows that the minimum axial load required to cause buckling remains
constant, which in this example is 486.75kN. For a constant shell length at this
minimum buckling load, the shell diameter increases as the number of half waves
along the shell decreases, as shown by points A and B in figure 5.13. Alternatively,
for the same number of half waves along the length of the shell, the axial load
increases as the diameter of the shell decreases, as shown by points C and D.
BUCKLING OF THIN CYLINDRICAL SHELLS: LITERATURE SURVEY 5.17

750
8 6 L/2r=4 L/2r=2

700
D
650

600

C
550
Ncr(kN)

500

Min = 486.75 kN A B
450
6 10 14 18 22 26 30
Length/number of half waves , L/m (mm)

Figure 5.13: Variation of the Critical Axial Load as a function of


the Cylinder Length to Number of Half Waves ratio

Since the number of lobes around the circumference of the shell is zero, this
analysis given by Timoshenko and Gere, represents a special case of buckling of a
thin cylindrical shell. Figure 5.14 graphically demonstrates this special case of a
cylinder buckling due to an axial pressure, as given by equation 5.56 with the
number of longitudinal half waves increasing from m=1 to m=36.

m=1 m=3 m=9 m=36

Figure 5.14: Special case of axial buckling with number


of half waves increasing from 1 to 36.
5.18 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

From equation 5.56 the length of waves into which the shell buckles can be
determined as follows:

d cr 2BD E d 1 2BD 2E
tS 0 (5.57)
dB t r
2 dB B 2 t 2 3
r B

Therefore, B 4
Et m (5.58)
r D
2 L

Solving for L/m gives:


2 2
L r t (5.59)
4
m 2
12 1

For steel =0.3, therefore equation 5.59 can be approximated by:

L /m 1.72 rt (5.60)

Equation 5.60 can be re-arranged to give the number of half waves, m, into which
a shell of known length, radius and thickness will buckle, as follows:
m = L / (1.72 rt) (5.61)

The number of half waves as a function of the t/r ratio as given by equation 5.61
has been plotted in figure 5.15 for shell lengths varying from 10r to 30r. The graph
shows that as the length of shell increases so does the number of half waves into
which it will buckle. It can also be seen that thicker shells buckle into fewer half
waves than do shells with a smaller t/r ratio, ie thinner shells.

1000

900

800
Number of half waves, m.

700

600

500

400 L=30r
300
L=20r
200
L=10r
100
0.02 0.06 0.10 0.14 0.18
Shell thickness to radius ratio, 100t/r

Figure 5.15: Number of half waves as a function of t/r


BUCKLING OF THIN CYLINDRICAL SHELLS: LITERATURE SURVEY 5.19

5.2.2 GENERAL CASE OF BUCKLING DUE TO AN AXIAL


PRESSURE

The analysis given in section 5.2.1 gives the result for the special case of buckling
when the deformed shape only has half waves along it’s longitudinal axis. There are
cases when the cylinders will buckle with m half waves in the longitudinal direction
as well as 2n half waves around it’s circumference.

Timoshenko and Gere have approached the general case of axial buckling in the
same way as the analysis for buckling due to a uniform external lateral pressure.
Referring to figure 5.3 given in section 5.1, the resultant internal forces due to an
axial pressure give the following three equations of equilibrium in the x, y and z-
directions. Assuming that all the forces except Nx are small, and neglecting the
products of these forces with the derivatives of the displacements u,v and w:

In the x-direction:

Nx N yx
r 0 (5.62)
x

In the y-direction:

Ny N xy 2
r r Nx v Qy 0 (5.63)
x x
2

In the z-direction:

Qx Qy 2
r r Nx w Ny 0 (5.64)
x
x2

Similarly, referring to figure 5.4, neglecting products of moments with the


derivatives of the displacements, the equations of moment equilibrium are the
same as given in section 5.1. These equations have been repeated below:

In the x-direction

M xy My My M xy
1
r rQ y 0 Qy (5.4 repeated)
x r x

In the y-direction

Myx Mx Mxy Mx
1
r r Qx 0 Qx (5.5 repeated)
x r x

In the z-direction:
M yx r(N xy N yx ) 0 M yx 0 (5.65)
5.20 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Substituting for Qx and Qy from equations 5.5 and 5.4, respectively, into equations

5.63 and 5.64 gives the following:

Equilibrium of forces in the x-direction remains unchanged:


Nx N yx
r 0 (5.57 repeated)
x

In the y-direction:
Ny N xy 2 My M xy
r r Nx v 1 0 (5.66)
x x
2 r x

In the z-direction:
2 2
Mx My 2
r 1 rNx w Ny 0 (5.67)
x
2 r 2 x2

The definitions of Nx, Nxy, Mx, My and Mxy are as given by equations 5.9, 5.10, 5.19,

5.20 and 5.21 respectively. Substituting for Nx and Nxy into equation 5.57 gives:

In the x-direction:

2 2 2
u 1 v w 1 u 0 (5.68)
x
2 2r x r x 2r
2 2

Timoshenko and Gere give the following definition for Ny:

t /2

Ny dz Et 1 v w u (5.69)
y 2 r r x
1
t /2

Substituting for Ny from equation 5.69, and for My and Mxy into equation 5.66 gives

the following:

In the y-direction:

2 2 2
1 u 2 1 v v w
r r .............................
2 x 2 2 2
x

2 3 3 2 2
........... v w r
2 w r
2
1 v r
2 v 0 (5.70)
3 2 2 2
2 x x x
BUCKLING OF THIN CYLINDRICAL SHELLS: LITERATURE SURVEY 5.21

Note that equation 5.70 differs from equation 5.22 only by the last term: ie
2
equation 5.70 has the extra term of r v/ x2.

Finally substituting for Mx, My, Nx and Ny into equation 5.68 gives the following

equation for the equilibrium of forces:

In the z-direction:

3 3 4 4 4
1 v r2 v r
3 w 1 w 2r w 0 (5.71)
r 3
x
2
x
4 r 4
x
2 2

The expressions for the displacements have been given in equations 5.25a,b,c in
section 5.1. Substituting the derivatives of these displacements into the above
three equations, gives the following three simultaneous equations, in matrix form:

0
2 1 n
2
n 1 A
2 2

n 1
2 2 2 2
0 1.5 1 n 1 n1 n B (5.72)
2
2
2 2 2 2 2
0 n1 n 2 1 n C

2 2
where, as before, =t /(12r ) and = m r/L

The solution to equation 5.72 can be found by setting the determinant equal to zero
and solving for . Timoshenko and Gere have ignored small quantities of higher
2 2
order containing terms in and . The solution for then becomes:
2
Nx 1 F (5.73)
Et G

where F and G are as follows:

4
2 4 2 2 4 2
F 1 n 2 3 n ..........

4 2 2 4 2 2 4 6
.......... 2 1 n 7 n 3 n 2n (5.74)

2
G 2 2
n 1
2
1 n
2 2
2 n
2 2
1 ...........
1 2

2 2 2
..........
2
n
2 2 2 (5.75)
1
5.22 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Both the numerator and denominator to equation 5.73 are unitless. Therefore, the
total axial force carried by a shell is given by:

Nx 2 r Et F (5.76)
TOTAL 1
2 G

2
Timoshenko and Gere state that since is a large number, equation 5.76 can be
approximated as follows:

2
2 2
3 n 2
Nx 2 Et (5.77)
12 r 2 2 2 2
2
TOTAL 1 n

Both equations 5.76 and equation 5.77 for the total critical axial load have been
plotted in figure 5.16, as a function of the longitudinal half waves, m. These curves
3
were calculated for a shell with a Young’s Modulus of 200x10 MPa, a Poisson’s ratio
of 0.3, a radius of 300mm, a thickness of 0.6mm, (t/r=0.002), a length of 30 times
the radius (L=9000mm) and the number of half waves around the circumference
2n=4, ie n=2.

15000
14 138 kN

12000 11 535 kN
Total Axial Load Capacity, NxTOT (kN)

9000

Equation
6000 5.77
Equation
5.76
3000

0
0 10 20 30 40 50 60 70 80
Number of Longitudinal half waves, m.

Figure 5.16: Total axial load as a function of


the number of longitudinal half waves.
BUCKLING OF THIN CYLINDRICAL SHELLS: LITERATURE SURVEY 5.23

From figure 5.15 it can be seen that the approximate equation 5.77 predicts a
higher axial load capacity than given by the full solution, equation 5.76 for m=19.
The difference for this particular cylinder is 22.5% (=100(14138-11535)/11535).

From figure 5.15 it was shown that shells with smaller thickness to radius ratios
buckle into a larger number of half waves than do shells with larger ratios.
Furthermore as the length of the shell increases, so does the number of longitudinal
half waves. For the shell under discussion the approximate number of half waves
into which the shell will buckle, at the minimum axial load capacity has been
calculated from equation 5.61 and found to be 390 waves. Thus the area of the
graph in figure 5.16 applicable to this shell lies beyond the graph where the two
equations coincide.

As the number of half waves around the circumference increases, so the two curves
approach each other. The case of n=2 results in the largest difference between the
approximate curve and the exact curve. For n=3 this difference is reduces to 9.7%
and for n=4 the difference is only 5%.

Therefore, in figure 5.16, it is not possible to be working within the range of fewer
longitudinal half waves, m, where there is a noticeable difference between the two
curves. Therefore, the approximation made by Timoshenko and Gere is
reasonable.

5.3 COMBINED AXIAL AND UNIFORM EXTERNAL


LATERAL PRESSURE

The problem of a thin cylindrical shell subjected to a combination of an axial and


uniform external lateral pressure is approached in the same way as before.
Equilibrium equations for the internal forces and moments acting in the x, y and z-
directions due to the applied loads are first assembled. It is to be expected that
these equations are a combination of the equilibrium equations for the case of
lateral pressure and the case of axial loads. They can be written as follows:

In the x-direction (this is the same as the equation for lateral pressure only):

N yx 2
Nx v w
r Ny 0 (5.1 repeated)
x x x
5.24 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

In the y-direction (this is the same as for axial pressure only)

Ny N xy 2
r r Nx v Qy 0 (5.63 repeated)
x x
2

In the z-direction (this is a combination of both cases)

Qx Qy 2 2
r Ny 1 v w r Nx w qr 0 (5.78)
x r 2
r x2

Similarly the equations of internal moment equilibrium are as follows:

In the x-direction (this is the same as the case of lateral pressure):

M xy My My M xy
1
r rQ y 0 Qy (5.4 repeated)
x r x

In the y-direction (this is the same as the case of lateral pressure):

Myx Mx Mxy Mx
1
r r Qx 0 Qx (5.5 repeated)
x r x

Substituting equations 5.4 and 5.5 into equations 5.1, 5.63 and 5.78 gives the
following:

In the x-direction:

Nx N yx 2
v w
r Ny 0 (5.6 repeated)
x x x

In the y-direction:

Ny N xy 2 My M xy
r r Nx v 1 0 (5.66 repeated)
x x
2 r x

In the z-direction:

2 2
Mx My 2 2
r 1 Ny 1 1 v 1 w rNx w qr 0 (5.77)
x
2 r 2 r r 2 x2
BUCKLING OF THIN CYLINDRICAL SHELLS: LITERATURE SURVEY 5.25

The definitions for Nx, Ny, Nxy, Mx, My, and Mxy have been given previously in

equations 5.9, 5.69, 5.10, 5.19, 5.20 and 5.21 respectively. Substituting these
equations into equations 5.6, 5.66, and 5.77 gives the following:

In the x-direction:

2 2 2 2
u 1 1 r
r w u v v w
2
r r q 0 (5.12 repeated)
x
2 x 2 2 2 x x x

In the y-direction:

2 2 2
1
r1 u v v w
2
r .............................
2 x 2 x
2 2

2 3 3 2 2
........... v w r
2 w r
2
1 v r
2 v 0
2 3 2 2 n 2
x x x

(5.70 repeated)

In the z-direction:

3 3 4 4
u v v 2 v 4 w w
r w 2 r r ..........
x 3 2 4 4
x x
4 2 2
.......... 2r
2 w w w r
2 w (5.78)
2 2 q 2 n 2
x x

where q
is due to the external lateral pressure and is given by:

2
=qr(1-v )/(Et)
q
(5.79)

and n
is due to the axial pressure and is given by:

2
n
=Nx(1-v )/(Et) (5.80)

2 2
and is as before: =t /12r

Using the definitions for the displacements as follows:

u=A Sin(n ) Cos(m x/L)

v=B Cos(n ) Sin(m x/L) (5.81a,b,c)

w=C Sin(n ) Sin(m x/L)


5.26 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Substituting the derivatives of the displacements given in equations 5.81,a,b,c in


equations 5.12, 5.70 and 5.78 gives the matrix given in equation 5.82.

Note that in the x-direction the term Sin(n )Cos(m x/L) cancels out; in the y-

direction the term Cos(n )Sin(m x/L) cancels out and in the z-direction the term

Sin(n )Sin(m x/L) cancels out.

0 A

0 M B and =m x/L as before (5.82)


0 C

where [M]=

2 2 1 2 1
r n rn q
r q
2 2

2 2 2
n1 n r
1 2
rn n
1.5 1 n 1
2 2 2 2
2r r

2
2 2 2 2
2 1 r n q
1 n
n 1 n 2
r 2 2 2 2 2 2 n
r r r

(5.83)

Furthermore, Timoshenko and Gere give the solution to the matrix in equation 5.82
by setting the determinant equal to zero. This results in the following solution:

H+R =K q
+P n
(5.84)

where:
2 4
H=(1- ) (5.85)

2 2 4 6 4 2 2 4 6 2 2 4
R=( +n ) - 2( +3 n + (4- ) n + n ) + 2(2- ) n +n (5.86)

2 2 2 2 2 2 4
K=n ( +n ) – (3 n +n ) (5.87)

2 2 2 2 2 2
P= ( +n ) + n (5.88)
BUCKLING OF THIN CYLINDRICAL SHELLS: LITERATURE SURVEY 5.27

Equation 5.84 shows that the lateral pressure and the axial pressure are linearly
related. This equation has been plotted in figure 5.17 for a cylinder with a Young’s
modulus of 200x103MPa, a Poisson’s ratio of 0.3, a wall thickness of 0.6mm, a
radius of 150mm (t/r=0.0004), a length of 7500mm (=50r) and 4 half waves
around the circumference ie n=2.

Three curves have been plotted for the number of longitudinal half waves,
increasing from m=15 to m=35 in increments of 10. Figure 5.17 shows that the
axial load capacity for a cylinder buckling into 35 longitudinal waves is less than the
load capacity for the cylinder to buckle in 25 longitudinal waves. Furthermore from
the difference in the slope of the three lines in figure 5.17, it can be seen that the
external lateral pressure has a greater influence on the axial pressure for a lower
value of m than a higher value.

7000
Total Applied Axial Load, Nx (kN)

6000 m=25

5000

4000
m=35
3000

2000

1000
m=15
0
0 1 2 3 4 5 6
External Lateral Pressure, qcr (MPa)

Figure 5.17: Total Axial Load, (Nx), as a function of


External Lateral Pressure, (qcr).

Figure 5.18 shows the interaction curve, as given by equation 5.84, for a thin
cylindrical steel shell with a length to radius ratio of 4 and a thickness to radius
ratio of 0.0025. The number of half waves along the length and around the
5.28 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

circumference have been varied to give the lowest possible failure combination of
the axial and lateral pressures.

m=7,n=14
800
m=3,n=13

600 m=3,n=14
Total axial load (kN)

400
m=3,n=15

200
m=3,n=16

0
0.00 0.02 0.04 0.06 0.08 0.10 0.12
Lateral Compressive Pressure (MPa)

Figure 5.18: Interaction Graph for a Thin Shell with L=4r.


CYLINDRICAL THIN SHELL EXPERIMENTS: LITERATURE SURVEY 6.1

CHAPTER SIX

CYLINDRICAL THIN SHELL EXPERIMENTS

6.1 LITERATURE SURVEY

6.1.1 TENNYSON

Tennyson(1968) conducted tests on thin cylindrical shells with cut outs. The shells
were made from photo-elastic plastic which were spun cast by pouring liquid epoxy
plastic into a cylindrical mould and rotated at high angular speed until the epoxy
plastic had cured. Tennyson conducted tests on 16 cylinders, each with only one
circular cut out located at mid height of the cylinder. The shells were fitted with end
plates to provide clamped constraints. A diagram drawn to
2r
scale of a shell with the largest cut out has been given in
figure 6.1. From a photograph of the shells given in

2rc Tennyson, the shells appear to have a height to diameter


ratio of 2. As can be seen, the holes in Tennyson’s tests
were small compared to the radius of the cylinder. The
shells were tested axially till elastic buckling occurred.
Each cylinder was loaded axially before the holes were cut
out, to determine the reference buckling load. The results
Figure 6.1: from Tennyson’s tests have been given in table 6.1 along
Diagram of thin
shell with cutout. with the dimensions of the cylinders and holes.

The radius of the cylinders is r, while the radius of the cut out is rc. In Table 6.1,

the reference failure load for cylinders without holes is N0 , and the failure load for

cylinders with holes is Nx. The ratio of shell thickness to radius (t/r) has been given

in the third column of the table, with the inverse of this ratio, ie r/t, in brackets.
The last column of the table gives the ratio of the buckling load of the cylinders
with cut outs to that of the reference buckling load for no cut outs.

A graph of the results from Tennyson has been given in figure 6.2 for the ratio of
Nx/No to the ratio of hole radius to cylinder radius. As Tennyson points out there is

a unique correlation between the test data which appears to be independent of the
cylinder thickness to radius ratio, t/r.
6.2 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

r (mm) t (mm) t/r (r/t) rc(mm) rc /r Nx (kN) Nx/N0

0.00302 0 0 1.557 1
140.97 0.43 12.7 0.09 0.868 0.597
(331)
15.88 0.113 0.801 0.514
0 0 1.958 1
5.82 0.042 1.686 0.861
8.33 0.059 1.357 0.693
0.00342 10.24 0.073 1.210 0.618
140.21 0.48
(292) 11.71 0.084 1.139 0.582
15.86 0.113 0.997 0.509
20.90 0.149 0.899 0.459
23.39 0.167 0.916 0.468
0 0 1.842 1
0.00617 3.18 0.047 1.406
67.31 0.42 4.78 0.071 1.166 0.632
(162)
6.35 0.094 1.005 0.546
12.7 0.189 0.828 0.449
Table 6.1: Test data from tests by Tennyson (1968)

As can be seen from the graph the larger the cut out the lower is the buckling load
of the cylinder. Tennyson gives a trendline through the data which gives an
apparent minimum for the ratio of Nx/No. This does not seem to be a reasonable

conclusion as it is doubtful whether a hole to shell radius of 0.5 gives the same
buckling load as a ratio of 0.2, and therefore the trend line has been omitted in
figure 6.2.

0.9

0.8

0.7

0.6

0.5
Nx/No

0.4
0 0.05 0.1 0.15 0.2 rc/r

Figure 6.2: Results from Tennyson(1968)


CYLINDRICAL THIN SHELL EXPERIMENTS: LITERATURE SURVEY 6.3

Tennyson states that the reason for the drastic reduction of buckling load for a
cylinder with a cut out is due to the growing imperfection in the shape of the
cylinder in the region of the cut out.

6.1.2 ALMROTH AND HOLMES

Almroth and Holmes (1972) conducted tests on 11 thin walled aluminium cylinders,
four with unreinforced cutouts and five with reinforced cutouts. In all cases two
rectangular cutouts were made in the cylinder at mid height and 180 apart on the
circumference. Before the rectangular holes were cut, the cylinders were subjected
to axial pressure to determine a reference buckling load No. Table 6.2 gives the test

data from Almroth and Holmes, converted to SI units. The length of the rectangular
cutouts has been given in Almroth and Holmes as a 30 or 45 arc in plan, as shown
in figure 6.3 drawn to scale. The segment length of the arc, S, has been calculated
and added to table 6.2. Their cylinders were 9 inches high with an outer diameter
of 12.75 inches, and a varying wall thickness. The average radius of each cylinder,
rav, has been given to the middle surface of the cylinder.

S
2r

Cut outs

2r
S 228.6
76.2

45

Plan Elevation

Figure 6.3: Diagram of cylinders with cut outs tested by Almroth and Holmes

The first four rows of data in table 6.2 are for cylinders with unreinforced cut outs
while the remaining seven rows of data are for cylinderrs with reinforced cut outs.
From table 6.2 it can be seen that the cut outs were all of approximately the same
dimensions, with only one being smaller than the rest. Consequently little
information can be gained from plotting their results as a function of the hole
dimensions. However, the wall thicknesses of the cylinders tested do vary, and
therefore the ratio of the buckling load to the reference buckling load, Nx/No, has
6.4 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

been plotted against the ratio of the wall thickness to the radius, t/r, for each
cylinder. The graph of their test results has been shown in figure 6.4. The data
points marked with a dot are for unreinforced cut outs, while the data points
marked with a star are for reinforced cut outs.

rav t 2
S 2
A (mm ) Ac (mm ) t/r Nx (kN) Nx/No
(mm) (mm) (mm)
161.83 0.38 232444.1 84.74 6456.8 0.0083 12.19 0.68
161.83 0.37 232444.8 127.10 9685.2 0.0067 11.30 0.55
161.84 0.33 232444.9 127.11 9685.9 0.0059 9.12 0.48
161.86 0.25 232444.1 127.13 9687.1 0.0044 3.59 0.64
161.83 0.37 232445.2 127.10 9685.2 0.0067 14.19 0.85
161.84 0.34 232457.7 127.11 9685.7 0.0061 12.68 0.77
161.84 0.35 232454.1 127.11 9685.6 0.0063 11.39 0.76
161.84 0.35 232453.5 127.11 9685.6 0.0063 11.57 0.85
161.86 0.24 232492.1 127.13 9687.2 0.0043 5.67 0.89
161.86 0.24 232491.8 127.13 9687.2 0.0044 4.58 0.75
161.86 0.24 232491.8 127.13 9687.2 0.0044 4.69 0.68
Table 6.2: Test data from Almroth and Holmes (1972)

1.0
Reinforced cut outs
reference buckling load, Nx/No

Unreinforced cut outs


Ratio of buckling load to

0.8

0.6

0.4
0.000 0.003 0.006 0.009
Wall thickness to cylinder radius ,t/r.

Figure 6.4: Test data from Almroth and Holmes (1972)

The data from Almroth and Holmes shows that cut outs which are reinforced around
the edges fail at approximately 68% to 88% of the reference buckling load,
whereas unreinforced cut outs fail at approximately 48% to 68% of the reference
buckling load.
CYLINDRICAL THIN SHELL EXPERIMENTS: LITERATURE SURVEY 6.5

In some instances, reinforcing the cut out edges can almost double the axial load
capacity of a cylinder with perforations. From figure 6.4 it can be seen that the
buckling load of cylinders with cut outs, reinforced or unreinforced, depends on the
ratio of wall thickness to cylinder radius, which is in contrast with the test results
from Tennyson. One possible explanation is the difference in the shape of the cut
outs and the different materials of the cylinders. Tennyson used plastic cylinders
whereas Almrtoh and Holmes used aluminium cylinders.

6.1.3 STARNES JR

Starnes Jr (1972) conducted tests on thin walled cylindrical shells made of Mylar™
polyester film and made of copper, with a single circular hole at mid height of the
cylinder. Mylar was chosen by this researcher for it’s ability to buckle elastically as
well as not being sensitive to handling, thereby allowing the same shell to be tested
several times with an increasing hole size. These shells were made from
rectangular sheets of Mylar™ which were lap jointed with a flexible adhesive.
Starnes Jr does not state in the report where the location of the hole was relative to
the lap joint, and it is assumed in this thesis that the hole was placed diametrically
opposite to the joint. The copper shells were manufactured by an electroforming
process. A fly cutter was used to cut the hole in the shell and subsequent surface
measurements showed no apparent bending around the hole due to cutting. The
Mylar™ shells were 203.2mm in diameter and 254mm long, while the copper
cylinders had a length and diameter of 203.2mm. The results from Starnes Jr tests
on copper shells have been given in table 6.3 and the results from the Mylar shells
have been given in table 6.4 in SI units. Starnes Jr does not give the value of Nx

and No for the copper shells however, the ratios of the stress Sx due to the applied

load to the reference buckling stress So have been given.

rc (mm) r (mm) rc/r Sx/So

4.06 101.63 0.0399 0.433


5.03 101.63 0.0495 0.395
6.10 101.63 0.0600 0.391
4.06 101.63 0.0399 0.507
3.05 101.68 0.0299 0.531
10.16 101.68 0.0999 0.398
Table 6.3: Test data for copper shells from
Starnes Jr (1972)
6.6 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

rc (mm) rc/r Nx (kN) Nx/No


0 0 0.994 1
0.64 0.00625 0.994 1
1.27 0.0125 0.994 1
1.91 0.01875 0.994 1
2.54 0.025 0.983 0.989
3.18 0.03125 0.894 0.9
4.06 0.04 0.705 0.709
5.08 0.05 0.583 0.586
6.35 0.0625 0.505 0.508
7.62 0.075 0.516 0.519
8.26 0.08125 0.449 0.452
8.89 0.0875 0.444 0.446
10.16 0.1 0.438 0.441
15.88 0.1563 0.416 0.418
21.34 0.21 0.394 0.396
26.04 0.2563 0.366 0.368
31.12 0.3063 0.338 0.34
40.64 0.4 0.283 0.284
51.44 0.5063 0.242 0.244
Table 6.4: Test data for Mylar shells
from Starnes Jr (1972)

A graph of the test results on the copper shells have been given in figure 6.5 with
the vertical axis in terms of Sx/So plotted as a function of the ratio of the cut out

radius to the cylinder radius. The results from the Mylar shells have been given in
figure 6.6 with the vertical axis in terms of Nx/No also plotted as a function of the

ratio of the cut out radius to the cylinder radius, rc/r. As can be seen from figure

6.5 and 6.6, the results lie on a similar shaped curve to the results from Tennyson.
reference buckling stress, Nx/No

0.6
Ratio of buckling stress to

0.5

0.4

0.3
0 0.02 0.04 0.06 0.08 0.1 0.12

Ratio of cut out radius to cylinder radius, rc/r

Figure 6.5: Test results for copper shells from Starnes Jr (1972)
CYLINDRICAL THIN SHELL EXPERIMENTS: LITERATURE SURVEY 6.7

1.0

Ratio of buckling load to reference 0.8

0.6
buckling load, Nx/No

0.4

0.2

0.0
0 0.1 0.2 0.3 0.4 0.5 0.6
Ratio of cut out radius to cylinder radius, rc/r

Figure 6.5: Test results for Mylar™ shells from Starnes Jr (1972)

6.1.4 SCUTELLA

L. Scutella (1998) conducted eighteen tests on thin walled cylindrical steel shells
with multiple perforations around the circumference and along the length of the
shell. The cylinders had a wall thickness of 0.8mm, a radius of 150mm and a length
of 600mm. The cylinders were made from cold formed steel sheets which were pre-
punched and then rolled and butt welded to form a cylinder. The cylinders were
fitted with end caps which were clamped on, giving constrained edge conditions.
The cylinders were then subjected to a uniform axial compression in a displacement
controlled test. A diagram of the layout of the holes on the flat sheet prior to rolling
has been given in figure 6.7.

The ratio of the hole area to the cylinder surface area (here after called cut out
area) was varied by increasing the hole sizes from an 8mm diameter hole to a
102mm diameter hole. The same ratio of cut out area was also achieved by
increasing the centre to centre spacing of the holes. The cut out area has been
calculated by the ratio of one hole area to the area of one diamond square. These
concepts have been diagrammatically shown in figure 6.8 (a) and (b).
6.8 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

600mm
45

942mm

Figure 6.7: Hole layout for flat sheet rolled into cylinders

Cut out area Cut out area Cut out area Cut out area
=5.3% =16.5% =36.6% =65.9%

Figure 6.8 (a): Increasing the cut out area by increasing the hole sizes
and maintaining the centre to centre spacing of the holes.

Cut out
diameter Cut out
=8mm diameter Cut out
=13mm diameter
=29mm

Figure 6.8 (b): Maintaining the cut out area at 5.3% by increasing the
hole sizes and increasing the centre to centre spacing of the holes.
CYLINDRICAL THIN SHELL EXPERIMENTS: LITERATURE SURVEY 6.9

Three sets of tests were done on a cylinder with no holes to determine a reference
buckling load, No , which was found to give an average value of 113.46kN. Table

6.5 gives the average results of the data for three tests of each cut out area and
hole arrangement, by Scutella. From the data for a cut out area of 0.05% there
appears to be little difference between a hole size of 8mm and 13mm, however
when the spacing and hole size was increased to 29mm there appears to be a drop
in the buckling capacity of the cylinder by approximately 37%.

2
rc (mm) Ac (mm ) Ac/A Nx (kN) Nx/No
8 50.3 0.05 96.42 0.85
12 113.1 0.112 77.72 0.68
13 132.7 0.05 96.46 0.85
29 660.5 0.05 60.04 0.53
51 2942.8 0.16 31.4 0.28
76 4536.5 0.37 13.04 0.12
102 8171.3 0.66 2.37 0.02
Table 6.5: Test data from Scutella (1998)

A graph of the test results from Suctella has been shown in figure 6.9. As can be
seen from figure 6.9, the data points tends to lie on a curve similar to those from
Tennyson and Starnes Jr. There is a very rapid decrease in the buckling capacity of
a cylinder with multiple perforations as the ratio of cut out area is increased from
0.05 to approximately 0.25.

0.9
Ratio of buckling load to reference

0.8

0.7

0.6
buckling load, Nx/No

0.5
0.4

0.3

0.2

0.1

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Ratio of cut out area to cylinder
surface area, Ac/A

Figure 6.9: Test results from Scutella (1998)


6.10 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Figure 6.10 shows the data points for a cylinder with a cut out area of 5% with an
increasing hole size and an increasing centre to centre hole spacing. Since the data
points appear to give a linear trend, a line has been drawn through these points
with the intercept on the y-axis set to 1 which represents the data point for a
cylinder with no cut outs.

1
reference buckling load, Nx/No.
Ratio of buckling load to

0.8

0.6

0.4
0 40 80 120
Centre to centre hole spacing (mm)

Figure 6.10: Test results from Scutella (1998)


for a cylinder with a cut out area of 0.05%

From the results by Scutella, given in figure 6.9 and 6.10, the buckling capacity of
a cylinder with multiple perforations appears to be a function of not only the cutout
area but also the centre to centre hole spacing. The trendline through the data
points in figure 6.10 has the following equation:
Nx/No = 1 - 0.004 s

where s is the centre to centre spacing of the holes.

6.2 DISCUSSION

In both the papers by Tennyson(1972) and Starnes Jr (1972), reference is made to


a parameter which the researchers suggest governs the buckling capacity of the
cylinder with cut outs. In Tennyson this parameter is as follows:

2 rc
4 12 1 (6.1)
8r t
CYLINDRICAL THIN SHELL EXPERIMENTS: LITERATURE SURVEY 6.11

where is Poissons ratio, rc is the cut out radius, r is the cylinder radius and t is the

cylinder thickness.

Since 8=2 2, equation 6.1 can be reduced to the following expression:

rc rc 2
K K (6.2)
2r t 2r t

2 1/4
where K = 0.5 [12(1-v )] (6.3)

A similar parameter is found in Starnes Jr as follows:

1 4 12 1 2 rc 2
(6.4)
2 rt

which can be reduced as follows:

rc 2
K (6.5)
rt

where K is as before and all other symbols are as defined previously.

The only difference between equation 6.2 and 6.5 is the value of 2 in the
denominator of equation 6.2. Multiplying the numerator and denominator in both
equations by gives the following:

rc 2
K for Tennyson’s expression (6.6)
2 rt

rc 2
K for Starnes Jr’s expression (6.7)
rt

Both equations 6.6 and 6.7 are expressions in terms of a constant multiplied by the
square root of the ratio of the cut out area to the area of cylinder material in plan.
However, in equation 6.7 the ratio has only half the cylinder material in the
denominator. Both these expression suggest that the buckling capacity is a function
of how much material has been removed from the cylinder by the cut out.
6.12 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

If both Tennysons data and Starnes Jr’s data are plotted as a function of the ratio
2
of the cut out area to the cylinder surface area ( viz ( rc /2 rL), the results plot as

shown in figure 6.11.

1 1
Tennyson(1968) Starnes Jr (1972)
0.8 0.8

0.6 0.6

0.4 0.4

-0 .1 8 2 5 - 0 .1 7 8 9
0.2 Nx /No = 0.1584(Ac /A) 0.2
Nx/No

Nx /No = 0.1574(Ac /A)


Nx/No

0 0
0.000 0.001 0.002 0.003 0.004 0.000 0.005 0.010 0.015 0.020
Ratio of cutout area to Ratio of cutout area to
cylinder surface area, Ac/A cylinder surface area, Ac/A

Figure 6.11: Test results from Tennyson (1968) and Starnes Jr (1972)
plotted as a function of the cut out area to the cylinder surface area.

For both sets of data a power series trendline has been plotted through the test
results. As can be seen there is a very good agreement between the two sets of
test results when using the ratio of cut out area to the cylinder surface area rather
than the ratio of cut out area to the cylinder plan area. This similarity between the
results may be due to the fact that both Tennyson and Starnes Jr conducted their
tests on plastic cylinders which could be manufactured to a very high standard.

In a similar manner plotting the test results of Almroth and Holmes (1972) and
Scutella(1998), as a function of the ratio of the cut out area to the cylinder surface
area, produces the graphs shown in figure 6.12.

As can be seen these curves also result in a power series plot, however the
agreement between the results is not very good. While both tests were conducted
on steel cylinders, the difference between the results may be due to a different
degree of wall imperfection in each case.
PERFORATED CYLINDRICAL SHELLS: EXPERIMENTAL SET-UP 6.13

1 1 Scutella (1998)

- 1 .2 3 0 6
0.8 0.8 Nx /No = 0.023(Ac /A)

0.6 0.6

0.4 Almroth and Holmes 0.4


(1972)

0.2 0.2
Nx/No

Nx/No
-0 .5 0 8 9
Nx /No = 0.1562(Ac /A)

0 0
0.00 0.02 0.04 0.06 0.08 0.10 0 0.2 0.4 0.6 0.8
Ac/A Ac/A
Figure 6.12: Test results from Almroth and Holmes(1972) and Scutella(1998)

The buckling capacity of a shell with perforations is therefore a function of the


amount of material that has been removed from the walls of the shell, and is not a
function of the plan area of the cylinder as initially suggested by Tennyson and
Starnes Jr. Furthermore, from figure 6.10 it can be seen that the reduction in
buckling carrying capacity of a cylinder with multiple perforations is also a function
of the centre to centre spacing of the cut outs.

From the test results given in figure 6.11 and 6.12, it is can be seen that the
reduction in buckling capacity of a cylinder with perforations, either multiple or
single, subjected to a uniform axial pressure plots as a power series given by
equation 6.8 as follows:

-B
Nx/No = C (Ac/A) (6.8)

where Nx is the buckling capacity of a shell with perforations

No is the reference buckling load of a shell without perforations

C, and B are empirically determined constants


Ac is the total area of the cut outs

and A is the surface area of the cylinder.


6.14 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

6.3 PERFORATED CYLINDER EXPERIMENTS

6.3.1 EXPERIMENTAL SET-UP

Following on from the experiments done by L Scutella, tests were done in the
laboratory on thin shells with multiple perforations subjected to a combined uniform
axial and lateral pressure. The shells were formed from thin cold-formed steel
sheets with a thickness of 0.8mm. The holes were pre-punched in the sheets prior
to rolling into a cylindrical shell. Figure 6.14 and 6.15 shows photographs of the
punching of the cut outs in the sheets and the rolled sheet once punching was
complete. The setting out of the holes started from the centre of the sheet to
minimise the effects of accumulated measurement errors. The centre to centre hole
spacing was kept constant at 157.5mm and the diameters of the cut outs were
51mm and 76mm. This gave a cut out open area of 16.5% and 36.6% respectively.
To ensure that no holes were located on the seam, the setting out was such that
the centreline of the sheet aligned with the three-quarter line of the hole group.
The dimensions of the sheet and hole centre to centre spacing have been drawn to
scale in figure 6.16.

Figure 6.14 : Photo of cut out Figure 6.15 : Photo of punched sheet
punching process rolled into cylindrical shape prior to
welding the seam
PERFORATED CYLINDRICAL SHELLS: EXPERIMENTAL SET-UP 6.15

Figure 6.16a shows the sheet with the 51mm diameter hole and figure 6.16b shows
the sheet with the 76mm diameter holes. These figures give a good visual
appreciation of the difference between cut out open area of the two groups of
cylinders investigated.

600mm
157.5mm
45

157.5mm

942mm

Figure 6.16a: 0.8mm thick sheet with 16.5% cut out open area
(51mm diameter pre-punched holes)

Figure 6.16b: 0.8mm thick sheet with 36.6% cut out open area (76mm
diameter pre-punched holes). No holes were located on the shell seam
6.16 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

After the holes were punched, the sheet was rolled into a cylindrical shape and the
seam was butt welded. To weld the seam, the rolled sheet was clamped into the
two end caps which had a diameter of 300mm. A diagram of the end clamps, which
were also used during the tests to ensure constraint end conditions of the cylinder,
is shown in figure 6.17.

Loading sphere circular


end caps

circular clamps
cylinder
(not shown)
fitted over
hold cylinder in
Side view end caps
place

Spaces between cross


beams closed off with
plates welded to the
beams and cap

25mmx45mm
deep cross
beams for
Figure 6.17: Top View carrying axial
and Side view of load
cylinder end caps Top view

The circular clamps and end caps were ridged on the inside to ensure a good grip
on the cylinder wall and thereby stop any possibility of slippage. The end caps were
closed with a plate welded to the sides of the cross beams and the cap. To apply
the axial pressure to the cylinder, one end cap was fitted with a solid half steel ball
and notched for seating of the loading rod, as shown in figures 6.18a and b. The
bottom end cap was also notched for seating on a small round stainless steel ball.

Loading rod
connected
to Instron

Loading
ball

Top of
end cap

Figure 6.18a : Loading ball Figure 6.18b: Photo of Loading


and rod ball on end cap
PERFORATED CYLINDRICAL SHELLS: EXPERIMENTAL SET-UP 6.17

A dummy PVC tube of 250mm diameter (50mm smaller than the perforated
cylinders) was placed inside the perforated cylinders. Polystyrene filler plugs were
then fixed with double sided tape to the inner dummy tube through the cut outs. A
plastic sheath was fitted over the perforated cylinder and clamped to the end caps
with the cylinders, to ensure that the whole assembly was water tight. This
assembly has been shown in diagrammatically figure 6.19 below.

End cap

Perforated
cylinder

Dummy
PVC tube

Polystyrene
filler plugs

plastic sheath

End cap

Figure 6.19: Perforated cylinder assembled with


plastic sheath, dummy tube, plugs and end caps

To ensure the dummy PVC tube did not carry any of the applied axial load, the top
of the tube was notched by 30mm to match the alignment of the cross beams of
the end caps. This allowed the end caps on the perforated cylinders to move the
required vertical distance of 25mm as well as maintaining the dummy tube in place
while lifting the assembly inside the loading bin. This has been shown
diagrammatically in figure 6.20.

End cap and


cross beams

Cross beams fit


inside notches

Dummy PVC tube


with notched top
edge

Figure 6.20: Notched dummy PVC tube to accommodate end cap.


6.18 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

To apply the lateral compressive pressure to the outside of the cylinder, the whole
assembly as shown in figure 6.19 was placed inside a large container which had a
removable top lid. This container, the loading bin, was fitted with a small stainless
steel ball in the centre of its base, thereby ensuring the cylinder assembly could be
placed centrally. The loading bin was first placed in the Instron machine and then
the cylinder assembly was lifted into place so that the notched bottom end cap
fitted the stainless steel ball at the base of the bin. The loading bin was then filled
with water up to a short distance from the top, leaving an air gap. The lid of the
loading bin was fitted with a rubber O-ring, and bolted to the main body of the
container. Through the lid were two valves, to which an air supply and pressure
gauge were fitted. Thus the top air gap inside the loading bin was pressurised from
the compressed air supply, and the amount of pressure applied was read off the
gauge. In this manner a constant lateral pressure was applied to the cylinder
assembly during the testing process. A photograph of the loading bin and top
valves has been shown in figure 6.21. The lid also had an opening with a rubber O-
ring through which the loading rod from the Instron machine was placed.

Opening for loading rod

Valve for compressed


air supply

Top lid bolted to


container

Valve for air pressure


gauge

Construction joint

Valve for draining water


after completed test

Figure 6.21: Photograph of container for cylinder assembly


PERFORATED CYLINDRICAL SHELLS: EXPERIMENTAL RESULTS 6.19

6.3.2 EXPERIMENTAL RESULTS

Once the perforated cylinder, dummy PVC tube, polystyrene fillers, plastic sheath
and end caps had been assembled and placed in the loading bin, the perforated
cylinders were subjected to axial displacement controlled tests in the Instron. For
each test a constant lateral pressure was applied first and then the displacement
controlled test started. For each group of cylinders, ie 51mm diameter cut outs and
76mm diameter cut outs, the lateral pressure was varied from 10kPa to 40kPa in
increments of ten. The data from Scutella(1998) has been used as the reference
buckling capacity for perforated cylinders with only applied axial pressure. Table 6.6
gives a summary of the results from the combined axial and lateral pressure tests.
The results from Scutella(1998) have been included in Table 6.6 for the case of
zero lateral pressure. The full set of test results have been plotted graphically in
Appendix G.1 for the 16.5% cut out area and in Appendix G.2 for the 36.6% cut
out area.

Applied Lateral Maximum Applied Total Axial Load (kN)


Pressure (kPa) 51mm Cutouts 76mm Cutouts
0 31.4 (Scutella) 13.04 (Scutella)
10 26.27 11.51
20 32.25 9.36
30 17.13 12.78
40 18.55 6.21

Table 6.6: Test results for cylinders with multiple perforations

As a reference point, the capacity of the perforated cylinders with a 16.5% and
36.6% cut out area due to a lateral pressure only was also determined. This was
done using the assembled system in the loading bin and applying a lateral pressure
in increments of 5kPa up to a maximum of 60kPa. Inside the perforated cylinders
were placed four displacement transducers around the circumference, which
measured the inwards movement of the cylinder walls. The applied lateral pressure
was plotted as a function of the inward wall movement as shown in appendix G.3
for the 16.5% and 36.6% cut out areas. Figure 22 shows a photograph of the
cylinder with a 16.5% cut out area after failure due to a lateral pressure only.
6.20 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Figure 6.22: Buckled perforated cylinder due


to lateral pressure only (16.5% cut out area)

The failure load for the cylinders with 16.5% and 36.6% cut out areas were 42kPa
and 47kPa, respectively.

The results from the perforated cylinder tests have been plotted in figure 6.23 for
the perforated cylinders with 16.5% cut out area and in figure 6.24 for the
cylinders with 36.6% cut out area.

35 14

30 12

25 10
Total Axial Load (kN)

Total Axial Load (kN)

20 8

15 6

10 4

5 2

0 0
0 10 20 30 40 50 0 10 20 30 40 50
Applied Lateral Pressure (kPa) Applied Lateral Pressure (kPa)

Figure 6.23: Test results for Figure 6.24: Test results for
combined pressure on perforated combined pressure on perforated
cylinders with 16.5% cut out area cylinders with 36.6% cut out area
PERFORATED CYLINDRICAL SHELLS: EXPERIMENTAL RESULTS 6.21

From the theory of thin walled shells subjected to a combined axial and lateral
pressure, given in chapter 5.3, it was shown that the relationship between the axial
and lateral pressure is linear. Furthermore, from the finite element analyses given
in section 6.4 of this chapter, it was shown that the relationship between the axial
and lateral pressure for a cylinder with multiple perforations consists of two straight
lines as shown in figure 6.23 and 6.24.

On closer inspection of the test results for the 16.5% cut out area, the data point
for an applied lateral pressure of 20kPa would appear to be too high. Therefore,
ignoring this point, a best fit straight line can be drawn through the remaining data
points for an applied lateral pressure up to 40kPa. The equation of this line is given
as follows:

Nx16.5 = 31.4 - 0.382 q (6.9)

where Nx16.5 is the total applied axial load in kN acting on a cylinder with cut out

area of 16.5%, q is the applied lateral pressure in kPa and 31.4 is the total axial
load capacity of a cylinder with 16.5% cut out area with zero lateral pressure.

The second best fit line has been drawn through the data point for an applied
lateral pressure of 40kPa and the reference pure lateral pressure, 47kPa. This line is
given by the following equation:

Nx16.5 = 124.55 - 2.65q (6.10)

Similarly from the results for the cylinder with 36.6% cut out area, the data point
for an applied lateral pressure of 30kPa appears too high and has therefore also
been ignored. The best fit straight lines through the remaining points have been
shown in figure 6.24. For the data points up to an applied lateral pressure of 40kPa,
the equation of the straight line is as follows:

Nx36.6 = 13.04 - 0.172 q (6.11)

where Nx36.6 is the total applied axial load in kN acting on a cylinder with cut out

area of 36.6%, q is the applied lateral pressure in kPa and 13.04 is the total axial
load capacity of a cylinder with 36.6% cut out area with zero lateral pressure.

The second straight line through the data point for an applied lateral pressure of
40kPa and the reference buckling pressure of 42kPa for an applied lateral pressure
only is as follows:
6.22 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Nx36.6 = 130.41 – 3.105 q (6.12)

Using these equations, interaction curves can be drawn for the capacity of a
cylinder with multiple perforations subjected to a combination of an applied lateral
pressure and total applied axial load as shown in figure 6.25. The areas under the
curves represent the load combinations which can safely be sustained by cylinders
with a cut out area of 16.5 and 36.6% .
35

30

25 Safe combined loads for


cut out area of 16.5%
Total axial load Nx , (kN)

20

15
Safe combined loads for
10 cut out area of 36.6%

0
0 10 20 30 40 50
Applied lateral pressure q, (kPa)

Figure 6.25: Interaction graphs for cylinders with multiple


perforations (cut out areas of 16.5% and 36.6%)

From figure 6.25 it can be seen that the influence of the lateral pressure is to
reduce the applied axial load up to a point when the lateral pressure is close to the
value of the cylinders lateral failure pressure. At this point the applied axial load is
approximately half the value of the axial load for zero lateral pressure acting.
Beyond this point the lateral pressure has a stronger influence on the axial load by
rapidly reducing it’s value to zero.

From equation 6.9, the axial load which a cylinder with 16.5% cut out area can
carry when subjected to a lateral pressure of 40kPa is as follows:
Nx16.5 = 31.4 - 0.382*40 = 16.12 kN

Similarly for a cylinder with a cut out area of 36.6%, the total axial load which can
be carried with a combined lateral pressure of 40kPa is as follows:
PERFORATED CYLINDRICAL SHELLS: EXPERIMENTAL RESULTS 6.23

Nx36.6 = 13.04 - 0.172*40 = 6.16 kN

Plotting the ratio of these values to the reference buckling loads (31.4 for the
16.5% cut out area and 13.04 kN for the 36.6% cut out area) as a function of the
cut out area of the cylinder, has been shown in figure 6.26.

From the work by Tennyson(1968), Starnes Jr (1972), Almroth and Holmes (1972)
and Scutella (1998) it was shown that the curve for the axial load on cylinders with
perforations has the shape of a power series. It has been assumed in this thesis
that this same shape curve will be applicable to a cylinder with multiple
perforations, which has a combination of an applied axial and lateral pressure.
Therefore, plotting a power series trendline through the points in figure 6.26 gives
the curve of the total axial load on a cylinder with multiple perforations and a
constant lateral pressure of 40kPa as a function of the cut out area. An additional
set of curves have also been plotted for a constant lateral pressure of 20kPa and
5kpa.

0.9
Ratio of axial load to reference buckling, Nx/ No

Constant lateral
pressure:
0.8
5kPa
0.7
20kPa

0.6 40kPa

0.5

0.4

0.3

0.2

0.1

0.0
0 0.1 0.2 0.3 0.4
0.165 0.366
Ratio of cut out area, Ac/A

Figure 6.26: Axial load as a function of cut out area


for a constant lateral pressure
6.24 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Table 6.7 gives the data points used to plot the curves in figure 6.26. The equations
of the resulting power series trendlines have also been given in table 6.7 for each
curve of constant lateral pressure.

Lateral Cut out Equation Calculated Nx/No Equation of power


Pressure area Number Nx series trendline
(kPa)
(kN) Nx/ No =

5 0.165 6.9 29.15 0.257 (-1.0947)


0.0357 q
0.366 6.11 12.14 0.107
20 0.165 6.9 23.37 0.206 (-1.1194)
0.0274 q
0.366 6.11 9.64 0.085
40 0.165 6.9 15.77 0.139 (-1.1824)
0.0165 q
0.366 6.11 6.13 0.054
Table 6.7 : Data points for curves in figure 6.25

Photographs of the buckled cylinders with 16.5% and 36.6% cut out area subjected
to a combined total axial load and external lateral pressure have been shown in
figures 6.27 and 6.28 respectively.

Figure 6.27a: Perforated Cylinder Figure 6.27b: Perforated Cylinder


with 16.5% cut out area subjected with 16.5% cut out area subjected
to a total axial load of 26.27kN and to a total axial load of 17.13kN and
10kPa lateral pressure 30kPa lateral pressure
PERFORATED CYLINDRICAL SHELLS: EXPERIMENTAL RESULTS 6.25

Figure 6.27c: Perforated Cylinder


with 16.5% cut out area subjected
to a total axial load of 18.55kN
and 40kPa lateral pressure

Figure 6.28a: Perforated


Cylinder with 36.6% cut out
area subjected to a total axial
load of 11.51kN and 10kPa
lateral pressure

Figure 6.28b: Perforated Cylinder


with 36.6% cut out area subjected
to a total axial load of 9.36kN and
20kPa lateral pressure
6.26 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

6.4 FINITE ELEMENT ANALYSES

6.4.1 DESCRIPTION

Since there is no existing theory for the buckling capacity of cylinders with multiple
perforations with which to compare the test results presented in chapter 6.3.2,
finite element analyses were performed on cylinders with similar ratios of cut out
area. The thin walled shells considered in the finite element analyses were cylinders
with 50mm diameter cut outs and cylinders with 76mm diameter cut outs. The
arrangement and spacing of the holes was as given in figure 6.16 of chapter 6.3.1.
In addition to cylinders with multiple perforations, a solid shell with no cut outs was
also analysed and the results used as a reference for cylinders with multiple
perforations.

The finite element (FE) software package used for these analyses was the student
version of Ansys 5.6/Mechanics which includes structural and thermal non-linear
capabilities.

For all the cylinders analysed, the most suitable element chosen from the FE
software package was the quadrilateral element “shell63” which has both bending
and membrane capabilities. Loads on this element can be in-plane node or edge
loads, or positive surface loads acting into the element. Stress stiffening and large
deflection capabilities are also included. The element is defined by four nodes, four
thicknesses specified at each node and orthotropic material properties. The
geometry, node locations and local co-ordinate system of this element have been
shown in figure 6.29.
L

y
K,L

z K I

J
x Triangular option
of ‘shell63’

Figure 6.29: Geometry and node location of element ‘shell63’


PERFORATED CYLINDERS : FINITE ELEMENT ANALYSIS 6.27

The external lateral pressures were applied as equivalent element loads at the
nodes as this produced more accurate stress resultants on the flat elements
representing the curved surface of the cylinders.

The easiest method of entering the model geometry into the FE package was to use
the solid modelling option. In this method a repeatable segment of the cylinder was
modelled with the cut outs of the required diameter. The element mesh density was
then generated and the segment copied 360 around the circumference of the
cylinder. The resulting portion of the cylinder was then copied several times to
obtain the required cylinder length. Figure 6.30 shows this procedure.

Segment copied around


Repeatable segment the circumference to Z

make up a portion of the


Y X
with cut outs and
element mesh cylinder Portion of meshed
generated cylinder copied to
make up the
required cylinder
Figure 6.30: Procedure for element mesh generation length.
of cylinder model with multiple perforations.

Once the FE cylinder models had been generated, the cylinders were analysed for
non-linear buckling with initial wall imperfections, to simulate a more realistic
model. To obtain the initial imperfections, an eigenvalue buckling analysis was first
performed on each FE cylinder model and several modes from each analysis were
expanded. The eigenvalue buckling analysis represents the theoretical buckling
strength of an ideal linear elastic structure. The wall displacements of each mode
shape were entered separately into the model data base as an updated geometry of
the cylinder, and the model analysed first for axial loads and then lateral pressures.
Each mode shape had to be entered as an initial imperfection, since it was not
possible to state with any certainty which mode shape would result in the lowest
buckling load.

Once the geometry of the FE models had been updated, both axial and lateral non-
linear buckling analyses were performed separately on each updated FE cylinder
6.28 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

model for all the eigenvalue modes expanded. Since the yield stress of the steel
sheets used for making the cylinders was given as 180MPa by BHP, the loads
corresponding to a nodal von Mises stress of 180MPa was chosen as the buckling
failure load of the cylinder. From these non-linear analyses, a graph was plotted of
the buckling failure load for each eigenvalue mode shape representing the initial
imperfection. The mode resulting in the lowest buckling load was then used as the
‘correct’ mode shape in the final non-linear buckling analysis of a cylinder with
multiple perforations subjected to a combined axial and lateral pressure. An
example of the buckling failure loads corresponding to the mode shapes has been
shown diagrammatically in figure 6.31, and does not represent any real analysis.

120 100

100
80

80
60
Buckling Load (kN)

Buckling Load (kN)

60
40
40

20
20
lowest buckling load lowest buckling load
0 0
0 5 10 15 20 25 0 5 10 15 20 25
Eigenvalue Mode Shape Eigenvalue Mode Shape

Axial Loads only Lateral pressures only

Figure 6.31: Diagrammatic plot of Eigenvalue mode corresponding


to Lowest Buckling Stress.

From these analyses two different mode shapes resulted in the lowest buckling load
of the cylinders, one for an axial load and one for a lateral pressure. Therefore, a
non-linear analysis with a combined axial and lateral pressure was performed on
the FE cylinder models for both mode shapes as an initial wall imperfection. The
method of load application in the FE analyses was similar to that of the perforated
cylinder experiments in the laboratory. The external lateral pressures were applied
first and kept constant while the axial load was applied incrementally until the nodal
PERFORATED CYLINDERS : FINITE ELEMENT ANALYSIS 6.29

von Mises stress reached 180MPa. This resulted in an interactive curve of the axial
load vs the lateral pressure for each cylinder.

In the non-linear analyses, the loads were applied gradually using ramped substeps
so that an accurate solution could be obtained. The difference between ramped and
stepped loads has been shown in figure 6.32. The lateral pressures were applied on
the cylinder first, in three substeps and the subsequent axial loads were applied in
10 substeps. To ensure convergence, 5 equilibrium iterations were calculated at
each substep.

Full load applied


at substep 1
Load

Load
Loadstep 1
Substep 3

Loadstep 1 Loadstep 1
Substep 3 Substep 1

Time Time

(a) Stepped Loads (b) Ramped Loads

Figure 6.32: Application of load increments:


(a) Stepped Loads, (b) Ramped Loads

To obtain the load-displacement graph, three nodes were monitored for each
analysis and the output from these nodes was transferred to an excel spreadsheet.
For each node the variable to be monitored was specified before the analysis was
started. In the case of an axial load, the circumferential nodes were monitored for
vertical displacements while the nodes on the body of the cylinder were monitored
for the von Mises stresses. For lateral pressures the nodes were monitored for an
inwards displacement.
6.30 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

6.4.2 CYLINDER WITH 16.5% OPEN AREA

Figure 6.33 shows the final arrangement of the FE cylinder model with 50mm
diameter cut outs.
The cylinder was modelled with 96 Nodes
around the circumference. A total of 5274
nodes were used with 4392 elements to
analyse the cylinder. The global axes
have been shown at the base of the
cylinder. To simulate the capped end
conditions of the laboratory experiments,
the base nodes were fully fixed in all
directions and top nodes were free to
move down vertically.

The axial load was applied at each


circumferential node and a ramped load
controlled analysis undertaken. Once the
Z analysis had reached the total load, the
Z nodal von Mises stresses were checked
Y X
X and the load corresponding to a 180MPa

Figure 6.33: Finite Element stress was accepted as the failure


Model of cylinder with 50mm buckling load. For the cases when the
diameter cut outs
stresses were not exactly 180MPa, a
linear interpolation was undertaken.

Before an eigenvalue bucking analysis could be done, a static analysis was first
performed on the cylinder with the option of prestress set to on. The subsequent
eigenvalue buckling analysis performed was the subspace buckling analysis. A total
of 26 modes were extracted with a subspace working size of 5 modes. This resulted
in the least computer time at no cost of accuracy. The extracted mode shapes have
been shown in Appendix H1, from an end view of the cylinder.

To determine the failure load of each mode shape the applied axial load was
multiplied by the number of circumferential nodes and by the load factor for each
mode. For the FE cylinder with 50mm diameter cutouts, the graph of Axial Non-
Linear Buckling loads for each mode shape has been shown in figure 6.34. Similarly
the graph of Lateral Non-Linear Buckling Pressure for each mode shape has been
PERFORATED CYLINDERS : FINITE ELEMENT ANALYSIS 6.31

shown in figure 6.35.

75
22
70 25
Non Linear Failure Load (kN)

12 21
65 23&24
18 Mode 4
Y

60 26 Z X

7&8
55 16

50
2&3 5&6
45
1
4 Lowest Buckling Load
40 from mode shape 4.
85 90 95 100 105 110 115 120 Non-Linear Buckling Load
Eigenvalue Buckling Load (kN) = 44.25kN
Eigenvalue Load =
95.11kN

Figure 6.34: Graph of Non-linear Axial Buckling Load


for each Eigenvalue buckling mode shapes.

115
4
110 5&6
Y
105 Mode 23 & 24Z
Non Linear Buckling (kPa)

X
2&3
100 18

95
8 21 25
90 1 13
22
26&27
85 Lowest Buckling Load
23&24 from mode shape 24.
80 Non-Linear Buckling
Load = 82.85kPa
85 90 95 100 105 110 115 120
Eigenvalue Load =
Eigenvalue Buckling (MPa) 113.29kN

Figure 6.35: Graph of Non-linear Lateral Buckling Load for


each Eigenvalue buckling mode shape.
6.32 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

The maximum displacement of the wall in applying the initial geometric


imperfection was set equal to one wall thickness, which was 0.8mm. From figures
6.34 and 6.35 it can be seen that different mode shapes result in the lowest
buckling load for an axial load compared to a lateral pressure application. While it is
not possible to predict which modes shape produce the lowest buckling failure,
there is a general trend that the first few mode shapes produce lower buckling
loads in an axial load application. Similarly the higher mode shapes produce the
lower failure loads for a buckling analysis due to a lateral pressure. In the case of
the cylinder with 50mm diameter holes, mode shape 4 produced the lowest failure
load in an axial buckling analysis while mode shape 23 and 24 produced the lowest
failure load in a lateral pressure buckling analysis.

Once these modes were established, the non-linear buckling analyses due to a
combined axial and lateral pressure were performed for both mode shapes 4 and
24. Note that mode 23 has the same shape as mode 24 and therefore there was no
difference in which shape was applied as the initial imperfection. The resulting
interaction curve for a cylinder with 50mm diameter holes has been shown in figure
6.36.

60
Axial load dominates

50

Mode shape 24
Total Axial Load (kN)

40
Mode shape 4

30

20
Lateral
pressure
10 dominates

0
0 20 40 60 80 100
Lateral Pressure (kPa)

Figure 6.36: Combined Axial and Lateral Load for mode shapes 4 and 24

From figure 6.36 it can be seen that analyses on both mode shapes are needed to
produce the final interaction curve. The graph can be divided into two distinct
PERFORATED CYLINDERS : FINITE ELEMENT ANALYSIS 6.33

sections, the first being dominated by the axial load and the second section being
dominated by the lateral pressure. Since mode shape 4 produced the lowest failure
buckling load for an axial load only, it is reasonable that this mode shape as an
initial geometric imperfection gives the lower results in the first section of the
interaction curve given in figure 6.36. The second section of the interaction plot
comes from an analysis of Mode shape 24 as an initial geometric imperfection since
this mode gives the lower values of a combined loading.

Since the results from the FE analysis are higher than the values produced from the
laboratory tests, another set of analyses was performed on the FE model using the
same mode shapes but with a wall displacement of 1.6mm as the maximum initial
imperfection scale. The interaction curve for this imperfection scale has been shown
in figure 6.37. However since it has already been shown that mode shape 4
dominates the axial loading and mode shape 24 dominate the lateral loading, only
these final two curves have been shown.

20 Mode shape 4
18
16 Mode shape 24
14
Total Axial Load (kN)

12

10

8
6

4
2

0
0 10 20 30 40 50 60 70 80 90
Lateral Pressure (kPa)

Figure 6.37: Interaction curve for FE model with 50mm


diameter holes and a wall displacement of 1.6mm

From figure 6.37 it can be seen that the axial capacity of the cylinder has dropped
while there has been very little change in the capacity of the cylinder in carrying a
lateral pressure.
6.34 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

The combined graphs of the two scales of wall imperfections has been shown in
figure 6.38.

45
Imperfection
40 scale = 1 wall
thickness
35

30
Total Axial Load (kN)

25 Imperfection
scale = 2 wall
20 thicknesses

15

10

0
0 10 20 30 40 50 60 70 80 90
Lateral Pressure (kPa)

Figure 6.38: Interaction plot for cylinder with 50mm diameter


cut outs and wall imperfection scales of 1 and 2.

As can be seen from figure 6.38, doubling the scale of wall imperfections has the
effect of reducing the axial capacity of the cylinder while there is very little effect on
the lateral pressure of the cylinder.
FINITE ELEMENT ANALYSIS: 76mm DIAMETER CUT OUTS 6.35

6.4.3 CYLINDER WITH 36.6% OPEN AREA


Figure 6.39 shows the final arrangement of the FE cylinder model with 76mm
diameter cut outs.
The cylinder was modelled with 70 nodes
around the circumference. A total of 2826
nodes were used with 2448 elements to
analyse the cylinder. The global axes have
been shown at the base of the cylinder. To
simulate the capped end conditions of the
laboratory experiments, the base nodes
were fully fixed in all directions and top
nodes were free to move down vertically.

The axial load was applied at each


circumferential node and a ramped load
controlled analysis undertaken. Once the
analysis had reached the total load, the
Z
Z
nodal von Mises stresses were checked and
Y X the load corresponding to a 180MPa stress
X
was accepted as the failure buckling load.
Figure 6.39: Finite Element
Model of cylinder with 76mm For the cases when the stresses were not
diameter cut outs exactly 180MPa, a linear interpolation was
undertaken.

Before an eigenvalue bucking analysis could be done, a static analysis was first
performed on the cylinder with the option of prestress set to on. The subsequent
eigenvalue buckling analysis performed was the subspace buckling analysis. A total
of 26 modes were extracted with a subspace working size of 5 modes. This resulted
in the least computer time at no cost of accuracy. The extracted mode shapes have
been shown in Appendix H2, from an end view of the cylinder.

To determine the failure load of each mode shape the applied axial load was
multiplied by the number of circumferential nodes and by the load factor for each
mode. For the FE cylinder with 50mm diameter cutouts, the graph of Axial Non-
Linear Buckling loads for each mode shape has been shown in figure 6.40. Similarly
6.36 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

the graph of Lateral Non-Linear Buckling Pressure for each mode shape has been
shown in figure 6.41.

30
Non Linear Failure Load (kN)

13
28 22 ModeY 1
18 Loadfactor
Z X

20 5.396
26
6

3
24

2 5
22
1 Lowest Buckling Load
from mode shape 1.
20 Non-Linear buckling
30 35 40 45 50 Load = 21.98kN
Eigenvalue Buckling Load (kN) Eigenvalue Load =
33.61kN

Figure 6.40: Graph of Non-linear Axial Buckling Load for each


Eigenvalue buckling mode shapes.

85
Non Linear Buckling (kPa)

2
80 1 Mode 18&19
Y
Loadfactor
Z X

7.347
75 8
13
5
25
22
70

3 20
18
65
Lowest Buckling Load
from mode shape 18.
60 Non-Linear buckling
30 35 40 45 50 Load = 66.73kPa
Eigenvalue Buckling (MPa) Eigenvalue Load =
45.178kN

Figure 6.41: Graph of Non-linear Lateral Buckling Load for each


Eigenvalue buckling mode shape.
FINITE ELEMENT ANALYSIS: 76mm DIAMETER CUT OUTS 6.37

The maximum displacement of the wall in applying the initial geometric


imperfection was set equal to one wall thickness, which was 0.8mm. From figures
6.40 and 6.41 it can be seen that different mode shapes result in the lowest
buckling load for an axial load compared to a lateral pressure application. While it is
not possible to predict which modes shape produce the lowest buckling failure,
there is a general trend that the first few mode shapes produce lower buckling
loads in an axial load application. Similarly the higher mode shapes produce the
lower failure loads for a buckling analysis due to a lateral pressure. In the case of
the cylinder with 76mm diameter holes, mode shape 1 produced the lowest failure
load in an axial buckling analysis while mode shape 18 and 19 produced the lowest
failure load in a lateral pressure buckling analysis.

Once these modes were established, the non-linear buckling analyses due to a
combined axial and lateral pressure were performed for both mode shapes 1 and
18. Note that mode 19 has the same shape as mode 18 and therefore there was no
difference in which shape was applied as the initial imperfection. The resulting
interaction curve for a cylinder with 50mm diameter holes has been shown in figure
6.42.

30 Axial load
dominates
Mode
25
shape 18
Total Axial Load (kN)

20
Lateral
pressure
15 dominates

10
Mode
shape 1
5

0
0 10 20 30 40 50 60 70 80
External Lateral Pressure (kPa)

Figure 6.42: Combined Axial and Lateral Load for mode shapes 1 and 18

The shape of this plot is similar to the plot for the cylinder with 51mm diameter cut
outs. There are two distinct sections to the curves, the first being dominated by the
axial load and the second dominated by the lateral pressure.
6.38 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Since this plot produced results greater than those from the laboratory tests, the
analysis was redone for an imperfection of twice the wall thickness. the results of
these analyses have been shown in figure 6.43 for mode shape 1 and mode shape
18.

10 Mode shape 1

8
Total Axial Load (kN)

6
Mode shape 18

0
0 10 20 30 40 50 60 70
Lateral Pressure (kPa)

Figure 6.43: Interaction curve for FE model with 76mm


diameter cut outs and a wall displacement of 1.6mm

The combined graphs of the two scales of wall imperfections has been shown in
figure 6.44.

25
Imperfection
scale = 1 wall
20 thickness
Total Axial Load (kN)

15

Imperfection
10 scale = 2 wall
thicknesses

0
0 10 20 30 40 50 60 70
Lateral Pressure (kPa)

Figure 6.44: Interaction plot for cylinder with 76mm diameter


cut outs and wall imperfection scales of 1 and 2
FINITE ELEMENT ANALYSIS: 76mm DIAMETER CUT OUTS 6.39

6.4.4 SOLID CYLINDER

Figure 6.45 shows the final arrangement of the FE cylinder model with no cut outs.

The cylinder was modelled with 72 Nodes


around the circumference. A total of 4392
nodes were used with 4320 elements to
analyse the cylinder. The global axes have
been shown at the base of the cylinder. To
simulate the capped end conditions of the
laboratory experiments, the base nodes were
fully fixed in all directions and top nodes were
free to move down vertically.

The axial load was applied at each


circumferential node and a ramped load
controlled analysis undertaken. Once the
Z analysis had reached the total load, the nodal
Z von Mises stresses were checked and the load
Y X X corresponding to a 180MPa stress was
Figure 6.45: Finite element accepted as the failure load. For the cases
model of cylinder with when the stresses were not exactly 180MPa, a
no holes
linear interpolation was undertaken.

Before an eigenvalue bucking analysis could be done, a static analysis was first
performed on the cylinder with the option of prestress set to on. The subsequent
eigenvalue buckling analysis performed was the subspace buckling analysis. A total
of 26 modes were extracted with a subspace working size of 5 modes. This resulted
in the least computer time at no cost of accuracy. The extracted mode shapes have
been shown in Appendix H3, from an end view of the cylinder.

To determine the failure load of each mode shape the applied axial load was
multiplied by the number of circumferential nodes and by the load factor for each
mode. The graph of Axial Non-Linear Buckling loads for each mode shape has been
shown in figure 6.46. Similarly the graph of Lateral Non-Linear Buckling Pressure
for each mode shape has been shown in figure 6.47.
6.40 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

160 5

150 3 9 19
Y
Non Linear Failure Load (kN)

11
Mode shape 18 Z X
140 6 Factor=80,2200
10
22
130 1&2

17
120 25

110 18
Lowest Buckling Load
100 from mode shape 18.
550 570 590 610 630
Non-Linear buckling
Load = 106.96kN
Eigenvalue Buckling Load (kN) Eigenvalue Load =
577.58kN

Figure 6.46: Graph of Non-linear Axial Buckling Load for each


Eigenvalue buckling mode shape.

320
3 Y

280 Mode shape 5 Z X


Non Linear Buckling (kPa)

Factor=77,4023

240
1

200

160
10
16 25
6 18 20 Lowest Buckling Load from
120
11 22 mode shape 5.
Non-Linear buckling
5
80 Load = 85.42kPa
550 570 590 610 630 Eigenvalue Load =
Eigenvalue Buckling (MPa) 557.3kN

Figure 6.47: Graph of Non-linear Lateral Buckling Load for each


Eigenvalue buckling mode shape.
FINITE ELEMENT ANALYSIS: 76mm DIAMETER CUT OUTS 6.41

As in the analyses of the 51mm and 76mm diameter cut out areas, the maximum
displacement of the wall in applying the initial geometric imperfection was set equal
to one wall thickness, which was 0.8mm. Again, different mode shapes result in the
lowest buckling load for an axial load compared to a lateral pressure application.
Unlike the analyses for the perforated cylinders, the lower mode shapes produce
the lower failure loads for a buckling analysis due to a lateral pressure. In the case
of the solid cylinder, mode shape 18 produced the lowest failure load in an axial
buckling analysis while mode shape 5 produced the lowest failure load in a lateral
pressure buckling analysis.

Once these modes were established, the non-linear buckling analyses due to a
combined axial and lateral pressure were performed for both mode shapes 18 and
5. The resulting interaction curve for the solid cylinder has been shown in figure
6.48.

Axial load
350 dominates

Mode
300 shape
18
Total Axial Load (kN)

250

200

150 Lateral
pressure
dominates
100
Mode
shape
50 5

0
0 20 40 60 80 100 120 140
Lateral Pressure (kPa)

Figure 6.48: Combined Axial and lateral Load for mode shapes 18 and 5

The shape of this plot is similar to that for the perforated cylinder analyses with two
distinct sections to the graph, ie axially dominated and laterally dominated.
Although no tests were done on solid cylinders, the finite element analyses on solid
cylinders was undertaken to form a reference for the perforated cylinders.
Therefore the analysis was repeated for an imperfection scale of two wall
thicknesses, ie 1.6mm. These results have been plotted in figure 6.49.
6.42 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

250
Mode
shape 18
200
Total Axial Load (kN)

150

Mode
100 shape 5

50

0
0 20 40 60 80 100 120 140
Lateral Pressure (kPa)

Figure 6.49: Interaction curve for a solid cylinder and a wall


displacement of 1.6mm

The combined graphs of the two imperfection scales have been shown in figure
6.50.

160

140 Imperfection
scale = 1 wall
120 thickness
Total Axial Load (kN)

100

80 Imperfection
scale = 2 wall
60 thicknesses

40

20

0
0 20 40 60 80 100 120
External lateral pressure (kPa)

Figure 6.50: Interaction plot for a solid cylinder subjected to a combined


axial and external lateral pressure for wall imperfection scales of 1 and 2.
COMPARISON BETWEEN FINITE ELEMENT ANALYSES AND LABORATORY TESTS 6.43

6.5 COMPARISON WITH LABORATORY TESTS

6.5.1 CYLINDER WITH 16.5% OPEN AREA

The interaction curve from the laboratory tests on the cylinder with 51mm diameter
cut outs has been shown in figure 6.51. Included in this graph are the curves from
the finite element analyses for a geometric imperfection of 1 and 2 wall
thicknesses.

45

40 FE analysis:
WI = 0.8mm
35

30
Laboratory
25 test results

20
Total Axial Load (kN)

15

10

5 FE analysis:
WI = 1.6mm
0
0 20 40 60 80 100
External Lateral Pressure (kPa)

Figure 6.51: Interaction curves from finite element analyses


and laboratory tests on cylinder with 16.5% open area.

From figure 6.51 it can be seen that the wall imperfection of the cylinders tested in
the laboratory lies in between 1 and 2 times the wall thickness as predicted by the
finite element analyses. This is only true for the section of the graph where the
axial load dominates. For the region of the graph where the lateral pressure
dominates, the wall imperfection scale of the cylinders in the laboratory tests lies
below the region predicted from the finite element analyses. The lateral pressure
from the laboratory test results is approximately 50% of that predicted by the finite
element analyses. It is therefore reasonable to assume the safe working loads of
the cylinder, as shaded in figure 6.51, lies in the area bounded by the laboratory
tests and the finite element analyses for a geometric imperfection of 2 wall
thicknesses.
6.44 DES NORTJE; PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

From the test results the equation of the line for the region of the graph dominated
by the lateral pressure has been given in chapter 6.3.2 and has been repeated as
follows:

Nx16.5 = 124.55 - 2.65q (6.10 repeated)

where Nx16.5 is the total axial load in kN applied on a cylinder with 16.5% open area,

and q is the external lateral pressure in kPa.

The equation of the straight line for the region of the interaction dominated by the
total applied axial load is as follows:

Nx16.5 = 17.29 + 0.0017q (6.13)

where the symbols are as before.

6.5.2 CYLINDER WITH 36.6% OPEN AREA

Similarly, the interaction curves from the laboratory tests on the cylinder with
76mm diameter cut outs has been shown in figure 6.52. Also included in this graph
are the curves from the finite element analyses for a geometric imperfection of 1
and 2 wall thicknesses.

25

FE Analysis
WI = 0.8mm
20

15 Laboratory
test results
Total Axial Load (kN)

FE Analysis
10 WI = 1.6mm

0
0 10 20 30 40 50 60 70
External Lateral Pressure (kPa)

Figure 6.52: Interaction curves from finite element analyses


and laboratory tests on cylinder with 36.6% open area.
COMPARISON BETWEEN FINITE ELEMENT ANALYSES AND LABORATORY TESTS 6.45

Figure 6.52 is similar to figure 6.51 since the laboratory test results on the cylinder
with 76mm diameter cut outs also lies between the two curves from the finite
element analyses. From the laboratory test results, the region of the graph
dominated by the lateral pressure appears to be approximately 64% lower than the
results from the finite element analyses. Since this result is similar to the case of
the cylinder with 51mm diameter cut outs it is again reasonable to assume the safe
working loads of the cylinder lie within the area bounded by the laboratory test
results and the finite element analyses for a geometric imperfection of 2 wall
thicknesses. This area has been shaded in figure 6.52.

Equation 6.12 given in chapter 6.3.2 describes the straight line of the graph for the
region dominated by the lateral pressure in figure 6.52. This equation has been
repeated below:

Nx36.6 = 130.41 - 3.105q (6.12 repeated)

where Nx36.6 is the total axial load in kN applied on a cylinder with 36.6% open

area, and q is the external lateral pressure in kPa.

The equation for the straight line in figure 6.52 dominated by the total axial load
applied is given as follows:

Nx36.6 = 9.08 – 0.0135q (6.14)

6.5.3 INTERACTION PLOTS FOR CYLINDERS WITH


MULTIPLE PERFORATIONS.

Using equations 6.10, 6.12, 6.13 and 6.14, the interaction plots of the safe working
loads for cylinders with 16.5% open area (51mm diameter cut outs) and 36.6%
open area (76mm diameter cut outs) have been shown in figure 6.53. Included in
figure 6.53 is the interaction plot for a cylinder with no cut outs and a geometric
imperfection of two wall thicknesses.

From figure 6.53 it can be seen that a cylinder with an open area of 16.5% can
sustain only approximately 16% of the total axial load of a solid cylinder. Similarly,
a cylinder with 36.6% open area can carry only approximately 8% the total axial
load of a cylinder with no cut outs.
6.46 DES NORTJE; PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

110
Solid Cylinder
100 16.5% Open Area
90 36.6% Open Area
Total Axial Load (kN)

80

70

60

50

40

30

20

10

0
0 10 20 30 40 50 60 70 80 90 100
External lateral pressure (kPa)

Figure 6.53: Interaction curves for cylinders with 16.5% and 36.6% Open
Area: Geometric Imperfection = 2 x wall thickness

For a cylinder with no cut outs, the region of the graph dominated by the lateral
pressure has the following equation:

Nx = 306.45 – 3.496q (6.15)

where Nx is the total axial load in kN, applied on a cylinder with no cut outs, and q

is the external lateral pressure in kPa.

Closer inspection of figures 6.51, 6.52 and 6.53, suggests that the portion of the
graphs dominated by the axial load could be estimated by a horizontal line. For the
cylinders with 16.5% and 36.6% open area these lines have been drawn horizontal
are as shown in figure 6.53. It is therefore reasonable to assume these lines will
also be horizontal for cylinders with other ratios of open areas.

Furthermore, from figures 6.51, 6.52 and 6.53, the slopes of the lines representing
the region of the curve dominated by the lateral pressure, appear to be very
similar. From equations 6.10, 6.12 and 6.15 the slopes of these lines are all
negative and are 2.65, 3.105 and 3.496 respectively. Taking an average of the
three slopes results in a slope of approximately –3.08.
COMPARISON BETWEEN FINITE ELEMENT ANALYSES AND LABORATORY TESTS 6.47

Since the general shape of the interaction curves have now been established, the
interaction curves of cylinders with multiple perforations subjected to a combination
of an axial load and an external lateral pressure can easily be plotted. To produce
these graphs, only two points are needed: one is the capacity of the cylinder with
an axial load only, and the other point is the capacity of the cylinder with an
external lateral pressure only.

The interaction curves for cylinders with, say, 5% and 10% open area will be drawn
using the method described above.

From the literature survey it was shown that the shape of the curve for the total
applied axial load as a function of the percentage open area of the cylinder, is a
power series. Therefore, using the value for a cylinder with no cut outs, and the
points for cylinders with 16.5% and 36.6% open area, the power series curve has
been shown in figure 6.54 for the assumption of a geometric imperfection of 2
times the wall thickness. Since it was shown in section 6.5.1 and 6.5.2 that the real
cylinder values lie approximately within the range of an imperfection of 2 times the
wall thickness, this would be a reasonable value to use in general.

120

100
Total Applied Axial Load (kN)

80

60
48.86kN
40

27.05kN
20

0
0 10 20 30 40
Percent Open Area

Figure 6.54: Capacity of cylinders with a geometric imperfection


of 2 x wall thickness, subjected to an axial load
6.48 DES NORTJE; PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

The results from the work by Starnes Jr (1972) have been given in figure 6.5 of the
literature survey, chapter 6.1. From this figure it can be seen that a cylinder with a
cut out value of rc/r (the ratio of the cut out radius to the cylinder radius) of 0.025

(2.5%) can carry the same axial load as a cylinder with no cut outs. Therefore in
figure 6.54, it has been assumed that a cylinder with 2% open area can sustain the
same load as a cylinder with no cut outs.

The equation of the power series given in figure 6.54 is as follows:


(-0.8527)
A = 192.72 * C (6.16)

where A is the total axial load in kN and C is the percentage open area of the
cylinder as defined by figure 6.7a in chapter 6.1.

Using equation 6.16, the total axial load which cylinders with 5% and 10% open
area can sustain are 48.86kN and 27.05kN respectively.

The capacities of the cylinders with no cut outs, a 16.5% open area and a 36.6%
open area due to an external lateral pressure only, has been shown in figure 6.55.
It has been assumed that a similar relationship exists for the lateral pressure as for
the axial load and therefore a power series trendline has been shown through these
points. The equation of this curve is as follows:
(-0.2558)
q = 104.18 * C (6.17)

where q is the external lateral pressure in kPa and C is the percentage open area of
the cylinder.

Using equation 6.17, the external lateral pressure which cylinders with 5% and
10% open area can sustain are 69.02kPa and 57.81kPa respectively.

Now the interaction curves for cylinders with 5% and 10% open area can be
plotted, as all the points needed for the curves have been determined. The results
have been shown in figure 6.56 with the curves for cylinders with 16.5% and
36.6% open areas. The plot for a cylinder with no cut outs has also been shown.
COMPARISON BETWEEN FINITE ELEMENT ANALYSES AND LABORATORY TESTS 6.49

90

80

External Lateral Pressure (kPa)


70
69.02kPa

60
57.81kPa

50

40

30
0 10 20 30 40
Percent Open Area

Figure 6.55: Capacity of cylinders with a geometric imperfection of


2 x wall thickness, subjected to an external lateral pressure only.
Total Axial Load (kN)

110

100

90

80

70 Slope = -3.08
60
5% open area
48.86 50

40
10% open area
30
27.05
20

10

0
0 10 20 30 40 50 60 70 80 90 100
57.81 69.02

External Lateral Pressure (kPa)

Figure 6.56: Interaction curves for cylinders with 5%, 10%, 16.5% and
36.6% Open Area: Geometric Imperfection = 2 x wall thickness.
CONCLUSIONS 7.1

CHAPTER SEVEN
CONCLUSIONS

7.1 SILO WALL PRESSURES

7.1.1 STATIC PRESSURES

Until Janssen’s paper in 1895, engineers calculated the vertical pressure on the
bottom of a silo in accordance with a hydrostatic pressure distribution. However
Janssen proved that the vertical pressure distribution in a silo varies from zero at
the top increasing exponentially towards a maximum asymptote as the depth
increases. The Reimberts(1976) conducted experiments, which demonstrated the
vertical pressure distribution varies hyperbolically from zero at the top to a
maximum asymptote at great depths in the silo. Both the Janssen and Reimbert
theories for the horizontal pressure acting normal to the walls of the silo tend
towards the same maximum asymptote. However, the Reimbert theory gives higher
values of horizontal pressures in the upper regions of the silo, and consequently
also in squat silos. In the case of a squat silo or in the upper regions of the silo, the
horizontal pressure distribution is best described by the Reimbert theory.

Conversely, the Janssen theory for the vertical pressures in the silo tends towards a
larger value than the Reimbert theory at the same depth. The Janssen vertical
pressure distribution gives the highest vertical pressure acting on the bottom of the
silo. This difference between the two theories has been shown in figure 2.5 in
chapter 2.

The prediction of the static vertical pressure in the hopper has been made by
several researchers and the results of their research varies greatly. Walker(1960)
assumed the principal stress planes were vertical and horizontal, consequently
Walker states that there can be no shear on the vertical planes. Hence Walker gives
the vertical pressure distribution in the hopper as a hydrostatic variation increasing
linearly from the value of the vertical pressure at the transition to a maximum
value at the discharge gate.

The Walker static normal wall pressure in the hopper is given as a constant times
the vertical pressure, where this constant is a function of only the wall friction angle
and the material friction angle, and is not influenced by the hopper half angle.
7.2 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

In direct contrast to Walker, Jenike(1968) states that the normal wall pressures in
the hopper have been shown experimentally to increase linearly from zero at the
vertex to an intermediate value at some height in the hopper. Beyond this
intermediate value the Jenike normal wall pressure distribution increases linearly to
a maximum value, equal to a constant times the Janssen vertical pressure at the
transition. This constant is a function of the hopper half angle and the wall friction
angle and does not depend on the material friction angle. This has been shown in
figure 2.17 of chapter 2.

Conversely, it has been shown in this thesis by the equilibrium of forces acting on a
horizontal slice of material in the hopper that the resulting differential equation of
the vertical pressure is a boundary value problem. This boundary value problem
was solved using the Runge-Kutta method of numerical integration as given in
Appendix A. The resulting distribution of the vertical pressure increases from a
minimum at the discharge gate to a maximum at the transition. The shape of this
curve has been shown in figure 2.12. This pressure distribution is in agreement
with the Jenike distribution at the discharge gate only, while giving larger values at
the transition, as shown in figure 2.22.

The equilibrium of a slice method shows that the hopper half angle and wall friction
angle have a small effect on the vertical pressure. The assumption of a stress ratio
increasing from 0.3 to 0.9 affects the pressure within the hopper, while not
affecting the value at the gate or transition. This has been shown in figure 2.13c
and 2.13d. It is believed in this thesis that the equilibrium slice method gives a
better indication of the static vertical pressure distribution in a converging channel.

7.1.2 DYNAMIC PRESSURES

Walters(1973a) solves the equilibrium of forces on a horizontal slice in the


cylindrical section of the silo during material flow. Walters assumed the average
vertical stress across the elemental slice was related to the vertical stress at the
wall by a distribution factor F. This distribution factor was determined from the
Mohr circle given in figure 2.24 and is a function of the material friction angle and
the wall friction angle. The Walters distribution factor varies from 1 up to a
maximum value of 3, and shows that the vertical pressure at the wall is higher than
at the centre of the cylindrical section. The Walters theory for the dynamic vertical
pressures in the cylindrical section of the silo predicts values which are
approximately 3 times smaller than the Janssen static vertical pressures. However
CONCLUSIONS 7.3

the Walters pressures normal to the wall during material flow are approximately 2.6
times larger than the Janssen static horizontal pressures.

The Walters (1972b) theory for the vertical pressure distribution in the hopper
during material flow is based on the equilibrium of forces on a horizontal elemental
slice of material. The resulting pressure distribution varies linearly from some non-
zero value at the discharge gate increasing to a maximum value at the transition.

The Jenike(1961) radial stress theory for vertical pressures in the hopper is an
exceedingly lengthy calculation requiring extensive algebraic manipulation.
Ultimately Jenike arrives at a linear pressure distribution increasing from zero at
the discharge gate to a maximum value at the transition. The normal pressures
acting on the hopper wall during material flow are given by Jenike(1961) is a
constant times the stress in the material. This constant is a function of the material
friction angle, the hopper half angle and the inclination of the major principal stress
to the vertical.

7.1.3 SWITCH PRESSURES

All researchers conclude there is a third pressure, the switch pressure, which
travels up through the silo. This switch occurs the instant the discharge gate is
opened and the stress state changes from an active state to a passive state. Switch
pressures are generally quoted as a constant times the Janssen static horizontal
pressures in the cylinder.

The Jenike(1973b) switch pressure varies from approximately 2 times the Janssen
pressure in the lower regions of the cylinder to approximately 3 times the Janssen
pressure in the upper regions of the cylinder. Jenike showed that for the same H/D
ratio of the silo, a smaller wall friction angle resulted in a higher switch pressure in
the cylinder. However, keeping the wall friction angle constant and decreasing the
H/D ratio tends to decrease the switch pressure in the cylinder. This has been
shown in figure 2.45 of chapter 2.

Walters(1973) theory for switch pressures predicts the switch to vary from
approximately 5.2 times the Janssen static pressure in the lower regions of the
cylinder up to approximately 3.1 times the Janssen pressure in the upper regions of
the cylinder.
7.4 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Decreasing the H/D ratio of the silo by a factor of 2, for the same friction angles
(wall and material), has the effect of decreasing switch pressure by 11.5% in the
lower regions to 39% in the higher regions of the cylinder. By reducing the material
friction angle by 37.5% (keeping the H/D ratio and the wall friction angle constant)
the magnitude of the switch pressures is decreased by approximately 44%.
Conversely a reduction in the wall friction angle of 25% has the effect of increasing
the switch pressure by 179% at the lowest level to 136% at the highest level in the
silo.

Both the Jenike and Walters theories for determining the switch pressures during
flow, predict the worst case would occur in a tall silo (a high H/D ratio) with smooth
walls (a low wall friction angle). A “short” silo with a rough wall would experience
lower switch pressures, but consequently may not undergo mass flow of the
material. Furthermore, the Walters theory predicts that a material with a high
internal angle of friction will result in higher switch pressures on the wall.

Therefore, the magnitude of the switch pressures on the walls of a steel silo
increases over a period of time since the walls become smoother from the wear of
the material. However, in a concrete silo the walls become rougher due to the wear
of the flowing material and therefore the magnitude of the switch pressures
decreases.

The area of influence of the switch is given as approximately one silo diameter in
both the Jenike and Walters theories. Therefore, these high pressures are localised
as well as being transient effects on the walls. However, the magnitudes vary
greatly ranging from approximately 2 to 8 times the Janssen static pressures.

From the experimental results in this thesis it was shown that the switch pressure
travels up the silo at a speed of approximately 22.69m/s (81.6km/h). Therefore in
a 30m tall silo it takes approximately 1.3 seconds for the switch to reach the top.
Assuming the silo has a diameter of 10m, the effects of the switch will be felt on
the walls for approximately 0.44 seconds. The experimental results show the switch
pressures to be between one to four times greater than the experimental static
pressures. This is smaller than the results from the tests by Blair-Fish and Bransby
(1973) as well as van Zanten and Mooij (1977).
CONCLUSIONS 7.5

7.1.4 STRESS RATIOS

In all the theories, the greatest influence on the magnitude of the static and
dynamic wall pressures for a given bulk material comes from the assumption of a
suitable stress ratio. There are no theories predicting which stress ratio should be
used in calculating the static and dynamic pressures. The recommended minimum
and maximum limits are however given as the active and passive stress ratios,
respectively.

Both Janssen and Jenike recommend a value of between 0.3 and 0.4 for static
filling conditions. Any value less than 0.3 is unrealistically low and should not be
used in determining the static pressures for design purposes.

It was shown from the experimental results in this thesis that the stress ratio is not
constant throughout the depth of the silo. For static conditions in the upper regions
of the silo the stress ratio is the highest and can be approximated by K = 1–
Sin 2 m
. The ratio decreases to a minimum value in the lower regions of the silo,

and should not be less than the stress ratio for active pressures.

However for dynamic conditions, no pattern to the varying values of measured


stress ratios was determined. The stress ratios measured varied from a minimum
value of 0.31 up to a maximum value of 3.2. This confirms the recommendation to
use the passive stress ratio as the maximum limit in the calculations of the dynamic
pressures in the silo.

7.2 ANTI-DYNAMIC TUBE FRICTION DRAG

Both Pieper(1969) and McLean(1985) report that the pressure on an object


submerged in the material is approximately equal to the Janssen static vertical
pressure at the same depth during filling of the silo. However McLean’s formula
(equation 4.1) gives the force during flow of the material as 2.5 times the static
value, whereas Pieper’s test results show the flow force to be nearly four times the
static value. Thus Pieper’s results give values that are approximately 1.6 times
larger than predicted by McLean.

Ooms and Roberts (1985) showed the friction force on the anti-dynamic tube to be
directly related to the stress ratio in the material, the wall friction angle of the tube,
the tube diameter and the tube height. They derived an expression for the static
7.6 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

vertical pressure in a silo with a tube installed, by considering the equilibrium of an


elemental slice. By increasing the tube to silo diameter ratio, dt/D from one quarter

to a half, the vertical pressure in the silo decreases from 84% to 61% of the
Janssen static pressure. Thus, the static vertical pressure in a silo, and hence the
horizontal static pressures on the walls, can be reduced by introducing a larger tube
to silo diameter ratio. This has been shown in figure 4.7 of chapter 4.

Ooms and Roberts(1985) also derived an equation to predict the frictional drag on
the outside of the tube during material flow. In their analysis of the equilibrium of
forces on a horizontal elemental slice, they assumed a constant vertical pressure
acting across the slice. They showed that the frictional drag force increases with
increasing height of the tube, as this is directly related to the tube surface area.
The relationship between the height of the tube and the vertical drag force was
shown to be approximately linear at depths greater than 1.4 times the silo
diameter.

Furthermore, from their theory it was shown that the frictional drag force should be
increased by a factor of 1.5 when the material flows inside and outside the tube at
the same time. This situation occurs when portholes have been provided at the
base of the tube for material flow to take place.

Kaminski and Zubrzycki(1985) state that the flexibility of the tube supports has a
large influence on the measured value of the vertical drag force. Thus when the
tube is allowed to move with the flowing material, the magnitude of the frictional
drag force on the tube will be decreased.

For optimum tube performance and material flow through the holes in the walls of
the tube, Kaminski and Zubrzycki derived three relationships. The first relationship
gives the minimum and maximum limits for the ratio of tube diameter to the
hydraulic radius of the silo. Their second relationship defines the limits for the total
area of the tube perforations as a function of the tube surface area. The third
relationship defines the flow characteristics of the tube perforations as a function of
the silo and discharge gate diameters. These have been shown in figure 4.10 of
chapter 4.

Schwedes and Schulze (1991) derived an expression for calculating the vertical
stress inside the tube during flow. This expression is the same as the Janssen
equation with the exponential term approximately equal to one. Using their
CONCLUSIONS 7.7

expression multiplied by the internal surface area of the tube, results in a friction
force approximately 2 times larger than given by Ooms and Roberts.

In this thesis an expression was derived for the frictional drag on the tube walls
during material flow. It was shown that the drag force is directly related to the
stress ratio in the material. From the experimental results of the drag on the tube,
the upper limit for the stress ratio as given by Blight(1993) should be used. This
limit is stated as follows:

K = 1+Sin m
. (7.1)

A similar approach to Ooms and Roberts in deriving the expression for the drag
force was adopted, however the distribution of the vertical pressure across the
elemental slice was not assumed constant. The shape of the variation has been
shown in figure 4.14 and is given by the following equation:

v 1 0.3 Cos 2 x (7.2)

where x is the distance from the tube wall, is the vertical pressure at distance x

from the tube wall and v is the average Janssen static vertical pressure acting on

the slice. The expression derived in this thesis for the calculation of the total drag
force on the tube is as follows:

h
FF 2 rKt tJ e Ch 1 (7.3)
C

where the constants J and C are as follows:

R2 r2
J (7.4)
3 1
R r
2 4

2 Ks sR Kt tr
C (7.5)
3 1
R r
2 4

Furthermore, it was shown in this thesis that the rate of discharge of the material
influences the magnitude of the drag force acting on the tube. It was shown
experimentally that the relationship between the shear on the walls of the tube and
the plan area of the discharge gate for the experimental model in this research is as
follows:
7.8 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

= 101.85 A (7.6)

where is the shear on the tube walls and A is the plan area of the discharge gate.

7.3 PERFORATED CYLINDERS

From the test results by Tennyson(1968), Starnes Jr(1972) and Scutella(1998) it


was shown in this thesis that the buckling capacity of a thin walled cylindrical shell
with perforations subjected to an axial load can be estimated by a power series as
follows:
-B
Nx/No = C (Ac/A) (7.7)

where Nx is the buckling capacity of a shell with perforations

No is the reference buckling load of a shell without perforations

C, and B are empirically determined constants


Ac is the total area of the cut outs

and A is the surface area of the cylinder.

The buckling capacity of a shell with perforations is therefore a function of the


amount of material that has been removed from the walls of the shell, and is not a
function of the plan area of the cylinder as initially suggested by Tennyson(1968)
and Starnes Jr(1972).

Using the test results from Scutella(1998) it was shown in this thesis that the
reduction in buckling capacity due to an axial load of a cylinder with multiple
perforations is a function of the centre to centre spacing of the cut outs. The
relationship between the buckling capacity of the cylinder and the centre to centre
spacing of the holes for a constant open area, was shown to be linear. This was
shown in figure 6.10 of chapter 6.1.

7.4 INTERACTION CURVES

The finite element analyses were undertaken to enable a comparison with the test
results of the cylinders in the laboratory. Finite element models were set up with
the same open areas as the cylinders used in the laboratory experiments. However
the method of applying an initial wall imperfection
CONCLUSIONS 7.9

From the finite element analyses it was shown that a cylinder with an open area of
16.5% can sustain only approximately 16% of the total axial load of a solid
cylinder. Similarly, a cylinder with 36.6% open area can carry only approximately
8% the total axial load of a cylinder with no cut outs. Similarly it was shown that
cylinders with 16,5% and 36.6% open area can sustain approximately 54% and
46%, respectively, of the external lateral capacity of a cylinder with no cut outs.

Furthermore, from the finite element analyses it was shown that the interaction
curves for thin walled cylindrical shells with multiple perforations subjected to a
combination of axial and lateral pressures consists of two straight lines. Thus the
graph is divided into two regions, the first region being dominated by the axial load.
As the lateral pressure is increased from zero there is very little discernible
reduction in the axial capacity of the cylinder. The second, smaller region of the
graph is dominated by the external lateral pressure. As the external lateral pressure
approaches the value of the lateral capacity of the cylinder, the axial load carrying
capacity of the cylinder is reduced. This has been shown in figure 6.53.

The straight line describing the region of the interaction curve dominated by the
external lateral pressure was shown to have an average negative slope of –3.08,
and passing through the value of the lateral capacity for no axial load applied.

Since the general shape of the interaction curves have been established in this
thesis, the interaction curves of cylinders with multiple perforations subjected to a
combination of an axial load and an external lateral pressure can easily be plotted.
To produce these graphs, only two points are needed: one is the capacity of the
cylinder with an axial load only, and the other point is the capacity of the cylinder
with an external lateral pressure only.
APPENDIX A: IMPLEMENTATION OF THE RUNGE-KUTTA METHOD A.1

CHAPTER 8
APPENDIX A

8.1 IMPLEMENTATION OF THE RUNGE-KUTTA


METHOD

8.1.1 THE RUNGE-KUTTA EQUATIONS

The Runge-Kutte method was chosen to solve to the differential equation 2.31 as
derived in chapter 2 since it has a local truncation error that is proportional to the
5
step size raised to the power of five, ie h . This is a smaller error than in the Taylor
3
and improved Euler methods, which have a truncation error of h , and the simple
2
Euler method, which has a truncation error of h .

The Runge-Kutte method involves a weighted average of the values of (z, ) taken
at different points over the interval zn z zn+1, and is given by:

h
K n1 2 K n2 2 K n3 K n4
n 1 n 6

where Kn1= (zn; Vn)

Kn2= (z n +h/2 ; Vn+K n1 h/2)


(A.1 a,b,c & d)
Kn3= (z n +h/2 ; Vn+K n2 h/2)

Kn4= (z n +h ; Vn+K n3 h)

8.1.2 EQUILIBRIUM SLICE METHOD

The initial value of the stress at the level of the transition, ie z=0, has been
calculated from the Janssen equation. The following values were taken from a
fictitious silo, for the calculations of this example.

Silo diameter = 2m

Depth to transition = 7m

Hopper depth = 2.6m

Hopper half angle = 15 and Tan 15 = 0.267949

Wall fricition angle = 20 and Tan 20 = 0.363970


A.2 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

3
Material bulk density = 17 kN/m
Static stress ratio = 0.4

Step size = 2.6/10 = 0.26. The choice of suitable step sizes has been discussed at
a later stage in this chapter.

The equation 2.32, given in chapter 2, has been re-written as follows:

2 v 2
r0 zTan r0 zTan C v (A.2)
z

The first calculation was done at the step z=0, as follows:


2
K01 = (z0; V0) = (r0-z0Tan - C01 V0 .

The constant C01 is dependent on the depth within the hopper and is calculated as
follows:

1 Tan 20
C 01 2 Tan20 1 z 0 Tan20 1 1 z 0 Tan 20 =1.495417
0.4 0.4
2 2
Therefore K01 = (r0-zTan - C01 V0 = 17(1) -1.495417*50.77889 = -58.935624

The calculation of the second constant K02 was done at step z = 0 +h/2 = 0 .13 as
follows:

2
K02 = (z 0 +h/2 ; V0+K 01 h/2) = (r0-(z0+h/2)Tan - C02 ( )
V0+0.13K01

The constant C02 is calculated as follows:

1 Tan 20
C 02 2 Tan20 1 0.13 Tan20 1 1 0.13 Tan 20 =1.524836
0.4 0.4

2
Therefore: K02 = (r0-(z+h/2)Tan - C02 ( V0+0.13K01 )
2
=17(1-0.13Tan20 ) -1.524836*(50.77889-0.13*58.935624) = -48.746751

Since K03 is also evaluated at the step interval z0+h/2, the value of C03 will be equal

to C02 and the constant K03 becomes:

2
K03 = (z 0 +h/2 ; V0+K 02 h/2) = (r0-(z0+h/2)Tan - C03 ( )
V0+0.13K02

2
K03=17(1-0.13Tan20 ) -1.524836*(50.77889-0.13*48.747651) = -50.766478
APPENDIX A: IMPLEMENTATION OF THE RUNGE-KUTTA METHOD A.3

The last constant K04 was calculated at the step interval, z0+h = 0 .26, as follows:

2
K04= (z 0 +h ; V0+K 03 h)= (r0-(z0+h)Tan - C04 ( V0+0.26K03)

1 Tan 20
where C 04 2 Tan20 1 0.26 Tan20 1 1 0.26 Tan 20
0.4 0.4

Therefore C04=1.554255 , and

2
K04=17(1-0.26Tan20 ) -1.554255*(50.77889-0.26*50.766478) = -41.408287

The vertical pressure at the next step level z1 can now be calculated from:

2 h
r0 z 0 Tan V1 V0 K 01 2 K 02 2 K 03 K 04
6

=50.77889 - 0.26(58.935624 + 2*48.746751 + 2*50.766478 + 41.408287)/6

V1 = 37.806173 kPa.

The calculation of the vertical pressure at the next step level, ie z=0.26, is carried
out in the same way as before, with z=0.26 and z+h/2=0.26+0.13=0.39.

Thus after calculation the following values were found:

C11 = 1.554255 K11 = -44.739010

C12 = 1.583674 K12 = -36.640465

C13 = C12 K13 = -38.307774

C14 = 1.613093 K14 = -30.897003

Now the vertical pressure at the third step level can be calculated as before:

2 h
r0 z1 Tan V2 V1 K11 2 K 12 2 K 13 K 14
6

V2 = 28.033098 kPa

The full set of calculations was set up on a spreadsheet, and the results of the
vertical pressure in the hopper were plotted on a graph as shown in figure A.1:
A.4 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Vertical pressure (kPa)


0 10 20 30 40 50
0

Depth below transition (m)


0.5

1.5

2.5

Figure A.1: Static vertical pressure in a hopper

To determine the most suitable step size to minimise the error in the calculation of the
Runge-Kutta method, four step sizes were considered. These have been tabulated below
along with the results for each step size.

Depth Vertical Pressure in the hopper (kPa)


(m) h=0.1300 h=0.01300 h=0.00325 h=0.00162
0 50.77888 50.77888 50.77888 50.77888
0.13 43.73370 43.58646 43.57504 43.57315
0.26 37.57858 37.39880 37.38471 37.38238
0.39 32.29345 32.08983 32.07378 32.07112
0.52 27.75899 27.53880 27.52138 27.51849
0.65 23.87033 23.63958 23.62128 23.61824
0.78 20.53873 20.29936 20.28059 20.27747
0.91 17.67536 17.43740 17.41849 17.41535
1.04 15.22011 14.98380 14.96501 14.96190
1.17 13.11039 12.87833 12.85988 12.85681
1.3 11.29507 11.06930 11.05135 11.04837
1.43 9.73046 9.51256 9.49523 9.49236
1.56 8.37938 8.17054 8.15394 8.15118
Table A.1: Effect of step size on accuracy
APPENDIX A: IMPLEMENTATION OF THE RUNGE-KUTTA METHOD A.5

From Table A.1 it can be seen that there is convergence to two decimal paces for
step size 0.00325 and step size 0.00162. However, for engineering purposes a step
size of 0.00325 is considered acceptable as the error is only:
(43.57504 - 43.57315) / 43.57315 = 0.004 %

Therefore in the subsequent calculations of the equilibrium slice method, a step size
of 0.00325m has been used.
APPENDIX A: IMPLEMENTATION OF THE RUNGE-KUTTA METHOD A.5

A.2 JENIKE RADIAL STRESS THEORY FOR DYNAMIC


PRESSURES IN THE HOPPER

The equation for the angle of the major principle stress, with respect to the angle
, is given by equation 2.178 as follows:

2
Cos Sin m Cos 2 s Cos m
.............
2s Sin m Cos2 Sin m

sSin m 1 Sin m Cot Sin2 cos 2 1


........... 1 (2.178 repeated)
2sSin m Cos2 Sin m

Equation 2.178 has been solved for a silo with a material friction angle, m, of 20 ,

a stress value of s=10, and an initial value of =0.01 (ie an infinitesimally small
angle off the axis of the hopper). Equation 2.178 cannot be solved for a value of
= 0 , since Cotangent 0 is undefined.
APPENDIX B: CALIBRATION CONSTANTS B.1

APPENDIX B: CALIBRATION CONSTANTS

8.2.1 PRESSURE CELL CALIBRATION

TUBE 1 TUBE 2
9 9

Voltage output
8
Voltage output

8
7 7
6 6
5 5
4 1/0.2725= 4 1/0.2635=
3 3.669 kPa/V 3 3.795 kPa/V

2 2
1 1
Applied pressure (kPa) Applied pressure (kPa)
0 0
0 5 10 15 20 25 30 35 0 5 10 15 20 25 30 35

TUBE 4 TUBE 5
9 9 Voltage output
Voltage output

8 8
7 7
6 6
5 5
4 1/0.2578= 4 1/0.2633=
3.879 kPa/V 3.798 kPa/V
3 3
2 2
1 1 Applied pressure (kPa)
Applied pressure (kPa)
0 0
0 5 10 15 20 25 30 35 0 5 10 15 20 25 30 35

PLATE 6 PLATE 7
9 9
8 8
Voltage output

Voltage output

7 7
6 6
5 5
4 4
1/0.2651= 1/0.2432=
3 3.774 kPa/V 3 4.112 kPa/V
2 2
1 1
Applied pressure (kPa) Applied pressure (kPa)
0 0
0 5 10 15 20 25 30 35 0 5 10 15 20 25 30 35
B.2 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

PLATE 8 PLATE 12
9 9

Voltage output
Voltage output
8 8
7 7
6 6
5 5
4 4
1/0.2449= 1/0.2581=
3 3 3.874 kPa/V
4.083 kPa/V
2 2
1 1
Applied pressure (kPa) Applied pressure (kPa)
0 0
0 5 10 15 20 25 30 35 0 5 10 15 20 25 30 35

Ball Cell 1 MTPM 3


0.4 5

Voltage Output
Voltage output

4
0.3

3
0.2

1/0.0236= 2 1/0.7272=
42.37 kPa/V 1.375 m/V
0.1
1

Applied pressure (kPa) Distance (m)


0 0
0 3 6 9 12 15 18 0 1 2 3 4 5 6

MTPM 10
5
Voltage output

2 1/0.703=
1.422 m/V

Distance (m)
0
0 1 2 3 4 5 6 7
APPENDIX B: CALIBRATION CONSTANTS B.3

8.2.2 ANTI-DYNAMIC TUBE SUPPORT CALIBRATION

Support Beam Middle Rod


160 160
Mass (kg)

Mass (kg)
140 140

120 120

100 100

80 80

60 393.06 kg/V 60
163.01 kg/V
40 40

20 20
Votlage Output (V) Votlage Output (V)
0 0
0 0.1 0.2 0.3 0.4 0 0.2 0.4 0.6 0.8 1

Bottom Rod
160
Mass (kg)

140

120

100

80
295.26 kg/V
60

40

20
Votlage Output (V)
0
0 0.1 0.2 0.3 0.4 0.5
APPENDIC C: CHECK LISTS C.1

APPENDIX C: CHECKLISTS

Pre Static Test Check List


Gate: 180mm Tube: 120mm Static23
Flow down inside of tube 08 Oct 99

1) Check all data acquisition cables and connections

2) Check signals from sensors, switches, strain gges & mtpm

3) Fully wind up multi-turn potential meter pulleys


Ensure pulley ropes are secured during filling

4) Measure height of tube from top of silo

5) Edit input file for static test.


Chnnls Combin Buffers Input freq { Max input freq = 500 000 }
11 0 1 1000
Test Duration = No of buffers * 512/(input freq * 0.095369)
= 5.3686 secs = 0.001491 hrs =
No of data rows = 46 Times no of runs: if using "KENHO"

6) Note calibration constants


Chnnl 1 Chnnl 2 Chnnl 3 Chnnl 4 Chnnl 5 Chnnl 6
Str Gge Tube Mtpm Tube Tube Plate
96.154 3.795 1.375 3.879 3.798 3.774 Cal cnstnt

Chnnl 7 Chnnl 8 Chnnl 9 Chnnl 10 Chnnl 11 Chnnl 12


Plate Plate Switches Mtpm Str Gge Str Gge
4.112 4.083 1 1.422 78.74 95.238 Cal cnstnt

7) Note positions from top of silo


Chnnl 1 Chnnl 2 Chnnl 3 Chnnl 4 Chnnl 5 Chnnl 6
Str Gge Tube Mtpm Tube Tube Plate

Chnnl 7 Chnnl 8 Chnnl 9 Chnnl 10 Chnnl 11 Chnnl 12


Plate Plate Switches Mtpm Str Gge Str Gge

8) Note method of filling silo Flexible hose

3
9) Density measurements: Bucket self weight =87.92g & Vol = 0.001345m
Depth
Weight
C.2 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Pre Dynamic Test Check List


Gate: 180mm Tube: 120mm Dynmic23
Flow down inside of tube 08 Oct 99
1) Level off top surface of sand

2) Measure height of sand from top of silo

3) Remove stopper from pulley and set to zero position


Slow pulley down so it cannot overshoot itself.

4) Ensure cables can move freely

5) Check switches are working

6) Edit input file


Chnnls Combin Buffers Input freq { Max input freq = 500 000 }
11 0 185 10000
Test Duration = No of buffers * 512/(input freq * 0.095369)
= 99.31949 secs = 1 min 39 secs

No of data rows = 8510 Use "SILO" program

7) Note calibration constants


Chnnl 1 Chnnl 2 Chnnl 3 Chnnl 4 Chnnl 5 Chnnl 6
Str Gge Tube Mtpm Tube Tube Plate
96.154 3.795 1.375 3.879 3.798 3.774 Cal cnstnt

Chnnl 7 Chnnl 8 Chnnl 9 Chnnl 10 Chnnl 11 Chnnl 12


Plate Plate Switches Mtpm Str Gge Str Gge
4.112 4.083 1 1.422 78.74 95.238 Cal cnstnt

8) Time to empty silo using stop watch


APPENDIC C: CHECK LISTS C.3

Pressure Cell Orientation Test 23


08 Oct 99
support beam
top of
silo

a-d tube

180mm
diameter
opening

bucket
elevator
OUTPUT FROM DYNAMIC TESTS
TEST -1 TEST 1
Test
10 -1: Ball cell placed 1.1m from top Test
8 1: Ball cell placed 1.1m from top

8.89

Pressure (kPa)
8 6 5.89
Pressure (kPa)

4 3.73
6

4.83 2
4
0
2
-2
Static
0 pressure
-4 =5.89 kPa
Static
-2 -6
pressure
=3.64 kPa
-4 -8
0 10 20 30 Time
40 (seconds)
50 60 70 80 0 20 40 (seconds)
Time 60 80

TEST 1 TEST 2
Test
3 1: Mtpm 3 placed 1.1m from top Test
14 2: Cell 2 placed 2.11m from top
12 11.93
Distance (m)

Pressure (kPa)

10
8
2 6
Cylinder Hopper 4
2
Constant flow 0 Static
rate in the pressure
1 -2 =8.56 kPa
cylinder
Accelerated -4
flow rate in
the hopper -6
-8
0 -10
0 10 20 30 (seconds)
Time 40 50 60 70 0 20 Time (seconds)
40 60 80

Test 2: Cell 4 placed 0.93m from top Test


14 2: Cell 5 placed 1.61m from top
12
11.49 12.26
12
Pressure (kPa)
Pressure (kPa)

10
10
8
8
6 6

4 4
2.13
2 2

0
0
-2
-2
Static -4 Static
pressure
-4 -6 pressure
=4.79 kPa
=6.69 kPa
-6 -8
0 20 40 (seconds)
Time 60 80 100 0 20 Time
40 (seconds)
60 80 100
TEST 2 TEST 3
Test
8 2: Cell 6 placed 0.6m from top Test
15 3: Cell 2 placed 2.28m from top

12 12.36
Pressure (kPa)

Pressure (kPa)
6.67
6
9

4 4.13 6

3
2 Static
0 pressure
=10.95 kPa
0 -3

Static -6
-2 pressure
=2.94 kPa -9

-4 -12
0 20 40(seconds)
Time 60 80 100 -10 0 10 (seconds)
Time 20 30 40

Test 3: Mtpm 3 placed 1.14m from top 24 Test 3: Cell 4 placed 1.39m from top
Distance (m)

Pressure (kPa)

20 20.13
2
16
Cylinder
12

8
1 2.35
4 Static
pressure
0 =7.19 kPa
Hopper
-4

0 -8
-10

10

20

30

40

50

Time (seconds) -15 0 15 Time


30 (seconds)
45 60 75 90

Test
25 3: Cell 5 placed 1.11m from top Test 3: Cell 6 placed 0.69m from top
15 14.67
22.18
Pressure (kPa)

20
Pressure (kPa)

15 10

10 5.18
5
5
1.66

0
0
Static Static
-5 pressure pressure
=7.81 kPa =4.33 kPa

-10 -5
-15 0 15 30 (seconds)
Time 45 60 75 -10 10 Time30
(seconds)
50 70 90
TEST 4
Test
25 4: Cell 2 placed 2.2m from top Test 4: Mtpm 3 placed 0.28m from top

Distance (m)
21.41
Pressure (kPa)

20
3 Cylinder
15
10.4
10

5 2

0
Hopper
-5 Static
pressure 1
-10 =16.8 kPa Change of flow rate
in cylinder due
-15 to side discharge

-20 -1
0 20 Time
40 (seconds)
60 80 100 0 10 Time
20 (seconds)
30 40 50

Test
6 4: Cell 4 placed 1.04m from top Test
4 4: Cell 5 placed 0.28m from top

3 3.26
Pressure (kPa)

Pressure (kPa)

4.24
3 2
2.46
1
0.53
0 0

-1

-3 Static -2
pressure Static
=7.53 kPa -3 pressure
=5.75 kPa
-6 -4

-5

-9 -6
0 20 40 (seconds)
Time 60 80 100 0 20 Time
40 (seconds)
60 80 100

TEST 5
Test
12 5: Cell 2 placed 2.55m from top Test
14 5: Cell 4 placed 2.17m from top
Pressure (kPa)

11.25
Pressure (kPa)

10 12
11.25

10
8
8
6
6
4 5.24
4

2
2

0 0 Cables
cut
-2 -2
0 5 Time
10 (seconds)
15 20 25 30 0 10 20 (seconds)
Time 30 40
TEST 5
Test
14 5: Cell 5 placed 1.26m from top Test
16 5: Cell 7 placed 2.17m from top

Pressure (kPa)
Pressure (kPa)

12.86 13.90
12
12
10
7.89 8
8
6.58
Static
6 4
pressure
=5.92 kPa
4
0
2
-4
0 Cables
cut
-2 -8
0 10 20 (seconds)
Time 30 40 50 60 0 10 20 (seconds)
Time 30 40 50 60

Test
6 5: Cell 8 placed 0.33m from top Test 5: Mtpm 10 placed 0.76m from top
Distance (m)
Pressure (kPa)

5.33
5 3.0
Cylinder
4 electrical
interference
3 2.93
2.0 Change in
2 flow rate

1
1.0 Hopper
0

-1

-2 0.0
0 20 40 (seconds)
Time 60 80 100 0 20 40

TEST 6
Test 5: Cell 12 placed 1.26m from top Test
8 6: Cell 12 placed 0.67m from top
12 11.31
Pressure (kPa)

Pressure (kPa)

6
8
7.42 4.49
4

4
2.28
2

0
0
Cables
cut
-4 Cell passes
-2
through the
transition
-8 -4
0 20 40 (seconds)
Time 60 80 -10 0 10 20 30 40 Time
50 60 70 80
(seconds)
APPENDIX D: DYNAMIC TEST RESULTS D.1

APPENDIX D:
DYNAMIC TEST RESULTS

A table of the results for the first five dynamic tests is shown in table D.1, showing
the values of the first and second measured peak pressures. The total pressures
acting on the silo wall during flow are equal to the dynamic value plus the static
value, which have also been given in table D.1.

To calculate the ratio of the dynamic to static pressure at the transition, the value
of the second peak in the output curve was used for the dynamic pressure, and an
average value of 7.10 kPa was used for the static pressure. This average value has
been calculated from the results of six static tests.

T est Depth Static Dynamic pressures T otal pressure


No from top pressure First Sec ond First Sec ond
of silo (m) peak peak peak peak
-1 1.1 3.64 4.83 8.89 8.47 12.53
1 1.1 5.57 3.73 5.89 9.3 11.46
2 0.6 3.15 4.13 6.67 7.28 9.82
0.93 4.73 2.13 11.49 6.86 16.22
1.61 6.65 - 12.26 18.91
2.11 8.06 - 11.93 19.99
3 0.69 4.04 5.18 14.67 9.22 18.71
1.11 6.5 1.66 22.18 8.16 28.68
1.39 6.64 2.35 20.13 8.99 26.77
2.28 7.93 - 12.36 20.29
2.54 3.88 - 3.88
4 0.28 2.57 0.53 3.26 3.1 5.83
1.04 4.53 2.46 4.24 6.99 8.77
2.2 5.33 10.4 21.41 15.73 26.74
5 1.26 5.66 7.89 12.86 13.55 18.52
2.17 8.29 5.24 11.25 13.53 19.54

Table D.1: Dynamic Test Results


D.2 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN SILOS

TEST -1 TEST 1
Test
10 -1: Ball cell placed 1.1m from top Test
8 1: Ball cell placed 1.1m from top

Pressure (kPa)
8.89
Pressure (kPa)

8 6 5.89

4 3.73
6

4.83 2
4
0
2
-2
Static
0 pressure
-4
=5.89 kPa
Static
-2 -6
pressure
=3.64 kPa
-4 -8
0 10 20 30 Time
40 (seconds)
50 60 70 80 0 20 Time
40 (seconds)
60 80

TEST 2
Test
3 1: Mtpm 3 placed 1.1m from top Test
14 2: Cell 2 placed 2.11m from top
12
Distance (m)

11.93
Pressure (kPa)

10
8
2 6
Cylinder Hopper 4
2
0 Static
Constant flow
pressure
1 rate in the -2 =8.56 kPa
cylinder
Accelerated -4
flow rate in
-6
the hopper
-8
0 -10
0 10 20 30 (seconds)
Time 40 50 60 70 0 20 Time (seconds)
40 60 80

Test 2: Cell 4 placed 0.93m from top Test


14 2: Cell 5 placed 1.61m from top
12
11.49 12.26
12
Pressure (kPa)
Pressure (kPa)

10
10
8
8
6 6

4 4
2.13
2 2
0
0
-2
-2
Static -4 Static
-4 pressure
-6 pressure
=4.79 kPa
=6.69 kPa
-6 -8
0 20 40 (seconds)
Time 60 80 100 0 20 Time
40 (seconds)
60 80 100
APPENDIX D: DYNAMIC TEST RESULTS D.3

TEST 2 TEST 3
Test
8 2: Cell 6 placed 0.6m from top Test
15 3: Cell 2 placed 2.28m from top

12.36

Pressure (kPa)
Pressure (kPa)
12
6.67
6
9

4 4.13 6

3
2
Static
0
pressure
0 -3 =10.95 kPa

Static -6
-2 pressure
=2.94 kPa -9

-4 -12
0 20 40(seconds)
Time 60 80 100 -10 0 10 (seconds)
Time 20 30 40

Test 3: Mtpm 3 placed 1.14m from top 24


Test 3: Cell 4 placed 1.39m from top
Pressure (kPa)
Distance (m)

20 20.13
2
16
Cylinder
12

8
1 2.35
4
Static
pressure
0 =7.19 kPa
Hopper
-4

0 -8
Time (seconds)
-10

10

20

30

40

50
0

-15 0 15 Time
30 (seconds)
45 60 75 90

Test
25 3: Cell 5 placed 1.11m from top Test 3: Cell 6 placed 0.69m from top
15 14.67
Pressure (kPa)

22.18
Pressure (kPa)

20

15 10

10
5.18
5
5
1.66

0
0
Static Static
-5 pressure pressure
=7.81 kPa =4.33 kPa
-10 -5
-15 0 15 30 (seconds)
Time 45 60 75 -10 10 Time30
(seconds)
50 70 90
D.4 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN SILOS

TEST 4
Test
25 4: Cell 2 placed 2.2m from top Test 4: Mtpm 3 placed 0.28m from top
Pressure (kPa)

21.41

Distance (m)
20
3 Cylinder
15
10.4
10

5 2

0
Hopper
-5 Static
1
pressure
-10 Change of flow rate
=16.8 kPa
in cylinder due
-15 to side discharge

-20 -1
0 20 Time
40 (seconds)
60 80 100 0 10 Time
20 (seconds)
30 40 50

Test
6 4: Cell 4 placed 1.04m from top Test
4 4: Cell 5 placed 0.28m from top
Pressure (kPa)

Pressure (kPa)

3 3.26
4.24
3 2
2.46
1
0.53
0 0

-1

-3 Static -2
pressure Static
-3
=7.53 kPa pressure
-6 -4 =5.75 kPa

-5

-9 -6
0 20 40 (seconds)
Time 60 80 100 0 20 Time
40(seconds)
60 80 100

TEST 5
Test
12 5: Cell 2 placed 2.55m from top Test
14 5: Cell 4 placed 2.17m from top
Pressure (kPa)

11.25
Pressure (kPa)

10 12
11.25

10
8

8
6
6
4 5.24
4
2
2

0 0
Cables
cut
-2 -2
0 5 Time
10 (seconds)
15 20 25 30 0 10 20 (seconds)
Time 30 40
APPENDIX D: DYNAMIC TEST RESULTS D.5

TEST 5
Test
14 5: Cell 5 placed 1.26m from top Test
16 5: Cell 7 placed 2.17m from top

Pressure (kPa)
Pressure (kPa) 12.86 13.90
12
12
10
7.89 8
8
6.58
6 4 Static
pressure
=5.92 kPa
4
0
2
-4
0 Cables
cut
-2 -8
0 10 20 (seconds)
Time 30 40 50 60 0 10 20 (seconds)
Time 30 40 50 60

Test
6 5: Cell 8 placed 0.33m from top Test 5: Mtpm 10 placed 0.76m from top
Distance (m)
Pressure (kPa)

5.33
5 3.0
Cylinder
4 electrical
interference
3 2.93
2.0 Change in
2 flow rate

1
1.0 Hopper
0

-1

-2 0.0
0 20 40(seconds)
Time 60 80 100 0 20 40

TEST 6
Test 5: Cell 12 placed 1.26m from top Test
8 6: Cell 12 placed 0.67m from top
12 11.31
Pressure (kPa)

Pressure (kPa)

6
8
7.42 4.49
4

4
2.28
2

0
0
Cables
cut
-4 Cell passes
-2
through the
transition
-8 -4
0 20 40 (seconds)
Time 60 80 -10 0 10 20 30 40 50 60
Time 70 80
(seconds)
D.10 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN SILOS

TEST 15
Test
5 15:Tube 2 placed 2.34m from top Test
6 15: Plate 6 placed 2.34m from top

4 4.3
5
3
4 4.3
2
1 3
0.5
0
2
-1
Pressure (kPa)

1
-2

Pressure (kPa)
-3 0
-4
-1
-5
-2
-6
-7 -3
-2 0 2 4 6 8 (seconds)
Time 10 12 -2 0 2 4 6 8 (seconds)
Time 10 12

18Test 15: Tube 4 placed 1.41m from top 14Test 15: Plate 7 placed 1.41m from top
16.33
15 12 12.2

12 10

9 8

6 6

3 2.4 4
2.4
0 2

-3 0

-6 -2
-5 0 5 10 15 20 Time30
25 (seconds)
35 40 -5 0 5 10 15 20 Time
25 30(seconds)
35 40

20Test 15: Tube 5 placed 0.74m from top 10Test 15: Plate 8 placed 0.74m from top
18
Pressure (kPa)

17.5 8
16
7.2
Pressure (kPa)

14
6
12
10 4
8
2.1
6 2

4
2.1 0
2
0 -2
-2
-4 -4
-10 0 10 20 30 Time50
40 (seconds)
60 70 -10 0 10 20 30 40
Time50 60 70
(seconds)
D.12 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN SILOS

TEST 17
Test 17: Tube 5 placed 0.88m from top Test 17: Plate 8 placed 0.88m from top

Pressure (kPa)
1.9 2
Pressure (kPa)

1.4

1 1

0 0

-1 -1
cells flow down the tube
cells flow down the tube
-2 -2
-10 10 30 50 70 90 110
Time (seconds) -10 10 30 Time
50 (seconds)
70 90

Test 17: Tube 4 placed 1.37m from top


3 Test 17: Mtpm 10 placed 0.88m from top
Pressure (kPa)

Distance (m)

2.3
2
0.0365m/s
2
Average flow
in the tube

1
1

0 0
-20 0 20 40 60 Time
80 (seconds)
100 120 0 20 40 60 80 (seconds)
Time 100 120

TEST 18
Test
8 18: Tube 2 placed 2.3m from top Test 18: Tube 4 placed 1.52m from top
12

10 10.2
Pressure (kPa)
Pressure (kPa)

6
4.8 8
4
6

2 4
1.3

0 2

0
-2
-2
-4
-4

-6 -6
-50 50 150 250 (seconds)
Time 350 -50 50 150 250 (seconds)
Time 350
APPENDIX D: DYNAMIC TEST RESULTS D.13

TEST 18
Test
10 18: Tube 5 placed 0.85m from top Test
4 18: Plate 8 placed 0.85m from top
8.6
8
Pressure (kPa)

Pressure (kPa)
3
6
2
4 1.7
1
2

0 0

-2
-1
-4
-2
-6
-3
-8

-10 -4
-50 0 50 100 Time (seconds)
150 200 250 -50 0 50 100 150Time200
(seconds)
250

TEST 19
Test 19: Tube 2 placed 2.31m from top
2 Test
1 19: Tube 4 placed 1.65m from top

0
Pressure (kPa)

Pressure (kPa)

-1
1.49
-2

-3

1 -4

-5

-6

-7

-8

0 -9
-50 50 150 250
Time (seconds) -50 50 150 250(seconds)
Time 350 450

30 19: Tube 5 placed 0.92m from top


Test 2
Test 19: Plate 6 placed 1.26m from top

26.7 1.65
Pressure (kPa)
Pressure (kPa)

20
1

10

0
0

-10 -1
-20 20 60 Time
100 (seconds)
140 -50 50 150 250 (seconds)
Time 350
D.14 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN SILOS

TEST 20
Test
4 20: Tube 2 placed 2.18m from top Test
6 20: Tube 4 placed 1.16m from top

Pressure (kPa)
Pressure (kPa)

4.8
3 4
2.3
2 1.8 2

1 0

0 -2

-1 -4

-2 -6

-3 -8

-4 -10
-5 5 15 25 35 45 Time
55 (seconds)
65 75 0 10 20 Time
30(seconds)
40 50

Test
6 20: Tube 5 placed 0.57m from top 2Test 20: Plate 6 placed 1.16m from top
Pressure (kPa)

Pressure (kPa)

5
4.7
1
4 0.7

3 0

2
-1
1

0 -2

-1
-3
-2

-3 -4
-5 5 15 25 Time
35 (seconds)
45 55 0 10 20 Time
30 (seconds)
40 50
D.6 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN SILOS

TEST 10
Test
14 10: Tube 2 placed 2.37m from top Test
12 10: Plate 3 placed 1.47m from top
13.3 11.4
12

Pressure (kPa)
10
10
8
8

6 6

4 4

2
1.28 2
0
0
-2

-4 -2
-5 0 5 10 Time
15 (seconds)
20 25 30 -10 10 30 50 Time70(seconds)
90

Test
7 10: Tube 4 placed 0.71m from top 10 10: Tube 5 placed 1.47m from top
Test

8.4
Pressure (kPa)

6
Pressure (kPa)

8
5.4
5
6
4
4
3.22
3
2
2
0
1
0.6
0 -2

-1 -4
0 1 2 3 4 5 6 Time
7 (seconds)
8 9 10 -2 0 2 4 6 8
Time10 12 14
(seconds)

8
Test 10: Plate 6 placed 2.37m from top 4
Test 10: Plate 7 placed 0.71m from top
Pressure (kPa)

7
Pressure (kPa)

3
6 6.1 2.7

5 2

4
1
3

2 0

1
-1
0

-1 -2
-5 0 5 10 (seconds)
Time 15 20 -2 0 2 4 Time
6 (seconds)
8 10
TEST 10

Test 10: SG 11 & 12: Drag force on tube


240

219.5
Force (kg)

210

180

150

120 121.2

90

60

30

-30
-4 0 4 8 Time (seconds)
12 16 20
2.2 Multi-turn potential meter : Channel 10
Distance travelled (m)

2
1.8
1.6
1.4
1.2
1
0.8
0.6
0.4
0.2
0
Data point number
APPENDIX D: DYNAMIC TEST RESULTS D.11

TEST 16
Test 16: Tube 2 placed 2.41m from top
18 Test 16: Mtpm 10 placed 0.09m from top
Pressure (kPa)

15 15.2 3.5

12 Cylinder

9 2.5

6
Flow rate=
3 1.5 0.037 m/s

0
Hopper
-3 0.5

-6

-9 -0.5
-5 5 15 Time
25(seconds)35 -10 10 30 50 (seconds)
Time 70

Test 16: Tube 4 placed 1.8m from top Test 16: Plate 7 placed 1.8m from top
12
20
Pressure (kPa)

19.8
Pressure (kPa)

9.9
16 9

12
6
8

4 3
2.7

0
0
-4

-8 -3
-10 0 10 20 30 Time
40 (seconds)
50 60 -10 0 10 20 30 Time
40 (seconds)
50 60

Test
28 16: Tube 5 placed 1.17m from top Test
24 16: Plate 7 placed 1.17m from top
Pressure (kPa)

Pressure (kPa)

24 23.6 20 18.9
20
16
16

12 12

8 8
2.5
4
4 1.7
0
0
-4

-8 -4
-10 0 10 20 30 40 (seconds)
Time 50 60 70 -10 0 10 20 30 Time
40 (seconds)
50 60 70
APPENDIX E: TEST RESULTS: ANTI-DYNAMIC TUBE FRICTIONAL DRAG E:1

8.5 APPENDIX E: TEST RESULTS


ANTI-DYNAMIC TUBE FRICTIONAL DRAG

240 Test number 10


219.5
210 Force (kg)
180
150
120
121.2
90
60
30
0
-30 Time (seconds)
-4 -2 0 2 4 6 8 10 12 14 16 18 20

270 Test number 11


252.7
240
210
180 183.1

150
120
90
60
30
Time (Seconds)
0
-10 0 10 20 30 40 50 60 70 80 90

270 Test number 12


Force (kg)

240 245.6
210
180
150 157.1
120
90
60
30
Time (seconds)
0
0 10 20 30 40 50 60 70 80 90
E:2 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

300 Test number 13

Force (kg)
278.4
250

200 193.8

150

100

50

Time (seconds)
0
-10 0 10 20 30 40 50 60 70 80 90 100

Test number 14
450
Force (kg)

400 414.3: Middle


section broke off
350
300 296.6: Top
250 254.4
200 197.4: Bottom
150
100
50 74.6
0
-50
-100
Time (Seconds)
-150
-4 -2 0 2 4 6 8 10 12 14 16

300 Test number 15


287.6: Top
Force (kg)

250

200 201.6: Middle

150

100
88.7: Bottom
50

0 Time (seconds)
-10 0 10 20 30 40 50 60 70 80
APPENDIX E: TEST RESULTS: ANTI-DYNAMIC TUBE FRICTIONAL DRAG E:3

360 358.0 Test number 16

Force (kg)
320
280
240 245.5: Top

200 212.5 211.6: Middle


160
120
80
66.5: Bottom
40 52.6
0 Time (seconds)
-10 0 10 20 30 40 50 60 70 80

Test number 17
100
91.9
90
Force (kg)

Top
80 85.9
70
Middle
60
50 46.3
40
30 Bottom
20
10
Time (seconds)
0
-10 0 10 20 30 40 50 60 70 80 90 100 110 120

120 Test number 18

Top 108.3
100
93.7 98.5

80
Middle
60
52.8
40 38.5
Bottom
20

0
-30 0 30 60 90 120 150 180 210 240 270 300 330 360
E:4 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

110 Test number 19


100 Force (kg)
96.2
90
80 Top
70
64
60
50
40 Bottom
30
20
10
Time (seconds)
0
-25 25 75 125 175 225 275 325

80 Test number 20
74.8
Force (kg)

70

60

50

40

30

20

10
Time (seconds)
0
-10 0 10 20 30 40 50 60 70

200 Test number 21

177.5
160

120 134.3
Force (kg)

80

40

Time (seconds)
0
-10 0 10 20 30 40 50 60 70 80
APPENDIX E: TEST RESULTS: ANTI-DYNAMIC TUBE FRICTIONAL DRAG E:5

140 Test number 22


127.9

Force (kg)
120
118.3
100

80

60

40

20
Time (seconds)
0
-10 0 10 20 30 40 50 60 70 80

120
110.9 Test number 23

100 104.8

80

60

40

20
Time (seconds)
0
0 10 20 30 40 50 60 70 80

120 Test number 24


Force (kg)

100 103.4

80

60

40

20
Time (seconds)
0
-10 0 10 20 30 40 50 60 70 80
APPENDIX F: SHELL THEORY: FLEXURAL RIGIDITY OF A SHELL F.1

8.6 APPENDIX F: SHELL THEORY

8.6.1 UNIFORMLY COMPRESSED CIRCULAR RING

qcr

z
x

Figure F.1: External lateral pressure on a Long Cylindrical shell

For thin cylindrical shells which have a length to diameter ratio greater than fifty
the following equation can be used as the theoretical buckling load due to a uniform
external lateral pressure:

3E I
qcr (F.1)
3
r

8.6.2 FLEXURAL RIGIDITY OF A SHELL

To determine the expression for the flexural rigidity of a shell, the beam equation is
used as a starting point,

M E
(F.2)
z I R

In this equation, is the stress across the face of an element, z is the distance
from the neutral axis to the most extreme fibre on the element, M is the bending
moment in the structure, I is the second moment of area, E is Youngs modulus of
elasticity and R is the radius of curvature of the neutral axis of the element in
bending.

A cross section through a simple element of thickness t, is shown in figure F.1 .


F.2 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

M
x x

z y

Cross section Stress across Strain across


through an the face of the the face of the
element element element

Figure F.1: Stress distribution across an element

Since the shell is a thin structure, the strain of the element can be considered as a
two dimensional problem. Therefore the strains of the element in the x and y
directions are given by:

z 1 z 1
x x y and y y x (F.3)
Rx E Ry E

These equations can be re-arranged in terms of the stresses in the x and y


directions as follows:

Ez Ez
x x y and y y x (F.4)
2 2
1 1

where is the curvature of the shell and is equal to 1/R

t /2

M x dy xz dy dz (F.5)
t /2

Substituting equation F.4 into F.5 gives:

t /2
Ez E t3 t3
Mx x y z 2 dz x y
2 2
1 1 24 24
t /2

And therefore:

E t3
Mx x y D x y (F.6)
2
12 1

where D is the flexural rigidity of the shell


APPENDIX G: PERFORATED CYLINDER TEST RESULTS: 51mm CUTOUTS G.1

8.7 APPENDIX G
PERFORATED CYLINDER TEST RESULTS

The results of the perforated cylinder tests have been given on the following four
pages. These have been grouped according to the diameter of the cut outs, ie first
the 51mm diameter cut outs (16.5% open area) followed by the 76mm diameter
cut outs (36.6% open area). The graphs have been deliberately placed on the next
pages so that the reader can obtain an overall view of the tests undertaken. Each
group of cylinders was subjected to a lateral pressure of 10kPa increasing to 40kPa
in increments of 10, which constitutes four tests per cut out size. Therefore pages
G.2 and G.3 shows the test results of the cylinder with 51mm diameter cut outs,
while pages G.4 and G.5 show the results for the 76mm diameter cut outs.
G.2 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

8.7.1 CYLINDERS WITH 16.5% OPEN AREA

30
Lateral pressure 10 kPa

25 26.27kN

20
Total Axial Load (kN)

15

10

0
0 5 10 15 20 25
Vertical displacement (mm)

35
32.25 kN Lateral pressure 20 kPa
30

25
Total Axial Load (kN)

20

15

10

0
0 5 10 15 20 25
Vertical displacement (mm)
APPENDIX G: PERFORATED CYLINDER TEST RESULTS: 51mm CUTOUTS G.3

18
17.13 kN Lateral pressure 30kPa
16

14

12

10
Total Axial Load (kN)

0
0 5 10 15 20 25
Vertical Displacement (mm)

20
Lateral pressure 40kPa
18.55kN

16
Total Axial Load (kN)

12

0
0 2 4 6 8 10 12 14 16
Vertical Displacement (mm)
G.4 DES NORTJE: PhD THESIS : THE ANTI-DYANMIC TUBE IN MASS FLOW SILOS

8.7.2 CYLINDERS WITH 36.6% OPEN AREA

12
11.51 kN Lateral pressure 10kPa

10

6
Total Axial Load (kN)

0
0 4 8 12 16 20
Vertical displacement (mm)

10
9.36 kN Lateral pressure 20 kPa

8
Total Axial Load (kN)

0
0 5 10 15 20 25
Vertical Displacement (mm)
APPENDIX G: PERFORATED CYLINDER TESTS: 76mm DIAMETER CUTOUTS G.5

14
Lateral Pressure 30 kPa
12.78 kN
12

10

8
Total Axial Load (kN)

0
0 4 8 12 16 20
Vertical Displacement (mm)

7
Lateral pressure 40 kPa

6.21 kN
6

5
Total Axial Load (kN)

0
0 4 8 12 16 20
Vertical Displacement (mm)
G.6 DES NORTJE: PhD THESIS: THE ANTI-DYANMIC TUBE IN MASS FLOW SILOS

8.7.3 LATERAL PRESSURE ONLY

CYLINDER WITH 16.5% OPEN AREA

60

50
47kPa

40
Lateral Pressure (kPa)

30

20

10

0
0 1 2 3 4 5 6 7 8
Inwards wall displacement (mm)

CYLINDER WITH 36.6% OPEN AREA

60

50

42kPa
40
Lateral Pressure (kPa)

30

20

10

0
0 1 2 3 4 5 6 7 8
Inwards wall displacement (mm)
APPENDIX H: FINITE ELEMENT ANALYSIS: EIGENVALUE BUCKLING: 50mm CUT OUTS H.1

8.8 APPENDIX H FINITE ELEMENT ANALYSIS:

EIGENVALUE BUCKLING MODE SHAPES

8.8.1 CYLINDER WITH 16.5% OPEN AREA

Mode 1 Y Mode 2 & 3YZ X


Z X
factor=12.4647 factor=12.7623

Y
Mode 4 Z X
Y
Mode 5 & 6Z X
factor=13.2104 factor=13.5373

Y Y
Mode 7 & 8Z X Mode 9 & 10Z X
factor=13.6054 factor=14.2684
H.2 DES NORTJE: PhD THESIS: THE ANTI-DYANMIC TUBE IN MASS FLOW SILOS

Mode 11 YZ X Y
Mode 12 Z X
factor=14.4791
factor=14.5045

Y Y
Mode 13 & 14
Z X
Mode 15 & 16
Z X

factor=14.8181 factor=14.8464

Mode 17 Y
Y
Mode 18 Z X
factor=14.9217
Z X
factor=15.4534

Y
Y
Mode 20 & 21
Z X Mode 23 & 24
Z X

factor=15.4953 factor=15.7345
k
APPENDIX H: FINITE ELEMENT ANALYSIS: EIGENVALUE BUCKLING: 50mm CUT OUTS H.3

Mode 26 & 27
Y
Z X
factor=16.1719

Eigenvalue Buckling
Mode 4: 3-D view

Eigenvalue Buckling
Mode 24: 3-D view
H.4 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

8.8.2 CYLINDER WITH 36.6% OPEN AREA

ModeY 1 ModeY 2&3


Loadfactor
Z X Loadfactor
Z X
5.396 5.726

Mode
Y
4 ModeY 5&6
Loadfactor
Z X Loadfactor
Z X
6.077 6.390

Mode Mode 9&10


Y 7&8 Y
Loadfactor
Z X
Loadfactor
Z X
6.528 6.970
APPENDIX H: FINITE ELEMENT ANALYSIS: EIGENVALUE BUCKLING: 76mm CUT OUTS H.5

ModeY 11&12 Mode Y13&14


Loadfactor
Z X Loadfactor
Z X
7.115 7.197

Mode Y15&16 ModeY


17
Loadfactor
Z X
Loadfactor
Z X

7.208 7.317

Mode 18&19
Y
Mode Y20&21
Loadfactor
Z X
Loadfactor
Z X
7.347
7.567

Mode 22&23
Y Mode
Y 24
Loadfactor
Z X
Loadfactor
Z X
7.951 7.959
H.6 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Mode
Y 25
Loadfactor
Z X
7.963

Eigenvalue Buckling
Mode 1: 3-D view

Eigenvalue Buckling
Mode 18: 3-D view
APPENDIX H: FINITE ELEMENT ANALYSIS: EIGENVALUE BUCKLING: SOLID SHELL H.7

8.8.3 SOLID SHELL

Y
Y
Mode shape 1&2Z X
Mode shape 3&4Z X
Factor=77,1074 Factor=77,1401

Y
Y
Mode shape 5 Z X Mode shape 6 Z X

Factor=77,4023 Factor=77,5164

Y
Y
Mode shape 10 Z X
Mode shape 9 Z X
Factor=78,1101
Factor=78,0958
H.8 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Y Y
Mode shape 13
Mode shape 11 Z X factor=79,9256 Z X
factor=79,6821

Y Y
Mode shape 15 Z X
Mode shape 17Z X
Factor=80,0300 Factor=80,1845

Y Y

Mode shape 18 Z X Mode shape 19Z X

Factor=80,2200 Factor=87,690

Y
Y
Mode shape 20 Z X Mode shape 22Z X

Factor=87,6958 Factor=87,7832
APPENDIX H: FINITE ELEMENT ANALYSIS: EIGENVALUE BUCKLING: SOLID SHELL H.9

Y
Y
Mode shape 24 Z X
Mode shape 25Z X
Factor=87,8753 Factor=87,9731

Y Y
Mode shape 27 Z X Mode shape 29 Z X
factor=96,1378 Factor=96,4148
REFERENCES 9.1

CHAPTER NINE
REFERENCES

Almroth BO and Holmes AMC (1972): “Buckling of Shells with Cutouts, Experiment
and Analysis”, International Journal of Solids Structures, Vol 8, 1972, pg
1057-1071

Arnold PC (1991): “On the Influence of Segregation on the Flow Pattern in Silos”,
Bulk Solids Handling, Vol 11, Number 2, may 1991, pg 447-449.

Arnold PC, McClean AG and Roberts AW (1985): “The Design of Storage Bins for
Bulk Solids Handling”, The Best of Bulk Solids Handling, Selected Articles
1981-1985, VolA/86, pg 7

Arnold PC, McClean AG and Roberts AW (1989): “Bulk Solids: Storage, Flow and
Handling”, TUNRA Bulk Solids Handling Research Associates, University of
Newcastle, Australia, Second edition, December 1989.

Australian Standards (1990): “AS 3774 – 1990 : Loads on Bulk Solids


Containers”; First edition; Standards Association of Australia, Standards
House, 80 Arthur Street, North Sydney, NSW, ISBN 0 7262 6159 9

Bishara AG, El-Azazy SS and Huang T (1981): “Practical Analysis of Cylindrical


Farm Silos Based on Finite Element Solutions”; ACI Journal, November/
Decmeber 1981, pg 456-462.

Blair-Fish PM and Bransby PL (1973): “Flow patterns and Wall Stresses in a Mass
Flow Bunker”; Journal of Engineering for Industry, Transactions of the ASME,
February 1973, pg 17-27

Blight GE, Fliss L and Schaffner RH (1989); “Design and Performance of Two Types
of Cement Storage Silos”; Journal of the Civil Engineer in South Africa, Vol 31,
number 1, January 1989, pg 11-21

Blight, G E (1993): “Loading Applied to Silos by Retained Granular Materials and


Powders”; GDE Lecture Notes, The University of the Witwatersrand, South
Africa, 1993
9.2 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Briassoulis D (1991): “Limitations in the Range of Applicability of the Classic Silo


Theories”; ACI Structural Journal, July-August 1991, pg437-444

Brown RL and Richards JC (1970): “Principles of Powder Mechanics”; First Edition,


1970, Pergamon Press

Drescher A, Cousens TW and Bransby PL (1978): “Kinematics of the Mass Flow of


Granualr material through a Plane Hopper”; Geotechnique, Vol 28, No 1,
1978, pg 27-42

Gere JM and Timoshenko (1996): “Mechanics of Materials”; Third SI Edition,


Chapman & Hall, London, 1996, ISBN 0-412-36880-3

Hasra and Bazur (1980): “Considerations in the Design of Silos for Process
Industry”; Proceedings of the International Conference on Design of Silos for
Strength and Flow, University of Lancaster, U.K, 2-4 Sept, 1980

Jenike AW (1961): “Gravity Flow of Bulk Solids”; Bulletin 108, Utah Engineering
Experiment Station, University of Utah, Salt Lake City, October 1961

Jenike AW (1964): “Storage and Flow of Solids”; Bulletin 123, Utah Engineering
Experiment Station, University of Utah, Salt Lake City, November 1964

Jenike AW (1967): “Quantitative Design of mass Flow Bins”; Journal for Powder
Technology, Vol 1, 1967, pg 237-244

Jenike AW (1968): “Bins Loads”; Journal of the Structural Division, Proceedings of t


the American Society of Civil Engineers, April 1968, pg 1011-1040

Jenike AW and Johanson JR (1969): “On the Theory of Bin Loads”; Journal of
Engineering for Industry, Transactions of the ASME, May 1969, pg 339-344

Jenike AW, Johanson JR and Carson JW (1973a): Bin Loads – Part 2: Concepts”,
Journal of Engineering for Industry, Transactions of the ASME, February 1973,
pg 1-5

Jenike AW (1973b): “Bin Loads – Part 3: Mass-Flow Bins”; Journal of Engineering


for Industry, Transactions of the ASME, February 1973, pg 6-12
REFERENCES 9.3

Jenike AW, Johanson JR and Carson JW (1973c): Bin Loads – Part 4: Funnel Flow
Bins”; Journal of Engineering for Industry, Transactions of the ASME, Series
B, Vol 95, Number 1, February 1973, pg 13-16

Johanson JR (1982): “Controlling Flow Patterns by use of an Insert”; Bulk Solids


Handling, Vol 2, No 3, Sept 1982.

Kmita J (1991): “An Experimental Analysis of Internal Silo Loads”; Bulk Solids
Handling, Vol 11, Number 2, May 1991, pg 459-467

Lansgton PA, Tuzun U and Heyes UK (1995): “Computer Simulations of Hopper


Flows - A Design Tool or a Research and Teaching Facility”; Bulk Solids
Handling, Vol 15, No 3, July/September 1995, pg 385-389

Luong MP (1993): “Flow Characteristics of Granular Bulk Materials”; Journal of


Particle-Particle-System Characteristics, Vol 10, 1993, pg 79-85

McLean AG (1985): “The Design of Silo Side Discharge Outlets for Safe and Reliable
Operation”; Bulk Solids Handling, Vol 5, Number 1, February 1985, pg 185-
190

Molenda M, Horabik J and Ross IJ (1993): “Loads in Model Grain Bins as Affected by
Filling Methods”; American Society of Agricultural Engineers, Vol 36, number
3, May/June 1993, pg 915-919.

Nielsen J and Andersen E Y (1982): “Loads in Grain Silos”; Bygningsstatiske


Meddelelser, Danish Society for Structural Science and Engineering, Vol 53,
number 4, 1982, pg 123-135. ( In English).

Nielsen J and Askegaard V (1977): “Scale Errors in Model Tests on Granular Media
with Special Reference to Silo Models”; Powder Technology, Volume 16, 1977,
pg 123-130

O’Neil, P V: “Advanced Engineering Mathematics”; Fourth Edition, PWS Publishing,


Boston, Massechusets, 1995, ISBN 0-534-94320-9

Ooms M and Roberts AW (1985): “The Reduction and Control of Flow Pressures in
Cracked Grain Silos”; Bulk Solids handling, Vol 5, No 5, October 1985.
9.4 DES NORTJE: PhD THESIS: THE ANTI-DYNAMIC TUBE IN MASS FLOW SILOS

Pieper K (1969): “Investigation of Silo Loads in Measuring Models”; Journal of


Engineering for Industry, Transactions of the ASME, May1969, pg 365-372

Ravenet J (1983): “The Development of Industrial Silos Throughout the World


During the Last 100 years”; Bulk Solids Handling, Vol 3, number 1, March
1983, pg 127-140

Reed AR and Duffell CH (1983): “ A Review of Hopper Discharge Aids”; Bulk Solids
Handling, vol 3, No 1, March 1983

Reimbert M and A (1976): “Silos Theory and Practice”; First edition, Trans Tech
Publications, Germany, 1976, ISBN 0-442-22684-5

Richards PC (1977): “Bunker Design - Part 1: Bunker Outlet Design and Initial
Measurements of Wall Pressures”; Journal of Engineering for Indistry,
Transactions of the ASME, November 1977, pg 809-813

Roberts AW (1995): “100 Years of Janssen”; Bulk Solids Handling, Vol 15, No 3,
July/September 1995.

Rombach G and Eibl J (1995): “Granular Flow of Materials in Silos: Numerical


Results”; Bulk Solids Handling,Vol 15, number 1,January/March 1995,pg 65-
70

Schulze D (1996): “Silos- Design Variants and Special Types”; Bulk Solids Handling,
Vol 16, Number 2, April/June 1996

Schwedes J and Schulze D (1991): “Examples of Modern Silo Design”; Bulk Solids
Handling, Vol 11, number 1, March 1991, pg 47-52

Scutella L (1998): “Local Buckling of Perforated Cylindrical Shells under Uniform


Axial Compression”; Honours Thesis, University of Western Australia,
Department of Civil Engineering, October 1998

Starnes JH Jr(1972): “Effect of a Circular Hole on the Buckling of Cylindrical Shells


Loaded by Axial Compression”; AIAA Journal, Vol 10, Number 11, November
1972, pg 1466-1472
REFERENCES 9.5

Suzuki M, Akashi T and Matsumoto K (1985): “Flow Behaviour and Stress


Conditions in Small and Medium Silos”; Bulk Solids Handling, Vol 5, number 3,
June 1985, pg 63-72.

Tennyson RC (1968): “The Effects of Unreinforced Circular Cutouts on the Buckling


of Circular Cylindrical Shells under Axial Compression”; Journal of Engineering
for Industry, Transactions of the ASME, November 1968, pg 541-546

Timoshenko SP and Gere JM (1963): “Theory of Elastic Stability”; Second edition,


McGraw-Hill Book Compnay, Singapore, 1963, ISBN 0-07-085821-7

Van Zanten DC and Mooij A (1977): “Bunker Design-Part 2: Wall Pressures in Mass
Flow”; Journal of Engineering for Industry, Transactions of the ASME,
November 1977, pg 814-818

Walker DM (1966): “An Approximate Theory for Pressures and Arching in Hoppers”;
Chemical Engineering Science, 1966, Vol 21, pg 975-997.

Walters JK (1973a): “A Theoretical Analysis of Stresses in Silos with vertical walls”;


Chemical Engineering Science, 1973, Vol 28, pg 13-21.

Walters JK (1973b): “A Theoretical Analysis of Stresses in Axially-Symmetric


Hoppers and Bunkers”; Chemical Engineering Science, 1973, Vol 28, pg 779-
789.

You might also like