Anda di halaman 1dari 259

Mechanics of Mine Backfill

By

Matthew Helinski

This thesis is presented for the

Degree of Doctor of Philosophy

The University of Western Australia

School of Civil and Resource Engineering

December 2007
DECLARATION FOR THESES CONTAINING PUBLISHED WORK AND/OR WORK PREPARED FOR
PUBLICATION

The examination of the thesis is an examination of the work of the student. The work must have been
substantially conducted by the student during enrolment in the degree.

Where the thesis includes work to which others have contributed, the thesis must include a statement that
makes the students contribution clear to the examiners. This may be in the form of a description of the
precise contribution of the student to the work presented for examination and/or a statement of the
percentage of the work that was done by the student.

In addition, in the case of co-authored publications included in the thesis, each author must give their signed
permission for the work to be included. If signatures from all the authors cannot be obtained, the statement
detailing the students contribution to the work must be signed by the coordinating supervisor.

Please sign one of the statements below.


1. This thesis does not contain work that I have published, nor work under review for publication.
(Note: A number of journal and conference papers have been published on various aspects of the work,
as listed Page v at the start of the Thesis. However, these are not part of the thesis per se.)

Signature:

Thought the publications listed are not part of the thesis, the work included in them forms a central part of
the thesis. The candidate, Mr Helinski, is first author on all of the publications, and can claim a contribution
of > 70% to each of them.

Signature: (Martin Fahey, coordinating supervisor)


2. This thesis contains only sole-authored work, some of which has been published and/or prepared for
publication under sole authorship. The bibliographical details of the work and where it appears in the thesis
are outlined below.

Signature.........................................................................................................................................................
3. This thesis contains published work and/or work prepared for publication, some of which has been co-
authored. The bibliographical details of the work and where it appears in the thesis are outlined below.
The student must attach to this declaration a statement for each publication that clarifies the contribution of
the student to the work. This may be in the form of a description of the precise contributions of the student to
the published work and/or a statement of percent contribution by the student. This statement must be signed
by all authors. If signatures from all the authors cannot be obtained, the statement detailing the students
contribution to the published work must be signed by the coordinating supervisor.

Signatures.........................................................................................................................................................

Signatures.........................................................................................................................................................
Mechanics of Mine Backfill Matthew Helinski
The University of Western Australia

ABSTRACT

Mine backfilling is the process of filling large underground mining voids (stopes)
with a combination of tailings, water and small amounts of cement, to promote regional
stability. Stopes are often in excess of 20 m 20 m in plan dimensions and 40-50 m tall,
and can be filled within a week. Barricades are constructed in all tunnels (drives) that
access the stope to contain the backfill material. In recent years, a significant number of
failures of mine backfill barricades have occurred, resulting in the inrush of slurry
backfill into the mine workings. In addition, sampling has shown material strengths in
situ to be far greater than equivalent mixes cured in the laboratory (indicating the
potential for reducing the cement content). The purpose of this thesis is to apply soil
mechanics principles to the mine backfill deposition process with the intent of providing
some insight into these issues.

In many cases, filling, consolidation and cement hydration all take place at a similar
timescale, and therefore, to understand the cemented mine backfill deposition process it
was necessary to appropriately couple these activities. Developing appropriate models
for these mechanisms, and coupling them into a finite element code, forms the core of
this thesis.

Firstly, the fundamental processes involved in the cementing mine backfill deposition
process are investigated and represented using theory founded on basic physical
observations.

Using this theory, one- and two-dimensional finite element models (called CeMinTaCo
and Minefill-2D, respectively) are developed to fully couple each of the individual
mechanisms.

A centrifuge experiment was undertaken to investigate the interaction between


consolidation and total stress distribution in a cementing soil. The results of this
experiment were also used to verify the performance of Minefill2D. Due to scale
effects, the centrifuge experiment was unable to fully couple the interaction of the
cement hydration and consolidation timescales. To achieve this, a full scale field
experiment was undertaken. The simulated behaviour achieved using Minefill-2D (with
independently derived material properties) provided a good representation of the
consolidation behaviour.

i
Mechanics of Mine Backfill Matthew Helinski
The University of Western Australia

Finally, a sensitivity study carried out using Minefill-2D is presented. This study
enables some useful suggestions to be provided for managing the risk of excessive
barricade stress, and for preparing laboratory samples to more appropriately represent in
situ curing conditions.

ii
Mechanics of Mine Backfill Matthew Helinski
The University of Western Australia

ACKNOWLEDGEMENTS

Firstly I would like to acknowledge the support of my wonderful family throughout the
period of my studies. My wife Libby, who after initially being somewhat apprehensive
about my decision to return to university, has provided me with undivided support
throughout this period. Jessica who was undesirably juggled during my early years of
study always had a wonder smile to greet me with and Lucy, our recent addition, who I
am equally proud of. To my parents, grandparents and sister who have provided me
with the wonderful gift of education and support throughout my life, I am forever
grateful of this.

To my supervisor Professor Martin Fahey, a true professor in the way he can make the
most complicated aspect of soil mechanics appear so clear and simple through the
application of his fundamental knowledge. This is something I aspire to. My supervisor
Professor Andy Fourie, whose guidance and friendship during my research was
essential in developing this project. Andy, I feel very fortunate that you arrived in
Australia and supported me when you did. Also, thanks to Dr Mostafa Ismail who
assisted me in the laboratory component of this work and Professor Jack Barrett who
helped shape this project during the early stages.

Thanks to all of my university colleagues, in particular James Schneider and James


Doherty and Shambu Sharma, I am extremely appreciative of your supervision and
guidance throughout this thesis.

Also my industry colleagues Cameron Tucker, Mat Revell and Tony Grice, I appreciate
all of your support and encouragement with this work.

Thanks to all of the academic and support staff in the Civil Engineering department in
particular Binaya, Clair and Natalia (who sadly passed away during this thesis) for their
ongoing patience with my chaotic style in the laboratory as well as Tuarn, John, Shane,
Phil, Bart, Don and Neil for their assistance with centrifuge testing.

Finally, this work would not have been possible without the wonderful post graduate
scholarship foundation at The University of Western Australia. The late Robert John
Gledden for establishing the Gledden trust that provided the majority of financial
support throughout this work. Merriwa for their wonderful top-up scholarship that

iii
Mechanics of Mine Backfill Matthew Helinski
The University of Western Australia

provided both financial and equipment support. Mrs N. Shaw who established the
F.S.Shaw scholarship in memory of her late husband, which provided much needed
funding towards the latter stages of this research. And the UWA travel scholarship
which allowed me to attend the Minefill 07 conference in Canada.

iv
Mechanics of Mine Backfill Matthew Helinski
The University of Western Australia

DECLARATION

I hereby declare that, except where specific reference is made in the text to work of
others, the contents of this thesis are original and have not been submitted to any other
university.

During the compilation of this thesis some of the work has been published in various
journals and conference proceedings. I acknowledge the contribution of my co-authors
in preparing these publications. Details of these publications are as follows:

Journal publications

Helinski, M. Fahey, M. and Fourie, A.B. (2007) Numerical modelling of cemented mine
backfill deposition, ASCE Journal of Geotechnical and Geoenvironmental Engineering,
Vol. 133, Issue 10, 1308-1319.

Helinski, M., Fourie, A.B. Fahey, M. and Ismail, M. (2007). The self desiccation
process in cemented mine backfill. Canadian Geotechnical Journal. Vol. 44, No. 10,
1148-1156.

Helinski, M. Fahey, M. Fourie, A.B. (2007) An effective stress approach to modelling


mine backfilling, CIM technical paper, Issue No. 5, August, Vol.2.

Fourie, A.B. Helinski, M. Fahey, M. (2007) Using effective stress theory to characterise
the behaviour of backfill. CIM technical paper, Issue No.5, August, Vol. 2

Conference publications

Helinski, M., Fourie, A.B. and Fahey, M. (2006) Mechanics of early age cemented paste
backfill. Paste 06, Limerick April 3-7, Australian Centre for Geomechanics, ISBN 0-
9756756-5-6.

Helinski, M. Coltrona, A.B. Fourie, A.B. and Fahey, M. (2007) Influence of tailings
type on barricade loads in backfilled stopes. Paste 07, Perth, March 13-15, Australian
Centre for Geomechanics, ISBN 0-9756756-7-2. 95-104.

Helinski, M. Fahey, M. Fourie, A.B. (2007) An effective stress approach to modelling


mine backfilling, Minefill 07, 9th International Symposium on Mining with Backfill,
Montreal, Paper # 2478.

v
Mechanics of Mine Backfill Matthew Helinski
The University of Western Australia

Fourie, A.B. Helinski, M. Fahey, M. (2007). Using effective stress theory to


characterise the behaviour of backfill. Minefill 07, 9th International Symposium on
Mining with Backfill, Montreal, Paper # 2480.

Helinski, M. Tucker, C. Grice, A.G. (2007) Water management in hydraulic fill


operations. Minefill 07, 9th International Symposium on Mining with Backfill,
Montreal, Paper # 2479.

Matthew Helinski

December 2007

vi
Mechanics of Mine Backfill Matthew Helinski
The University of Western Australia

TABLE OF CONTENTS

Abstract...........................................................................................................................i
Acknowledgements...................................................................................................... iii
Declaration.....................................................................................................................v
Table of Contents ....................................................................................................... vii
List of Figures................................................................................................................x
List of Tables ............................................................................................................ xvii
Chapter 1 Introduction..............................................................................................1.1
1.1 Significance of consolidation to mine backfill ...................................................1.4
1.2 Project methodology...........................................................................................1.5
Chapter 2 Background & literature review ............................................................2.1
2.1 Introduction.........................................................................................................2.1
2.2 Mine backfill literature .......................................................................................2.1
2.2.1 Influence of consolidation on barricade stresses .....................................2.1
2.2.2 Influence of consolidation on in situ strengths........................................2.7
2.2.3 Influence of consolidation on exposure stability.....................................2.9
2.2.4 Summary ...............................................................................................2.10
2.3 Consolidation....................................................................................................2.10
2.3.1 Consolidation behaviour of cementing soil...........................................2.11
2.4 Structured soil...................................................................................................2.13
2.4.1 Modelling structured soil behaviour......................................................2.16
2.5 Cementation......................................................................................................2.17
2.5.1 Cementation behaviour..........................................................................2.20
2.6 Summary...........................................................................................................2.20
Chapter 3 Behaviour of cementing slurries .............................................................3.1
3.1 Introduction.........................................................................................................3.1
3.2 Strength and stiffness..........................................................................................3.1
3.2.1 Uncemented material response................................................................3.1
3.2.2 Stress-strain behaviour of cemented fill ..................................................3.2
3.2.3 Hardening ................................................................................................3.3
3.2.4 Damage due to yielding during hydration (dD) ......................................3.6
3.2.5 Unconfined compression strength (qu) ....................................................3.7
3.2.6 Stiffness ...................................................................................................3.7
3.2.7 Stress-strain behaviour: summary ...........................................................3.8
3.3 Permeability........................................................................................................3.9
3.3.1 Uncemented permeability........................................................................3.9
3.3.2 Cemented permeability..........................................................................3.10
3.4 Self desiccation.................................................................................................3.11
3.4.1 Cementation reactions ...........................................................................3.12
3.4.2 Impact on pore pressure ........................................................................3.14
3.4.3 Analytical model ...................................................................................3.15
3.4.4 Experimental demonstration of effect of self desiccation .....................3.17
3.4.5 Material properties influencing self desiccation ...................................3.18
3.4.6 Experimental derivation of parameters .................................................3.21

vii
Mechanics of Mine Backfill Matthew Helinski
The University of Western Australia

3.5 Temperature......................................................................................................3.25
3.6 Material characterisation technique..................................................................3.27
3.7 Conclusion ........................................................................................................3.28
Chapter 4 One-dimensional consolidation modelling.............................................4.1
4.1 Introduction.........................................................................................................4.1
4.2 Model development ............................................................................................4.1
4.2.1 Modelling the behaviour of uncemented tailings: the MinTaCo Program4.1
4.2.2 Modelling the behaviour of cemented tailings: the CeMinTaCo
Program ...................................................................................................4.3
4.2.3 CeMinTaCo governing equations ...........................................................4.5
4.3 Numerical implementation .................................................................................4.9
4.4 Model verification ............................................................................................4.11
4.4.1 Compressibility .....................................................................................4.11
4.4.2 Self desiccation......................................................................................4.11
4.5 Sensitivity study................................................................................................4.12
4.5.1 Influence of cementation .......................................................................4.12
4.5.2 Influence of permeability ......................................................................4.13
4.5.3 Typical damage scenario .......................................................................4.15
4.5.4 Strain requirements................................................................................4.19
4.5.5 Comparison with data from in situ monitoring of filled stopes ............4.20
4.6 Conclusion ........................................................................................................4.22
Chapter 5 Two-dimensional consolidation analysis (Minefill-2D) ........................5.1
5.1 Introduction.........................................................................................................5.1
5.2 Programming requirements ................................................................................5.2
5.3 Programming Methodology................................................................................5.4
5.3.1 Introduction .............................................................................................5.4
5.3.2 The finite element method.......................................................................5.4
5.3.3 Boundary conditions..............................................................................5.14
5.3.4 Solution to the global equations ............................................................5.18
5.4 Material Behaviour ...........................................................................................5.20
5.4.1 Influence of cementation on governing equations ................................5.20
5.4.2 Constitutive model, [K G ([D(t , e, Cc )])] ...............................................5.22
5.4.3 Permeability model, [G (t , e, Cc )], [nG (t , e, Cc )] .................................5.26
5.4.4 Self desiccation, Q(t,e,C) .....................................................................5.27
5.5 Model Verification............................................................................................5.27
5.5.1 Comparison with analytical/numerical solutions ..................................5.27
5.5.2 Comparison with CeMinTaCo ..............................................................5.29
5.5.3 Stope mesh details .................................................................................5.30
5.5.4 Comparison with in situ measurements.................................................5.32
5.5.5 Investigation of the arching mechanism................................................5.34
5.6 Conclusion ........................................................................................................5.36
Chapter 6 Centrifuge modelling ...............................................................................6.1
6.1 Introduction.........................................................................................................6.1
6.2 Experimental Apparatus .....................................................................................6.2
6.3 Calibration ..........................................................................................................6.4
6.4 Experiment..........................................................................................................6.7
6.4.1 Material ...................................................................................................6.7

viii
Mechanics of Mine Backfill Matthew Helinski
The University of Western Australia

6.4.2 Experimental procedure ..........................................................................6.8


6.4.3 Experimental results ................................................................................6.9
6.5 Numerical back analysis ...................................................................................6.11
6.5.1 Material characterisation .......................................................................6.11
6.5.2 Numerical back analysis........................................................................6.13
6.6 Conclusion ........................................................................................................6.15
Chapter 7 Sensitivity study .......................................................................................7.1
7.1 Introduction.........................................................................................................7.1
7.2 Comparision of hydraulic fill and paste fill ........................................................7.1
7.2.1 Experimental results ................................................................................7.2
7.2.2 Modelling ................................................................................................7.3
7.2.3 Comparison of hydraulic fill and paste fill..............................................7.6
7.3 Consolidating fill ................................................................................................7.6
7.3.1 Influence of stope geometry ....................................................................7.7
7.3.2 Influence of permeability ........................................................................7.8
7.3.3 Influence of cementation .........................................................................7.9
7.3.4 Influence of filling rate..........................................................................7.11
7.3.5 Consolidating fill: discussion ................................................................7.11
7.3.6 Consolidating fill: conclusion ...............................................................7.13
7.4 Non-consolidating fill.......................................................................................7.13
7.4.1 Influence of stope geometry ..................................................................7.14
7.4.2 Influence of permeability ......................................................................7.15
7.4.3 Influence of cementation .......................................................................7.16
7.4.4 Filling rate .............................................................................................7.17
7.4.5 Non-consolidating fill: discussion.........................................................7.17
7.4.6 Non-consolidating fill: conclusion ........................................................7.19
7.5 Development of effective stress during curing.................................................7.20
7.5.1 Comparision between consolidating and non-consolidating fill ...........7.20
7.5.2 Development of effective stress in consolidating fill............................7.21
7.5.3 Development of effective stress in non-consolidating fill.....................7.22
7.5.4 Curing of fill: discussion and conclusion ..............................................7.23
7.6 Conclusion ........................................................................................................7.25
Chapter 8 Concluding remarks and recommendations for future work..............8.1
8.1 Concluding remarks............................................................................................8.1
8.2 Main outcomes....................................................................................................8.1
8.3 Recommendations for future work .....................................................................8.4
Chapter 9 References.................................................................................................9.1

ix
Mechanics of Mine Backfill Matthew Helinski
The University of Western Australia

LIST OF FIGURES

Figure 1.1 Schematic of a typical mine tailings based backfill system (contributed
by Cobar Management Pty Ltd).

Figure 1.2 Schematic showing a typical stope filling situation.


Figure 1.3 Photograph showing a failed barricade (from Revell and Sainsbury,
2007).

Figure 2.1 Stress distribution down the centreline of a stope assuming drained
and undrained filling.

Figure 2.2 The impact of drained and undrained filling on barricade stress.
Figure 2.3 Conversion from vertical total stress to horizontal stress.
Figure 2.4 Gibson's(1958) consolidation chart with typical minefills.
Figure 2.5 Comparison between structured and unstructured compression
behaviour.

Figure 2.6 Comparison between structured and unstructured yield surfaces.


Figure 2.7 Powers illustration of the Cement hydration process (from Illstron et al.
1979).

Figure 2.8 Relationship between void ratio and binder content to achieve critical
porosity and typical mine backfill range.

Figure 3.1 Incremental yield stress as it is defined in this thesis.


Figure 3.2 Relationship between void ratio and qu for CSA hydraulic fill.
(a)
Figure 3.2 Relationship between void ratio and qu for Cannington paste fill from
(b) Rankin (2004).

Figure 3.3 Normalised qu against time for CSA hydraulic fill and Cannington
paste fill.

Figure 3.4 Incremental small strain shear stiffness against qu for CSA hydraulic
(a) fill.

Figure 3.4 Young's modulus (at large strains) against qu for Cannington paste fill.
(b)

x
Mechanics of Mine Backfill Matthew Helinski
The University of Western Australia

Figure 3.5 Comparison between one-dimensional compression experiments and


the model results.

Figure 3.6 Comparison between eff and permeability.


Figure 3.7 Particle size distribution curves.
Figure 3.8 Pore water pressure (u) and effective stress changes in triaxial samples
hydrating under constant total stress and undrained boundary
conditions.

Figure 3.9 Typical result from bender element test.


Figure 3.10 Typical pore water pressure (u) and effective stress changes in a
triaxial sample (CSA hydraulic fill material with 5% cement) hydrating
under constant total stress and undrained boundary conditions (with
periodic re-establishment of back pressure, to minimise effective stress
change).

Figure 3.11 The development of bulk stiffness Ks with time for CSA hydraulic fill:
experimental data (symbols) and Equation 3.31 (lines).

Figure 3.12 Rate of pore water pressure (u) reduction with time after initial set for
various cement contents for CSA hydraulic fill.

Figure 3.13 Normalised apparent water loss rate plotted against time for different
cement contents for CSA hydraulic fill: experimental data compared
with Equation 3.32.
Figure 3.14 Comparison of experimental reduction of pore water pressure (u)
against time and adjusted theoretical solution for CSA hydraulic fill.

Figure 3.15 Predicted and measured reduction in pore water pressure (u) for KB
paste backfill.

Figure 3.16 Temperature variation across stope half-space after 20 hours.


Figure 3.17 Hydration test setup.

Figure 4.1 Schematic representation showing the relationship between a, and x


in the convective coordinate system.

Figure 4.2 Schematic representing pore water continuity across an element a.

xi
Mechanics of Mine Backfill Matthew Helinski
The University of Western Australia

Figure 4.3 Schematic showing mesh used in CeMinTaCo finite difference


approximation.

Figure 4.4 Comparison between the self desiccation pore pressure reduction in a
hydration test and CeMinTaCo output.

Figure 4.5 Idealisation of the base boundary conditions used to represent a stope
in CeMinTaCo.

Figure 4.6 CeMinTaCo output illustrating the influence of the cement induced
mechanisms on the pore pressure response.

Figure 4.7 Variation in permeability against time


Figure 4.8 Pore pressure against time for the different cases analysed.
Figure 4.9 Pore pressure isochrones for the different permeability cases analysed.
Figure 4.10 e against v for different damage parameters.
Figure 4.11 CeMinTaCo output for different damage parameters.
Figure 4.12 Development of material strength against time for different damage
parameters.

Figure 4.13 CeMinTaCo output for different damage parameters in free draining
Figure 4.14 material
Development of material strength for different damage parameters with
free draining fill.

Figure 4.15 Axial strain levels for different filling scenarios.


Figure 4.16 Comparison between CeMinTaCo and in situ pore pressure
measurements.

Figure 5.1 Element geometry adopted for plane-strain displacement and pore
pressure finite element calculations in this thesis.

Figure 5.2 8 noded isoparametric element (taken from Potts and Zdravkovi,
1999) showing (a) the parent element and (b) the global element.

Figure 5.3 Element geometry adopted for axi-symmetric displacement and pore
pressure finite element calculations in this thesis.

Figure 5.4 (a) shear stress against axial strain and (b) shear stress against mean
stress for a triaxial test.

xii
Mechanics of Mine Backfill Matthew Helinski
The University of Western Australia

Figure 5.5 Tangent shear stiffness normalised by small strain shear stiffness
against shear stress normalised by the peak shear strength.

Figure 5.6 Illustration of one-dimensional consolidation problem.


Figure 5.7 Comparison between Minefill-2D and the analytical solution for one-
dimensional consolidation analysis.

Figure 5.8 Illustration showing the one-dimensional self weight consolidation


problem used in the Minefill-2D verification.

Figure 5.9 Comparison between Minefill-2D and Plaxis for a self weight
consolidation problem.

Figure 5.10 Numerical simulation undertaken to verify the performance of the self
desiccation mechanism.

Figure 5.11 Comparison between Minefill-2D and the analytical solution for self
desiccation.

Figure 5.12 Numerical geometry for comparison between Minefill 2D and Darcy's
law for a falling head permeability test.

Figure 5.13 Comparison with Minefill-2D and Darcy's law for the flow through the
surface layer of the fill.

Figure 5.14 Comparison between Cemintaco and Minefill 2D.


Figure 5.15 Comparison between Cemintaco and Minefill 2D with a modified
"initial set" point.

Figure 5.16 Finite element mesh used to represent (a) coarse mesh, (b) medium
mesh and (c) a fine mesh.

Figure 5.17 Calculated pore pressure in the centre of the stope floor for different
mesh shapes.

Figure 5.18 Calculated barricade stress in the centre of the stope floor for different
mesh shapes.

Figure 5.19 Vertical total stress contours at the completion of filling for the (a)
coarse mesh, (b) medium mesh and (c) the fine mesh.

Figure 5.20 Comparison between Minefill-2D and in situ measurements.

xiii
Mechanics of Mine Backfill Matthew Helinski
The University of Western Australia

Figure 5.21 Illustration showing the boundary conditions adopted for the (a) fixed-
BC and (b) free-BC case in the "arching" analysis.

Figure 5.22 Comparison between u and v in a stope with fixed and free vertical
displacement boundary conditions.

Figure 5.23 v contours for a stope with (a) fixed vertical displacement boundary
conditions and (b) with free vertical displacement boundary conditions

Figure 5.24 Total vertical stress along the stope centreline for the fixed and free
BC.

Figure 6.1 Schematic showing a section through the sample container.


Figure 6.2 (a) Photograph of strain gauged cylinder that was used to represent the
stope walls and (b) the inside of the cylinder showing the rough

Figure 6.3 cylinder walls.


Photograph of the false base and loadcells that were used in the
experiment.

Figure 6.4 Experimental apparatus positioned in a strong box on the UWA


geotechnical centrifuge.

Figure 6.5 Change in pressure and stress during Stage 1 loading.


Figure 6.6 Incremental change in u during Stage 2 loading.
Figure 6.7 Incremental load / stress distribution in second stage of loading.
Figure 6.8 Relationship between vertical effective stress and void ratio from the
Rowe cell test.

Figure 6.9 Relationship between void ratio and permeability from Rowe cell.
Figure 6.10 Comparison between measured and calculated pore pressure in Stage 1.
Figure 6.11 Comparison between the measured and calculated load distribution in
Stage 1.

Figure 6.12 Evolution of Go and qu against time for the kaolin with 25% cement
mix.

Figure 6.13 Comparison between measured and calculated pore pressure in Stage 2.
Figure 6.14 Comparison between the measured and calculated load distribution in
Stage 2.

xiv
Mechanics of Mine Backfill Matthew Helinski
The University of Western Australia

Figure 7.1 Particle size distribution of backfills tested.


Figure 7.2 Evolution of permeability against time for different mine backfills.
Figure 7.3 Evolution of cohesion against time for different mine backfills.
Figure 7.4 Minefill 2D results of barricade stress against time for different backfill
types.

Figure 7.5 Development of pore pressure against time for different mine backfills.
Figure 7.6 Pore pressure isochrones for different mine backfills.
Figure 7.7 Influence of drawpoint permeability on pore pressure at the base of a
stope with consolidating fill.

Figure 7.8 Pore pressure isochrones for consolidating fills with various drawpoint
permeabilities.

Figure 7.9 Barricade stress against time for different drawpoint permeabilities
with consolidating fills.

Figure 7.10 Pore pressure against time for consolidating fills with different
permeabilities.

Figure 7.11 Barricade stress against time for consolidating fills with different
permeabilities.

Figure 7.12 Pore water pressure against time for consolidating fill with different
binder contents.

Figure 7.13 Barricade stress against time for consolidating fills with different
binder contents.

Figure 7.14 Comparison between applied shear stress and cohesion for a boundary
element.

Figure 7.15 Contour of cohesion at the end of filling for the (a) the 3.0% cement
and (b) the 1.5% cement case.

Figure 7.16 Total vertical stress calculated for the (a) 3.0% cement and (b) the
1.5% cement case.

Figure 7.17 Influence of filling rate on consolidating fill pore pressures.

xv
Mechanics of Mine Backfill Matthew Helinski
The University of Western Australia

Figure 7.18 Influence of filling rate on consolidating fill barricade stress.


Figure 7.19 Relationship between pore pressure and barricade stress in a
consolidating fill.

Figure 7.20 Pore pressure against time for non-consolidating fills with different
drawpoint permeabilities.

Figure 7.21 Barricade stress against time for non-consolidating fills with different
drawpoint permeabilities.

Figure 7.22 Pore pressure profile at the end of filling for (a) kdp=10kstope and (b)
kdp=0.1kstope.

Figure 7.23 Pore pressure against time for non-consolidating fills with different
permeabilities.

Figure 7.24 Barricade stress against time for non-consolidating fills with different
permeabilities.

Figure 7.25 Pore pressure against time for non-consolidating fills with different
cement contents.

Figure 7.26 Barricade stress against time for non-consolidating fills with different
cement contents.

Figure 7.27 Barricade Stress and pore pressure against time for non-consolidating
fill with a bonded and unbonded interface

Figure 7.28 Pore pressure against time for non-consolidating fills with different
filling rates.

Figure 7.29 Barricade stress against time for non-consolidating fills with different
filling rates.

Figure 7.30 Development of effective stress within an element of consolidating and


non-consolidating fill against time.

Figure 7.31 Development of effective stress against time in a consolidating fill.


Figure 7.32 Development of effective stress against time in a non-consolidating fill.
Figure 7.33 Development of effective stress against time at different elevations in a
non-consolidating fill.

xvi
Mechanics of Mine Backfill Matthew Helinski
The University of Western Australia

LIST OF TABLES

Table 5.1 Material properties adopted for CeMinTaCo - Minefill-2D comparison.


Table 5.2 P.F.-A material properties (from Helinski et al. 2007) adopted for back
analysis of in situ test results.

Table 5.3 Material properties adopted in the investigation of the arching


mechanism.

Table 6.1 Material properties for kaolin with 25% cement content.

Table 7.1 Comparison of hydraulic fill and paste fill properties.


Table 7.2 Material Properties.

xvii
Mechanics of Mine Backfill Matthew Helinski
Introduction The University of Western Australia

CHAPTER 1
INTRODUCTION

Many underground hard-rock mining operations adopt an open stoping technique,


which involves the extraction of ore in large underground blocks called stopes. Stope
sizes are dictated by geotechnical conditions, but the extraction of an orebody generally
requires many, often hundreds of stopes. In order to maintain geotechnical stability on
both a local and regional scale, significant pillars containing valuable ore must be left
between stopes, but this can significantly reduce the quantity of ore recovered. To
increase recovery, many mines choose to re-fill previously mined stopes.

The filling process involves placing mine waste materials into previously mined stopes,
in order to provide a number of services to the mining operation. These services
include:

Providing support to the surrounding rockmass

Creating a working surface for production activities

Disposal of mining waste products.

The primary reason for adopting a mine backfill strategy is to increase the quantity of
ore that may be recovered and reduce the amount of ore dilution that occurs during
stoping. In addition, mine backfill can be used to improve ground conditions by
replacing poor natural host rock with an engineered material, in regions that are
sensitive to mining.

Mine backfilling can be carried out in a number of different ways including hydraulic
fills, which use the coarse component of the tailings stream; paste backfill, whose
source is full-stream tailings, and rockfill, which may be generated from quarried rock
or mining waste rock. The main advantage of hydraulic fill and paste fill over rockfill is
their ability to be transported hydraulically. This provides benefits regarding material
handling costs, but once placed within a stope the transportation water can reduce
stability if not managed appropriately.

1.1
Mechanics of Mine Backfill Matthew Helinski
Introduction The University of Western Australia

This thesis is focused on tailings-based backfill, which includes both hydraulic and
paste fills. These are grouped under the term tailings-based backfill throughout this
thesis. Figure 1.1 provides an illustration of the processes involved at a typical tailings-
based backfill site. This figure illustrates the interaction of mining, processing and
filling activities. The relevant features of Figure 1.1 with respect to the backfill process
are as follows (the numbers refer to the key in Figure 1.1):

3. Shows ore being extracted from a stope.

27. Shows the concentrator, where the ore is crushed and the commodity
extracted.

36. Shows the backfill plant, which is where the tailings from the concentrator
are post-processed to generate the desired backfill product (this often
includes cycloning, thickening, filtration and cement mixing).

23. Shows the underground backfill delivery borehole/ pipeline system.

6. Shows the backfill slurry being deposited into a stope.

4. Shows a stope being blasted, which leads to the adjacent backfill mass being
exposed vertically, and the overlying backfill mass being exposed
horizontally.

The focus of this work is to improve the understanding of the processes involved during
the deposition of mine backfill into a stope. Figure 1.2 presents a schematic showing
backfill being deposited into a stope. This figure shows a typical stope with a
drawpoint that would previously have been used to extract ore from the stope. Fill
would be deposited into the stope from the top and containment barricades (also
referred to as bulkheads) would be constructed in the drawpoint to retain the fill from
flowing out of the stope and into the active mining environment. These barricades are
typically constructed from permeable bricks or sprayed fibrecrete. The terminology
presented in Figure 1.2 will be used throughout this thesis.

Figure 1.1 illustrates that, during the extraction of ore around existing fill masses, the
material is exposed vertically and horizontally. In order to maintain stability during
these exposures, small proportions of binder (cement) are often added to the tailings

1.2
Mechanics of Mine Backfill Matthew Helinski
Introduction The University of Western Australia

mix. As conveyed throughout this thesis, the addition of cement to the fill mass results
in additional complexities when attempting to characterise the deposition behaviour.

Traditionally the use of tailings-based backfill has involved the removal of fines from
the tailings stream (referred to in the industry as classification) to create a coarse
material, which is transported through pipelines in a turbulent manner with large
amounts of water (Thomas and Holtham 1989). The purpose of removing the fines is to
increase the hydraulic conductivity of the material so, when placed, transportation water
can drain away freely. Recently there have been significant developments in the field of
tailings-based mine backfill with the introduction of full-stream tailings into the
underground environment. This type of fill is transported with lower water contents than
hydraulic fills and the presence of fines allows the material to be transported at slower
velocities without creating segregation in the reticulation system. This material is
termed paste backfill and is comprehensively addressed by Landriault (1995, 2006).

The main focus of this thesis is to investigate the processes and mechanisms associated
with the placement and consolidation of tailings-based backfill, and develop a sound
methodology for understanding the mechanisms, and interaction of mechanisms, that
occurs during the filling process. The thesis examines the influence of these
mechanisms on total stresses applied to containment barricades as well as the
development of effective stress within cemented mine backfill during the hydration
process. Only by applying a rigorous approach to these problems, can certainty be
developed regarding these areas of concern.

In the mining industry there has, in the past, been an acceptance of operational rules of
thumb, which were developed through experience at other sites, or through experience
at the site of interest. These rules of thumb can result in a successful outcome, but as
noted by Baldwin (2004) each tailing stream is unique and the resulting paste
geomaterial behaviour can and does vary dramatically from one operation to another.
Therefore, without a fundamental understanding of the processes involved, it is
impossible to determine when subtle characteristics associated with a given situation
will result in the rules of thumb becoming invalid. In this situation, a variation from
the expected behaviour can result in catastrophic consequences such as the failure of

1.3
Mechanics of Mine Backfill Matthew Helinski
Introduction The University of Western Australia

containment barricades and inrush of fill material, or the failure of a fill mass during
exposure.

Between the period of December 2003 and December 2004, there were seven tailings-
based fill barricade failures worldwide; in 2006 there were at least five barricade
failures. Due to confidentially reasons barricade failures are often not reported in
literature. One example is that reported by Revell and Sainsbury (2007). These authors
present a number of different examples of barricade failures. Included in this document
are two examples where rapid filling rates resulted in excessive loading of barricade
structures. These failures resulted in a very large energy release with paste fill flowing
up to 250 m from the original bulkhead location. Figure 1.3 presents a photograph of a
failed barricade from this paper. In addition, this paper discussed various barricade (or
bulkhead) designs and presents details relating to the structural failure mechanisms.
Another hydraulic fill barricade failure, that was documented in a coroners report
(Coroners Report, 2001), is that at the Bronzwing mine in Western Australia where
three miners were killed. The overall conclusion of this investigation suggested that
inadequate stope drainage was the aspect that resulted in excessive loads on the
barricade structure.

In addition to addressing the problem of barricade loads, by developing a model that


rigorously captures the most important mechanisms associated with cemented mine
backfill processes, filling conditions can be varied throughout the whole spectrum of
tailings-based fill types without having to adopt simplifying assumptions to achieve a
result. This approach provides an understanding of which parameters control the
behaviour and which fundamental relationships need to be satisfied in order to deem
various rules of thumb appropriate.

1.1 SIGNIFICANCE OF CONSOLIDATION TO MINE


BACKFILL

Tailings-based backfill that is transported hydraulically, is a multi-phase system


consisting of a solid phase and water phase. Due to the large compressive stiffness of
the pore water, relative to the soil structure, the application of load (due to the accretion
of overlying material) creates an increase in the pore fluid pressure. Should there be no
dissipation of pore fluid pressure, there would be no increase in effective stress and the

1.4
Mechanics of Mine Backfill Matthew Helinski
Introduction The University of Western Australia

situation is said to be undrained. If the pore pressure is dissipated immediately, any


applied load would be immediately placed on the soil matrix in the form of effective
stress, and the situation is said to be drained.

Various mine sites have fill types that possess different mineralogy and particle size
distribution, while filling rates and stope sizes vary from site to site. As a result, the
degree of consolidation that occurs during filling also varies.

To complicate the consolidation process further, most mine backfills contain a small
component of cement, which acts to modify the consolidation properties by increasing
the stiffness and reducing the permeability.

It is important to appreciate that throughout this thesis, consolidation is defined as the


transfer of stress from the water phase to the soil phase. This should not be confused
with the compression of the soil matrix or the bonding of material, through cement
hydration, as has been noted throughout mine backfill literature. An appropriate
definition of consolidation is particularly important in relation to stiff cemented soil
where only small amounts of compression is required to mobilise large stresses.

Given the theoretical developments that are generated in this work the results will be
applied to mine backfill problems such as;

Providing a rigorous technique for estimating loads placed on barricade


structures

Providing an understanding of in situ curing conditions relative to those in the


laboratory

1.2 PROJECT METHODOLOGY

The overall objective of this thesis was to develop a means of representing the cemented
mine backfill deposition process. As will be outlined throughout this thesis, the
cemented mine backfill process involves the interaction of a number of complex
mechanisms, and in order to achieve the desired outcome simplification of many of
these mechanism was required.

The methodology adopted is to develop an understanding of the consolidation behaviour


of cementing soils through an assessment of past research in relevant fields. On this

1.5
Mechanics of Mine Backfill Matthew Helinski
Introduction The University of Western Australia

basis an assessment was undertaken of the applicability and limitations of previous


work to the mine backfill situation. Using previous work, a rational approach to
simulating the cemented mine backfill deposition process is formulated.

The second component of this project focuses on the utilisation of existing literature as
well as experimental testing, to formulate suitable models to represent various
mechanisms that occur during the mine backfill deposition process. Areas investigated
include the characterisation of strength, stiffness and permeability of the material prior
to cementation, but unlike in conventional soil mechanics, a major focus is placed on
the evolution of these properties due to cement hydration. Hydrating cement acts to
develop a structure that increases the material stiffness and yield strength, yet the
application of stress can potentially destroy this cemented structure. Furthermore, the
growth of cement hydrates has been shown to reduce the material permeability. In
addition, the cement hydration reaction creates a net volume reduction, which is termed
self desiccation. In later chapters this volume change is shown to have a significant
influence on the overall consolidation behaviour and must be incorporated into the
understanding of the mine backfill deposition process.

Cemented mine backfill placement essentially involves the interaction of three time-
dependent relationships, which include filling rate, consolidation rate and hydration
rate. In order to appropriately link these mechanisms, they are coupled numerically. The
first numerical stage included incorporation of the cementing soils material
characteristics into the one-dimensional, large strain, tailings consolidation program
MinTaCo (Seneviratne et al., 1996). This modified model was renamed CeMinTaCo,
and is used in a sensitivity analysis to investigate the interaction of mechanisms during
typical mine backfill runs. From this sensitivity analysis, the most significant
characteristics associated with the consolidation of cemented mine backfill were
identified.

A new two-dimensional (plane strain and axi-symmetric) fully coupled consolidation


model named Minefill-2D is then presented. Minefill-2D incorporates all of the relevant
cemented mine backfill mechanisms, as well as the ability to simulate the progressive
accretion of material and any stress redistribution onto the surrounding stiff rockmass.
This model is verified against a number of established analytical solutions as well as

1.6
Mechanics of Mine Backfill Matthew Helinski
Introduction The University of Western Australia

laboratory experiments (including both element testing and centrifuge modelling) and in
situ data.

Having gained appropriate confidence in Minefill-2D, a sensitivity study was


undertaken. This study investigates the fundamental difference in mechanisms between
full-stream tailings backfill (paste fill) and classified-tailings backfill (hydraulic fill). A
sensitivity study was undertaken for both full-stream and classified-tailings backfills, to
highlight what are likely to be important, and less important, characteristics when
dealing with the various fill types.

The results of the sensitivity study are used to provide some guidance when designing
and managing filling operations.

1.7
Mechanics of Mine Backfill Matthew Helinski
Background & Literature Review The University of Western Australia

CHAPTER 2
BACKGROUND & LITERATURE REVIEW

2.1 INTRODUCTION

The purpose of this chapter is to provide an overview of the literature that is relevant to
this thesis. Firstly, there is a discussion regarding the state of the art in mine backfill.
This provides an overview of existing methods being applied in the mining industry and
highlights a number of areas where further research can deliver an improved
understanding.

This is followed by a review of previous work that is considered to be relevant to this


thesis. Specifically, topics addressed include consolidation theory, structured-soil
mechanics, and cement technology. In addition to the overview of previous work, at the
end of each section there is a brief description of how the literature has been applied in
the context of mine backfill throughout this thesis.

2.2 MINE BACKFILL LITERATURE

In conventional surface tailings disposal, consolidation is important since this


mechanism dictates the settlement of the tailings mass and therefore the quantity of
material that can be placed into a storage facility. In addition, the stress history of
tailings is important when determining the undrained shear strength of the material. As a
result, authors such as Gibson (1967), Williams (1988), Toh (1992), Seneviratne et al.
(1996) and Newson et al. (1996) have thoroughly investigated the consolidation
behaviour of surface tailings facilities.

In an underground environment, the degree of settlement is not of major importance and


since the material is often cemented, the undrained shear strength is less likely to be
influenced by stress history. Therefore the relevance of consolidation might be
questioned.

2.2.1 Influence of consolidation on barricade stresses

In the underground backfill environment, the surrounding rockmass is significantly


stiffer than the material being deposited. Rankin (2004) investigated numerically the

2.1
Mechanics of Mine Backfill Matthew Helinski
Background & Literature Review The University of Western Australia

development of stress in the backfill during filling and found that much of the vertical
stress can be redistributed to the surrounding rockmass. This work neglected pore
pressures in the calculation of the stress distribution and as shown later in the thesis, this
approach can lead to a gross oversimplification of the process.

To understand how stress is redistributed to the surrounding rockmass requires an


understanding of the effective stress. And in order to understand effective stress, an
understanding of the consolidation process is required.

To demonstrate the significance of consolidation (or specifically the influence of pore


water pressure) on the stress distribution in a backfill mass, Helinski et al. (2006)
undertook a series of numerical simulations using the finite difference program FLAC
(Fast Lagrangian Analysis of Continua). This analysis involved filling a plane strain
stope (20 m wide by 50 m high), with fully hydrated cemented paste backfill, assuming
either fully drained or undrained conditions. The fully drained case neglected the
influence of pore water pressure, while the undrained case assumed that the material is
placed without any consolidation. Neglecting the influence of pore water pressure in the
fully drained case is considered a valid representation of the best case scenario as the
water table is assumed to have been drawn to the bottom of the stope, and with a low air
entry suction value (for non-plastic tailings), this would result in atmospheric pressures
existing throughout most of the fill mass.

Material properties adopted in this analysis were those quoted by Rankin et al. (2001)
for Cannington paste fill. These included a unit weight of 20 kN/m3,Youngs modulus
(E) of 60 MPa and Poissons ratio () of 0.25. These parameters equate to a shear
modulus (G) of 24 MPa and a drained bulk modulus (K) of 40 MPa. The friction angle
() adopted was 25 and, to demonstrate the influence of bond strength (or lack
thereof), an artificially high cohesion (c) of 25 MPa was adopted.

Figure 2.1 presents the calculated total vertical stress down the centreline of the stope at
the end of filling, when the material has been placed in both a drained and an undrained
manner. Also presented in Figure 2.1 is the total vertical stress assuming no stress
redistribution to the surrounding rockmass.

Figure 2.1 demonstrates that there is a significant difference between the total vertical
stress down the centreline of a stope, if filling is carried out under fully consolidated

2.2
Mechanics of Mine Backfill Matthew Helinski
Background & Literature Review The University of Western Australia

(drained) or fully unconsolidated (undrained) conditions. With fine-grained material


(with a high air entry suction value) drawing the water table down through the fill mass
may cause suctions to develop within the fill mass. The development of suctions would
increase the amount of arching in the drained case, increasing the influence of pore
pressure on the arching process.

Even with fully-hydrated cemented backfill, if there is no consolidation there is very


little stress redistribution to the surrounding rockmass, even with an inflated value of
cohesion. In order to mobilise any shear stress at the fill/rock interface, shear strains are
required. To generate shear strain the material must settle, but under undrained
conditions the compressive stiffness of the bulk material is that of water, which is very
high, and therefore, very little settlement occurs. It is not until water is squeezed out of
the fill mass that soil compression (settlement) can occur and shear strains at the
interface can occur, generating the arching effect. The lower vertical stress for the
drained case in Figure 2.1, is due to this arching effect.

It is interesting to note that, in the lower 10 m of the stope in the drained case, there is a
linear increase in vertical total stress. The reason for this is that, towards the stiff base of
the stope, the total amount of settlement (vertical deformation) is reduced, resulting in
less shear strain being mobilised at the interface and therefore less shear stress. As a
result less stress redistribution or arching occurs towards the base even in a drained
situation. In addition, as the name implies, a certain vertical height is required to
accommodate the arch in the material, so that arching only takes effect above a
certain distance from the base.

The results from this numerical study can be extended to the total horizontal stress
placed on the barricade for the two extreme cases. It should be noted that in both cases
the pore pressure immediately behind the barricade is set to zero. The total horizontal
stress calculated at the barricade location is plotted against height up the barricade in
Figure 2.2 for both the drained and undrained cases. This Figure illustrates that if filling
is carried out undrained, barricade stresses are very high ( 800 kPa) while drained
filling results in much lower barricade stresses ( 80 kPa).

Figure 2.1 demonstrates that without consolidation, there is little stress redistribution
and high vertical total stresses. In addition to these high total vertical stresses within the

2.3
Mechanics of Mine Backfill Matthew Helinski
Background & Literature Review The University of Western Australia

stope, the undrained filling case also results in higher horizontal total stress for a given
vertical total stress. This may be understood with reference to Figure 2.3, which
illustrates Rankins lateral earth pressure theory. This theory relates the vertical total
stress ( v ), vertical effective stress ( v ), pore pressure (u) to the horizontal total stress

( h ) in accordance with Rankins lateral earth pressure coefficient K0. For most
uncemented tailings K0 is typically in the range of 0.3-0.5.

If all of the self-weight stress were supported by the water phase, the horizontal total
stress would be equal to the vertical total stress. As consolidation occurs and the self-
weight stress is transferred off the water phase and onto the soil structure, the horizontal
total stress becomes less than half of the vertical total stress, for typical Ko values.

Consolidation also influences the horizontal effective stress within a stope. In order to
generate frictional shear strength between the fill mass and the surrounding rockmass,
horizontal effective stress is required. If all of the self weight vertical stress is being
carried by the water phase, the horizontal total stress and pore pressure would be equal.
In this case, the horizontal effective stress would be zero. Consolidation reduces the
pore pressure, increasing the horizontal effective stress and allows some frictional shear
strength to exist at the rock/fill interface.

Gibson (1958) investigated analytically the amount of consolidation in a deposit of a


saturated soil where the thickness of the deposit is increasing with time. He derived a
relationship between the development of excess pore pressure (uex), the coefficient of
consolidation (cv) , the filling rate (m) and the duration that filling has been ongoing (t),
in a one-dimensional situation.

Applying this analytical solution, Gibson (1958) developed a chart to relate a non-
dimensional time term (T) to the excess pore pressure (uex). In this solution, uex is
represented by a gradient of uex against depth relative to the gradient of uex that would
be created with no consolidation. This ratio is defined as (du/dz)/. If this ratio is equal
to unity, the total stress and pore pressure are equal (meaning that there is no
consolidation) while if equal to zero there is no excess pore pressure (i.e. complete
consolidation).

The non-dimensional time constant is defined as;

2.4
Mechanics of Mine Backfill Matthew Helinski
Background & Literature Review The University of Western Australia

m 2 .t
T=
cv (2.1)

Gibsons chart to relate the degree of consolidation that occurs during placement to the
non-dimensional time constant is reproduced as Figure 2.4. The relevance of this figure
to tailings backfill is discussed further later.

The situation analysed is well suited to the case of tailings-based backfill placement into
an open stope. While it might be argued that the placement of minefill into a stope does
not strictly conform to the one-dimensional assumption due to the potential for arching
to the surrounding rockmass, neglecting arching during the early stages of placement
can provide a reasonable representation of most situations.

The other point of question may relate to the base boundary conditions. In most
situations it is reasonable to consider free-draining conditions at the barricade location.
But, as illustrated in Figure 1.2, the drawpoint flow area is often significantly smaller
than that of the stope. Furthermore, as filling progresses away from the base, the
distance to this boundary increases, making drainage towards the top boundary more
likely. As a result, it may be more appropriate to ultimately assume an impermeable
condition at the base. Gibson (1958) analysed both permeable and impermeable
boundary conditions at the base.

Qiu and Sego (2001) undertook a series of laboratory experiments on full-stream gold
and copper tailings. One component of this work included large strain consolidation
testing of this material to determine cv. The tests indicated cv values of 14 m2/yr and 25
m2/yr for the gold and copper tailings, respectively, over the density range typical for
paste backfill. Vick (1983) suggests that cv for classified tailings (which would be
representative of hydraulic fill) ranges between 1500 m2/yr and 300,000 m2/yr and cv for
full-stream tailings ranges from 3 m2/yr to 300 m2/yr.

To develop an understanding of the degree of consolidation likely to occur during filling


in typical mine backfill operations, these suggested cv values were combined with a
filling rate of rise of 5 m/day (a typical filling rate) and a 40 m high stope to calculate
the respective dimensionless time factor (T). The location of the various material types
is superimposed onto Gibsons consolidation chart in Figure 2.4.

2.5
Mechanics of Mine Backfill Matthew Helinski
Background & Literature Review The University of Western Australia

Figure 2.4 therefore indicates that during the placement of hydraulic fill it is unlikely
that excess pore pressures would develop unless very high filling rates are adopted. In
contrast, for full-stream tailings (paste fill), it indicates that it is very unlikely that
consolidation would occur. This suggests that, following the logic outlined earlier, loads
applied to barricade structures can be extremely high in full-stream tailings backfill in
an uncemented state. It should be noted that Gibsons chart only considers excess pore
pressures, neglecting hydrostatic pore pressures. However, as discussed later in this
document, the presence of hydrostatic type pore pressures (generated as a result of a
flow restriction through the drawpoint area) also make the influence of pore water
pressure relevant to a hydraulic filling scenario.

To the authors knowledge, the only case where the influence of pore pressure has been
incorporated into the analysis of mine backfill deposition mechanics is in the work
described by Kuganathan (2002). This work presented an analytical solution for
estimating barricade stresses in hydraulic fill stopes, which incorporates steady state
seepage-induced pore pressures1 within the stope drawpoint. This solution is based on a
limit state analysis, which incorporates the influence of pore pressures only within the
drawpoint of a stope. Being a limit state method, this approach assumes the mobilisation
of the ultimate material strength at the rock/fill interface and there is no consideration of
the influence of pore pressures on the stress distribution within the stope, nor is there
consideration of excess pore pressures that may be created during filling. Therefore,
while this technique may provide a reasonable indication of barricade stresses under
some conditions, these limitations can lead to oversimplification in many cases
(particularly cases involving fine-grained full-stream tailings backfill).

Another approach that has previously been adopted to establish an understanding of the
stresses placed on barricades is direct stress measurement using diaphragm-type earth-
pressure cells. Revell (2002) and Belem et al. (2004) used total pressure cells to
measure the loads placed on paste fill retaining structures during filling. Based on the

1
Steady-state seepage pore pressures will be properly defined later. Briefly, this is the pore pressure
profile that is reached when steady-state seepage is established within the stope, in equilibrium with the
boundary total head conditions (the boundaries being the top fill surface and the barricade).

2.6
Mechanics of Mine Backfill Matthew Helinski
Background & Literature Review The University of Western Australia

measurements, it was concluded that the amount of load transferred to the wall reduced
as the cement hydration proceeded.

Clayton and Bica (1993) demonstrated that the stress measurements from earth-pressure
cells are typically lower than the true values; this is called under-registration. The
degree of under-registration is related to the relative stiffness between the cell and the
surrounding soil. As hydration proceeds, the stiffness of a cemented soil increases
significantly, and therefore the degree of under-registration in such circumstances could
also change. Therefore, questions may arise as to whether the reduction in stress
measured by Revell (2002) and Belem at al. (2004) is a true reduction in barricade
stress or if it is simply a result of the change in under-registration due to the progress of
hydration.

Stone (personal communications 2007) suggests that the biggest problem facing the
mining industry with regard to mine backfill is uncertainty over barricade loads. The
need for further research into loads being placed on mine backfill retaining structures
has also been recognised by Le Roux (2004) who suggests that there is a need for a
renewed focus on barricade design with high rates of filling. McCarthy (2007)
suggests that mine backfill barricades are a problem in the mining industry and that the
problem is technically complex and there is a need for more research in this area.

2.2.2 Influence of consolidation on in situ strengths

It has been documented that in situ backfill strengths are often significantly greater than
those measured in the laboratory, for the same mix (Cayouette, 2003, Revell 2004). It
has also been well established that the application of effective stress to a cementing soil
prior to or during hydration can increase the material strength (Blight and Spearing
1996, Consoli 2000, Rotta 2003). The improvement in strength is said to be a result of
soil matrix compression (which leads to an increased density) as well as an
improvement in the intimacy of the contact points. This topic has been researched
experimentally by a number of different authors such as Blight and Spearing (1996), Le
Roux at al. (2002) and Belem et al. (2006).

2.7
Mechanics of Mine Backfill Matthew Helinski
Background & Literature Review The University of Western Australia

The work by Blight and Spearing (1996) focused on investigating the influence of
closure strains2 on the strength of mine backfill. This work demonstrated that as the
strain rate is increased there is an associated increase in strength. As this work was
carried out in the context of strains induced by ground movement in classified tailings
backfill, a strain criterion was adopted and there was no need for an effective stress
consideration. The loading rate adopted by these authors is less likely to be relevant to a
large open stoping situation where the application of total stress during filling is usually
dependent on the self weight of the overlying fill mass.

Le Roux et al.(2002) investigated the impact of applying load in a one-dimensional


situation to a sample during curing. The results indicated that higher cement contents
reduced the degree of compression for the same loading sequence. The deficiency with
this work was that the rate of loading adopted was in accordance with the rate of
application of self weight total stresses in a typical filling situation rather than that of
effective vertical stress. In order to develop a rational approach to the loading rate that
should be applied, it is the rate of development of effective stress in the field that must
be matched, rather than the total stress. As demonstrated in this thesis, this is
particularly relevant during the early stages of loading where the pore pressures in a
stope can increase at the same rate as the total stress, resulting in no change in effective
stress during the early stages of hydration.

Belem (2006) undertook a series of column filling tests. The columns used in this
work were 3 m high, and 300 mm square in plan. Three columns were tested: the first
had impermeable vertical boundaries, the second had an impermeable top half and a
permeable bottom half and the third had completely permeable boundaries. These
columns were filled with paste backfill containing cement. At the end of the tests, the
resulting cemented material was investigated. This investigation indicated that drainage
created a significant increase in density which resulted in an increase in material
strengths. These tests were meant to represent possible conditions within a stope, but
due to the significant reduction in drainage path length associated with the boundary

2
Closure strains refer to the inward movement of the walls of a stope after the fill is placed and the
hydration process is occurring.

2.8
Mechanics of Mine Backfill Matthew Helinski
Background & Literature Review The University of Western Australia

the rate of effective stress development is not considered representative of the field
situation1.

Therefore, in order to improve the understanding of the mine backfill deposition process
and to start to develop a rational approach for understanding the interaction of cement
hydration and the application of effective stress, an understanding of the mine backfill
consolidation process is required. This aspect has been recognised by Le Roux (2004)
who suggests that it is necessary to understand the influence of hydrating cement on
the consolidation properties of the paste fill and how this influences the final material
properties and affects the performance of the material.

2.2.3 Influence of consolidation on exposure stability

During the removal of stopes adjacent to a cemented fill mass (as illustrated in Figure
1.1 point number 3), there is a reduction in confining stress, which reduces the stability
of the fill mass. If the fill mass has insufficient cohesive strength, fill failure can occur,
resulting in dilution of the ore being mined in the adjacent stope.

A number of authors have proposed analytical solutions for estimating vertical exposure
strength requirements. The most commonly adopted analytical solutions include the
upper bound mechanisms by Mitchell and Wong (1982) and those based on the limit
state arching theory of Marston (1930) and Terzaghi (1943) (such as those presented by
Winch 1999 and Aubertin et al. 2003). The interesting point about these two commonly
adopted limit state techniques is the strength parameters used to represent the contact
between the fill mass and the surrounding rockmass. In the upper bound mechanism
proposed by Mitchell and Wong.(1982), it is assumed that the interface is cohesive,
while the technique presented by Winch (1999) and Aubertin et al. (2003) is based on a
frictional contact strength only.

During the filling process, the interaction between the development of shear strength
and the application of shear stress may modify the cohesive properties at this interface.
Such interaction is dependent on the material properties, boundary conditions, filling

1
The limitations of using scaled models (whether on a geotechnical centrifuge or at 1g) for this type of
investigation is that the timescale of consolidation/drainage is reduced by reducing the scale, but the
timescale for hydration is not. This is explained in much more detail in Chapter 6.

2.9
Mechanics of Mine Backfill Matthew Helinski
Background & Literature Review The University of Western Australia

rate and the consolidation rate, but without appropriately understanding the filling
process it is difficult to establish an understanding of this important characteristic.

2.2.4 Summary

This section has presented previous work in the field of mine backfill and demonstrated
the need for a soil mechanics approach to understanding the mine backfill deposition
behaviour. The remainder of this chapter focuses on work carried out in the fields of
consolidation, structured-soil and cementation which is considered relevant to the
cemented mine backfill application.

2.3 CONSOLIDATION

A solution for the process of consolidation in a saturated soil was developed by


Terzaghi, on the basis of a number of simplifying assumptions, such as:

the pore water is infinitely stiff relative to the soil skeleton, such that an
application of external stress results in generation of a pore pressure equal to
that stress

the permeability (hydraulic conductivity) and compressibility of the soil does


not change during a loading increment

Darcys law applies i.e. the rate of water flow in the soil depends on the
permeability and the gradient in the total head.

For slurries consolidating under self-weight stresses, many authors, such as Carrier
(1982), pointed out that the assumption of constant permeability and compressibility of
the matrix is not correct, due to the large changes in void ratio that occur as
consolidation proceeds, resulting in significant change in compressibility, and even
more significant change in permeability.

The Terzaghi model was extended by Gibson (1958) to simulate the progressive
sedimentation of material, as discussed in the previous section. This model was based
on small-strain theory and restricted to material properties that remain constant
throughout the consolidation process. The main advancement in this work was to take
account analytically of an increase in drainage path length as well as the inclusion of
self-weight stresses.

2.10
Mechanics of Mine Backfill Matthew Helinski
Background & Literature Review The University of Western Australia

In order to take account of the significant nonlinearity in large-strain consolidation


problems, researchers began investigating finite strain consolidation models. This
included work by Mikasa (1965) and Gibson et al. (1967, 1981). The methodology uses
a Lagrangian coordinate system where the boundaries move in accordance with the
evolving soil dimensions. Through applying this theory numerically, the authors were
able to rigorously account for variations in geometry as well as variations in
permeability and stiffness during compression. Based on this work, Gibson et al. (1981)
demonstrated that conventional small-strain theory had the potential to significantly
underestimate excess pore pressures. Tan and Scott (1988) continued the comparison of
small- and large-strain theory, suggesting that Terzaghis small-strain solutions were
only applicable for strains less than 20%.

With modern developments in computational efficiency, numerical programs


specifically focused on solving the large-strain consolidation equations for the purpose
of understanding the mine tailings deposition process, have been generated (Williams et
al., 1989, Tao, 1992 and Seneviratne et al., 1996).

In addition to the numerical and analytical developments, significant research effort has
been dedicated to understanding material parameters including oedometer testing,
constant and falling head permeability tests as well as Rowe cell testing that captures
volume and permeability changes under various confining stresses. Based on results
from this type of testwork, authors such as Carrier et al. (1983) developed models to
represent the large strain compressive behaviour of soil.

2.3.1 Consolidation behaviour of cementing soil

The development of a one-dimensional consolidation model for soil undergoing


cementation (CeMinTaCo) is described in detail in Chapters 3 and 4. However, a brief
overview is given here, to show where it fits with respect to other models in the
literature.

The basis for this model is the large-strain consolidation equation derived by Gibson et
al. (1981). The governing equation is derived through combining equilibrium of the soil
and water phases using Darcys law to represent flow. Combining these equations and
maintaining continuity, the following governing equation was derived to represent one-
dimensional large strain consolidation of uncemented slurry.

2.11
Mechanics of Mine Backfill Matthew Helinski
Background & Literature Review The University of Western Australia

u k 1 + eo u k v
+ (1 + eo ) v + =
t
{ 144244 e a
3 144w44
1 + e a a {
t
(A ) 42444 443 (B )
(D ) (C ) (2.2)

where a is a Lagrangian coordinate, u is pore pressure, k is hydraulic conductivity


(permeability), and v and v are the vertical effective and total stresses, respectively.
In this equation, term A is the rate of change in pore pressure as a result of a rate of
application of total stress (term B), term C is the volumetric strain, which is dictated by
the hydraulic conductivity (k) of the material, and term D is the current stiffness of the
material.

During the consolidation of uncemented mine tailings from an initial slurry state, both
the stiffness (compressibility) and the permeability of the material change as the void
ratio reduces. When cement is added, these changes still occur but a number of other
mechanisms associated with cement hydration are introduced. These include the
development of cement-induced stiffness and strength, a reduction in permeability and
the introduction of a new mechanism that is referred to as self desiccation.

The modified governing equation is presented here:

u e v
1 (t , C c , e, v )
t
14 K e
4 4 4w4 42 4 4 4 4 44 3
(A)


+ v (t , C c , e , v )(1 + e0 )
( ) ( )
k e eff 1 + e0 u k e eff (2.3)
+
14e 4 4 42 4 4 4 43 a w 1 + e a a
14 44 4 44 42 4 4 4 4 4 44 3
(D) (C)
V (t , C c ) v
+ v (t , C c , e , v ) sh =
14e 4 4 4 42 4 4 4 4 t 43 { t
(E) (B)

In this equation, the modified terms are identified using the same labels as the
equivalent unmodified terms (in Equation 2.2) but there is an additional term (term E)
that does not have an equivalent term in the unmodified equation.

Due to the soil matrix stiffness potentially approaching that of water, the change in pore
pressure (term A in Equation 2.2) resulting from a change in total stress, requires

2.12
Mechanics of Mine Backfill Matthew Helinski
Background & Literature Review The University of Western Australia

v
modification to take account of the cemented matrix stiffness ( ) and the bulk
e
modulus of water (Kw) to satisfy strain compatibility.

The volumetric strain term (term C in Equation 2.2) requires modification to take into
account the fact that the permeability (k) is affected not just by the normal void ratio
reduction due to consolidation, but also by the formation of cement gel in the void
space.

The stiffness term (term D in Equation 2.2) requires modification to express the material
v
stiffness ( ) as a function of cement hydration, current stress state and previous
e
stress excursions.

An additional term (term E) needs to be introduced to take account of volumetric


Vsh
changes ( ) associated with the cement hydration process (the self-desiccation
t
process referred to above).

The derivation of the modified material models is presented in Chapter 3 while the
derivation of the modified governing equation for one-dimensional consolidation of a
cementing soil (Equation 2.3) is presented in Chapter 4.

2.4 STRUCTURED SOIL

The term structured soil refers to any soil that has a true cohesive strength. This may
include artificially-cemented soils or natural soils that may have been cemented through
natural processes, such as calcite precipitation.

Clough et al. (1981) pioneered the research into cemented soils by investigating
experimentally the impact of cementation on both naturally-cemented and artificially-
cemented soils. This work demonstrated that cementation creates true cohesion but has
little impact on the friction angle. It was also shown that density, grain-size distribution,
and grain shape all have a significant influence on the behaviour of cemented soils.
Clough et al. (1989) continued this work on cemented soils with experimental
investigations into the cyclic loading of cemented soils.

2.13
Mechanics of Mine Backfill Matthew Helinski
Background & Literature Review The University of Western Australia

In the late 1970s - early 1980s during the construction of offshore oil and gas
platforms on the North West shelf of Australia, piling problems were encountered. As
discussed in Jewell et al. (1988), the major reason for these problems was associated
with a lack of understanding regarding the behaviour of structured soils.

In response to these problems, a significant focus was placed on understanding the


behaviour of structured soils. Leroueil and Vaughan (1990) were the first to present a
comprehensive framework for understanding structured soils, with a particular focus on
the behaviour of residual soils. They discuss how the shape of the yield surface varies
with curing stresses. The different failure modes that would occur through loading along
different stress paths were also discussed. In this framework it was shown that shearing
would produce a localised failure plane, while compressive stresses would destroy the
structure more uniformly, leading to volumetric contraction. The authors also
introduced the concept of the structure-permitted space, which is illustrated in Figure
2.5. This is a region in volumetric compression space (e - p) where only material with
structure can exist.

Coop and Atkinson (1993) presented a framework for understanding the behaviour of
cemented carbonate sand. This work showed that when sheared, the materials in either a
cemented or uncemented state approach the same critical state. Also, they suggest that
the cemented friction angle is slightly lower than the uncemented equivalent. This is
attributed to surface coating. They demonstrate that, as with overconsolidated soils,
shearing under low stress results in high stress ratios with post-yield strain softening,
while shearing under high mean stress results in the cementation having little influence
on the ultimate strength. It was also shown that cementation can significantly increase
the elastic compressive stiffness.

Cuccovillo and Coop (1997) investigated the yielding and pre-failure deformation of
cemented soils. They demonstrated that a cemented soil showed two major yield points
in loading, these being the stress where the material stiffness starts to reduce below the
small strain stiffness (termed Y1 yield stress) and the point where significant reduction
in stiffness is observed (termed Y2 yield stress). Y1 is said to be the first onset of
structural breakdown. The rate of stiffness change at Y2 was shown to be a function of

2.14
Mechanics of Mine Backfill Matthew Helinski
Background & Literature Review The University of Western Australia

the state of the soil fabric. It was shown that a loose calcarenite reduces in stiffness
significantly more than dense silica sandstone.

Asaoka et al. (2000) investigated the significance of structure in clayey soil on the
consolidation/ compression behaviour. It was concluded that prior to yield, the higher
stiffness of the structured soil (relative to the destructured soil) resulted in faster rates of
consolidation. However, as the material was loaded beyond its yield point (Y2), the
stiffness of the structured soil was less than that of the destructured soil (under the same
stress) and, as a result, consolidation of the structured material occurred over a longer
period with greater settlement.

More recently, there has been renewed interest in the behaviour of soils cemented
artificially for the purpose of soil improvement. This research effort has been primarily
driven by Consoli and his co-workers, as detailed below.

Consoli et al. (2000) showed that the stress state during curing plays a significant role in
the mechanical behaviour of the soil. Schnaid et al. (2001) investigated the triaxial
behaviour of cemented soils experimentally. They showed that the shear strength of a
cemented soil can be appropriately determined from the unconfined compressive
strength and the uncemented friction angle. For the confining stresses used in the
testwork, the secant modulus is unaffected by the confining stress, suggesting that the
stiffness is more a measure of the bond strength.

Rotta et al. (2003) investigated isotropic yielding of a cemented soil. This work
involved an experimental investigation of material cemented at different densities,
under different stress levels. They introduced the concept of the incremental yield
strength, which is considered to be the contribution of cementation to the increase in
yield stress of the material. The focus of this work was to develop a rational approach to
determining the material characteristics that influence this incremental yield strength. It
was demonstrated that this value is dependent on the degree of cementation and the
material state.

Finally, Consoli et al. (2006) combined the results of Schnaid et al. (2001) and Rotta et
al. (2003) to develop a unified framework for understanding the strength of cemented
soils. This work demonstrated that the incremental isotropic yield stress and initial bulk

2.15
Mechanics of Mine Backfill Matthew Helinski
Background & Literature Review The University of Western Australia

modulus can be linearly related to the unconfined compressive strength for a cemented
material.

Yin and Fang (2006) undertook experimental studies into the influence of cement-
treated clay columns on the consolidation of an otherwise untreated clay mass. These
results indicated that the presence of the cement-treated material played a significant
role in accelerating the consolidation of the mass. This was primarily attributed to the
increase in material stiffness.

A number of different constitutive models have been developed to represent the


behaviour of structured soils (Gens and Nova, 1993; Lagioia and Nova, 1995; Rouainia
and Wood, 2000; Kavvadas and Amorosi, 2000). Comparison with experimental data
indicates that these models provide a good representation of the soil behaviour. But in
order to provide this representation, significant mathematical detail is required. While
these types of models are considered superior to those presented in this thesis, at the
time of writing this thesis, such complexity was considered a second order effect in the
context of this work.

Liu and Carter (2002, 2005) present a modification of the well known Cam-Clay model
to represent the behaviour of structured soil. This model is referred to as the Structured
Cam-Clay model. The concept behind this model is that the yield surface for the
cemented soil is simply an expansion of the original yield surface for the uncemented
soil, and the magnitude of this expansion is dependent on the degree of cementation.
This concept is illustrated in Figure 2.6. As the soil is loaded, the model allows for the
degradation of the structured yield surface and hardening of the uncemented yield
surface. This can occur at the yield point (virgin yielding) as well as for loading
excursions within the yield surface (sub-yielding), but beyond a particular stress level
(Y1 as defined by Cuccovillo and Coop, 1997). This logic is the same in both the
compression and shearing stress paths.

2.4.1 Modelling structured soil behaviour

In order to characterise the material response for one-dimensional compression


modelling, a modified version of the Structured Cam-Clay model was adopted. This
model provides the flexibility to represent the compression behaviour initially in
accordance with the uncemented Cam-Clay model and to take account of the presence

2.16
Mechanics of Mine Backfill Matthew Helinski
Background & Literature Review The University of Western Australia

of cementation through an increased yield surface. The Structured Cam-Clay model also
allows for the yield surface to be degraded (potentially back to that of the uncemented
soil) in accordance with the destructuring functions suggested by Carter and Liu (2005).

To represent the size of the yield surface, the methodology presented by Rotta et al.
(2003) and Consoli et al. (2006) is used. The concept of an incremental yield stress is
incorporated to separate the cemented yield stress from the uncemented yield stress. The
magnitude of the incremental yield stress is said to be a function of the degree of
cementation and material state in accordance with the function suggested by Rotta et al.
(2003). Furthermore, based on the findings of Consoli et al. (2006), the unconfined
compressive strength and small-strain bulk stiffness is said to be proportional to the
incremental yield stress along the isotropic and one-dimensional stress paths. Details of
this approach are presented in Section 3.2.

2.5 CEMENTATION
Combining Sections 2.3 and 2.4, a rational methodology for the consolidation analysis
of a cemented soil can be developed. However, in order to appropriately represent the
cemented mine backfill process, a model that can appropriately represent the
consolidation of a cementing soil is required. In this work, previous work in the field of
cement and concrete research was taken into account.

Pioneering work in the field of cement research is described in a series of publications


by Powers and Brownyard (1947). These publications were the culmination of a
research program, by the Portland cement association, aimed at understanding the
behaviour of Portland cement paste. This work was later summarised in a single concise
document by Brouwers (2004).

This, and subsequent work by Powers (1958, 1979), developed the first model for
understanding cement hydration. When the cement particles are undergoing hydration, a
number of chemical reactions take place. These reactions result in the growth of
hydrates that effectively act to connect particles. An illustration of this process (taken
from Illstron et al., 1960) is presented in Figure 2.7. Powers and Brownyard (1947)
suggest that the cemented structure is made up of solid particles and cement gel. With
this, they introduce the concept of non-evaporable water, which is bound within the

2.17
Mechanics of Mine Backfill Matthew Helinski
Background & Literature Review The University of Western Australia

solid material, and gel water, which exits between gel particles. Due to the large surface
area of the very fine gel particles, van der Waals forces bind gel water within the gel
phase.

Based on their research, Powers and Brownyard (1947) determined that the specific
gravity of the unhydrated cement is 3.17 while that of the hydrated gel solids is 2.43 and
that of the gel including pores is 1.76. This indicates that, when combined with water in
the hydration process, the unhydrated cement volume increases by approximately 80%
(including gel products).

It was also found that the weight of chemically-bound water (termed non-evaporable
water) used in cement hydration is 23% of the weight of unhydrated cement.
Furthermore, it was found that due to the change in volume stoichiometries during the
cement reactions, there is a net reduction in volume from the unhydrated cement and
water to the final hydrated cement product. This change in volume (in cm3) was shown
to be 27% of the weight of non-evaporable water. In conventional concrete literature,
this volume reduction is termed self desiccation as it acts to desaturate lean concrete
mixes. This mechanism is important with respect to conventional concrete mechanics
because of the way that it influences shrinkage cracking (termed autogenous
shrinkage). Researchers such as Powers and Brownyard (1947), Hua et al. (1995),
Koenders and Van Breugel (1997), Bentz (1995) and Brouwers (2004) investigated the
impact of this mechanism on the shrinkage of conventional concrete.

Illstron (1979) presents the rate of hydration for the various compounds that make up
cement paste. This work indicated that the four main compounds react at vastly different
rates, ranging from C3S, which achieved 70% of its ultimate strength after 28 days, to
C2S which has achieved only 5% of its strength after 28 days.

Rather than simulating the rate of hydration for each individual compounds, other
researchers have developed empirical relationships to represent the rate of hydration for
the overall cement product (Rastrup, 1956, Guo, 1989 and Sideris, 1993). These authors
term the rate of hydration maturity and essentially fit various curves to the
development of cement related characteristics (maturity) against time.

The growth of cement hydrates increase the material stiffness and strength, due to the
bonding of particles. This has been well documented by authors such as Powers and

2.18
Mechanics of Mine Backfill Matthew Helinski
Background & Literature Review The University of Western Australia

Brownyard (1947), Illstron (1979) and Sideris et al. (2004). The strength achieved is
said to be a function of the cement content as well as the water-cement ratio. For a
saturated soil, this is consistent with the findings of the cemented soil literature, which
suggests that the strength is dependent on the density of the cemented soil (Consoli. et
al. 2006).

During the hydration process, some of the water is converted from a free liquid to either
solid or gel product, which forms a component of the soil structure and occupies some
of the previous void space. As already mentioned, Powers and Brownyard (1947)
suggest that the hydrated solid and gel is 80% greater than the original unhydrated
cement volume. Just as with a reduction in void space from soil compression, the
infilling of voids by cement gel acts to reduce the permeability of the material. This
aspect has been researched in the concrete literature for the purpose of reinforced
concrete corrosion resistance (Garboczi and Bentz,1995, Breysse and Gerard, 1997, and
Bentz et al. 1998).

Garboczi and Bentz (1995) present the concept of critical porosity. This is defined as
the point where the cement has enlarged sufficiently that the combination of the cement
solids and the gel structure create a seal across the entire cross-sectional area of the
sample. Since the cement gel is composed of platelets with very large surface areas, this
situation produces a significant reduction in permeability; in fact, at this point the
material is considered to be impermeable.

Based on the volumetric changes suggested by Powers and Brownyard (1947), Figure
2.8 was developed to illustrate the relationship between void ratio and cement content
required to achieve the critical porosity (as described by Garboczi and Bentz, 1995).
Also presented in Figure 2.8 is the range over which typical mine backfills are
produced. This indicates that, due to the relatively low density and small cement
contents, it is very unlikely that a typical mine backfill would ever approach critical
porosity. Therefore, while the presence of cementation acts to reduce the permeability
(and should be taken into account), this reduction is not expected to render the material
impermeable.

2.19
Mechanics of Mine Backfill Matthew Helinski
Background & Literature Review The University of Western Australia

2.5.1 Cementation behaviour

In order to represent the cementation process, research in the field of concrete


technology (which was presented in the previous section) was investigated. A brief
overview of how this logic is applied has been provided in this section.

While each of the maturity functions (discussed in the previous section) appears to
provide a reasonable fit to experimental data, the exponential relationship suggested by
Rastrup (1956) was adopted to represent the maturity of cement throughout this thesis.
This work provides a simple exponential relationship between cement hydration and
time. Investigations presented later in this thesis indicate that the rate of change of
strength, stiffness, permeability and volume can all be appropriately represented using
the same maturity relationship. Details of this approach are provided in Section 3.2.

Combining the total volumetric changes recommended by Powers and Brownyard


(1947) with the maturity function presented by Rastrup (1956), a function is developed
to represent the rate of volumetric change that occurs with time due to the self
desiccation mechanism. This approach is detailed in Section 3.4.

To represent the influence of cement hydration on permeability, the relationship


between void ratio and permeability suggested by Carrier et al (1983) is combined with
the volumetric changes during cement hydration as recommended by Powers and
Brownyard (1947) and the maturity relationship of Rastrup (1956). Details of this
relationship are provided in Section 3.3.

2.6 SUMMARY

This chapter has presented an overview of previous work in the field of mine backfill. In
addition, some simple examples were presented to demonstrate the significance of some
of the assumptions inherent in existing solutions. Specifically, these examples
demonstrate that gaining an understanding of the degree of consolidation that occurs
during placement is essential in attempting to determine the stress distribution around a
stope.

This chapter continues with a description of literature relevant to understanding the


consolidation that takes place during the deposition of cemented mine backfill. The
fields covered included consolidation, structured soil and cement hydration. Following a

2.20
Mechanics of Mine Backfill Matthew Helinski
Background & Literature Review The University of Western Australia

description of background literature, a brief description is provided in each of these


areas to explain how this previous work is applied to the cemented mine backfill
problem throughout this thesis.

The next chapter will expand on this brief description and develop these ideas to form a
unified framework to represent the individual mechanisms relevant to the cemented
mine backfill deposition process.

2.21
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

CHAPTER 3
BEHAVIOUR OF CEMENTING SLURRIES

3.1 INTRODUCTION

During the placement of mine backfill, the material initially behaves in accordance with
the uncemented material characteristics, but as cement hydrates the behaviour of the
material changes. Due to the increase in stiffness, reduction in permeability, and the self
desiccation mechanism, the change in material behaviour can have a significant
influence on the consolidation response. This chapter presents a description of the
behaviour of a tailings material undergoing simultaneous consolidation and
cementation, and develops equations to characterise this behaviour. These equations
will form the basis of the numerical models presented later in the thesis.

3.2 STRENGTH AND STIFFNESS


Terms A, D and E in the governing consolidation equation (Equation 2.3) are all
influenced by the stiffness of the material. Therefore, it is important that an appropriate
model be developed to represent the evolution of the material stiffness throughout the
filling process. This model must take account of the response of the material prior to the
formation of any cementation as well as the evolution of the material properties with
time. Due to the interaction of filling and cement hydration, this model must be capable
of representing the formation of cement bonds as well as the possible breakdown of
these bonds as a result of stresses that exceed the current bond strength.

3.2.1 Uncemented material response

As the eventual goal is to adopt the Structured Cam-Clay model to represent the
behaviour of the material in a cemented state, a convenient method of representing the
compression behaviour of the uncemented soil is through the Cam-Clay model. This
relationship is represented by Equations 3.1 and 3.2.

v (i )
e = . ln
v (i 1)
(3.1)

3.1
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

v (i )
e = . ln
v ( i 1)
(3.2)

where e is the change in void ratio in the current time increment, and are the
conventional Cam-Clay parameters describing the gradient of the compression curve,
and v(i-1) and v(i) are the vertical effective stresses at the start and end of the
increment respectively. The normally consolidated relationship (Equation 3.1) applies
to loading of the material on or above the uncemented compression line while the elastic
compression relationship (Equation 3.2) applies for material undergoing compression at
stress levels below the uncemented compression line (i.e. material that has been
overconsolidated).

3.2.2 Stress-strain behaviour of cemented fill

To incorporate the effect of cementation on strength, a convenient starting point is the


Structured Cam-Clay model developed by Liu and co-workers at Sydney University
(Liu et al., 1998; Liu and Carter, 2002; and Carter and Liu, 2005). In this approach,
structure (which could include cementation) has the effect of increasing the isotropic
(or one-dimensional) compression yield stress in a manner analogous to
overconsolidation. In the case of cemented mine backfill, the yield stress in
compression is a function of void ratio and cementation, and consequently it changes
with time as both of these change. A considerable extension to the Structured Cam-
Clay model is required in this case, to deal with the combined effects of growth of
structure (e.g. cement gel) as hydration occurs and damage to this structure due to
possible yielding with increasing effective stress.

The principle that has been adopted to determine the cement contribution to
compression resistance follows the concept suggested by Rotta et al. (2003). Rotta et al.
(2003) propose the concept of the incremental isotropic yield stress (py) and define
this as being the difference between the primary yield stress and the isotropic curing
stress. The concept of incremental yield stress has been illustrated in Figure 3.1, which
presents a plot of mean effective stress (p) versus void ratio (e) for soil in uncemented
and cemented states during isotropic compression. As mine backfill material is placed
and consolidated along the normally consolidated line, it is assumed that the curing

3.2
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

stress can be appropriately represented by the yield stress of the material in an


uncemented state. Therefore, the incremental yield stress is said to represent the
difference between the yield stresses for the uncemented and cemented material.

Rotta et al. (2003) suggested that py is a measure of the bond strength, and Consoli et
al. (2006) continued this concept to show that py is proportional to the unconfined
compressive strength (qu). Given these findings, it is considered reasonable to assume
that this proportionality would continue along the one-dimensional compression stress
path. Therefore, the incremental one-dimensional vertical yield stress (vy, the
preconsolidation stress) for a cemented soil has been assumed to be adequately
represented by the superposition of an uncemented and a cemented component (the py
contribution). Furthermore, based on the findings of Consoli et al. (2006) it has been
assumed that vy can be linearly related to qu.

During the mine backfill deposition process, previously placed fill is subjected to
mechanical processes that cause changes to the one-dimensional yield stress (vy).
These include:

conventional soil hardening due to void ratio reduction (dH);

increases in yield stress with time due to cementation (dHyd);

a potential reduction in strength due to plastic deformation effectively,


yielding of the cement bonds as compression occurs simultaneously with
hydration (dD).

These characteristics are accounted for by cumulating the changes in each of these
individual characteristics over a particular timestep (dt):

vy dH + dHyd dD
=
dt dt (3.3)

3.2.3 Hardening

Hardening in the model results from void ratio reduction as for conventional
uncemented soil and also from the hydration process.

3.3
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

The conventional soil strain hardening term (dH) applies to an increase in the
uncemented yield strength as a result of the compression of the soil matrix. This
hardening is a function of the uncemented soil properties and :

e p
dH = exp

(3.4)

where and are as defined above (Equations 3.1 and 3.2), corresponds to the void
ratio on the normal consolidation line at a vertical effective stress of 1 kPa, and ep is
the plastic change in void ratio.

The additional, hardening (strength increase) that occurs due to hydration (dHyd) is
assumed to be a function of cement content, tailings density at the time of the hydration
increment and the time from the start of hydration. It has been well documented
(Leroueil and Vaughan, 1990; Consoli et al., 2000; Li and Aubertin, 2003; and Rotta et
al., 2003) that an increase in either cement content or density increases the strength of a
cemented soil. Rotta et al. (2003) developed an empirical equation to relate the
incremental isotropic yield strength (py) to both cement content and void ratio:

X.C c + Y e
p' y = exp
Z.C c + W (3.5)

where Cc is cement content (weight of cement per unit weight of solids), e is void ratio
and X, Y, Z and W are dimensionless constants and py is in kPa. As it stands, this
equation implies that there is some cement component of strength even when the
cement content is zero. In order to ensure that there is zero cement component of
strength at zero cement content, the Y constant has been replaced by a cement power
term and the function adopted is shown in Equation 3.6. In addition, a constant
multiplier A (with units of kPa) has been introduced.

X.C c + C c 0.1 e
p' y = A. exp

Z.C c + W (3.6)

Assuming that vy, qu and py are all proportional (as discussed earlier), data on any
of the stress paths may be used to derive the constant terms (X, Y, Z and W), and by
adjusting the constant A (according to the ratio of proportionality between the various

3.4
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

cemented strength components) the cement component of strength along other stress
paths may be determined. An example of this is given in Figures 3.2(a) and 3.2(b),
which show how Equation 3.6 can be fitted to a series of unconfined compression test
results on CSA hydraulic fill and Cannington mine paste fill data (Rankin, 2004),
respectively. This indicates that Equation 3.6 can appropriately represent the
development of cementation in a range of typical cemented mine backfill materials. For
one-dimensional compression analysis the most direct method of determining all
constants (A, X, W and Z) is through regression analysis on a series of one-dimensional
compression tests.

The maturity relationship adopted to represent the progress of hydration with time is
an exponential relationship originally presented by Rastrup (1956) and republished by
Illston (1979). This relationship is presented as:

d
m = exp
(3.7)
t
*

where m is the degree of maturity (0 at the start of the process, 1 at the end), d is a
maturity constant (day-1/2), t* is the time (in days) since initial set. While it is
acknowledged that Equation 3.7 is not dimensionally independent, to maintain
consistency with previous work, the published form was preserved with t* always
specified in units of days and d in terms of day1/2.

A series of unconfined compression tests was carried out at different stages of hydration
to assess the development of this bond strength with time. Figure 3.3 shows qu
(normalised by dividing by the maximum qu) against the time in hours for both
Cannington paste fill (PF) (Rankin, 2004) and CSA hydraulic fill (HF).

Based on regression of the two data sets in Figure 3.3, maturity constants (d) of 0.9
day1/2 and 2.6 day1/2 provided a very close match to the hydraulic fill and the finer paste
fill, respectively, with the duration until initial set (to) being 4 hours (0.16 days) in both
cases.

The maturity relationship (Equation 3.7) may be combined with the strength increment
relationship (Equation 3.6) to determine the increment of bond strength (dHyd) over a
given time interval (t):

3.5
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

d X.C c + C c 0.1 e
dHyd = exp exp d .A. exp
* (3.8)
t + t t Z.C + W
*
c

It should be noted that the change in bond strength is incremented in accordance with
the material state at t*, the time at the start of that increment. This ensures that any strain
hardening of the soil matrix is accounted for when evaluating hardening due to
hydration over the next time increment.

3.2.4 Damage due to yielding during hydration (dD)

As hydration may be occurring simultaneously with an increase in effectives stress (due


to consolidation) the newly-forming structure may experience damage due to the
evolving yield stress being exceeded. This damage may occur as a result of loading
within the yield surface as well as loading on the yield surface (virgin yielding).

The damage relationship adopted here follows the approach used in the Structured Cam-
Clay model (Liu et al., 1998; Liu and Carter, 2002; and Carter and Liu, 2005). In this
approach, incremental plastic volumetric strain induces damage, which is manifest as a
reduction in the size of the yield surface.

The function adopted to account for plastic volumetric strain in the model follows the
work of Carter and Liu (2005) where the plastic component of strain may be represented
by Equation 3.9.

dvp

= 1 *
( )
dpc + b dpc
* * 3

(1 + e) ps (3.9)

where the asterix (*) indicates uncemented properties, is the stress ratio (q/p), M is
the stress ratio at critical state, pc is the applied effective stress, b is a constant
representing the structural breakdown, and is a measure of the kinematic hardening
given by:

p pu
= c
p
s p
u (3.10)

where pu is the stress at which kinematic hardening or destruction occurs, and ps is the
isotropic stress on the yield surface.

3.6
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

Given the plastic strain increment, the associated reduction in compressive yield
strength may be determined as:

(1 + e )b ps dvp ps
dD = ln c
(* * )1 + bln ps po
c
po
(3.11)

where c is the constant separating the limiting compression lines of the cemented and
uncemented soils (for the very low cement contents commonly used in mine backfill
this term is equal to zero), and po is the stress required to place the uncemented soil in
the same state on the normal consolidation line.

While changes in Poissons ratio (due to cementation) can modify the stress path in one-
dimensional compression, CeMinTaCo has been simplified by replacing all mean stress
terms in Equations 3.9 3.11 by vertical effective stress (v), and setting the stress ratio
term () to zero.

3.2.5 Unconfined compression strength (qu)

As explained earlier, it is assumed that the incremental one-dimensional yield stress is


linearly related to qu. In this analysis, the total one-dimensional yield stress and the
equivalent uncemented one-dimensional yield stress are monitored. By subtracting the
latter from the former, and applying the constant of proportionality, qu is determined.
This assumption has little impact on the overall consolidation behaviour. However, it is
common to use qu when referring to material strengths in the mining industry therefore
it is useful to gain an understanding of the impact of the various mechanisms on qu
throughout the filling process.

3.2.6 Stiffness

The previous section addressed the change in yield stress due to strain hardening,
cement hydration and damage. As with strength, these characteristics also influence the
material stiffness and in order to undertake consolidation analysis it is therefore
essential to characterize any material stiffness changes that occur. In an approach
similar to that used for strength, it is assumed that the stiffness of the cemented soil is a
combination of the stiffness due to the uncemented soil skeleton and that due to the

3.7
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

cementation. The uncemented stiffness is determined in accordance with Equations 3.1


and 3.2, while the cemented component of stiffness is assumed to be proportional to the
cemented component of strength (vy, qu or py).

A series of experiments was carried out to assess the validity of the assumption that the
cement induced stiffness and cement induced strength could be linearly related. These
experiments involved the measurement of small strain shear stiffness (Go) using bender
elements prior to unconfined compression testing of these specimens. Figure 3.4(a)
shows the results of incremental shear stiffness (Go(inc)) relative to qu for a variety of
combinations of CSA hydraulic fill mixes. Figure 3.4(b) illustrates the relationship
between qu and Youngs Modulus for Cannington Paste fill as published by Rankin
(2004).

Figures 3.4(a) and 3.4(b) indicate that for a range of material strengths and material
types there appears to be a linear correlation between the cement component of stiffness
and cemented strength (qu). Therefore, if 'vy or qu are known, it is reasonable to
assume that a constant of proportionality may be applied to calculate the cement
component of the stiffness.

These Youngs modulus or shear modulus values can be converted to constrained


modulus values using Poissons ratio and these may be combined with the uncemented
constrained modulus values to give total values of constrained modulus for every stage
of hydration.

Assuming that the strength and stiffness are proportional, it may be useful to utilise non-
destructive stiffness measurement techniques such as bender elements (Baig et al.,
1977) to assess the rate of cementation development with time. Experiments on CSA
hydraulic fill indicate that the maturity factor (d in Equation 3.7) to represent the
development of stiffness with time was 1.0 which is very close to that determined for
the development of strength with time (0.9).

3.2.7 Stress-strain behaviour: summary

This section has presented the details of how the various constitutive aspects have been
related in order to characterize the mechanical response of the cemented mine tailings
during filling. Combining these aspects, the stiffness term in Equation 2.3 can be

3.8
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

determined as a function of material density (i.e. void ratio e), cement content (Cc),
hydration duration (t) and effective stress (v) as illustrated in Equation 3.12.

v'
= f (e, Cc , t , v' )
e (3.12)

In order to demonstrate its applicability, the proposed constitutive relationship was used
to simulate a series of one-dimensional compression experiments on cemented CSA
hydraulic fill. In these experiments the specimens were prepared at different densities
and allowed to cure for different periods of time prior to loading them in one-
dimensional compression. The material constants d, X, W and Z (from Equation 3.8)
were determined from qu and bender element experiments. The parameters *and *
(from Equation 3.9) were determined from one-dimensional compression tests on
uncemented material and through modifying the terms A (from Equation 3.6) and b
(from Equation 3.9 and 3.11) the one-dimensional compression response could be
adequately represented. A comparison between the experimental results and the
proposed model is presented in Figure 3.5. Note that different initial void ratios were
used in the three tests, which explains why the 16-day result plots above the 5-day
result, initially. Once the cementation bonds are broken, all three results tend to
converge to the same compression line.

Figure 3.5 indicates that, given suitable experimental results, the proposed constitutive
relationship provides a good representation of the material behaviour when subject to
curing, compression and cementation breakdown.

3.3 PERMEABILITY

Term C in Equation 2.3 is highly dependent on the material permeability. This is the
term that controls the rate water is expelled from the system. As a result, permeability
can have a significant influence on the overall consolidation behaviour.

3.3.1 Uncemented permeability

During the compression of a soil matrix, the void volume reduces, which can lead to a
reduction in permeability. Carrier et al. (1983) developed a relationship that has been
shown to provide a good representation of the relationship between void ratio and

3.9
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

permeability for mineral waste materials. This function was adopted for this study and is
presented as equation 3.13

d
c k (e ) k
k=
(1 + e) (3.13)

where k is the permeability, e is the void ratio and ck and d k are constants.

3.3.2 Cemented permeability

The hydration of cement is associated with a growth of cement products. These products
are in the form of solid cement hydrates as well as cement gel. This product growth fills
some of the void volume, which further reduces the permeability. Due to the relatively
low permeability of the cement gel itself, it is suggested that in addition to the growth of
cement solids the entire gel volume should be taken into account in determining the
reduction in permeability. Powers and Brownyard (1947) suggest that when combined
with water, the solids and gel volume created (after full hydration) are 80% greater than
the initial unhydrated cement volume.

In order to account for this characteristic, Equation 3.13 is modified to be in terms of


effective void ratio (eeff). The calculation of effective void ratio is based on the void
space determined in the conventional manner as well as that calculated in accordance
with the hydrating cement products. This concept is illustrated in Equations 3.14 and
3.15.

eeff = f (e, C c , t )
(3.14)

c k (eeff ) dk

k=
(1 + e) (3.15)

The rate at which the growth of cement products takes place is simulated using the
maturity relationship (Equation 3.7).

A series of permeability experiments was conducted to assess the applicability of this


relationship to cementing tailings. These tests were conducted in a permeability cell
where the cell pressure was maintained at 520 kPa. The hydraulic gradient was
established by setting the back pressure at the top of the sample to 510 kPa and that at

3.10
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

the base to 490 kPa. Testing at these elevated back pressures ensured full saturation
throughout the test. As curing progressed, permeability measurements were taken at
regular intervals. At the completion of the test, the sample dimensions were measured
and, based on the initial dry weight, the void ratio could be calculated. The results of
tests on cemented hydraulic fill with cement contents of 2%, 5% and 10% are presented
in Figure 3.6, which shows the calculated effective void ratio against measured
permeability for the different experiments, along with the model estimate.

Figure 3.6 indicates that a reasonable fit to the measured data may be achieved using the
proposed method. However, it is suggested that the constant terms (in Equation 3.15)
representing compression and those representing cement growth may vary. Further
work may be required to develop the understanding of the contribution of these two
mechanisms to the reduction in permeability.

3.4 SELF DESICCATION


The process known as self desiccation has been well documented with respect to its
impact on concrete behaviour. The basis of this process is that following cement
hydration, the resulting hydrated volume is less than the combined volume of the
unhydrated constituents (cement and water). Researchers such as Powers and
Brownyard (1947), Hua et al. (1995), Koenders and Van Breugel (1997), Bentz (1995)
and Brouwers (2004) have investigated the impact of this mechanism on the shrinkage
of conventional concrete. Most conventional concrete masses have lean (low) water
contents and are placed in thin horizontal layers resulting in low total vertical stress.
Therefore, the volume reduction of the cement/water constituents can result in
development of negative pore pressure and desaturation of the mixture hence the term
self desiccation.

However, cemented mine backfills have much higher water contents (and lower cement-
water ratios) than conventional concrete and can be subjected to high self-weight total
stresses due to rapid rates of rise in typical stope-filling operations. In fine-grained
(paste) fills, this can result in high positive pore pressures. As a result, the processes
involved in self desiccation act, in these circumstances, to reduce the build-up of
positive pore pressure rather than desaturating the material and creating negative pore
pressures. Consequently there is no self desiccation per se. Thus, when reference is

3.11
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

made to self desiccation in this thesis, this refers to reduction in pore pressure resulting
from cement hydration, rather than desaturation (desiccation) of the material. The term
self desiccation has been used to preserve consistency with the mechanism of
hydration-induced water volume reduction, rather than implying any actual drying out
(desaturation). In saying this, given the appropriate conditions, this mechanism does
have the potential to generate negative pore pressures and potentially desaturate the
mass.

The aims of the work described in this section are to show that the self-desiccation
process can have a significant effect on the behaviour of the backfill (during the
hydration process), to derive a model for describing the process and to devise a
laboratory testing procedure to enable the model parameters to be determined for any
fill/cement combination.

3.4.1 Cementation reactions

The reactions associated with cement hydration involve the chemical combination of
cement and water. If it is assumed that an enclosed volume of the soil-cement-water
slurry prior to cement hydration contains a water volume of Vw, and an unhydrated
cement volume of Vcu. After hydration, the hydrated cement volume is Vch, such that Vch
Vcu = Vhyd. In this reaction, the increase Vhyd is less than the volume of water used
in hydration, Vwh. In keeping with the terminology used in the concrete literature, this
loss of volume is denoted Vsh (the chemical shrinkage volume).

For the purposes of the calculations that follow, the total water volume used in
hydration (Vwh) can conveniently be thought of as being composed of two parts an
amount converted directly into solid volume equal to Vhyd, and a volume equal to Vsh
that is lost from the system as if removed via an internal water sink.

Vwh = Vhyd + Vsh (3.16)

Thus, using this approach, Vsh represents an apparent water volume lost from the
system due to the hydration reaction, whereas Vhyd represents water volume that is
substituted by solid volume, and hence has no overall effect on total volume or water
pressure.

3.12
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

All of this assumes that the soil voids can compress to accommodate the lost volume
(Vsh) in a completely unrestrained way, and Vsh, when integrated over the total
volume, would give the total shrinkage. This would be the case in a slurry where the
soil matrix has zero stiffness, and the voids can compress without any change in
effective stress or any change in the pressure in the remaining water. Thus, it is only in
this case that Vsh in Equation 3.16 represents the apparent unrestrained water volume
lost via the internal sink, and, when integrated, gives the overall slurry shrinkage.
Conversely, in the hypothetical case of a soil matrix with infinite stiffness (assuming a
fully-saturated state and no inflow of water allowed), no overall volume change can
occur, and the chemical shrinkage can only be accommodated by a volume expansion of
the remaining water equal to Vsh, leading to a drop in pore pressure equal to the bulk
modulus of water multiplied by Vsh/(Vw Vwh). For the general case of a soil matrix of
finite stiffness, some of the volume loss is accommodated by soil matrix compression
(and hence some increase in effective stress), and some by expansion of the remaining
water (and hence some reduction in pore pressure). This will be discussed further in the
next section.

It should be remembered that mine backfill slurries are fully saturated, typically with
initial water content (mass of water per unit dry mass of soil) of 100% or greater, and
cement content (mass of cement per unit dry mass of soil) typically 2 5%. Therefore,
the water-cement ratio is much greater than for conventional concrete (for 100% water
content, and 2% and 5% cement content, the water-cement ratio would be 50 and 20,
respectively, in mass terms, corresponding to about 160 and 62, respectively, in volume
terms). Thus, the actual volume of water involved in hydration is relatively small, and
hence volumetric strains in the water can be calculated relative to the original total
volume of water (Vw), rather than the final volume (VwVwh). In fact, in the numerical
implementation of the equations, all calculations are performed in incremental fashion,
so that volumes are continually updated, and strains are therefore calculated using the
appropriate water volume.

It should also be noted that, following the convention used in soil mechanics,
compressive stresses and strains are considered positive, while tensile stresses and
strains are considered negative. Thus, volume reduction is considered positive (and
hence chemical shrinkage Vsh is positive).

3.13
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

Powers and Brownyard (1947) found experimentally that, for a fully hydrated system,
Vsh for a cement paste could be related to the mass of chemically-combined water (Wn)
through Equation 3.17:

Vsh = 0.279 Wn (3.17)

where Vsh is the volume reduction in cm3 and Wn is in gram (and thus the constant
has units of cm3/g). From a series of laboratory experiments, they found that Wn could
be related to the proportion of the compounds that make up the cement product and the
mass of unhydrated cement (Wc) in accordance with Equation 3.18.

Wn/Wc = 0.187XC3S + 0.158XC2S + 0.665XC3A + 0.213XC4AF (3.18)

where X is the proportion by mass of the subscript compound in the cement. For the
proportions contained in most General Portland cements, they established empirically
that the Wn can be approximately related to Wc via Equation 3.19:

Wn/Wc = 0.23 (3.19)

Combining Equations 3.17 and 3.19 allows the shrinkage volume (in cm3) that would
occur in a General Portland cement paste over the full hydration period to be
determined. This relationship is shown in Equation 3.20 as a function of the original
mass (in g) of cement:

Vsh = 0.064 Wc (3.20)

3.4.2 Impact on pore pressure

As mentioned above, the apparent unrestrained volume change in the water resulting
from hydration (Vsh) could occur under undrained conditions only if the soil skeleton
were of zero stiffness, and this volume change would result in an equal compression of
the soil skeleton. To calculate the actual volume changes and pore pressure changes, it
is necessary to consider the water and soil matrix stiffnesses, and use principles of strain
compatibility and stress equilibrium to calculate the actual behaviour.

The change in pore pressure is a function of the difference between the shrinkage
volume (Vsh), as defined in Equation 3.16, and the actual reduction in void volume

3.14
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

(Vv) resulting from soil matrix compression. The difference between these two
volumes is denoted Vrel:

Vrel = Vv Vsh (3.21)

Effectively, strain compatibility requires that the water must expand by Vrel to
accommodate the fact that the actual void volume reduction Vv is less than Vsh (and
hence Vrel as defined by Equation 3.21 is negative, signifying expansion).

Initially, when the soil matrix has a very low stiffness, the void volume reduction (Vv)
is close to Vsh, and hence there is very little pore pressure change. However, in the
case of soil containing cement, an increase in stiffness comes about not just as a result
of ongoing compression, as with uncemented soils, but also due to the formation of
cement bonds so that as hydration proceeds the pore pressure reduction can be quite
substantial.

A major difference between self desiccation and evaporative desiccation is that, with
evaporation, the water sink is at the (top) boundary, which sets up internal hydraulic
gradients to feed water to the evaporation process. However, in the hydration process,
internal sinks are set up within every pore throughout the material. Therefore the
mechanism is purely intrinsic and the incremental reduction in pore pressure is
dependent on the hydration time and not on any length scale in the problem.

3.4.3 Analytical model

An analysis relating fundamental material properties to the reduction in pore pressure is


discussed below. The analysis assumes that:

the material is in an undrained state (with respect to water flows across the
external boundary);

the soil compressibility is linear, corresponding to the current small-strain bulk


modulus, at any stage of hydration (though changing with ongoing hydration);

soil particles are incompressible;

the water bulk modulus (Kw) is constant;

the material is fully saturated at all stages of the process; and

3.15
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

the water density is independent of pressure.

In the experimental work described later, full saturation at all stages has been assured by
using high initial back pressure, such that positive pore pressure exists at all stages, even
following the pore pressure reduction resulting from the hydration process.

Equilibrium of the pore water system requires:

Vrel
u = v . K w = Kw
Vw (3.22)

where u is the change in pore pressure in the current increment, v is the increment of
volumetric strain in the water required to maintain strain compatibility with the soil
skeleton, Kw is the bulk modulus of the water, Vrel is the relative pore volume
reduction (as defined by Equation 3.21), and Vw is the total volume of the pore water.
Since Vrel is negative (expansion), both v (expansion) and u (pore pressure
reduction) are also negative.

The change in soil bulk volume is proportional to the change in effective stress. With a
constant total stress, the change in effective stress () is equal in magnitude and
opposite in sign to the change in pore pressure:

= u (3.23)

The incremental volumetric strain in the soil matrix (v-soil) is a function of the change
in effective stress () as well as the bulk modulus of the soil matrix (Ks):

Vv
v soil = =
VT Ks (3.24)

where VT is the total volume of the combined soil and water (bulk volume) and Vv is
the actual change in the void volume due to compression (which is identical to the
change in the bulk volume VT, since the soil particles are taken to be incompressible).

Combining the behaviour of the pore water (Equation 3.22) with the behaviour of the
soil matrix (Equation 3.24) via Equation 3.23 gives:

Vv V
.K s = rel .K w
VT Vw (3.25)

3.16
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

Equation 3.21 may then be substituted into Equation 3.25 to derive a relationship
between the actual incremental contraction of the pore volume (Vv) and the chemical
shrinkage volume (Vsh):

Vsh .K w
Vw
Vv = (3.26)
K s Kw
+
VT Vw

which indicates, as expected, that Vv = Vsh when Ks = 0. By combining Equations


3.26, 3.24 and 3.23 and rearranging, a relationship can be obtained between the
incremental change in pore pressure (u), the bulk stiffnesses of the soil (Ks) and water
(Kw), the porosity of the material (a function of Vw and VT) and the incremental change
in volume associated with the hydration reaction (Vw):

Vsh K wK s V Kw
u = . = sh .
Vw .VT K s K w VT (n + K w K s ) (3.27)
+
VT Vw

This gives u 0 as Ks 0, and u Kw{Vsh/(n.VT)} as Ks , as expected.

3.4.4 Experimental demonstration of effect of self desiccation

A series of preliminary tests was carried out to demonstrate the validity and relevance of
the self-desiccation concept as applied to cemented tailings backfill. These experiments
used a silty silica sand hydraulic fill (HF) from the CSA mine and a silt-sized paste fill
(PF) from Kanowna Belle (KB) mine. These materials were mixed with 5% General
Portland cement from the Kandos Cement Plant (Kandos, NSW, Australia) and
Cockburn Cement (Perth, WA, Australia), respectively. In this thesis, the cement
content is defined as the mass of dry cement divided by the total dry mass of solids.
Particle size distribution curves for the two tailings materials are presented in Figure
3.7. The specific gravity (Gs) of CSA material is 2.81 while that for the KB material is
2.72.

The tests were carried out on samples set up in a triaxial cell in the conventional
manner. The specimens were prepared using the dry sand preparation technique
explained by Ismail et al. (2000). This technique involves preparing the specimens dry

3.17
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

before purging with CO2 and back-pressure saturating the samples. While this sample
preparation technique does not represent the mine filling process, the technique was
adopted to ensure consistency between samples. Both samples were prepared at a void
ratio of 0.80. After saturation in the triaxial cell, the cell pressure was increased to 550
kPa and the back pressure to 500 kPa. The purpose of the high back pressure was to
ensure positive pore pressure and full saturation throughout the test, given the large pore
pressure reductions expected from the self-desiccation process. For both tests, the
Skempton B-value was checked and found to be greater than 0.95. At this point, the
back pressure valve was closed and the pore pressure was monitored with time. The
results of the tests are presented in Figure 3.8, which shows the applied total stress, the
pore pressure and the effective stress, plotted against time from the start of the test.

As may be seen in Figure 3.8, even with cement content as low as 5%, the hydration
process creates a significant reduction in pore pressure, with both material types. With a
constant total stress, this pore pressure reduction is associated with an effective stress
increase of equal magnitude. The figure shows that the rate and final amount of pore
pressure reduction for the CSA HF specimen are significantly greater than for the finer
KB PF specimen. Consequently the impact of the self-desiccation process depends on
material type, as well as other factors discussed below.

These tests provide a graphic illustration of the changes in pore pressure that result from
the cement hydration process and the resulting change in effective stress, even where
the samples are subjected to undrained boundary conditions. Therefore, it is apparent
that this self-desiccation phenomenon needs to be considered when analysing the
behaviour of cemented mine backfill, particularly where filling rates are rapid and fine-
grained (i.e. low permeability) tailings are used, resulting in the hydration process
occurring under undrained conditions.

3.4.5 Material properties influencing self desiccation

Equation 3.27 indicates that the reduction of excess pore pressure is sensitive to the
water and soil bulk moduli (Ks and Kw) as well as to the rate of water consumption and
the total volume of water consumed during the hydration process.

3.18
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

Material stiffness
Due to the growth and strengthening of hydrates, the soil matrix undergoes an increase
in stiffness during hydration. A non-destructive test that is often used in soil mechanics
to monitor the small strain shear stiffness (Gmax, also called Go) of a soil matrix
involves the measurement of shear wave velocity (Dyvik and Olsen, 1989, Baig et al.
1997, Fernandez and Santamarina, 2001). This technique consists of generating a shear
wave pulse at one end of a sample using a piezoceramic bender element, and
measuring the arrival time at the opposite end of the sample using a second bender
element.

Figure 3.9 shows an example of the data from one of the tests carried out in this study.
In this case, the transmitting bender element is excited by a single sine wave pulse,
nominally of 10 V amplitude, and the arrival of this shear wave at the other end of the
sample is picked up by the receiver bender element. Based on the time of transmission
and the length of the sample, the shear wave velocity (Vs) can be obtained. From this
value of Vs and the bulk density of the material (), the value of Gmax may be inferred
using Equation 3.28.

Gmax = .Vs2 (3.28)

This test can be carried out at intervals during the hydration process to monitor the
development of the shear modulus with time. The corresponding small strain effective
bulk modulus Kmax can be related to Gmax via the Poissons ratio ():

2( 1 + )
K max = Gmax
3( 1 2 ) (3.29)

In this paper, this value of Kmax is assumed to be equivalent to the soil matrix stiffness
Ks mentioned earlier (e.g. Equation 3.24), which is equivalent to assuming that soil
matrix stiffness is linear over the range of strain relevant to this work. Santamarina et
al. (2001) and Jamiolkowski et al. (1994) suggest that a small strain drained Poissons
ratio of 0.1 to 0.15 is appropriate for many soils, and thus, for the interpretation of the
results in this paper, a small strain Poissons ratio of 0.125 has been adopted. It should
be noted that varying the Poissons ratio over this range has minimal impact on the
results.

3.19
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

At the University of Western Australia, bender elements are fitted as standard in triaxial
setups, allowing shear wave velocity (and hence Gmax) to be determined routinely.
Measuring the compression wave velocity Vp (and the equivalent Emax) could also be
used to indicate the progress of hydration, but doing so in the triaxial apparatus was not
possible with the equipment available.

Water consumption during hydration


The process that determines the rate at which pore water volume is consumed is very
complex and is made particularly difficult to quantify theoretically due to:

the hydration of cement involving at least 8 different chemical reactions;

each reaction consuming different volumes of water;

each reaction producing a different hydrate volume;

each reaction commencing at a different time after the start of hydration;

each reaction occurring at a different rate;

only cement surfaces exposed to pore water reacting;

the cement being made up of different proportions of each constituent;

the reactions being dictated by the random collision of various cement


constituents;

not all of the total cement content in the mix may react.

Cement technology researchers have developed detailed microscopic models to predict


this process for the purpose of concrete shrinkage predictions (Bentz, 1995). These are
complex models that involve the input of many fundamental cement properties, and
further discussion of them is beyond the scope of this work.

The other complicating factor associated specifically with mine backfill is that in
addition to different cement types, different tailings mineralogy and chemicals
contained in the tailings after processing may have an impact on the chemical reactions
that take place. Therefore, it is suggested that the most practical method of determining
the net volumetric change and the rate at which this change occurs is through direct
experiment with each cement/tailings combination. Furthermore, it is suggested that

3.20
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

rather than adopting the total volume change (6.4% of the unhydrated cement weight,
Equation 3.20) as determined by Powers and Brownyard (1947), the volume change
should be defined as a variable (Eh) for each particular cement/tailings combination.
Therefore, Equation 3.20 is re-written as Equation 3.30, where Eh is defined as the total
volumetric change Vsh per unit mass of cement Wc:

Vsh = Eh. Wc (3.30)

3.4.6 Experimental derivation of parameters

By measuring the incremental pore pressure reduction and monitoring the material
stiffness, the rate of water volume consumption may be back-calculated using the
proposed analytical solution (Equation 3.27). The rate of hydration (d in Equation 3.7)
and hydration efficiency (Eh) are considered fundamental material properties. Therefore,
once determined, these parameters may be incorporated into a coupled analysis model
to account for the impact of this mechanism on the consolidation and filling process.

Experimental procedure
A series of pore pressure reduction experiments was carried out to verify the proposed
theory, provide examples of how the experimental process may be conducted and
demonstrate how the relevant material parameters may be derived. The material used in
these experiments was again silty sand (hydraulic fill) from the CSA mine and silt
(paste fill) from Kanowna Belle. These materials were mixed with General Portland
cement from the Kandos Cement Plant (Kandos, NSW, Australia) and the Cockburn
Cement Company (Perth, WA, Australia), respectively, in various proportions.

The experiments were conducted in a triaxial cell, with the specimens being prepared
using the dry sand preparation technique explained by Ismail et al. (2000). During
saturation, the amount of water added to the system was measured (as this would be the
volume of water subject to the volumetric changes).

As was shown in a previous section, (Figure 3.8) the hydration process results in a
reduction in water volume, which leads to a reduction in pore pressure and a
corresponding increase in effective confining stress. This increase in effective stress
could lead to yielding of the hydrating matrix, which could invalidate the assumption
that the small strain stiffness Kmax is the relevant bulk stiffness Ks for the soil. Also,

3.21
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

depending on the initial back pressure, the reduction in pore pressure could lead to air
coming out of solution, thereby changing the bulk modulus of the pore fluid.

To avoid any of these potential problems associated with a decreasing pore pressure, the
back pressure and cell pressure were initially set at values well above those
recommended by Bishop and Henkel (1962) for complete saturation. To avoid the
possibility of yielding due to increasing effective stress, the effective stress was kept
low by regularly restoring the back pressure to its original value by opening the
drainage valve at various stages during the test. This restoration of back pressure means
that, strictly speaking, the experiment was not conducted under undrained conditions,
but since the material properties are only determined during the undrained stages (i.e.
while the back pressure valves are closed) the application of Equation 3.27 remains
valid. At different stages during the tests, shear wave velocity measurements were
made, using bender elements, to monitor the evolution of stiffness (Gmax) with time.

Figure 3.10 presents the results of one of the tests on the CSA HF (hydraulic fill)
material. This shows the actual pore pressure behaviour i.e. reduction in pore pressure
while the drainage valve was shut, followed by restoration of the initial back pressure in
the brief intervals when the valve was opened. From this, the cumulative pore pressure
change was determined to be of the order of 800 kPa for this test. The actual effective
stresses during the test are also shown, with the procedure adopted limiting the effective
stress to a maximum of less than 200 kPa.

An identical test on an uncemented specimen was carried out prior to those on the
cemented specimens to assess the compliance of the system. The system indicated a
pore pressure change of less than 5 kPa for the uncemented specimen over a 3 day
period, indicating that the system was free of any leaks.

Stiffness development
From the measurements of Gmax made as hydration proceeded, Equation 3.29 was used
with a Poissons ratio of 0.125 to determine the soil matrix small strain bulk modulus
Kmax, which was taken to be equivalent to Ks. Figure 3.11 shows how this calculated
value of Ks increased with time during the various experiments.

3.22
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

The curves shown fitted to the data in Figure 3.11 are based on the exponential
maturity relationship (for cement hydration) published in Illston et al. (1979). For this
application, this relationship takes the form:

d
K s = K s i + K s f . exp
(3.31)
t
*

where Ks-i is the initial bulk modulus and Ks-f is the increase in bulk modulus at the
completion of the hydration period, d is a maturity constant, and t* is the time since the
commencement of hydration. The curves in Figure 3.11 were obtained using d = 0.9
day1/2 with all cement contents. The time until initial set was found to be reasonably
consistent at about 4 hours for all these tests.

Pore pressure reduction


The test procedure used in these tests involved opening the drainage valve at regular
intervals during the test, thereby re-applying the initial back pressure. The data from
pore pressure reduction in the undrained phases that followed each of these re-
applications of back pressure can be combined to form a continuous pore pressure
reduction curve for each test, as shown in Figure 3.10. By dividing the pore pressure
reduction into one-hour increments, the incremental rate of pore pressure reduction was
determined for the duration of the test. Figure 3.12 shows this incremental reduction
rate plotted against time for the CSA hydraulic fill material with different cement
contents.

While there is some fluctuation in the pore pressure measurements, it can be seen that
the rate of reduction diminishes with time over the duration of the test (from the start of
initial set). Figure 3.12 also indicates that the rate of pore pressure reduction increases
with an increase in cement content. It should be noted that the pore pressure fluctuation
is temperature sensitive (fluctuating on a 24 hour cycle), suggesting that these tests
should have been conducted in a temperature-controlled environment.

Pore water volume decrease


In order to incorporate the self-desiccation mechanism into a finite element (or other
numerical) computer code, the incremental water volume change with time is required.
By substituting the instantaneous bulk modulus and the pore pressure reduction over a

3.23
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

given time period into the analytical solution (Equation 3.27) and rearranging, the pore
water volume change over that given time period may be back calculated.

Direct measurement of water volume consumption was also attempted in these


experiments. However, the volume measuring system used proved to have insufficient
resolution and accuracy, given that the volumes involved were of the order of tenths of a
cm3 per day. An improved system of volume-change measurement is a priority for
inclusion in future experiments.

After calculating the rate of water volume change throughout the experiment as
described above, the results can be divided by the relevant cement mass to determine a
rate of volume change per unit mass of cement. The results of this analysis are presented
in Figure 3.13 as the rate of water volume consumption per unit mass cement (Vw/Wc)
plotted against time for tests with three different cement contents.

The maturity model presented in Illston et al. (1979) was combined with Equation
3.30 (for total water consumption) to estimate the total water volume change after a
given hydration time (t). This relationship has been differentiated and divided by the
mass of unhydrated cement (Wc) to derive a function for the rate of volume change per
unit mass of cement. This function is presented as Equation 3.32:

d d
.exp
(Vw Wc ) (Vsh Wc ) 1
= = Eh . 1. 5 (3.32)
t t 2 (t *) t*

The same maturity constant (d = 0.9 day1/2) as that found for the rate of stiffness
development was substituted into Equation 3.32, and the efficiency term (Eh) was
adjusted to achieve the best fit to the experimental data, resulting in a best-fit value of
Eh = 0.035 cm3/g. The derived curve is compared with the experimental data in Figure
3.13. In this case, the fit was obtained by taking t as applying from the start of the test,
rather than the initial set; slightly different parameters would be obtained if the latter
had been used.

Cumulative pore pressure reduction


Combining the experimentally derived terms for hydration efficiency (Eh = 0.035
cm3/g) with the maturity constant (d), the rate of pore pressure change can be
determined. This rate may be integrated over a given time period to predict the

3.24
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

cumulative pore pressure drop. The experimental results are compared with the
analytical solution in Figure 3.14 for CSA material, with d = 0.9 day0.5 and Eh = 0.035
cm3/g.

Figure 3.14 indicates that the predicted pore pressure reduction due to cementation can
be estimated accurately using the proposed analytical solution with appropriate values
of d and Eh. The values of d and Eh have been shown to be unique for a given
cement/tailings combination over the range of typical cement contents. The value of Eh
determined for the CSA fill (0.035 cm3/g) is somewhat less than the value of 0.064
cm3/g suggested by Powers and Brownyard (1947) for cement paste.

Kanowna Belle paste fill experiments


Experiments were carried out using silt sized Paste backfill material from the Kanowna
Belle (KB) mine (with cement contents of 2% and 5%) to assess the applicability of the
proposed approach to a different type of minefill. The experimental technique used was
identical to that described for the CSA test work. From the results of these experiments,
values of d and Eh of 2.5 day1/2and 0.055 cm3/g, respectively, were determined. These
values were substituted into Equations 3.31 and 3.32 before combining them in
Equation 3.27 to predict the cumulative drop in pore pressure with time and this
prediction is compared with experimental results in Figure 3.15. It can be seen that the
analytical solution compares well with the experimental results in this figure. It should
be noted that, again, the maturity constant (d) representing the rate of hydration appears
similar for both the rate of pore water volume consumption as well as the development
of shear stiffness with time. For the KB Paste backfill the Eh term of 0.055 cm3/g
corresponds closely to the value of 0.064 cm3/g suggested by Powers and Brownyard
(1947) for cement paste, whereas a significantly lower value (0.035 cm3/g) appears
relevant for the CSA hydraulic fill.

3.5 TEMPERATURE

As cement hydration is an exothermic reaction, in a bulk filling situation hydration of


cemented mine backfill can lead to temperature increases. However, as cement contents
in minefill are often very low, temperatures typically range from 20 30C in a typical
cemented mine backfill situation. Temperature increases greater than 5C act to reduce
the water density, increasing the water volume. This was taken into consideration, as a

3.25
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

volume increase has the potential to negate volumetric reductions from self desiccation.
Assessment of the magnitude of volumetric change over this range indicates the
potential for a maximum 0.1% increase in water volume throughout a typical filling
period. In comparison with volumetric changes associated with the self desiccation
process, this change was shown to be a second order influence and was therefore not
addressed in this thesis.

Turcry et al. (2002) demonstrated that thermal volumetric changes can simply be
superimposed onto chemical volumetric changes to achieve a net volumetric change.
Therefore, to incorporate temperature variation into the analysis of the influences of any
volumetric changes, it could simply be incorporated as an independent mechanism in
the analysis.

Temperature variations also have the potential to influence the rate of hydration.
Therefore, rather than superimposing the influence of temperature changes at the
analysis stage, consideration was given to incorporating the influence of temperature in
the experimental process. This work is ongoing, but to investigate the appropriateness of
carrying out the hydration test with an insulated specimen a numerical analysis was
carried out to assess the appropriateness of a fully insulated assumption in a typical
mine backfill scenario. This study utilised the numerical code Temp/W1 to assess the
likelihood of heat transfer to the surrounding rockmass in a typical mine backfill
scenario. The analysis involved establishing an initial temperature of 30C throughout a
10 m wide, 40 m tall plane-strain stope with a boundary condition of 20C. The
material properties adopted included a thermal conductivity of 1 J/s/m/K and a heat
capacity of 3 MJ/m3/K, which are considered suitable for a saturated soil at a void ratio
of 1.0.

Figure 3.16 presents the calculated temperature profile laterally across the analysed
half-space after 20 days (a typical filling period). The results indicate that only the outer
1 m of the fill mass is significantly influenced by the heat exchange at the fill-rockmass
boundary, and the majority of the material remains in an insulated state. Based on this

1
Temp/W is a part of the GeoStudio suite of programs from Geo-Slope International Ltd, Calgary,
Alberta, Canada, www.geo-slope.com

3.26
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

result, it may be more appropriate to undertake hydration testwork (as discussed in


Section 3.6) in a fully insulated environment, which would address both water
volumetric changes (through a modified Eh in Equation 3.32) and the influence of
temperature on hydration rate (through a modified d term in Equation 3.7).

3.6 MATERIAL CHARACTERISATION TECHNIQUE

This chapter has addressed details of the mechanisms that are considered to be relevant
to the deposition and consolidation of cemented mine backfill. Being a complicated
interaction of different mechanisms, it is important to characterise the influence of
different materials on each of these mechanisms via fundamental material properties.
Keeping in mind the properties required, an experimental technique was devised to
capture most of the important material properties. This technique generally only
requires a hydration test, a triaxial test, and a Rowe cell test.

The hydration test is similar to that described in Section 3.4.6. Using the technique
described an understanding of the development of small-strain stiffness against time as
well as the self-desiccation characteristics can be obtained. In addition, a hydraulic
gradient can be established across the sample, at any time during hydration, to measure
the permeability of the material and assess how this changes with cement hydration.
Figure 3.17 presents an illustration of the experimental setup for a hydration test.

The hydration test can be combined with a conventional consolidated drained triaxial
test to determine the shear strength parameters such as cement induced bond strength
and frictional characteristics. If equipped with a local strain measurement system, it is
possible to use this experiment to determine non-linear elastic stiffness parameters and,
if the sample is strained to sufficient levels, the rate of cementation breakdown with
plastic strain can also be determined.

Finally one-dimensional compression (or Rowe cell) testing can be used to define the
stiffness and permeability characteristics during the compression of the material in
either a cemented or uncemented state.

3.27
Mechanics of Mine Backfill Matthew Helinski
Behaviour of Cementing Slurries The University of Western Australia

3.7 CONCLUSION

Background citations and experimental data have been presented to demonstrate the
mechanisms and material models that are used throughout this thesis. In addition,
experimental techniques have been presented for determining material properties that
are considered to most significantly influence the cemented mine backfill deposition
process. With this basis of understanding, the remainder of this thesis is focused on the
combination of these mechanisms and how they influence the overall filling behaviour.

3.28
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

CHAPTER 4
ONE-DIMENSIONAL CONSOLIDATION MODELLING

4.1 INTRODUCTION

In Chapter 2, Gibsons governing equations for one-dimensional consolidation were


introduced and the influence of cement hydration on these equations was discussed. In
Chapter 3, a description of the different processes that are expected to occur during
filling with cemented mine backfill was presented, and equations to describe these
processes were developed. This chapter is focused on incorporating these mechanisms
into Gibsons one-dimensional consolidation equations. These modified equations are
then solved using a modified version of the one-dimensional finite element tailings
consolidation program MinTaCo. This program has been renamed CeMinTaCo, and is
used in a sensitivity study to demonstrate the interaction of the different consolidation
mechanisms.

4.2 MODEL DEVELOPMENT

4.2.1 Modelling the behaviour of uncemented tailings: the MinTaCo Program

In previous work carried out some 10 years ago at UWA by others, a finite element
program was developed to model the consolidation and evaporation behaviour of mine
tailings, within the context of conventional tailings deposition in above-ground tailings
storage facilities (TSFs). This program, named MinTaCo (Mine Tailings
Consolidation) forms the basis of the new program. The new program has been named
CeMinTaCo, to indicate it deals with cemented mine tailings consolidation. A full
description of the MinTaCo model is provided by Seneviratne et al. (1996), and a
summary of some of its features that are relevant to cemented backfill is provided in the
following section:

MinTaCo is a one-dimensional model, which uses a large-strain formulation


with Lagrangian coordinates and Gibsons consolidation equations (Gibson
1967) to deal with the large volume changes. The form of the Gibson
consolidation equation was presented as Equation 2.2:

4.1
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

u k 1 + eo u k v
+ (1 + eo ) v + =
t
{ 144244 e a
3 144w44
1 + e a a {
t
(A ) 42444 443 (B )
(D ) (C ) (2.2bis)

where a is a Lagrangian coordinate, u is pore pressure, k is hydraulic


conductivity (permeability), 'v and v are the vertical effective and total
stresses, respectively. In this equation, term A is the rate of change in pore
pressure as a result of a rate of application of total stress (term B), term C is
the volumetric strain, which is dictated by the hydraulic conductivity (k) of the
material, and term D is the current stiffness of the material. As with tailings
placed into surface TSFs, paste fill placed underground may undergo large
settlements as it drains and consolidates. Incorporating a large strain
formulation into the model was regarded as important. MinTaCo provided this.

Fresh tailings slurry can be added at any desired rate, and this rate can be
changed at any stage during filling. Details of this aspect can be found in Toh
(1992) and Seneviratne et al (1996).

The properties of the fresh layers can be different from preceding layers.

The settled density of the material is taken as the starting point for
consolidation. Thus if very wet (low solids content) slurry were used, the
model is able to account for the generation of bleed water and update the
initial void ratio of the fill to account for this.

The base of the storage area can be perfectly permeable, perfectly


impermeable, or any state in between.

The input data required to run the program are: the material parameters; the filling
schedule and the drainage conditions. The material parameters are:

The specific gravity Gs of the tailings material.

The initial settled density of the tailings.

The ke (permeability void ratio) and e'v (void ratio effective stress)
relationships, which are generally very non-linear, are expressed using power
functions suggested by Carrier (1983):

4.2
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

b
e = a c ( v ) c
d
c k (e ) k
k=
1+ e (4.1)

where ac, bc represent soil compression constants and ck, dk represent


permeability constants. The large volume changes that occur during paste fill
consolidation referred to earlier mean that large void ratio changes occur and it
is essential to account for the effects of these changes on parameters such as
permeability.

The air entry suction value the point where desaturation starts to occur
during drying. This allows an important feature of a consolidating fill mass,
development of a partially saturated matrix, to be accounted for.

Estimates of shear strength and its variation with time may be made using the
Cam Clay model, so that values of the Cam Clay parameters (, , M, ) are
required for the material.

Drainage occurs in the vertical direction only (upwards or downwards,


depending on hydraulic gradient), and strains are vertical only.

The MinTaCo program has been used extensively for modelling a wide variety of
tailings filling operations. Some examples of its application may be found in Fahey and
Newson (1997) and Fahey et al. (2002).

4.2.2 Modelling the behaviour of cemented tailings: the CeMinTaCo Program

Section 2.2 showed that regardless of the ultimate cured strength, the loads applied to
(backfill) barricades during the filling process are highly dependent on the degree of
consolidation. Furthermore, it was demonstrated that in an uncemented state, very little
consolidation is likely to occur in paste fill during a typical filling sequence. As a result,
barricade stresses are more appropriately calculated assuming undrained conditions.
However, as discussed in Chapter 3, the addition of cement to backfill complicates the
rate of pore pressure change and in most cases where paste fill is used, the behaviour is
likely to be neither fully drained or fully undrained, but somewhere in between. Use of a
model such as that described in this thesis makes it possible to determine where a
particular fill is located between these two extremes.

4.3
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

Because it provides full coupling between filling rates, boundary drainage, and stiffness
and permeability changes within the tailings, the MinTaCo program appeared to provide
the ideal basis for the development of a program to model the consolidation behaviour
of backfill. However, to be of general application to the backfilling problem, additional
features were required, relating particularly to the effect of adding cement in various
quantities to the tailings.

During the consolidation of uncemented mine tailings from an initial slurry state, the
material behaviour that is most important are the changes in soil stiffness and
permeability that occur due to the reduction in void ratio as consolidation progresses.
The rate of increase of the fill stiffness is important as it governs, inter alia, the amount
of stress transfer due to arching that can occur. In the case of cemented mine backfill, a
number of other mechanisms associated with cement hydration are introduced, which
also need to be addressed during modelling. These include:

the development of an appropriate material stiffness that takes account of soil


volumetric changes, cement hydration and damage that may occur to cement
bonds during the filling process (and thus, changes are required in term D of
Equation 2.2);

changes in permeability, not only with changes in void ratio, but also with the
growth of cement gel in the voids (and thus, changes are required in term C in
Equation 2.2);

self desiccation of the hydrating cement paste the water volume changes
that occur due to chemical reaction in the hydration process (an additional term
in Equation 2.2);

the large increase in stiffness of the cementing soil matrix, leading to the matrix
bulk stiffness being comparable to that of water, which must be taken into
account in estimating pore pressure changes due to increases in total vertical
stress. (and hence changes are required in term A in Equation 2.2); as shown by
Black and Lee (1973), when the stiffness of a soil matrix becomes similar to
that of water, the change in pore pressure due to load application cannot be

4.4
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

estimated using conventional approaches but must account for the relative
stiffnesses.

4.2.3 CeMinTaCo governing equations

The following section provides the derivation of the governing equations for the large
strain consolidation of a cementing soil. The description provided follows that of
Gibson et al. (1981), but these equations are modified to take into account the influence
of cement hydration. In the derivation, convective forces are ignored and the motion of
pore fluid relative to the solids is assumed to be governed by Darcys law.

For this analysis, an updated Lagrangian coordinate system is adopted. Figure 4.1
presents the definition of this coordinate system over a consolidation timestep t to t+t.
With reference to Figure 4.1, a is the original element height, x is the equivalent height
of soil solids, and is the real coordinate system, which is required for the calculation of
hydraulic gradients. These terms can be related through:

da d
dx = =
(1 + eo ) (1 + e) (4.2)

As illustrated in Figure 4.2, the weight of fluid flow out of an element of thickness a
is given by:


[n w (w s )]a (4.3)
a

where n is the porosity, w is the velocity of water s is the velocity of soil and w is
the unit weight of water.

The change in water volume in an element of height a as a result of self desiccation


over time ( t ) is be given by:

a Vsh
(1 + e0 ) t (4.4)

where ( Vsh ) is the change in volume per bulk unit volume of solids as a result of self
desiccation.

4.5
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

Assuming the water volume is independent of the water pressure, Equations 4.3 and 4.4
can be combined to determine a relationship between the change in stored water in an
V f
element of height a with time ( ).
t

a Vsh V f
+ [n ( )]a =
(1 + e0 ) t a w w s
t (4.5)

Assuming laminar flow conditions, the fluid velocity ( w ) relative to the soil velocity

( s ) can be determined using Darcys law, which is defined as:

k uex
n( v w v s ) =
w (4.6)

where ( w s ) is the relative velocity between the soil and the fluid, k is the

uex
coefficient of permeability and is the excess pore pressure gradient.

Combining Darcys law with the coordinate transformation relationship (Equation 4.2)
Equation 4.5 becomes:

a Vsh k 1 + e0 uex V f
+ a =
(1 + e0 ) t a w 1 + e a t (4.7)

In most, if not all, stopes free-draining barricades are constructed at the base. This free
draining condition acts as a base drain and draws down the phreatic surface. The
gradual drawdown of the phreatic surface combined with the accretion of overlying
material makes the it difficult to define excess pore pressures. Due to these reasons, it is
inconvenient to undertake calculations in terms of excess pore pressures and it is more
suitable to perform calculations in terms of total pore pressure (u). When converting
from uex to u Equation 4.7 becomes:

a Vsh k 1 + e0 u 1 + e V f
+ + a =
(1 + e0 ) t a w 1 + e a 1 + eo
w
t (4.8)

In conventional soil mechanics, it is common to assume that the stiffness of the water
phase is significantly greater than that of the soil skeleton and as a result the volumetric

4.6
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

soil compression is considered to be equal to removed water. However, in the case of


cementing soil, the soil stiffness may be comparable to that of water. In order to take
account of the relative stiffness between the soil and water phases, Hookes law can be
applied to the water phase to derive pore pressure changes. If the soil particles are
assumed to be incompressible, this may be written as:

u V f Vs V f (1 + e0 ) Vs (1 + e0 )
= K w = K w
t t V f t V f a t e a t e (4.9)

where Vs is the change in soil volume in an element of height a .


Assuming a constant stiffness ( ) over the timestep ( t ), the constitutive relationship
e
may be represented as:

Vs s v e 1 u e 1
= = = v (4.10)
a t t t (e0 + 1) t t (e0 + 1)

where v and v are the change in vertical effective and total stress, respectively,

and s is the change in vertical strain in the element.

Substituting Equations 4.8 and 4.10 into Equation 4.9 and rearranging:

v u e v v Vsh
= 1 +
t t K w e e t
(4.11)
+
v
(1 + eo ) k 1 + eo u + k
e a w 1 + e a a

v
If Kw is significantly greater than and ( Vsh ) is equal to zero Equation 4.11 takes
e
the form of Gibsons large strain consolidation equation (Equation 2.2) for conventional
soils.


As explained in Chapter 3 the material stiffness v (1 + e0 ) permeability (k ) and
e
self desiccation induced volumetric shrinkage (Vsh ) can all be represented by

functions involving various combinations of effective stress (v ) , void ratio (e ) cement

4.7
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

content (C c ) , time (t ) and effective void ratio (e eff ) . Substituting these dependent

variables into the appropriate terms of Equation 4.11 gives Equation 2.3, which has
been implemented and solved using CeMinTaCo.

u e v
1 (t , Cc , e, v )
t K w e
14
4444 42444444 3
( A)

( ) ( )
k eeff 1 + eo u k eeff
+ v (1 + eo )(t , Cc , e, v ) + (2.3bis)
14e 444244443 a w 1 + e a a
1444444 424444444 3
(D) (C)
V (t , Cc ) v
+ v (t , Cc , e, v ) sh =
14e 444424444 t 43 { t
(E) (B)

In this equation, the modified terms are identified using the same labels used to identify
the equivalent terms in the unmodified equation (Equation 2.2), but there is also an
additional term (E), which does not have an equivalent term in the unmodified equation.

As the stiffness of the cemented soil matrix may be comparable to that of pore water,
the change in pore pressure (term A in Equation 2.2) due to a change in total stress, is
v
modified to take account of the stiffness of the cemented matrix ( ) and that of
e
water ( K w ). This formulation incorporates strain compatibility to achieve an
appropriate distribution of total stress.

The volumetric strain term (term C in Equation 2.2) is modified to take into account the
fact that the permeability (k) is affected not just by normal void ratio reduction due to
consolidation, but also by the formation of cement gel in the void space. This aspect
was addressed in Section 3.3.

The stiffness term (term D in Equation 2.2) is modified to take account of the fact that
v
the material stiffness ( ) is now a function of cement hydration, current stress state
e
and previous stress excursions as documented in Section 3.2.

4.8
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

In addition to these modified terms, an additional term (term E) has been included to

quantify the impact of volume changes ( Vsh ) associated with the self desiccation
t
mechanism, as discussed in Section 3.4.

4.3 NUMERICAL IMPLEMENTATION

The solution to Equation 2.3 is obtained through implementation into the one-
dimensional tailings consolidation program MinTaCo. This program, which has been
renamed CeMinTaCo in its modified form, solves these equations using an implicit
finite element formulation for the space variable, and an explicit finite difference time
marching scheme.

Figure 4.3 shows the geometric layout adopted in the finite difference solution to
Equation 2.3. Each finite element is represented using a single nodal point. Node i has
nodal points i-1 and i+1 in the adjacent elements with the vertical spacing between these
points being a1 and a2 respectively. Assume a time increment that is defined as t where
t = tj+1-tj. The finite difference representation of Equation 2.3 is given by:

{ } { [ ]}
S f a1 1K f + 2 D f ui +1, j +1 + S f 2 D f (a2 a1 ) 1K f (a1 + a2 ) ui , j +1 +
S f a2 {1K f 2 D f } ui 1, j +1 = (4.2)
S f .e i , j S .t S f Vshi , j
ui , j 1 + + v f
K w 1 + ei, j
( ) 2 a a (
1 + ei, j )
1 2

where the coefficients are defined as:

2t
1 =
a1a 2 ( a1 + a 2 )

t
2 =
4( a1 + a2 ) 2

( ) (
K perm = a1 ki +1, j +1 ki , j +1 + a2 ki , j +1 ki 1, j +1 )
d
Sf = v (1 + ei , j )
de i , j +1

4.9
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

ki , j +1 1 + ei , j
Kf =
w 1 + ei , j +1

1 + ei +1, j 1 + ei , j
a1 ki +1, j +1 ki , j +1 +
1 + ei +1, j +1 1 + ei , j +1
1
Df =
w 1 + ei , j 1 + ei 1, j
a2 ki , j +1 ki 1, j +1

1 + ei , j +1 1 + ei 1, j +1

where Kw is the bulk modulus of the water phase and Vsh is the volumetric change that
occurs over the time increment (t) due to self desiccation. This may be calculated as:

d
Vsh =
1
t .Wc Eh . .exp d
2 (t *)1.5 t* (3.32bis)

where Wc is the weight of cement per unit volume of material, t* is the time since the
commencement of hydration, Eh is the efficiency of hydration (as defined in Section
3.4) and d is the hydration maturity constant.

The duration of each consolidation timestep is initially estimated based on Terzagis


time factor for individual layers. This time factor is then used along with a user-defined
allowable strain level to make an initial estimate for the allowable time increment.
Using the defined timestep, a solution (based on a void ratio convergence criterion) is
sought through a maximum of 40 successive iterations. Should the solution not
converge, a smaller timestep is established and the calculation repeated. After
converging to an acceptable solution, the maximum induced strain is determined and
compared with a user-defined tolerance. Should the tolerance be exceeded, the time
increment is halved and the calculation repeated until the strains obtained are less than
the allowable strains.

Over each timestep, the material properties are assumed to remain constant, but at the
completion of each timestep, the material properties are updated in accordance with the
time increment and strains that occurred during that timestep.

4.10
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

4.4 MODEL VERIFICATION

4.4.1 Compressibility

In order to demonstrate its applicability in modelling compressibility, the proposed


approach has been used to simulate a series of one-dimensional compression tests on
cemented CSA hydraulic fill. In these experiments, the specimens were prepared at
different densities and allowed to cure for different times prior to loading. The material
constants d, X, W and Z (in Equation 3.5) were determined from the measured values of
qu and from Go values obtained from bender elements. The parameters * and * (in
Equation 3.1 and 3.2) were determined from one-dimensional compression tests on
uncemented material, and through modifying the terms A (in Equation 3.5) and b (in
Equation 3.11), the one-dimensional compression response could be adequately
represented, as was previously presented in Figure 3.5.

4.4.2 Self desiccation

To verify the self-desiccation aspect of the model, a hydration test was carried out, and
the CeMinTaCo program was used to reproduce the reduction in pore pressure induced
by self desiccation observed in the experiment. The test involved preparing a fully
saturated sample of CSA hydraulic fill, at a void ratio of 0.7 and cement content of 5%,
in the form of a triaxial test sample. This was mounted in a triaxial cell, and enclosed in
a latex membrane in the usual way. The sample was subjected to a total cell pressure of
850 kPa and an initial back pressure of 830 kPa. Then, the back-pressure valve was
closed, so that the hydration process could take place in a completely undrained state.

The results are shown in Figure 4.4 as a plot of measured pore pressure versus time. In
this case, the self-desiccation mechanism has reduced the pore pressure from the initial
back pressure value of 830 kPa to a final value close to zero. (This suggests that it might
have been appropriate to start from an even higher back pressure in this case, to ensure a
final pore pressure well above zero). The response fitted using the CeMinTaCo program
is also shown in Figure 4.4, which indicates that the program is capable of reproducing
the observed experimental results quite well. It should be noted that the CeMinTaCo
output was modified in accordance with the Poissons ratio to convert from a one-
dimensional situation to an isotropic situation.

4.11
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

This experiment shows how effective the self-desiccation mechanism can be in reducing
pore pressure (and hence in increasing effective stress). In this case, this occurs in the
absence of any external drainage effects, but in a real stope, it would combine with any
consolidation drainage to potentially produce a faster pore pressure reduction than
would otherwise be the case.

4.5 SENSITIVITY STUDY


A limited sensitivity study was undertaken to illustrate the effect of varying some of the
input parameters on the response when modelling the filling of a stope. The filling
strategy used in the study was based on a common paste fill schedule, and involved
filling an initial plug at 0.4 m/hr for 16 hours, followed by a 24-hour rest period, and
then completing filling the remaining 30 m at a rate of 0.4 m/hr. The base-case set of
input parameters used in the study is given in Table 4.1; these parameters correspond to
typical paste backfill properties, and except where otherwise indicated these parameters
are used in all the examples that follow.

Proper simulation of the three-dimensional geometry of a real stope and drawpoint


(illustrated schematically in Figure 1.2) would require a three-dimensional FE program,
or at least a two-dimensional or axi-symmetric program. However, since CeMinTaCo is
only one dimensional, a means of simulating the restriction to drainage resulting from
the reduced drawpoint cross-section was required. The method adopted is illustrated in
Figure 4.5, which consisted of introducing a 5 m thick layer with reduced permeability
(1/8 of the value adopted for the bulk of the material at a corresponding void ratio) at the
base of the stope. Note, however, that the material in this region had zero cement
content, so none of the effects of self desiccation apply in this region.

In the following sections, the effect of changing a number of parameters is investigated.


These parameters include cementation, permeability, and damage.

4.5.1 Influence of cementation

To illustrate the effect of cementation on the consolidation response, analysis was first
carried out using uncemented material. Then, the analysis was repeated with cement
content (Cc) of 5%, in one case with the self-desiccation mechanism disabled in the
model, and the other with it enabled.

4.12
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

Figure 4.6 presents the results of these analyses as plots of pore pressure (2 m above the
base drainage layer) versus time. Also shown is a plot of the total vertical stress versus
time at the same point in the profile. These plots indicate that for the case without any
cement, the increase in pore pressure is similar to the increase in total stress, and hence
very little effective stress is generated during filling. After the completion of filling, the
rate of dissipation of pore pressure is slow, and very little consolidation has occurred at
the end of the period modelled (250 hours). This indicates that, in the absence of
cementation, paste fill barricade stresses would be well represented by the undrained
case that was discussed on Section 2.2.

For the second case, it may be seen from Figure 4.6 that, even with the self-desiccation
mechanism disabled, pore pressures are significantly less than for the uncemented case
once hydration begins. The effect of adding cement is to produce an increase in soil
stiffness (after the start of hydration) and a reduction in permeability, as previously
discussed. Since the rate of pore pressure reduction (consolidation) is dictated by the
product of stiffness and permeability, this result indicates that the effect of the increase
in stiffness outweighs the effect of the reduction in permeability, and there is some
increase in the rate of pore pressure dissipation, even during filling. However, it should
be noted that there are still significant excess pore pressures present at the end of filling.
Due to the increased stiffness, these dissipate somewhat more rapidly after completion
of filling than in the previous case.

In the third case, with self desiccation enabled, there is a significant increase in the
degree of pore pressure reduction that takes place, indicating the potential importance of
the self-desiccation mechanism in promoting dissipation of pore pressure. It should be
noted that the dissipation of pore pressure would act to increase the effective stress,
promoting arching, which would further reduce the pore pressure. This will be shown
with reference to in situ measurements later, and also when two-dimensional modelling
results are presented in Chapter 7.

4.5.2 Influence of permeability

To illustrate the influence of permeability on the overall filling response, analyses were
carried out using the filling sequence and material properties described above, with Cc
of 5%, but using three different permeability relationships (i.e. three different values of

4.13
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

the permeability parameter ck in Equation 3.15). The resulting three permeability


functions (denoted k1, k2 and k3) are plotted against curing time in Figure 4.7. In this
plot, the change in permeability with time shown for each case is due only to cement
growth, whereas in a modelling situation, the change could be greater with the added
effect of void ratio reduction due to compression.

Figure 4.8 shows the pore pressure during and following filling for a point 7 m above
the base (i.e. 2 m above the lower-permeability drawpoint region) for the three cases
considered. Figure 4.9 shows the pore pressure profile down through the stope at the
end of filling. Also shown in Figure 4.9 is the final steady state pore pressure (SSPP)
for the k1 case (the SSPP lines for the other two cases are practically coincidental). In
Figure 4.8, the final steady state equilibrium has been reached at about 230 hours for the
k1 case, whereas changes were still occurring for the other two cases at this stage.

In Figure 4.9, the line labelled k1 SSPP refers to the steady state pore pressure that
results from maintaining a water table in the stope at 33.65 m height (the final filled
height) and zero pore pressure at the base of the drawpoint, with the permeability in the
lower 5 m being about 8 times less than the permeability in the stope proper for each
case. Note that the permeability in the stope is not uniform with depth, so the ratio of 8
refers to average values. In Figure 4.9, the difference between the pore pressure at the
end of filling and the SSPP line is the excess pore pressure at this stage. Thus, there are
excess pore pressures in the stope for each of the three cases at the end of filling, but
these are different for the three cases.

These results appear to be counterintuitive, since conventional consolidation theory


suggests that high permeability material should dissipate pore pressures more quickly
than low permeability material. Thus, in Figure 4.8, while the lowest permeability case
(k3) shows the highest pore pressure in the first filling stage, it shows pore pressures
very much less than the higher permeability cases during the rest period and at all times
thereafter. Examination of the pore pressure plot at the end of filling for this case in
Figure 4.9 shows that from the surface down to about 25 m above the base, the pore
pressure gradient corresponds to the total overburden stress gradient i.e. there is no
pore pressure dissipation in this region at this stage, and hence the effective stresses are
zero. However, below about 25 m, the pattern changes completely. In this region,

4.14
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

hydration and self desiccation are occurring, setting up high negative excess pore
pressures. Though this results in a very steep downward hydraulic gradient (from 25 m
down to 13 m), the low permeability prevents sufficient internal water flow from
dissipating these negative pore pressures. If later pore pressure isochrones for this case
were plotted, these would show the point of minimum pore pressure gradually moving
upwards until it reached the surface (as hydration progressed). Then, slow internal
flows would gradually move the pore pressures onto the steady state line.

For the highest permeability case (k1), the same internal volume change occurs due to
hydration, and thus the potential pore pressure reduction resulting from this is the same.
However, the higher permeability in this case means that high internal hydraulic
gradients are not sustainable, due to the ease of generating internal water flow to smooth
out the pore pressure profile. Thus, at the end of filling, the pore pressures for the k1
case are not very different from the SSPP values. There is still evidence of the self
desiccation process occurring in this plot i.e. the slight concave-upward curvature of
the pore pressure profile from about 10 to 20 m would not be present if self desiccation
was not influencing the process.

The intermediate permeability case (k2) shows behaviour similar to the k1 case,
commensurate with the fact that the permeability is not very much lower than for the k1
case, as shown in Figure 4.7.

4.5.3 Typical damage scenario

As discussed throughout this thesis, the filling process involves the interaction of three
time-dependent processes: the rate of filling, cement hydration and consolidation.
Researchers such as Le Roux at al. (2002) and Belem et al. (2006) have investigated the
mine backfill process experimentally. In this work, the authors apply effective stress to
curing cemented paste fill at a rate equivalent to the accumulation of total stress from
the fill self weight. Through adopting a total stress approach, rather than an effective
stress approach, the third of the time-dependent processes (consolidation) has been
neglected. This section presents a numerical investigation of the interaction of the three
time-dependent processes with specific emphasis on the influence that damage (from
excessive vertical effective stress on fragile cement bonds during the early stages of
hydration) has on the consolidation process.

4.15
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

The purpose of this analysis is to select a material that possesses an ultimate strength
that is suitable for eventual vertical exposure, and determine if this material can be
damaged during a typical filling sequence. The focus of the analysis is the early stages
of hydration, where fragile cement bonds may be damaged by the application of
compressive stress. The material properties from Table 4.1 were adopted, but in order to
produce a material with an ultimate qu that is equal to the final vertical effective stress,
the void ratio has been increased to 1.25. The property that dictates the response of the
material to damage is the damage coefficient (b) in Equation 3.11; in this analysis,
damage coefficients of 0.05 and 3.0 have been adopted. These values are considered
appropriate for representing the entire range of damage coefficients for mine backfill.

The impact of this variation on the structural breakdown has been demonstrated in
Figure 4.10, which presents the results of a simulated one-dimensional compression test
on a fully hydrated sample. Figure 4.10 indicates that prior to yield, the behaviour is
largely independent of b, but after yield the rate of cementation breakdown is very
sensitive to the value of b.

The analysis has considered two materials, one with the value of the permeability term a
given in Table 4.1, and the other with the value of a increased by two orders of
magnitude. These are meant to represent a paste fill example and a hydraulic fill
example, respectively.

Paste fill modelling


The results of the paste backfill modelling are presented in Figures 4.11 and 4.12.
Figure 4.11 shows the development of pore pressure (u), vertical total stress (v) and
vertical effective stress (v) against time, at a point 2 m above the base of the stope,
using both damage coefficients (b). The results of the simulation with b = 3 are plotted
as lines, while those with b = 0.05 are plotted as symbols. Figure 4.11 indicates that in a
typical paste fill situation, compressive damage is unlikely to influence the
consolidation behaviour.

The main reason for independence is as follows:

The application of total stress (due to the accumulation of overlying material) is


resulting in an equal increase in pore pressure, and therefore there is no change
in effective stress, and no possibility of damage being caused.

4.16
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

Due to the low coefficient of consolidation (associated with the fine grained
nature of paste fill) the cement hydration is required to achieve consolidation
(i.e. pore pressure reduction, rather than conventional consolidation).

As a result, the hydration timescale dictates the timescale of consolidation (or


application of effective stress). Therefore, softer/weaker material consolidates more
slowly, allowing bonds to mature appropriately, while stiff/strong material rapidly
develops sufficient strength to overcome the associated rapid application of effective
stress.

Figure 4.12 presents the application of vertical effective stress (v) and one-
dimensional yield stress (vy) against time for the element. Initially both v and vy
are very low. After reaching initial set vy increases at a significantly faster rate than
v therefore eliminating the likelihood of damage. This would typically be the case in a
paste fill situation. Also presented in Figure 4.12 is the calculated unconfined
compressive strength (qu) for the in situ material and that for material cured under no
stress (as would be the case in the laboratory). Comparison of these indicates that the
application of effective stress during filling can actually increase the material strength
(by reducing the final density) rather than damaging fragile cement bonds.

This behaviour is in accordance with that observed by Blight and Spearing (1996), who
investigated the effect of stope lateral strain (i.e. stope closure) on cementing backfill,
and showed that such strain (i.e. the resulting effective stress increase), resulted in
higher final strengths. It is also in accordance with many field observations (Revell,
2004, Cayouette, 2003), where the unconfined compression strengths obtained from
cores taken from filled stopes are often significantly greater than the control samples
taken from the material as it is filled and cured under zero confining pressure.

Hydraulic fill modelling


It has been demonstrated that in a typical paste fill situation the damage coefficient,
which represents the rate of cementation breakdown with strain, has little influence on
the consolidation and hydration behaviour because in this case it is the hydration
process that generates effective stress. However, in the case of hydraulic fill, the
material often has a much higher permeability. As a result (if it is assumed that

4.17
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

immediate consolidation occurs) the rate of development of effective stress is more


closely related to the filling rate.

In order to simulate a typical hydraulic filling situation, the same material properties as
those adopted for the paste fill example are adopted, but the permeability is increased by
two orders of magnitude. Furthermore, rather than a filling rate of 0.5 m/hr with a single
long rest period (as in the paste fill case) a constant filling rate of 3 m/day was adopted
without a rest period. This is expected to be indicative of typical hydraulic filling rates,
when taking account of fill and rest periods that are commonly adopted to avoid piping
type failures (Cowling et al. 1987).

The results of this modelling are presented in Figures 4.13. Figure 4.13 presents the
development of pore pressure (u), vertical total stress (v) and vertical effective stress
(v) against time for a point 2 m above the base of the stope. Again the results of
modelling with a damage coefficient b = 3 are presented as lines while those with a
damage coefficient b = 0.05 are presented as symbols. It is clear from this figure that,
as in the paste fill example, the damage coefficient has little influence on consolidation
in a typical hydraulic fill situation. This is most likely to be related to the fact that (as
indicated in Section 2.2 using Gibsons (1958) analytical solution) the combination of
permeability and stiffness immediately after placement of hydraulic fills results in
excess pore pressures dissipating rapidly. Hence, water pressures in a hydraulic fill
stope are dictated by the restriction at the drawpoint, rather than the dissipation of
excess pore pressures. This point was also evident in the high permeability example in
Section 4.4.2.

Figure 4.14 presents the development of one-dimensional yield stress (vy) and applied
vertical effective stress (v) against time in an element 2 m above the stope floor.
Unlike with the paste-fill case, the application of vertical total stress (from the accretion
of material) creates an immediate increase in vertical effective stress. This acts to
compress the soil, increasing vy even prior to the onset of cement hydration. This
creates some initial hardening of the material and, as with the paste-fill case, the onset
of cement hydration causes vy to increase significantly faster than v. Again this
reduces the likelihood of damage.

4.18
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

Also presented in Figure 4.14 is the calculated qu against time for material in situ and
that expected for the material cured under zero applied stress. As with the paste-fill
case, qu for material cured in situ is greater than that for material cured under zero
stress. But it is important to note that even with the reduced filling rate (relative to the
paste fill situation), the increase in qu of the in situ material relative to that cured under
zero stress is significantly greater. The increased ratio is a result of the soil compression
that occurred immediately after deposition. This compression increased the material
density, which leads to the higher strengths.

Summary
Overall, it can be concluded that, under most conditions, the interaction of effective
stress and the growth of fragile cement bonds during filling is unlikely to adversely
influence the consolidation behaviour or damage the fragile cement bonds in either a
paste or hydraulic fill situations. In fact, modelling indicates that in most cases this
interaction will actually lead to higher material strengths in situ due to soil compression.

4.5.4 Strain requirements

One component that is often incorporated into tailings consolidation models is large
strain consolidation theory. This is important when analysing the consolidation of large
storage facilities containing compressible tailings. But in a tailings-based backfill
situation material is either cycloned (to remove fine particles) or, if placed as full-stream
tailings, combined with cement to avoid the material liquefying after placement.

In addition to reducing the risk of liquefaction, the removal of fines from a tailings
stream (as in a hydraulic fill) increases the material stiffness (in an uncemented state)
and therefore reduces the amount of compression, while, as noted by Le Roux et al.
(2005) the presence of cementation in full-stream tailings backfill (paste fill) acts to
reduce the compression that occurs during the filling process. For these reasons large-
strain numerical formulation may be unnecessary when undertaking consolidation
analysis of mine backfill.

To investigate this aspect, the vertical strain calculated for the paste fill example
presented in Section 4.5.3 is plotted against time in Figure 4.15. This figure indicates
that vertical strain levels of 4% occurred for the cemented paste example.

4.19
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

As hydraulic fill can be placed without any cement, it is important to consider the likely
strains that would be generated for a typical uncemented hydraulic fill. The values of
uncemented compression parameters ( = 0.06, = 0.006) that were adopted for the
previous study are considered to be greater than the values relevant to a typical
hydraulic fill, and hence using these parameters would overestimate strains in a
hydraulic fill situation. Based on Rowe cell testing of hydraulic fills, compression
parameters of = 0.035, = 0.0035 are considered more appropriate. These
compression parameters were adopted, along with the other material properties for the
hydraulic fill example in Section 4.5.3, and zero cement content, in an analysis to
investigate the degree of compression likely to occur in a typical uncemented hydraulic
fill situation. The calculated axial strain is plotted against time in Figure 4.15, indicating
that a maximum strain of 6% occurred in this example.

The axial strains calculated for both the paste and hydraulic fill examples are both
significantly less than 20%, which was specified by Tan and Scott (1988) as the strain
levels requiring large strain formulation. This result indicates that the large strain
approach (involving a Lagrangian coordinate system and very small timesteps) adopted
in CeMinTaCo is unnecessary, and it would appear that a conventional Cartesian
coordinate system can provide a suitable representation of most mine backfill situations.
This is the approach adopted in the two-dimensional consolidation program, which is
presented in Chapter 5.

4.5.5 Comparison with data from in situ monitoring of filled stopes

To determine how well the CeMinTaCo program can reproduce the behaviour in an
actual mine backfilling situation, the program was used to simulate the deposition of a
fine-grained cemented paste fill at the Cayeli mine in Turkey. The properties of the as-
placed material adopted in the modelling were in accordance with those in the field.
These included a placed void ratio of 1.0, a cement content of 8%, and a fully hydrated
qu of 1 MPa. However, some of the other relevant material properties could not be
obtained, and thus, in order to gain a reasonable estimate of appropriate material
properties, those determined for a similar tailings material with the same cement content
were adopted. This material had grain size distribution, and mineralogical and cement

4.20
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

characteristics, similar to those encountered at Cayeli and was therefore expected to


have similar properties.

The values chosen in this way include the efficiency and rate of hydration (Eh = 0.032, d
= 1.5), the ratio of qu to vy (6), the uncemented compression parameters ( = 0.06,
= 0.009) and permeability parameters (k 510-8 m/s), the ratio between cemented
stiffness and strength (1300) and the damage coefficient (b = 0.5). The method adopted
to account for the flow restriction due to the drawpoint is the same as that used earlier in
the parametric study, and is illustrated in Figure 4.5.

The modelling was carried out by increasing the fill height at the same rate as that
adopted in the field. The filling rate was a constant rate of rise of 0.4 m/hr for the first
24 hours followed by a 9-hour rest period, and then filling the remainder of the stope at
a rate of 0.4 m/hr over a 100-hour period. Figure 4.16 shows a plot of pore pressure
versus time obtained from the modelling, compared to the field measurements. The
monitoring location for the in situ measurements was 1.0 m above the stope floor, and
the CeMinTaCo results plotted refer to the same elevation.

In Figure 4.16, the response during the initial stage of filling is linear (and, though not
shown, coincides with the total stress increase during this period). However, the onset
of initial set coincides with a reduction in the rate of pore pressure increase, such that
from about 20 hr onwards, the pore pressure is actually reducing for both the measured
and model values even as filling continues (up to 24 hr). When filling recommences (at
approximately 33 hr) the initial pore pressure behaviour appears to be reasonably well
modelled, but as filling continues the model and in situ results start to diverge. For the
in situ case, it is likely that, due to consolidation, some of the fill/rock interface strength
is mobilised, resulting in a stress redistribution to the surrounding rockmass (arching).
This reduces the total vertical stress imposed on the material at the monitoring point,
resulting in a lower pore pressure increase than would otherwise have occurred. In fact,
any tendency for pore pressure increase beyond this point is completely counteracted by
on-going drainage (and self desiccation), with the result that the pore pressure at the
monitoring point continues to reduce. Clearly, the one-dimensional CeMinTaCo model
is not able to account for the arching mechanism, and in this case it predicts an increase,
rather than a decrease, in pore pressure when filling recommences.

4.21
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

While many of the material parameters used in the modelling have not been derived
directly for the material being modelled, the ability of the model to reproduce the
significant characteristic of the filling process based on properties of similar material
illustrates that the model is capturing the most significant mechanisms associated with
mine backfill placement. However, it is also clear that a two-dimensional model (plane
strain or axi-symmetric), or even a full three-dimensional model, is required to capture
the complete behaviour.

4.6 CONCLUSION
Overall, this section has demonstrated that the mine backfill process is a complex
interaction of mechanisms. This interaction of mechanisms can actually create
circumstances that produce counterintuitive outcomes, as was demonstrated in the
permeability sensitivity study (Section 4.5.2). For example, contrary to rules of thumb
used in industry, filling a stope with low-permeability material can result in very low
pore pressures being present in the fill at the end of filling pore pressures much lower
than those in a free-draining fill. Therefore, in order to predict the overall response, the
individual mechanisms need to be fully coupled into a program such as CeMinTaCo. It
would be unwise to speculate about the impact of a single mechanism on the overall
response and it would be unwise to attempt to superimpose the impact of individual
mechanisms in an effort to understand the cumulative response.

This work has demonstrated that compressive yielding during placement is unlikely to
occur in a typical mine backfill situation. In addition, analysis of typical strains levels
during filling indicate that, under normal conditions, small-strain formulation is
sufficient to capture the consolidation behaviour of mine backfill.

Comparison with in situ monitoring demonstrated that the one-dimensional model


(CeMinTaCo) was capable of accurately representing the early age behaviour; but
diverged from in situ measurements during the later stages of filling. This is believed to
be a result of stress redistribution to the surrounding rockmass. To appropriately address
this mechanism, a two- or three-dimensional model is required. The development of a
new fully-coupled two-dimensional finite element model is presented in Chapter 5 of
this thesis.

4.22
Mechanics of Mine Backfill Matthew Helinski
One-dimensional Consolidation Modelling The University of Western Australia

4.23
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

CHAPTER 5
TWO-DIMENSIONAL CONSOLIDATION ANALYSIS
(MINEFILL-2D)

5.1 INTRODUCTION

In Chapter 4, a computer code (CeMinTaCo) for modelling the behaviour of backfill


material undergoing consolidation and cement hydration during and following
placement in a stope was presented. Being one-dimensional, the model can not, of
course, deal with any of the two-dimensional or three-dimensional aspects of the
behaviour in a real stope. For example, in comparing the numerical output from
CeMinTaCo with the in situ measurements, e.g. as presented in Section 4.5.6, one
drawback that becomes evident is the inability of the one-dimensional model to
appropriately capture the redistribution of stress to the surrounding rockmass (arching),
and thus it cannot take account of the reduced vertical stress that can result from
arching. It is likewise incapable of representing the stope drawpoint in a geometrically
correct fashion. Without being able to represent the stope drawpoint correctly, artificial
base boundary conditions are required to represent the drainage restrictions through the
drawpoint, such as those described in Figure 4.3. In addition, a one-dimensional model
cannot appropriately represent the horizontal stresses placed on barricades. In order to
take these aspects into account, a two- or three-dimensional model is required.

While the one-dimensional nature of CeMinTaCo has these drawbacks, it has shown
that some aspects of cemented mine backfill behaviour that were previously thought to
be important were, in fact, not so important after all in most situations. These were the
requirement to use large-strain formulation, and the possibility of yielding of cement
bonds as they formed due to the development of excessive confined vertical
compressive stresses.

With regard to the first aspect, the sensitivity study in Chapter 4 demonstrated that the
presence of cementation (in paste fill) or the removal of fines (in hydraulic fill) result in
strain levels that are far less than 20% (strains in the order of 6% were calculated). Tan
and Scott (1988) suggest that for strains less than 20%, a small-strain formulation

5.1
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

provides accurate solutions. Therefore, it is considered acceptable to perform the


calculations using the conventional Cartesian coordinate system rather than the
Lagrangian coordinate system that was used in CeMinTaCo.

With regard to the second aspect, a study into the influence of structural damage from
excessive compressive stress indicates that, provided the material has sufficient strength
to support the ultimate stress state, effective stress generated during the early stages of
hydration is unlikely to damage cement bonds as they form. Therefore, it is considered
appropriate to undertake calculations where compressive yield of the cemented structure
is neglected and a Mohr Coulomb yield surface is used to represent the behaviour of the
cemented material. It should be noted that compressive yield of the material in an
uncemented state is taken into account to address any compression that occurs prior to
the onset of hydration, which was shown to be relevant in Section 4.5.3.

This chapter begins by outlining the unique requirements of a model to represent the
mine backfill deposition process. This is followed by a description of the governing
equations and numerical formulations developed within the program Minefill-2D. An
overview of the material models adopted in this program is then provided. Minefill-2D
is then compared with various well-established analytical solutions and CeMinTaCo to
verify the performance of the model. Finally, a comparison against in situ
measurements is undertaken to verify the applicability of the program to the mine
backfill deposition process.

Many of the basic ideas incorporated into Minefill-2D are the same as those
incorporated into CeMinTaCo, and these ideas have been thoroughly explored in
previous chapters. Nevertheless, there is a considerable amount of repetition of this
material in this chapter; this has been done for the sake of completeness, and to make
this chapter more coherent.

5.2 PROGRAMMING REQUIREMENTS

The main features required in this model included the following:

fully coupled analysis i.e. full coupling between compression, water flow, and

the cementation processes;

5.2
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

calculations in terms of total water head;

ability to vary boundary conditions on the upper surface, to enable fresh material

to be added, and to deal with various water outflow or inflow conditions;

ability to take account of water accumulation above the fill surface (influencing

surface total stresses and pore pressures);

a constitutive model that takes account of cement hydration;

a permeability model that takes account of cement hydration;

a self-desiccation model;

strain softening of the fill mass due to interface shear during filling.

In order to achieve these requirements, a number of commercially-available programs,


including FLAC, Plaxis and AFENA, were examined. None of these programs proved
to be suitable due primarily to problems associated with simulating the accretion of soft
soil whose properties evolve with time under fully coupled conditions. Specifically this
related to dynamic waves generated during the consolidation of soft material disrupting
the calculation scheme (in FLAC), problems associated with undertaking calculations in
terms of total water head and problems associated with establishing boundary
conditions along the surface nodes and then including these surface nodes back into the
calculation scheme at a later stage. Because of these problems, it was decided that the
most appropriate approach would be to develop a new two-dimensional (plane-strain or
axi-symmetric) model, which specifically addressed the described criteria. This program
was coded in Visual Fortran 90 and named Minefill-2D.

5.3
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

5.3 PROGRAMMING METHODOLOGY

5.3.1 Introduction

This section presents the relevant numerical components incorporated into the program
Minefill-2D. It includes an introduction of the numerical technique, a description of the
calculation sequence, an outline of the governing equations, as well as a description of
how other numerical difficulties were addressed. Throughout this section, all
relationships are formulated for the condition of plane strain analysis. Section 5.3.6
presents a description of how the equations are converted for axi-symmetric analysis.

5.3.2 The finite element method

The numerical technique adopted is the finite element method. As described in Potts
and Zdravkovi (1999) the finite element method involves 6 main steps. These include:

Element discretisation

Primary variable approximation

Element equations

Global equations

Boundary conditions

Solution to the global equations

Element discretisation
The first step in the finite element formulation is to discretise the problem geometry into
small domains, (elements). These individual element are then connected by a series of
points (nodes). The most important feature of defining a suitable discretisation of the
problem geometry is to increase the number of elements in regions where unknowns
vary rapidly, such as displacements (shear strains) at the fill/rock interface.

In two-dimensional problems, triangular or quadrilateral elements are commonly


adopted. Throughout this thesis, quadrilateral elements were adopted with 8 boundary

5.4
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

nodes to represent displacement and 4 boundary nodes to represent pore pressures, as


illustrated in Figure 5.1.

Primary variable approximation


The primary unknown variables adopted in Minefill-2D are displacement and pore
pressure. Stresses and strains are determined as a function of the calculated
displacements field. In the two-dimensional analysis presented, global displacements u
and v are determined in the x and y direction respectively (typically the conventional
symbols adopted for displacements are u and v but to avoid confusion with pore
pressures and Poissons ratio these modified symbols were used).

Across each element, the displacements are assumed to vary in accordance with a
polynomial shape function, where the order of the polynomial depends on the number
of nodes in the element. Displacement ( u , v ) at any point within an element are defined
in accordance with nodal displacements and the matrix of shape functions [N d ] such
that:

u
v = [N d ]{u n1 ,u n 2 ,...,u n8 ,vn1 ,vn 2 ,...,vn8 } (5.1)

where u and v are the displacements at any point within the element and uni , vni are
the displacements (in the x and y directions respectively) at nodal points i.

If displacements vary quadratically across an element, strains and therefore effective


stresses vary linearly across the element. To ensure that effective stress and pore
pressure vary in the same way across an element, it is conventional to adopted 4-noded
elements to represent pore pressure variations. Therefore, pore pressures (u) within an
element are related to the four nodal pore pressures by:

[u] = [N p ]{un1,un 2 ,...,un 4 } (5.2)

where uni is the nodal pore pressure at location i and N p is the shape function.

To assist with the numerical formulation, most finite element programs (including
Minefill-2D) uses isoparametric elements, where the global coordinates (x, y) of a point
in an element are expressed as a function of the global nodal coordinates (xi, yi) and

5.5
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

local shape (interpolation) functions Ni(s,t) . For the elements shown in Figure 5.2, the
global coordinates of a point in the element can be expressed by coordinate
interpolation of the form:

8 8
x = N i (s, t )xi and y = N i (s, t ) y i (5.3)
1 1

where s and t are local coordinates which vary from -1 to 1 across each element. The
purpose of introducing this coordinate system is to allow a solution to be sought through
Gaussian integration. The term isoperimetric come from the fact that the geometry is
approximated using the same shape function as that for displacements, which simplifies
the element equations.

Element equations
a. Co-ordinate transformation

The element equations govern the deformation of each element. The formation of
element equations combines compatibility with equilibrium and the constitutive
relationship.

Using the primary variable approximation, presented previously, the compatibility


equations can be represented as:

N1 N 8
0 0 0 u1
x x x
0 N1 N 8 v&&1
y 0 0
= y y ... (5.4)
xy N1 N1 N 8 N 8 u
z y 0
x v&&
8
x y

0 0 0 0 0 8

which can be more conveniently expressed as:

{} = [B ]{d }n (5.5)

where [B ] contains the derivatives of the shape functions [Ni],and {d }n contains a list
of the nodal displacements for a single element. The shape functions [Ni] depend only
on the local coordinates (S and T). Therefore, in order to calculate the derivatives of
these shape functions relative to the global coordinate system (x and y), a chain rule is

5.6
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

required to relate the x and y derivatives to derivatives with respect to S and T. Using the
chain rule:

T T
N i N i N i N i
= J (5.6)
S T x y

where J is defined as the Jacobian matrix:

x y
S
J = S
x y (5.7)

T T

the global derivatives of the shape functions (used in Equation 5.4) can be obtained by
inverting Equation 5.6.

Time-dependent consolidation requires the material constitutive model to be combined


with equilibrium and Biots consolidation equations. Coupling these leads to the
development of the governing equations of finite element consolidation as follows.

b. Constitutive model

Assuming elastic behaviour over a given loading increment ( ), Hookes law states
that the relationship between stress and strain can be represented as:

{} = [D]{} (5.8)

[
where {} = 'x , 'y , xy ]T [ ]
and {} = x , y , xy T are the incremental

effective stress and strain vectors in plane strain respectively, and [D ] is the assumed

relationship between these vectors. [D ] is based on the drained Youngs modulus (E)
and the drained Poissons ratio () such that:

E (1 ) E
(1 + )(1 2) 0
(1 + )(1 2)

E E (1 )
[D ] =
0
(1 + )(1 2) (1 + )(1 2) (5.9)
E
0 0
3(1 2)

5.7
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

Terzaghis principle of effective stress states that:

{} = {}+ {u} (5.10)

where is the change in total stress, is the change in effective stress and u is
the change in pore pressure. Substituting Equation 5.8 into Equation 5.10 gives:

{} = [D]{}+ {u} (5.11)

where is the strain matrix.

c. Continuity

The equation for pore fluid continuity is:

x y
+ Q = v (5.12)
x y t

where x, and y are the components of pore fluid velocity in the coordinate directions
and Q is any source or sink term.

Assume the pore fluid flow is in accordance with Darcys law:

h
x k xx k xy x
= k k yy h (5.13)
y yx
y

or {} = [k ]{h}

where kij is the coefficient of hydraulic conductivity for the soil, [k ] is the hydraulic
conductivity matrix and h is the hydraulic head, defined as:

u
h= + (x.i gx + y.i gy )
w (5.14)

Vector {i g } = {i gx , i gy } is the unit vector parallel to, but in the opposite direction to,
T

gravity and w is the unit weight of water.

d. Equilibrium and governing equations

5.8
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

Rather that solving for force equilibrium, a more convenient formulation is brought
about through considering conservation of energy, (i.e. internal ( W ) and external
work ( L ) must be equal):

W L = 0
(5.15)

The internal work is given by the integration of the increment of total stress multiplied
by the increment of strain across the element:

1
{} {}dVol
T
W = (5.16)
2
Vol

Using Equation 5.11, Equation 5.16 can be split into soil matrix and pore pressure
terms, and re-written as:

[{} [D]{}+ {} {u}]dVol


1 T T
W = (5.17)
2
Vol

The work done by the incremental applied loads ( L ) can be divided into contributions
from body forces and surface tractions, and can be expressed as:

T T
L = {d } {F }dVol + {d } {T }dSrf (5.18)
Vol Srf

Combining the above, the equilibrium condition is satisfied during the consolidation
step when

[K E ]{d }+ [L]{u} = {RE } (5.19)

where d is the nodal displacements vector, and

T
[K E ] = Vol [B ] [D][B ]dVol (5.20)

which is termed the element stiffness matrix, and [B ] is the derivative of the shape
functions as discussed previously. [LE ] is termed the element volume matrix and is
defined as:

5.9
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

[LE ] = {m}[B ]T [N p ]dVol (5.21)


Vol

where [Np] is the matrix of shape functions for pore fluid pressure interpolation and

{m}T = {1 1 1 0 0 0}

Using the principle of virtual work, the continuity equation (5.6) can be written as:

T
{} {(u )}+ tv u dVol Qu = 0 (5.22)
Vol

Substituting Darcys law (Equation 5.13) into Equation 5.22 gives:


{h} [k ]{(u )}+ tv u dVol = Qu
T
(5.23)
Vol

where [k ] is the permeability matrix and {h} is the gradient of total water head. h
{}
takes account of both elevation head ig and total pore pressure {u} so that

calculations can be carried out in terms of total water head rather than excess pore
pressures. It is important for Minefill-2D to carry out analysis in terms of total water
head, as the combination of on-going filling and drainage (through base barricades)
make it difficult to define hydrostatic conditions. The sink term ( Q ) becomes
particularly important when accounting for self-desiccation volumetric changes.

v v
If is approximated as , equation 5.23 can be re-written as:
t t

[LE ]T {d }n [ E ]{u}n = [nE ] + Q (5.24)


t

where

[E ]T [k ][. E ] dVol
E =
Vol
w (5.25)

represents the element permeability matrix,

5.10
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

[E ] [k ]{i }dVol
T
nE = g (5.26)
Vol

represents flows due to gravitational forces, and

T
N N N
[E ] = p , p , p (5.27)
x y z

represents the derivative of the shape functions for pore water pressure interpolation.

Equations 5.24 and 5.19 are solved using a time-marching process such that if the nodal
pore pressure {u}n and displacements {d }n are known at time t1, then the solution for
nodal pore pressures and displacements is sought at time t2 = t1 + t. If a finite
difference approach is adopted, and assuming a linear interpolation in time, the resulting
equation is:

t2
[ E ]{u}n dt = [ E ][({u}n )2 + (1 )({u}n )1]t (5.28)
t1

Booker and Small (1975) demonstrated that, in order to form a stable solution to
Equation 5.28 , the value of must be greater than or equal to 0.5. To maintain an
implicit time-marching solution, a value of 1.0 was adopted for throughout this work.

Substituting equation 5.28 into 5.24 gives:

[LE ]{d }n t [ E ]{u}n = ([ E ]({u}n )1 + Q + [nE ])t (5.29)

Combining equations 5.13 and 5.23, the governing equations for finite element
consolidation analysis can be developed. These governing equations are presented as
Equation 5.30:

[K E ] [LE ] {d } n {RE }
[L ]T = (5.30)
E t [ E ] {u} n ([nE ] + Q + [ E ]{u}n )1 t

where [K E ] represents the element stiffness matrix, [LE ] represents the element volume
submatrix, [ E ] represents the permeability submatrix, {RE } is the vector of boundary
stresses, [nE ] represents flow due to gravitational forces, Q represents an internal sink

5.11
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

term, {u}n is the vector of nodal total pore pressure increments, {d}n is vector of
nodal displacement increments.

Global equations
The global system of finite element equations for the consolidation problem is simply
the assembly of terms from Equation 5.30, across the entire problem domain in
accordance with corresponding nodes.

[K G ] [LG ] {d } n {RG }
[L ]T =
G t [ G ] {u} n ([nG ] + Q + [ G ]{u}n )1 t (5.31)

where:

N
[KG ] = [K E ] (5.32)
i =1 i

N
[LG ] = [LE ] (5.33)
i =1 i

N
[G ] = [ E ] (5.34)
i =1 i

N
[RG ] = [RE ] (5.35)
i =1 i

N
[nG ] = [nE ] (5.36)
i =1 i

where N is the number of elements in the problem domain.

Extension to axi-symmetric conditions


To model the complexities of the stope drawpoint geometry would require the
development of a full three-dimensional version of the two-dimensional (plane strain)
model. This is beyond the scope of this thesis. The plane-strain Minefill-2D model is
capable of providing very good representation of the behaviour in many stopes,
especially where the length-to-breath ratio is high. However, in some situations, an axi-
symmetric model might provide a better representation of reality. In addition, Chapter 6

5.12
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

presents the results of an axi-symmetric centrifuge test, and clearly an axi-symmetric


model would be more appropriate for modelling this test.

All of the previous discussion relating to the development of Minefill-2D was based on
a plane-strain formulation. Axi-symmetric analysis follows the same calculation
methodology as that for plane strain conditions, but rather than analysing an element of
unit depth, the analysis is carried out on a one-radian slice through an axi-symmetric
geometry (i.e. through a cylinder) as illustrated in Figure 5.3. As is well documented in
the literature (Naylor et al., 1981, Potts and Zdravkovi ,1999, Smith and Griffiths,
1998), rather than calculations being carried out in terms of x and y coordinates (as is
the case in plane strain) calculations are carried out using a cylindrical coordinate
system (r, z, ).

As the calculations progress outwards from the centreline, the thickness of the slice is
always equal to the radius (since a one-radian slice is considered). In order to account
for the increase in thickness, a number of minor numerical adjustments are required.
These include the following:

The conversion of surface stress into nodal forces needs to be modified to

account for the increasing radius and therefore increased force applied to the

surface nodes with an increase in radius.

The terms in the stiffness matrix for both the undrained and consolidation

calculation steps need to be multiplied by the respective Gauss-point radius.

The volume submatrix [LE ] must be multiplied by the respective Gauss-point

radius to account for the increase in volume that occurs as the radius increases.

These modifications were made to Minefill-2D to allow the program to be used in either
plane-strain or axi-symmetric mode.

5.13
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

5.3.3 Boundary conditions

Initial conditions
The methodology adopted in Minefill-2D is to simulate the continuous placement of
material by activating discrete layers of material at a defined rate. The initial placement
of each layer involves an undrained step. As this step is assumed to occur over an
infinitely short time and as the material stiffness of each new layer is initially very low,
it is assumed that the layer being placed demonstrates completely undrained conditions
(i.e. within the new layer, the increase in total stress and pore pressure with depth are
equal).

The existing fill mass is loaded with a vertical stress that is equal to the weight of the
new layer. The response of the existing layer to this vertical force is calculated with an
undrained calculation step.

The reason for undertaking this undrained step is to provide a consistent match between
model geometry and applied self weight, and to independently establish the stress
distribution throughout the matrix in accordance with strain compatibility between the
water and soil phases.

The undrained loading step in Minefill-2D is undertaken using effective stress


parameters. When undertaking undrained analysis Equation 5.19 simplifies to:

[K E ]{d }n = {RE } (5.37)

or in global form:

[K G ]{d }nG = {RG } (5.38)

While the formation of the element stiffness matrix remains the same, the constitutive
matrix [D] is modified to [D ] , which accounts for the compressive stiffness of the
water phase through its bulk modulus Kw. Therefore, in an undrained situation the plane
strain stiffness matrix [D ] takes the form:

5.14
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

E (1 ) E
(1 + )(1 2 ) + K W (1 + )(1 2 )
+ Kw 0

E E (1 )
[D] = + Kw + KW 0
(5.39)
(1 + )(1 2 ) (1 + )(1 2 )
E
0 0
3(1 2 )

After the placement of each new layer, Equation 5.38 is solved to derive the nodal
displacements and, like with the consolidation case, Equation 5.4 is used to derive the
strains increment (). These strains are then used to determine the associated changes
in effective stress, total stress and pore pressure in accordance with

= D.
(5.40)

= D.
(5.41)

u = K w . (5.42)

As discussed previously, a fully-coupled analysis requires the pore pressures to be


calculated at the element nodal points rather than the integration points. Therefore, pore
pressures calculated at the integration points must be converted to nodal pore pressures
for the consolidation calculation phase. A number of stress-recovery techniques have
been proposed by researchers such as Zienkiewicz and Zhu (1992). These authors note
that if the variation of properties is linear across the element, then a straightforward
averaging technique would provide accurate results. As the flow calculations are carried
out using 4-noded elements, the shape functions vary linearly across the element and
linear interpolation is suitable. Therefore, in order to recover the nodal pore pressures,
the integration-point pore pore-pressure tensor is multiplied by the inverse of the shape
function to calculate the contribution of each integration point to the particular node.
The contributions of all integration points (from elements surrounding the particular
node) are averaged to recover the appropriate nodal point value. These values can then
be used during the consolidation calculation phase.

Phreatic surface control


During the deposition of saturated slurry material, the solids may immediately settle,
creating a free surface of water above the solids mass. Also, due to the bulk unit weight

5.15
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

being greater than that of water, undrained placement creates a hydraulic gradient that is
steeper than hydrostatic. This gradient causes upward water flow, which leads to water
ponding on the surface.

A surface water pond applies a total stress and (an equal) pore pressure at the fill
surface. The magnitude of the total stress and pore pressure is proportional to the depth
of the pond, and since they affect the total stress distribution throughout the fill mass
and the hydraulic gradients, it is necessary to accurately take account of the
accumulation of water above the fill surface.

In Minefill-2D, changes in surface ponding are accounted for by monitoring the water
exchange through the surface layer, in a manner almost identical to that employed in
CeMinTaCo (and in the original MinTaCo). The characteristics that are taken into
account in estimating the water exchange include flows through the surface layer,
volumetric changes that occur in the surface layer and any self-desiccation volumetric
changes. Based on the cumulative impact of these three mechanisms, the change in
phreatic surface elevation is determined.

During the placement of a new layer, any existing ponded water is transferred directly
to the surface of the new layer. Often when delivering mine backfill to a stope, the water
content required for transportation results in the solids settling (almost immediately) and
free water accumulating on the surface. If this is the case, any surplus water is added to
the existing surface water and consolidation calculations begin at what is defined as the
settled density.

All modelling described in this thesis was carried out assuming completely saturated
conditions. Therefore, the only relevance of the phreatic surface is that it defines a
surface on which the pore pressure is equal to atmospheric pressure. As a result, if base
drainage is occurring (without the addition of an equal quantity of water at the surface),
the phreatic surface would eventually be drawn below the fill surface. As discussed by
Fredlund and Rahardjo (1993), in order to maintain equilibrium, water above the
phreatic surface develops negative pore pressures that increase in magnitude with height
according to the unit weight of water. Full saturation assumes that the pore suctions at
the fill surface are always less than the air-entry suction of the fill, no matter how large
these suctions become. In real stope filling, desaturation can occur, though it is

5.16
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

generally where high cement contents are adopted, or it occurs late in the process, well
after the development of maximum barricade loads. Therefore, at this stage, extension
of the model to deal with unsaturated behaviour has not been attempted, but this is
certainly a desirable aspect that could be investigated in future work.

Boundary node control


Changes in boundary conditions occur during the transition from an undrained loading
calculation step to a consolidation calculation step. During this change in calculation
routine, the pore pressures at the controlled boundaries change from the value calculated
during the undrained step to that specified by the boundary condition. This change in
pore pressure can be accounted for by reassigning the nodal pore pressure values, but
this change in pore pressure must be reflected in a change in effective stress and
associated strains.

Also, as discussed in the previous section, the initial placement of material can create a
surface water pond or an upward hydraulic gradient which allows water to accumulate
on the fill surface. Any accumulated water would apply a total stress and pore pressure
to the surface nodes, with the magnitude being a function of the water depth. This
condition cannot be managed by eliminating the surface nodes from the calculation
scheme, as later they must be reintroduced into the calculation scheme after the
placement of the next layer.

In both cases, the initial and eventual nodal pore pressures are known, and therefore the
technique that is commonly adopted to apply known nodal displacement increments can
be utilised to address these issues. The following section presents how this logic can be
applied to the situation of changing nodal pore pressures along a boundary.

Assume the problem to be solved is:

k11 ... k1 j (
... k1n {u}1 .... + [G ] {u}ng )1t
...
... ... ... ... ...

ki1 ... kij ... k jn {u}j = .... + [G ] {u}ng
( )
j
t
(5.43)
... ... ... ... ... ...

k n1
... k nj

(
... k nn {u}n .... + [G ] {u}ng ) t
n

5.17
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

When progressing from an initial condition ( {u}Undrained ) to a condition specified by

the boundary condition ( {u}BC ) the change in pore pressure {u} j can be determined in

accordance with:

{u}j = {ucons } = {u}BC {u}Undrained (5.44)

This situation occurs when progressing from the undrained calculation step to the
boundary condition of atmospheric pressure at the barricade location or due to a change
in surface nodal pore pressure and total stress from a change in surface pond elevation.

Therefore, {u}j = {ucons } can be simply substituted into the jth row of the

consolidation matrix and all other terms in the jth row of the [k] matrix can be removed.
Also, as the value of {u}j is now known, it can be subtracted from both sides of the

equation such that:

k11 ... 0 ... k1n {u}1 .... + [G ] {u}ng( )1t k1 j .(.... + [G ]({ucons }) j t )

... ... ... ... ... ... ...

0 ... 1 ... 0 {u}j = {ucons }

... ... ... ... ... ... ...

k n1 (
... 0 ... k nn {u}n .... + [G ] {u}ng )n (
t k nj . .... + [G ]({ucons }) j t )

(5.45)

Given this procedure, the pore pressure change can be incorporated into the calculation
scheme such that it is appropriately reflected in changes to effective stress and strain in
addition to pore pressures. Furthermore, by maintaining the terms in the calculation
matrices, they can be included into the conventional calculation scheme when required.

5.3.4 Solution to the global equations

As will be described in Section 5.4, a non-linear constitutive equation is adopted to


represent the cemented mine backfill. In order to solve the non-linear constitutive
relationship, the visco-plasticity technique (Zienkiewicz and Cormeau, 1967) was
adopted. The associated numerical code is taken from Smith and Griffiths (1998).

5.18
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

The visco-plasticity technique allows the material to be loaded to a stress state that is
beyond the yield surface. If the yield function [Fy] is positive (i.e. yielding has occurred)
Q
this function is combined with the flow rule and viscosity parameter ( ) to

( )
determine the visco-plastic strain rate & vp in accordance with:

Q
& vp = .F y .
(5.46)

Ideally, should be determined experimentally, but as the main interest is in


determining steady-state stress and plastic strains, the transient stress path is not
important. It will be shown later (in Equation 5.48) that when determining stress states
and unbalanced forces, cancels, making the result independent of the chosen value.

The visco-plastic strain rate & vp is combined with a pseudo-time step (t cr ) to

determine the increment of plastic strain vp . The time step used in the time-marching
process (t) must be limited to maintain stability. For the Mohr-Coulomb yield surface,
which is adopted in Minefill-2D, it has been shown that the maximum stable timestep
(tcr) is given by:

2(1 2)
tcr =
(
G 1 2 + sin 2 ) (5.47)

where G is the shear modulus, is the Poissons ratio and is the friction angle.

The visco-plastic strains are combined with the material stiffness to evaluate the stress
increment .

Q
= D. .[F ]D tcr
(5.48)

where D is the elastic stiffness matrix.

In addition, the last term in Equation 5.48 is used to derive the unbalanced body
stresses. These body stresses are integrated over each Gauss point to determine
unbalanced body forces, which are redistributed to other nodes in the finite element
mesh during subsequent iterations. The solution is iterated until no Gauss-point stress

5.19
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

state violates the yield criterion (within a certain tolerance). The convergence tolerance
adopted throughout this thesis was 0.0001.

During each calculation step, it is assumed that the material properties remain constant
but these properties are updated after each increment of load or time. A flaw that exists
with this approach is the implementation of the strain-softening criteria into the visco-
plasticity solution scheme. Inherent in the visco-plasticity solution is the assumption of
perfect plasticity, and therefore body forces are calculated based on the assumption of
plastic strains not absorbing (or releasing) energy. But the degradation of material
strength with plastic strain does result in an energy release. Therefore, a rigorous
solution should take account of this energy release when calculating the body forces. By
updating the yield strength (taking account of this strength degradation) at the end of
each increment, and mapping the stress state back onto the yield surface during the
following timestep the energy release is, to some degree, taken into account. Detailed
and rigorous treatment of this aspect is beyond the scope of the thesis but future work
may consider applying a secant stiffness solution algorithm, which is better suited to
managing strain softening.

To minimise the likelihood of numerical drift, the maximum strain increment is


maintained below 0.001 in any calculation increment. If this tolerance is exceeded in the
undrained load calculation, the load increment is reduced, or if exceeded in the
consolidation calculation the consolidation timestep is reduced, and the calculation
repeated to ensure that this tolerance criterion is satisfied. Currently this process is
carried out manually, but future versions of Minefill-2D will combine the coefficient of
consolidation with the element size to derive a suitable timestep in a similar way to that
suggested by Yong et al. (1983) for the one-dimensional consolidation situation. This
approach would increase the efficiency of this process by maximising the timestep
increment without exceeding the strain criteria.

5.4 MATERIAL BEHAVIOUR

5.4.1 Influence of cementation on governing equations

Immediately after placement, and before the cemented mine backfill has reached initial
set, the material is assumed to behave in accordance with the uncemented material

5.20
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

properties. However, as demonstrated in Chapter 4 (using CeMinTaCo), the time-scale


associated with consolidation may be similar to that of the hydration process. As a
result, the cement hydration process must be fully coupled with the consolidation
process.

As discussed previously (in Chapters 3 and 4), the cement-hydration process influences
the consolidation behaviour. This influence was shown to be most significantly
associated with:

Stiffness: Aspects that can influence the material stiffness include the initial

uncemented density, cement hydration, and damage to the cement bonds due to

excessive strain. The evolution of these influences the constitutive matrix [D] ,

which influences the global stiffness matrix [K G ] .

Strength: This can be influenced by cement hydration as well as by destruction

of cement bonds due to excessive stress. Strength or yield stress influences the

constitutive matrix [D] and the global stiffness matrix [K G ] as it governs the

selection of stiffness properties.

Permeability: This can be influenced by material density, particle size

distribution and cement hydrate growth. These mechanisms interact to influence

the permeability matrix [G ].

Self desiccation: This refers to the volume changes that occur during the

hydration process (as discussed in Chapter 3) and can be taken into account

through the internal sink term Q.

Therefore, in order to take account of the cement hydration processes during the
consolidation process, the governing consolidation equations (Equation 5.31) are solved
such that the relevant terms are a function of time, material state (void ratio), and
cement content (t, e and Cc):

5.21
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

[K G (t , e, Cc )] [LG ] {d }nG {RG }


t.[G (t , e, Cc )] {u}nG ([nG (t , e, Cc )] + Q (t , e, Cc ) + [ G (t , e, Cc )]{u}nG )t

[L ]T =
G

(5.49)

Details of the cemented material relationships incorporated into Equation 5.49 were
discussed previously in Chapter 3, and are discussed further in the following section.

5.4.2 Constitutive model, [K G ([D(t , e, Cc )])]

The overall approach to the material model used in Minefill-2D is similar to that
described in Section 3.2, where the cemented and uncemented material behaviour is
superimposed to represent the overall response, and the small strain stiffness is linearly
related to the material strength while the secant stiffness is degraded in accordance with
the proximity of the loading surface to the yielded surface. Section 3.2 focused on a
model to represent one-dimensional loading, while the description that follows here
focuses on the response of the material under two-dimensional loading conditions. In
what follows, there is considerable repetition of material previously dealt with in
Chapter 3; this has been done for completeness, and to make it easier to follow the
developments.

In order to represent the behaviour of the material after placement but prior to initial
set of the cement, a power law was adopted to relate the void ratio to the applied
effective stress. This power law takes the form of:

b
e = ac ( ) c (5.50)

where e is the void ratio and is the mean effective stress, while ac and bc are curve
fitting constants. This function has been successfully applied to the compression
behaviour of uncemented mine tailings (Fahey and Newson, 1997 and Fahey et al.,
2002). Note that this differs from the Cam Clay approach to representing material
compression used in Chapter 3 (Equations 3.1 and 3.2). By differentiating Equation
5.50 and combining the result with well known elastic relationships, the uncemented
shear stiffness G(uncem) can be derived such that:

5.22
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

1
1
(e + 1) (e + 1) e bc

G(uncem ) = =
e 2(1 + ) 2ac bc (1 + ) ac
(5.51)

where is the drained Poissons ratio.

Once the initial set point is reached, it is also necessary to take account of the
influence of cementation on the behaviour. The material model used in Minefill-2D
assumes a Mohr-Coulomb failure criterion, where the size of the yield surface is
governed by the material state and degree of hydration. The shear stiffness is dependent
on the size of the yield surface and the mobilised stress relative to the yield stress. The
compressive stiffness is then related to the shear stiffness in accordance with an
assumed value of Poissons ratio. This is a much simpler approach than that used in
CeMinTaCo, where a model similar to the Structured Cam Clay model was used to take
account of possible yielding on the compression side of the yield surface, an aspect that,
as explained earlier, has not been incorporated into Minefill-2D.

In the Mohr-Coulomb yield surface, it is assumed that the friction angle is independent
of cementation, and during hydration, only the cohesive component evolves. This is
consistent with the findings of Clough et al. (1981) and Schnaid et al. (2001), who
suggest that cementation has little influence on the friction angle. The cohesive
component of strength increases as a result of cement hydration and decreases due to
damage in accordance with:

Hyd D
c = p
t S (5.52)

Hyd
where c is the change in the effective cohesion, is the change in c due to
t
D
hydration with time and is the degredation in c with time due to plastic shear
Sp

strain ( Sp ).

This differs from the approach taken in CeMinTaCo where, firstly, the damage term is a
function of the vertical compressive strain, and secondly, in addition to a damage term
degrading the cement contribution to the yield surface, a hardening term increases the

5.23
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

size of the uncemented yield surface. This hardening term is not included in Equation
5.52, but, as the cohesive strength is a function of the cement content and void ratio (in
accordance with Equation 5.53), any soil compression will manifest in an increase in
Hyd over subsequent timesteps.

Soil hardening is accounted for in the hydration function by calculating the hydration
component as a function of both void ratio (e) and cement content Cc, expressed as a
percentage. It was shown in Section 3.2 that the unconfined compressive strength (qu)
can be related to e and Cc by Equation (3.6). Since the friction angle is assumed to be
constant, cohesion (c) is linearly related to qu. This implies that a modified version of
Equation 3.6 can be used to relate Cc and e to c (in kPa). The modification involved the
introduction of another constant term (Ac) which is ratio between c and qu.

X .Cc + Cc 0.1 e
c = Ac exp
Z .C + W (5.53)
c

where Ac, X, Z and W are curve fitting constants.

As in CeMinTaCo, the exponential relationship proposed by Rastrup (1956) (Equation


3.7) is used to relate the cumulative evolution of cement hydration (or maturity, m)
against time. By differentiating Equation 3.7 and combining this with Equation 5.53 the
rate of development of cohesion with time can be represented. This function is

c 1 d d X .Cc + Cc 0.1 e
= . exp . A exp
t 2 (t *)1.5 t *
Z .C c + W

(5.54)

If the material is strained beyond yield, progressive breakdown of the bond strength can
occur. This mechanism has been accounted for by reducing c linearly as a function of
D
the plastic shear strain Sp (i.e the term in Equation 5.52).
Sp

The breakdown of cementation can be characterised using a triaxial test. An example


showing the stress-strain plot from a triaxial test on cemented paste fill is presented in
Figure 5.4 (a). Figure 5.4 (b) shows the assumed evolution of the Mohr-Coulomb yield
surface during this test.

5.24
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

Figure 5.4 (a) shows that as the material is loaded beyond yield, there is a progressive
reduction in the shear strength until it plateaus. The strength reduction is considered to
be a result of a progressive breakdown in the cement bonds until eventually all of the
bond strength is destroyed and the shear strength is purely a function of the frictional
strength and confining stress. The impact of cementation breakdown on the yield
surface is demonstrated in Figure 5.4 (b), where the gradient of the yield surface
remains constant (due to the assumption of a constant friction angle), but c is degraded
between the fully cemented and uncemented surfaces. The rate of breakdown of c is
assumed to be linearly related to the plastic shear strain according to the behaviour in
triaxial compression.

In Section 3.2.6, it was shown that the cement-induced component of stiffness can be
linearly related to qu. Therefore, the cement-induced component of stiffness can also be
linearly related to c, assuming a constant friction angle. This assumption is convenient
for modelling since, by evolving the cohesive intercept in accordance with Equation
5.52, the cement-induced component of stiffness can be linearly related to this value in
accordance with a constant rigidity term.

To represent the pre-yield response, a non-linear stiffness function was adopted. This
function degrades the material tangential stiffness linearly as the shear stress approaches
yield in accordance with:


Gt(cem) = G0(cem) 1 f mob (5.55)
max

where mob is the mobilised shear stress, max is the yield stress, Gt(cem ) is the cement
contribution to the tangential shear stiffness, G0(cem) is the cement component of the
small strain shear stiffness and f is a curve fitting constant.

If the uncemented soil contribution to the material stiffness is very low and f is equal
to 1, infinite strain is required to reach the ultimate shear strength (max). Thus the peak
is never reached and softening never occurs. For a model that has a similar drawback,
Fahey and Carter (1993) suggested the introduction of a constant term f, which if less
than 1, ensures that the material fails at finite strain. The actual value for f can be
derived from a triaxial stress-strain curve.

5.25
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

Figure 5.5 compares the adopted model with experimental data from local strain gauges
on a cemented backfill sample loaded in a triaxial compression. The comparison
suggests that for this particular material, the proposed model provides a reasonable
representation with f = 0.85.

Others (Fahey and Carter, 1993) suggest that the secant stiffness degrades in accordance
with a hyperbolic function while the tangential stiffness degrades in accordance with the
square of the hyperbolic function. While a more complex model may provide an
improved representation, the linear relationship was adopted because it provides
modelling convenience and appears to represent the experimental data reasonably well.

After the cement-induced component of stiffness is calculated, it is directly added to the


uncemented stiffness (from Equation 5.51) to determine an appropriate stiffness for the
cemented soil mass. The superposition of the cemented and uncemented properties
provides a convenient method of addressing the evolution from an uncemented material
to a fully-cemented material. If required, the approach is also suitable for simulating the
breakdown of the cementation due to excessive shear stress.

Modelling has assumed a constant Poissons ratio. Therefore, after an appropriate shear
stiffness is evaluated, this value can be used to derive the bulk modulus in accordance
with well documented elastic relationships. It is recognised that this approach neglects
the potential for strain localisation but with appropriately sized boundary elements, this
approach is considered to be reasonable for this thesis.

5.4.3 Permeability model, [G (t , e, Cc )], [nG (t , e, Cc )]

The permeability function adopted in Minefill-2D is the same as that presented in


Section 3.3. This model is a modified version of that originally suggested by Carrier et
al. (1983), but in this case the void ratio term is modified to account for both cement
hydrate growth as well as soil compression. This relationship was previously presented
as Equation 3.15 and is repeated below:

c k (eeff )
dk

k= (3.15 bis)
1+ e

5.26
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

5.4.4 Self desiccation, Q(t,e,C)

The process of self desiccation was addressed in Section 3.4, where a relationship
between the rate of volume change in an element Q (t , e, Cc ) , the cement weight per
element (Wc,), the efficiency of hydration (Eh) and a constant to represent the rate of
hydration (d) was presented. This relationship was implemented into Minefill-2D as an
internal sink term and is repeated below:

1 d d
Q(t , e, C c ) = E h .Wc . 1.5 .exp
(3.32 bis)
2 t* t
*

5.5 MODEL VERIFICATION

5.5.1 Comparison with analytical/numerical solutions

In the following section, Minefill-2D is compared with a range of analytical and


numerical solutions to verify its performance.

The first simulation involved an elastic, weightless, one-dimensional consolidation


problem, with 2-way drainage the most basic problem encountered in any
undergraduate textbook treatment of Terzaghis consolidation solution. This analysis
assumed a Youngs modulus of 100 MPa and a permeability of 1x10-6 m/hr. Figure 5.6
presents an illustration showing the problem geometry. The proportion of excess pore
pressure at the various elevations after 30 and 50 hours of consolidation are shown in
Figure 5.7. Also shown in Figure 5.7 is the well-known analytical solution to this
problem, for the same times. This demonstrates that the conventional consolidation
behaviour can be well represented by Minefill-2D.

To assess the performance of Minefill-2D with respect to self-weight consolidation, a


comparison between the Minefill-2D program and the commercially-available program
Plaxis (Vermeer and Brinkgreve, 1998) was undertaken. The development and
dissipation of excess pore pressures due to the deposition of a fresh layer of material
was simulated using these two programs. This analysis involved placing a 4 m thick
layer of material with a saturated unit weight of 19.5 kN/m3, a Youngs modulus of 100
MPa and a permeability of 1x10-6 m/hr. The problem is illustrated in Figure 5.8.

5.27
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

Immediately after placement, the base and top boundaries were set to atmospheric
pressure and consolidation was initiated. For the Minefill-2D case, the initial pore
pressure profile was determined using an undrained step (based on self-weight loading)
followed by the top and base boundary conditions being reset, in accordance the
technique described previously in this chapter.

Figure 5.9 indicates that the undrained calculation step (in Minefill-2D) provides an
accurate method of establishing the initial conditions, and the subsequent consolidation
calculations are consistent with results from Plaxis.

To assess the performance of the self-desiccation mechanism in Minefill-2D, the model


was compared with the analytical solution (Equation 3.27). The problem simulated was
what was referred to as a Hydration Test in Section 3.4.6. Modelling represented a
sample placed into a triaxial cell where the cell pressure was increased to 500 kPa with
the back pressure valves closed. The material properties included a void ratio of 1.05, a
cement content of 5%, an initial effective bulk modulus of 20 MPa and an ultimate bulk
modulus of 420 MPa. The rate of hydration was assumed to be in accordance with
Equation 3.7, with a hydration coefficient (d) value of -1.5 days1/2, and an efficiency of
hydration (Eh) of 6.4%. The initial set time was 12 hours. This example is illustrated
in Figure 5.10.

Figure 5.11 compares the variation in pore pressure (u) and the development of effective
stress () against time, for an element test as determined using Minefill-2D and the
analytical solution presented in Section 3.4.3 for this problem. The analytical solution
for undrained self desiccation (Equation 3.27) is presented as symbols while the
Minefill-2D results are presented as lines.

Figure 5.11 indicates that the self-desiccation component of the model is performing in
the appropriate manner. It should also be noted that cement-induced development of
stiffness against time for Minefill-2D was also compared with the measured results and
found to provide a match. The combination of these is evidence that the cementation
component of Minefill-2D is providing accurate results.

The final modelling carried out to assess the performance of Minefill-2D (against
analytical solutions) was a falling head permeability test. The purpose of this was to
ensure the algorithm controlling the elevation of the phreatic surface above the fill mass

5.28
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

is operating appropriately and also to ensure that flow calculations are performed
appropriately in terms of total water head.

Modelling involved establishing a 4-m thick saturated soil layer with 0.8 m of water
above the layer and atmospheric pressure specified along the base boundary. The
material was assigned a Youngs modulus of 11020 kPa (to ensure that volumetric
changes were minimal) and a permeability of 5x10-5 m/s. An illustration showing this
problem is presented in Figure 5.12.

Figure 5.13 shows a comparison between the elevation of the phreatic surface (above
the fill surface) calculated using Minefill-2D and that calculated according to Darcys
law, against time.

Figure 5.13 indicates that the change in phreatic surface elevation determined using
Minefill-2D is consistent with that calculated using Darcys law. This demonstrates that
the algorithm controlling the movements of the phreatic surface above the fill surface
and the component controlling flows due to gravitational forces, are providing accurate
results.

5.5.2 Comparison with CeMinTaCo

Because CeMinTaCo is a modification of the well tried tailings consolidation


program MinTaCo, which takes account of the influence of cementation on the
consolidation process, this provides an excellent basis to validate the performance of the
new finite element program Minefill-2D. Furthermore, because CeMinTaCo uses a
Lagrangian coordinate system and takes full account of yielding in one-dimensional
compression, this comparison provides an opportunity to investigate the significance of
these characteristics on the numerical results.

The material properties adopted for this comparison are those deemed typical for a
cemented paste backfill. These properties are presented in Table 5.1.

Using these properties, a one-dimensional filling situation was simulated, adopting an


impermeable base condition. This simulation involved the accretion of material at a
constant rate of 4 m/day for a period of 4 days. Figure 5.14 presents the development of
total vertical stress (v) and pore pressure (u) at a point 2.0 m above the base against
time.

5.29
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

The results presented in Figure 5.14 indicate that the pore pressure calculated using
CeMinTaCo is greater than that calculated using Minefill-2D. It is also clear that the
point where the curves representing the pore pressure and total vertical stress diverge
also differs for the two models.

Examination of the output from the two models indicates that during the very early
stages of strength and stiffness development, these material properties begin to change
(i.e. reach initial set) at a slightly different times in the two programs. In CeMinTaCo,
the cemented strength and stiffness are dependent on the difference between the
cemented and uncemented yield surfaces in one-dimensional compression, while in
Minefill-2D the cemented strength and stiffness are independent of the uncemented
yield surface. Therefore, in CeMinTaCo during the early stages of cement hydration,
any compression of the soil matrix hardens this material and softens the cemented yield
surface, effectively merging these surfaces. However, in Minefill-2D, hardening of the
uncemented soil simply acts to increase the density and therefore increase the influence
of cementation in the next timestep.

The overall influence of this behaviour is to effectively delay the initial set point, but
the calculated behaviour is essentially the same beyond this point. This is demonstrated
in Figure 5.15, which again presents the calculated pore pressure and vertical stress, for
the one-dimensional accretion of material at 4 m/day, but in this case the initial set point
for the Minefill-2D example has been delayed an additional 0.4 days. By making this
adjustment the results from the two programs are equal.

While this initial set point has been shown to have an influence on the result of the
modelling, and is therefore an aspect that should be addressed, it is suggested that
curing samples under effective stresses induced by self desiccation (as suggested in
Chapter 3), and adopting the measured initial set, Minefill-2D can provide a good
representation of the behaviour.

5.5.3 Stope mesh details

Element size
To investigate the influence of mesh size on the modelling results, and choose an
appropriate mesh size for the sensitivity study in Chapter 7, a convergence study was

5.30
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

undertaken. This study adopted typical paste backfill material properties and involved
identical simulations of the filling process using Minefill-2D in plane strain mode with
various mesh sizes. The stope geometry simulated was 10 m wide and 40 m tall with a 6
m high drawpoint, 5 m in length. Mesh sizes adopted in the analysis included:

A coarse mesh consisting of 5 elements horizontally across the stope width, 2


elements along the drawpoint length, 3 elements representing the drawpoint
height and 20 elements to represent the stope height. This mesh is presented in
Figure 5.16a.

A medium mesh consisting of 10 elements horizontally across the stope, 5


elements along the drawpoint, 6 elements representing the drawpoint height and
50 elements representing the stope height. This mesh is presented in Figure
5.16b.

A fine mesh consisting of 40 elements horizontally across the stope, 10 elements


representing the drawpoint length, 10 elements representing the drawpoint
height and 80 elements representing the stope height. This mesh is presented in
Figure 5.16c.

These meshes were used to simulate the same filling sequence, which consisted of
filling the stope at a constant rate of 0.5 m/hr for 12 hours, at which time filling was
suspended for 24 hours prior to a second filling at a constant rate of rise of 0.5 m/hr
from 6 m to the top of the 40 m high stope. In this analysis, all boundary nodes (apart
from those along the fill surface) were fixed against displacement in both the vertical
and horizontal direction. Nodes along the barricade boundary were maintained at zero
pore pressure.

The calculated pore pressure, at the centre of the stope floor is plotted against time for
each of the meshes in Figure 5.17, while the calculated barricade stress is plotted against
time for each mesh in Figure 5.18. These figures indicate that the coarse mesh produces
considerable different results to the other meshes, but the results from the medium and
fine mesh appear very close.

To investigate the influence of element size throughout the mesh the vertical total stress
(v) contours at the completion of filling for the coarse, medium and fine meshes are

5.31
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

shown in Figures 5.19 a, b and c respectively. Again, comparison of these images


indicates that the coarse mesh produces significantly different results to the fine and
medium mesh. Figures 5.19 b and c are very similar with the fine mesh indicating
slightly higher stresses than the medium mesh.

The accurate results with the medium sized mesh are due to the relatively low pore
pressure and effective stress gradients throughout the simulated geometry.

Considering the minor variation in results between the fine and medium mesh and the
significant difference in computational time, it is considered most suitable to adopt the
medium mesh in the remainder of the calculations in this chapter as well as the
sensitivity study presented in Chapter 7.

Interface behaviour
In Minefill-2D, the interface between the fill and surrounding rockmass was represented
using conventional elements where the boundary nodes (corresponding to the interface)
are fixed against displacement in both directions. The significance of deformation at this
interface is evident in Figure 5.19, where the v contours change sharply in direction
near the stope boundaries, indicating a stress discontinuity. Reducing the element size
reduces the region of influence such that it is difficult to identify shear planes in the fine
mesh. As shear planes are most likely to form immediately adjacent to the interface, the
boundary elements were reduced in size so that any yielding was concentrated at the
interface. Apart from the influence on the stress transfer to the surrounding rockmass,
the development of shear planes in these elements would have minimal influence on the
overall consolidation behaviour.

An alternative approach would have been to introduce interface elements to concentrate


any yielding along a defined plane. This may be considered in future developments.

5.5.4 Comparison with in situ measurements

To assess the ability of Minefill-2D to capture the important aspects associated with the
mine backfill deposition process, the model output was compared with in situ
measurements. In this analysis, the calculated pore pressure at the centre of a stope is
compared with actual pore pressures measurements during filling at site Paste Fill A

5.32
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

(PFA) in August 2007. This is the same material used in the sensitivity study presented
in Chapter 7.

The plan dimensions of the stope were 15 m x 18 m and 50 m tall. Using data collected
on the log sheets from the paste-fill plant, the volume of placed material was combined
with the actual stope cross sectional area (at various elevations) to determine the
elevation of the fill surface against time. This approach is considered valid as this
material typically does not experience significant volumetric changes during placement
and subsequent consolidation. The filling sequence consisted of filling the first 10 m at
a vertical rate of rise of 0.2-0.5 m/hr prior to a 24-hour rest period. After the rest period,
filling continued at a vertical rate of rise of 0.3-0.6 m/hr until the stope was filled. It
should be noted that the rate of rise adopted in the modelling was varied to match the
actual rate for each individual layer.

While the axi-symmetric geometry could have been more appropriate to represent the
stope geometry, the plane strain version of Minefill-2D was adopted to allow the
drawpoint to be represented. Based on a numerical comparison between an equivalent
axi-symmetric and plane strain stope, an 11 m wide plane-strain stope was deemed
appropriate to represent a stope of 15 m x 18 m plan dimensions.

Pore pressure (ucl) in the field was measured using a vibrating wire piezometer that was
installed on the floor of the stope prior to the commencement of filling. Readings were
taken on an automatic data logger at 2 minute intervals throughout the filling process.

The material used to fill the stope was PFA mixed with 3.1% cement and delivered at a
density of 75% solids by weight. Material properties used in the back analysis were
derived using a 1-D compression test, a hydration test and a triaxial test. Laboratory
testwork was actually carried out for PFA material at the same density (75% solids by
weight) but containing 3% cement, prior to field testing (Helinski et al., 2007). Due to
the close match of the tested material to the actual mix adopted, further testing was not
undertaken with 3.1% cement. The material properties adopted for the model are
presented in Table 5.2.

Figure 5.20 presents a comparison between the measured pore pressure and that
calculated using Minefill-2D. Also shown in Figure 5.20 is the total vertical stress (v)
that would be applied to the stope floor if no arching occurred.

5.33
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

Because of the low permeability and stiffness of the PFA material (in the uncemented
state) soon after placement, ucl initially increases at the same rate as v. This is an
indication that no consolidation is occurring, and that the fill mass is fully saturated and
it also provides confidence that the vibrating wire piezometer is fully saturated, and
measuring water pressures accurately.

Soon after the material reaches initial set, ucl diverges from v. However, as fill
continues to be deposited, and the increase in ucl due to the accretion of material is
greater than the reduction from consolidation, ucl continues to increase. After 35 hours,
filling stops, but as consolidation continues ucl decreases rapidly.

When filling resumes (at 59 hours), there is an increase in ucl, but as filling extends up
into the stope, some of the fill self weight is redistributed to the surrounding rockmass.
As a result, the incremental increase in v at the piezometer location reduces. This
results in ucl plateauing and then reducing at the later stages of filling.

Overall, Figure 5.20 indicates that using independently-determined material properties


Minefill-2D provides a very good representation of the pore pressure in the centre of the
stope floor during filling, which indicates that Minefill-2D appears to be representing
the consolidation behaviour of a cemented paste backfill accurately.

5.5.5 Investigation of the arching mechanism

A study was undertaken using Minefill-2D to investigate the significance of the arching
mechanism in a typical mine backfill scenario. Minefill-2D was used to simulate a 13 m
wide, 40 m tall plane-strain stope that was filled at a constant rate of rise of 0.4 m/hr.
The material properties adopted in this study where considered typical for a cemented
paste backfill and were the same for both scenarios analysed. These properties are
presented in Table 5.3. To investigate the significance of stress redistribution onto the
surrounding rockmass (arching), analysis was undertaken assuming boundary nodes
that were fully fixed in both directions (fixed BC) and boundary nodes that were fixed
against horizontal displacement but free to displace in a vertical direction (free BC).
The fixed-BC and free-BC cases are illustrated in Figure 5.21a and b, respectively.

The calculated pore pressure (u) and total vertical stress (v) on the stope centre line 1 m
above the stope floor are plotted against time during filling in Figures 5.22 for both

5.34
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

cases. Also shown in Figure 5.22 is the vertical total self-weight stress from the
overlying fill calculated assuming no arching.

Figure 5.22 shows that during the early stages of deposition both u and v are equal to
the vertical self-weight stress for both cases. This is because there is no consolidation,
and even with consolidation the ratio of fill height to width is insufficient to create any
noticeable stress redistribution away from the centre of the stope. After about 12 hours,
u diverges from v (indicating consolidation) but the ratio of consolidated fill height to
width is insufficient to create a stress redistribution. After 30 hours, 12 m of fill is in
place and, of this, the bottom 6 m has achieved a considerable amount of consolidation.
In the fixed-BC scenario, the rate of increase in v is less than that for the free-BC case,
because some of the applied vertical stress from the accretion of fill material is
redistributed to the fixed boundary (which represents stiff rockmass) as arching. It is
interesting to note that, even though the consolidation characteristics are the same for
both scenarios, the calculated u values also diverge at this point. This is because
arching, which is occurring in the fixed BC case, is reducing the amount of total stress
transferred to the stope floor during the accretion of additional material.

Finally, it is interesting to note that in the free-BC case, v is not equal to the total self-
weight stress, as would be expected in a true one-dimensional case. The reason for this
result is a stress redistribution that is occurring around the drawpoint opening. This
stress redistribution acts to spread some of the vertical stress around the drawpoint
opening, which actually reduces v in the centre of the stope floor. This will be
discussed in further detail below.

Contours of v at the end of filling for the fixed BC and free BC case are presented in
Figures 5.23a and b respectively. In addition, the v along the centreline for both cases
and that due to self-weight stress without arching is presented in Figure 5.24.
Comparison of these figures indicates that for the upper 10-15 m, the total vertical stress
for both scenarios is approximately the same. This is a consequence of reduced
consolidation in this area and an insufficient height-to-width ratio to generate arching.
But progressing further from the top of the stope, the contour plots are significantly
different. Figures 5.23a and 5.24 indicate that in the fixed-BC case, v (along the centre

5.35
Mechanics of Mine Backfill Matthew Helinski
Two-dimensional consolidation analysis (Minefill 2D) The University of Western Australia

line) remains relatively constant at approximately 200 kPa for the entire height of the
stope. This is an indication that significant arching is occurring.

Figures 5.23b and 5.24 indicates for the free-BC case no arching occurs throughout
the stope, resulting in an increase in stress with depth that is equal to the self-weight
stresses. Approaching the drawpoint, the rate of stress increase with depth reduces,
which is a consequence of arching around the relatively soft drawpoint opening. Even
though stress cannot be transferred vertically along the stope walls, a semi-circular
stress arch is established between the corner of the stope floor (opposite the
drawpoint) to above the drawpoint, as illustrated in Figure 5.23b.

This section has demonstrated that stress redistribution to the surrounding rockmass can
make a significant contribution to reducing the vertical stress in a typical mine backfill
stope. Two extreme cases were presented, but the trends are applicable for stopes of
different plan dimensions. As the stope plan area increases, the situation tends towards
the free-BC case,where little arching occurs, while a reduction in stope plan area would
cause the result to trend towards the fixed-BC case, or, in the case of even narrower
stopes than that presented, would promote additional stress redistribution leading to
even lower vertical stresses.

5.6 CONCLUSION

This chapter has presented the basis of the two-dimensional cemented tailings
consolidation program Minefill-2D. The program provides results that are consistent
with well-established analytical solutions to drainage-, consolidation- and cementation-
type problems. When compared with in situ monitoring data, Minefill-2D was shown to
provide a very good representation of the mine backfill deposition process. Overall, this
assessment provides confidence that Minefill-2D is performing appropriately and can be
used to represent the mine filling process.

Finally, a brief study was undertaken to investigate the arching mechanism in a typical
mine backfill situation. This study revealed that stress arching to the stiff surrounding
rockmass can make a significant contribution to reducing vertical total stresses in a
typical mine backfill stope.

5.36
Mechanics of Mine Backfill Matthew Helinski
Centrifuge Modelling The University of Western Australia

CHAPTER 6
CENTRIFUGE MODELLING

6.1 INTRODUCTION

A centrifuge modelling experiment was undertaken to demonstrate experimentally some


of the points that have been made in this thesis and to verify the performance of
Minefill-2D for representing the consolidation and arching processes in a cementing
backfill. In this experiment, a 3-D stope was represented by a cylindrical container i.e.
an axi-symmetric representation. This axi-symmetric geometry was chosen for practical
experimental reasons outlined later, but also because the axi-symmetric geometry could
be modelled numerically using Minefill-2D.

Centrifuge modelling (Schofield 1980) is an experimental technique commonly adopted


in geotechnical research. The concept behind centrifuge modelling is that by rotating a
scale model, the centrifugal forces increase the gravitational forces, increasing the stress
levels within the soil mass. By increasing the gravitational forces the model size can be
proportionally reduced while still creating an equivalent self-weight stress distribution.
In soil mechanics, it is important to reproduce similar stress conditions, as soil behaves
differently (particularly with respect to volume changes during shearing) under different
stress levels. In the language of centrifuge modelling, a centrifuge imposes an
acceleration level of N times the acceleration due to the earths gravity (g) i.e. an
acceleration of Ng. Provided the soil in the model is of the same density as that in the
prototype, an acceleration of Ng increases the unit weight by a factor of N, such that at a
depth z in the model, the vertical stress is the same as that at a depth Nz in the prototype.
The factor N is therefore thought of as being the length scaling factor of the model test.
Since the time-scale for consolidation depends on the square of the drainage path length,
the time for consolidation in the model is N2 times faster than in the prototype.

Therefore, the centrifuge is an ideal tool for investigating problems involving


consolidation. However, in this case, the interest is in consolidation and cement
hydration in mine backfill, where the time-scales of the two processes are similar, and
hence where these two processes are coupled. However, unlike for consolidation, the
time-dependency of cement hydration does not depend on any length scale, and hence

6.1
Mechanics of Mine Backfill Matthew Helinski
Centrifuge Modelling The University of Western Australia

the time for hydration in the prototype is not reduced in the model. Therefore, if the
time-scales of consolidation and hydration are similar in the prototype, they cannot be
similar in a centrifuge model in which the same materials are used, and therefore the
coupling between these processes cannot be studied directly in a centrifuge model.

The aim of the centrifuge experiment was to investigate the interaction between
consolidation and the total stress distribution for a consolidating soil undergoing
cementing in a model that represents an idealised stope. This test specifically focused on
how consolidation influenced the stress transferred through the soil mass and that
transferred as shear to the surrounding stiff container. This concept is fundamental to all
of the work discussed in this thesis.

The aim of this work had originally been to conduct an experiment that coupled the
time-dependent processes of loading, consolidation and cement hydration. Coupling of
these processes proved to be impossible due to the small scale of the model, as
explained above. Although the stress field could be scaled up in the experiment, the
consolidation time was still dictated by the actual drainage path length in the model.
This created very high hydraulic gradients that accelerated conventional drainage-type
consolidation, such that the time-scale for consolidation was very much less than that
for hydration. As a result, the material completely consolidated prior to the
commencement of hydration.

6.2 EXPERIMENTAL APPARATUS


The experimental apparatus is designed to capture the important aspects relating to the
distribution of stress around a stope during the placement of fill material. A schematic
showing a section view of the apparatus is presented in Figure 6.1.

The apparatus consists of a hollow cylinder (620 mm high and 180 mm in diameter with
an average wall thickness of 6 mm), machined on the inside to provide a rough interface
with the fill material. The cylinder is fitted with six axial and six hoop Wheatstone
bridge strain-gauge sets spaced at 100 mm centres along the vertical axis of the
cylinder. Each strain-gauge set consists of 4 gauges spaced at 90 intervals around the
cylinder circumference. The readings from each of the strain gauges in the set are

6.2
Mechanics of Mine Backfill Matthew Helinski
Centrifuge Modelling The University of Western Australia

averaged to provide the value at that particular elevation. The Wheatstone bridges are
completed in each case by external resistors.

The purpose of the strain gauges is to measure the hoop and axial strains in the cylinder.
As will be discussed later, the axial and hoop strains can be combined to determine the
axial and hoop stress in the cylinder. As the strain gauges are sensitive to any
temperature variations, each strain gauge site is fitted with a thermocouple for ongoing
temperature measurements. Photographs of the strain gauged cylinder are shown in
Figures 6.2 (a) and (b).

A floating base is inserted into the bottom of the cylinder. This base fits into a smooth
section of the cylinder to ensure a water-tight O-ring seal. The base rests on three
loadcells that measure the load carried by the base throughout the test. A drainage hole
is drilled into the centre of the base, with a filter placed immediately above this surface.
The drainage hole is connected to a pore pressure transducer to monitor the change in
pore pressure at the base location throughout the experiment. A photograph of the base
arrangement showing the floating base, loadcells and the base stand is presented in
Figure 6.3.

The purpose of the floating base resting on stiff loadcells is to provide a true measure of
the stress transferred vertically through the fill mass to the base. Due to the high
stiffness of the base loadcells, the displacement of the floating base is negligible,
relative to the soil stiffness. This boundary can therefore be considered as being rigid,
and the loads measured by the loadcells considered equal to those placed on a rigid
boundary, such as the base of a stope.

It would have been useful to also measure the stresses at discrete points within the fill
mass. However, the measurement of stress within a soil mass is very difficult, since this
generally requires inclusion of transducers with different relative stiffnesses. As
discussed in Sections 2.2, this is particularly problematic in cemented materials, where
the soil stiffness can become significantly greater than the diaphragm of an earth
pressure cell. These problems have been well documented by authors such as Clayton
and Bica (1993), and Take and Valsangkar (2001).

6.3
Mechanics of Mine Backfill Matthew Helinski
Centrifuge Modelling The University of Western Australia

During testing, the experimental apparatus was placed within an aluminium strong
box, which was mounted on the swinging platform of the geotechnical centrifuge.
Figure 4 presents a photograph of the apparatus in place on the centrifuge.

6.3 CALIBRATION
The process of calibrating the load cells and strain gauges attached to the cylinder walls
was carried out on the centrifuge, where known weights were accelerated to known
levels, to create a known force (or stress). Details of this calibration process are
presented below.

Base readings
All of the model testing discussed later was carried out at 100g. Therefore, the first
stage of calibration involved placing the empty cylinder on the centrifuge and
accelerating the centrifuge to 100g. The increment of strain measured on each
instrument when ramping up from 1g to 100g was due purely to the weight of the
apparatus itself. This value was deducted from all subsequent experimental results
recorded in the model tests. The empty apparatus was also subjected to different g-
levels to determine base-line readings to be deducted from calibration measurements at
different g-levels.

Loadcell calibration
For the calibration of the base loadcells, various weights were placed on the floating
base and the centrifuge accelerated to 100g. Knowing the weight and the radius to the
weight as well as the angular velocity, the force could be calculated. The force
increments were then used to form a linear correlation with the loadcell output.
Calibrating the loadcells using the centrifuge rather than at 1g meant that the same
logging system as that used in the experiment was used as well as taking into account
any friction loss between the O-ring and cylinder.

Strain gauges
One of the potential concerns regarding the strain gauge output was the influence of
temperature on the output. As only half Wheatstone bridge strain gauges were used in
either direction, all bridge outputs had to be adjusted for temperature changes. To
calibrate for the effect of temperature, the cylinder was filled with warm water (at a

6.4
Mechanics of Mine Backfill Matthew Helinski
Centrifuge Modelling The University of Western Australia

temperature of 40C) and left stationary (at 1g) overnight. During this period, the
cylinder temperature (as measured by the thermocouples mounted on the outer surface
of the cylinder at each site) and bridge output was logged. Due to the high thermal
conductivity of the aluminium cylinder, this temperature measurement was considered
representative of the average cylinder temperature.

Temperature measurements indicated that the water initially heated the cylinder to 40C
and overnight the cylinder temperature reduced to 20C. As the stress conditions
remained constant over this period the strain gauge half bridges could be calibrated for
temperature variations. This calibration was then used to correct the bridge output for
temperature variations throughout the calibration and testing stages.

The strain gauges used were manufactured from aluminium and as the cylinder was also
aluminium the temperature correction was minimal.

As strains can be linearly related to the strain gauge electrical resistivity (via a gauge
factor) and strains can be linearly related to applied stresses or forces the calibration and
the interpretation strategy adopted in the study was to simply relate the strain gauge
voltage output to an applied force or stress.

To calibrate the apparatus for axial force (F), a top cap was placed over the empty
cylinder and various weights were stacked onto this top cap. Once the weights were in
place, the centrifuge was accelerated to 50, 100, 150 and 200g to vary the axial force
applied to the cylinder. This process applied an axial force (F) without any change in
internal radial pressure (P). For calibration of radial stress the cylinder was filled with
water before applying 50, 100, 150 and 200g. After deducting the influence of the
cylinder self-weight stresses, this procedure provided an increase in radial pressure (P)
without changing the axial force (F).

When applying only an axial force (F) a voltage change was measured across the axial
and hoop bridges. If we denote the axial and hoop bridge voltage change as VAF and
VHA, respectively these can be related to F via calibration factors AF and HF in
accordance with:

V AF = F .( AF ) VHA = F .( H F ) (6.1)

6.5
Mechanics of Mine Backfill Matthew Helinski
Centrifuge Modelling The University of Western Australia

By accelerating the water-filled container an internal pressure (P) is generated which


induces axial and hoop bridge outputs of VAP and VHP respectively. After deducting the
apparatus self-weight voltage changes, P can be related to the axial and hoop bridge
outputs (VAP and VHP respectively) via calibration factors AP and HP in accordance with:

V AP = P.( AP ) VHP = P.(H P ) (6.2)

The changes in bridge voltages due to both an axial force and radial pressure are given
by:

V A = V AF + V AP VP = VHA + VHP (6.3)

Substituting 6.1 and 6.2 into 6.3

V A AF AP F
=
VH H F H P P (6.4)

Using this approach calibration factors AF, AP, HF and HP were derived.

To interpret the experimental results Equation 6.4 must be rearranged such that F and P
can be derived from the measured values of VA and VH. Inverting the matrix of
calibration factors and rearranging Equation 6.4 gives:

F 1 H P AP V A
= H
P AF H P H F AP F AF V H (6.5)

Therefore, the bridge outputs were simply adjusted for temperature variations and the
apparatus self-weight prior to direct substitution into Equation 6.5 to calculate the axial
force and radial stress throughout the experiment.

Pore pressure transducer


During the calibration with the water-filled cylinder, the standpipe pore-pressure
transducer was in place. From the calculated water pressure at the transducer location,
the calibration factor supplied with the transducer was verified.

The previously-described calibration routine provided suitable calibration factors for the
various instruments. This calibration yielded excellent consistency, providing
confidence in the performance of the data collection system.

6.6
Mechanics of Mine Backfill Matthew Helinski
Centrifuge Modelling The University of Western Australia

6.4 EXPERIMENT

Only one experiment was undertaken in this investigation. The aim of this experiment
was to investigate the interaction between consolidation and stress distribution in a
typical mine backfill situation.

6.4.1 Material

In a centrifuge model of thickness d representing a full-scale prototype of thickness D


(where d = D/N, and N is the acceleration multiplier), the time-scale for consolidation is
reduced by a factor N2 compared to the time-scale for the prototype. However, the
time-scale for hydration is unaffected by the g-level, so that the time until initial set, and
the total time for hydration, are the same for the model and the full-scale prototype.
Thus, if the time-scales for consolidation and hydration are similar in the full-scale
prototype, they are completely different for the model, assuming that the same material
is used in the model as in the prototype.

In an effort to prolong the consolidation time (with the aim being to investigate the
interaction of consolidation / stress arching and cement hydration), a material with a low
coefficient of consolidation was required, as the use of conventional mine tailings (even
a fine grained paste fill) would result in very rapid consolidation at the scale of the
centrifuge experiment. In order to prolong the consolidation rates, commercially-
available kaolin clay was adopted for this test.

The aim of the experiment was to investigate the interaction of the loading,
consolidation and cementation time-dependent processes. Therefore, cement was added
to the kaolin mix. Due to the high compressibility of the kaolin clay and the high water
contents required to achieve a flowable mix, 25% cement was required to achieve an
appropriate development of stiffness with time. Without sufficient cementation, it would
have been difficult to identify the influence of cementation on the overall consolidation
behaviour.

To maximise the strength gain for minimum cement content, and minimise the material
permeability, the water content was maintained at the minimum level that the mix
would be sufficiently free flowing to remove entrained air. The water content adopted

6.7
Mechanics of Mine Backfill Matthew Helinski
Centrifuge Modelling The University of Western Australia

was 62%, which corresponded to a placed void ratio of 2.20, assuming fully saturated
conditions.

As will be shown in this chapter, these high cement contents altered the coefficient of
consolidation such that consolidation was complete prior to the onset of initial set. Thus,
even with the use of the kaolin clay, the interaction of consolidation and hydration could
not be investigated in this experiment. Nevertheless, some interesting results are
presented.

6.4.2 Experimental procedure

The experiment involved filling the cylinder with the kaolin/cement/water mix in two
layers. The first of the layers was 250 mm high and the second 240 mm high.

Prior to the placement of the first layer, 5 mm of water was placed in the base of the
cylinder to assist with the removal of air during the filling process. The first layer was
placed in 5 sub-layers, each 50 mm high, with each layer being tamped 20 times to
ensure that all entrained air was removed. After placement of the first layer, water was
added above the material. This water filled the cylinder until it reached the overflow
valves at the top of the cylinder, where it was directed out of the strong box. The
water level in the cylinder was kept constant by continually adding water to maintain an
overflow, to make up for any water lost via evaporation. The high water level also
provided a back pressure within the soil, which assisted with ensuring full saturation
throughout the consolidation period. With the first layer in place, the centrifuge was
ramped up to 100g and this was maintained for 20 hours.

After 20 hours the centrifuge was stopped, resulting in a total stress reduction in the soil
at all depths. Initially the total stress reduction created an equivalent reduction in pore
pressure, and therefore there was no change in effective stress. But due to the short
drainage path, and increased material stiffness, stopping the centrifuge allowed the
negative pore pressures to increase to hydrostatic levels at 1g. This led to a reduction in
effective stress.

Should the total stress again be increased (through ramping up the centrifuge), the stress
distributed to the surrounding cylinder would be different to that before ramping down,
as there has been a change in shear stiffness at the soil / cylinder interface that was

6.8
Mechanics of Mine Backfill Matthew Helinski
Centrifuge Modelling The University of Western Australia

brought about due to cement hydration (this was demonstrated numerically by Rankin,
2004) . To remove the influence of any changes in stress distribution from the stress
cycling, the centrifuge was again ramped up to 100g where new baseline readings were
taken, which were subsequently used to determine the incremental change in stress due
to the application of the second layer.

A second layer of material (with the same mix proportions as the first) was then added
above the original material. This layer was 240 mm thick and was placed in the same
way as described for the first layer. The centrifuge was again ramped up to 100g and the
material was allowed to consolidate until equilibrium was achieved.

6.4.3 Experimental results

The results of the experiment have been divided into two sections, namely Stage 1
loading and Stage 2 loading. Stage 1 loading contains data collected during the
consolidation of the initial layer and Stage 2 loading contains data gathered during the
placement and consolidation of the second layer.

Stage 1 loading
During the placement of the first layer, the relationship between degree of consolidation
(as measured by the base pore pressure measurement) and the distribution of total stress
around the cylinder was investigated. These results are presented in Figure 6.5, which
shows the vertical force resulting from the soil and overlying water weight within the
cylinder (Total soil force), as well as the total force measured on the base load cells
(Base load cells force) and the cylinder (wall) axial load (Cylinder axial force)
plotted against time. Also plotted on the right axis of Figure 6.5 is the measured pore
pressure at the base of the cylinder (Base u), plotted against time.

Initially, after the material is placed and the centrifuge reaches the operating speed, the
measured pore pressure at the base of the cylinder was 600 kPa, which corresponds to
the total self-weight stress of the overlying material (assuming no arching). This
indicates that at this stage, the material is fully saturated and in an undrained state (i.e.
no consolidation has occurred). It can also be seen that initially the entire load is being
transferred through the saturated soil to the base loadcells, and no arching is occurring.
This is verified by the fact that no load is transferred axially through the cylinder (apart
from the cylinder self-weight).

6.9
Mechanics of Mine Backfill Matthew Helinski
Centrifuge Modelling The University of Western Australia

As consolidation takes place (indicated by the reduction in base u), there is a gradual
reduction in load measured on the base and an increase in the axial load on the
surrounding cylinder walls. This demonstrates that without consolidation there can be
little arching, but as consolidation occurs (even with a very low-strength clay material)
arching can create a significant amount of stress distribution onto the surrounding
cylinder.

At approximately 2.25 hours, the pore pressure at the base of the cylinder has reached
hydrostatic levels, and the system is therefore in equilibrium at this stage. As the pore
pressure reaches a constant value, there is no further change in effective stress and no
further change in any other measurements. This demonstrates that it is consolidation,
and not cement hydration bond strength, that is most important for arching.

Stage 2 loading
During the second loading stage, a 240-mm layer of slurry was placed over the original
layer. At the time of placement of the second layer, the original layer had been allowed
to cement to an unconfined compressive strength of approximately 300 kPa. The
incremental changes in pore pressure, base stress and axial stress that resulted from the
application of this second layer and subsequent ramping back up to 100g are presented
in Figures 6.6 and 6.7.

Figure 6.6 shows an incremental increase of 95 kPa in the pore pressure at the base of
the cylinder. The total vertical stress generated by the new layer was 105 kPa, so even
after developing a significant cemented strength, most of the applied load is still being
supported by the water phase and not being distributed to the surrounding cylinder
through arching.

Figure 6.7 shows the increment of force applied from the second layer (Applied force
increment) as well as the incremental change in force measured by the base loadcells
(Base loadcells force) and that measured as axial force in the cylinder (Cylinder
(wall) axial load). The results indicate that even with cemented material in the lower
part of the cylinder (which is expected to have qu in excess of 300 kPa), immediately
after placement almost all of the applied stress is transferred to the base. But as
consolidation takes place (which is indicated by a reduction in base u) force is
transferred off the base and onto the surrounding cylinder.

6.10
Mechanics of Mine Backfill Matthew Helinski
Centrifuge Modelling The University of Western Australia

Using the strain gauges at the #5 location, the total horizontal stress increment (h
#5) applied to the cylinder was monitored and is plotted, using the right hand axis, in
Figure 6.7. This strain-gauge set was installed 80 mm above the floating base, making
the location adjacent to the cemented layer (Layer 1). The measurements indicate that
immediately after the placement of the second layer, the increase in horizontal total
stress is 100 kPa, which is almost equal to the applied total vertical stress (105 kPa).
Therefore, even within the cemented mass, the horizontal stress increase is
approximately the same as the applied vertical stress if no consolidation takes place.
But, as consolidation takes place, h reduces significantly such that after complete
consolidation h reduces to approximately 30 kPa, or 30% of the applied total stress
increment.

6.5 NUMERICAL BACK ANALYSIS

As explained previously, it was not possible to conduct a centrifuge experiment in


which the time-scales of consolidation and hydration would be similar, and hence it was
not possible to study the entire interaction of mechanisms. Nevertheless, the experiment
has shown an interesting interaction between consolidation and the distribution of stress
around a stope-shaped container. The purpose of this section is to use the numerical
program (Minefill-2D) to simulate the consolidation behaviour, and replicate the
distribution of stress that was measured during the experiment.

6.5.1 Material characterisation

Consolidation in Stage 1 of the experiment was primarily conventional drainage-type


consolidation (rather than self desiccation), so back analysis of the results is most
sensitive to the consolidation characteristics of material in an uncemented state. For the
purposes of the numerical modelling, the one-dimensional consolidation properties of
the material were determined using a Rowe Cell consolidation test.

The Rowe Cell is a one-dimensional loading apparatus for measuring the consolidation
characteristics of soil. Load increments are applied via a flexible membrane, rather than
via a solid piston, as in a standard oedometer apparatus. With the Rowe Cell setup used
at UWA, the permeability can be directly determined at the end of each loading
increment. The test is carried out using one-way drainage, with the pore pressure

6.11
Mechanics of Mine Backfill Matthew Helinski
Centrifuge Modelling The University of Western Australia

response at the undrained boundary being measured directly. The test is performed
under elevated back pressure, which helps ensure full saturation (and also ensures fast
response in the base pore pressure transducer).

The material used in the element testing was identical to that used in the centrifuge
experiment. The material consisted of 75% commercially available kaolin clay and 25%
ordinary Portland cement, by dry weight, which was mixed to a water content of 62%.

Axial stresses of 220, 250, 300, 350, 400, and 500 kPa were applied with a constant
back pressure of 200 kPa. Throughout the consolidation period, the pore pressure at the
base (the undrained end of the sample) and the settlement of the sample were measured
to determine the confined modulus and permeability over a range of densities.

Direct permeability measurements were taken by establishing a hydraulic gradient


across the sample after completion of consolidation under axial effective stresses of 50,
100, 200 and 300 kPa.

The results of these tests are presented in Figures 6.8 and 6.9. Figure 6.8 presents the
relationship between applied vertical effective stress and void ratio, and Figure 6.9
presents the relationship between permeability and void ratio. Also presented in these
figures are the material relationships that were adopted in the numerical back analysis.
While the relationship between vertical effective stress and void ratio departs from the
experimental results at higher stress levels, over the range of vertical effective stresses
encountered in Stage 1 of the test (0-150 kPa) the relationship provides a good
representation. Note that normally a linear relationship in semi-logarithmic space (e
log v) would be used to describe such a relationship, but in this case the linear
relationship shown is more than adequate for the purpose required.

Stage 2 of the experiment involved the consolidation of both cemented and uncemented
layers. Therefore, the consolidation characteristics of Layer 1 (taking account of cement
hydration) at the time of the application of the second layer were required. To
characterise the cemented material properties, a hydration cell and triaxial test (after 77
hours of hydration) was carried out on kaolin with 25% added cement.

Using the experiments, the influence of cementation on the material behaviour was
determined. The relevant material properties are presented in Table 6.1. Figure 6.12

6.12
Mechanics of Mine Backfill Matthew Helinski
Centrifuge Modelling The University of Western Australia

presents the experimental measurements of small strain shear stiffness (Go) and
unconfined compressive strength (qu) as well as the relationship assumed in the back
analysis plotted against time. After 22 hours of hydration (the time when the second
layer is applied in the centrifuge test) the material is expected to have a Go of 180 MPa
and a qu of 300 kPa. In addition, an internal friction angle of 23 (Randolph and Hope,
2004) was adopted, assuming that the friction response was in accordance with that of
pure kaolin.

One aspect that complicates the derivation of appropriate cemented material properties
in Stage 2 is the significant density change that occurs due to conventional drainage-
type consolidation in Stage 1. Any density increase would lead to higher cement-
induced strength/stiffness than that measured in the element test where no (drainage-
type) consolidation took place. To address this complexity, the stiffness and strength
results were extrapolated to different densities using the exponential relationship
between void ratio and strength that was defined in Section 3.2.3, for Cannington Paste
backfill.

To account for the effect of cement hydration on the material permeability, the
relationship between void ratio and permeability for the uncemented material was
maintained, but the influence of cement hydrate growth was taken into account via the
effective void ratio term as described in Section 3.3.2.

6.5.2 Numerical back analysis

The program Minefill-2D, which was presented in Chapter 5, was used in axi-
symmetric mode for back analysis of the experiment results. The material relationships
are those described in Section 6.5.1.

Figure 6.10 presents a comparison between the measured base pore pressure and that
predicted using Minefill-2D during Stage 1. This figure indicates that the experimental
pore pressure reduction is only slightly more rapid than that calculated using Minefill-
2D. This provides confidence that Minefill-2D is providing an accurate representation
of the consolidation behaviour (albeit conventional drainage-type consolidation).

Figure 6.11 presents a comparison between the experimental and numerical results for
the distribution of vertical stress between the floating base and the surrounding cylinder.

6.13
Mechanics of Mine Backfill Matthew Helinski
Centrifuge Modelling The University of Western Australia

The experimental results are represented by solid lines while the numerical results are
represented by symbols. This comparison indicates that Minefill-2D provides a good
representation of the initial stress distribution as well as the redistribution of stress
during the consolidation period.

Initially, both experimental and numerical results indicate that without consolidation all
of the soil weight is transferred through the saturated soil to the floating base. But as
consolidation takes place, the force transferred to the base reduces and that transferred
through the surrounding cylinder (through arching) increases. The transfer of force off
the base and onto the surrounding cylinder continues as the pore pressure reduces (or as
consolidation takes place), but when the pore pressure plateaus, this force transfer stops.
It can be seen that the slightly faster consolidation rates measured in the experiment are
associated with a slightly faster increase in axial stress in the cylinder.

The calculated incremental change in pore pressure during Stage 2 is compared with the
experimental measurements in Figure 6.13. Again the calculated increase in pore
pressure due to the application of Layer 2 and the subsequent reduction in pore pressure
are well represented.

Figure 6.14 presents a comparison between calculated and measured incremental


changes in vertical forces acting on the floating base and transferred to the surrounding
cylinder when the centrifuge was restarted after Layer 2 was added. In addition, the
measured and calculated total horizontal stress at the strain gauge #5 level in the
cylinder is shown on the right axis. The measured results are presented as solid lines
while the calculated results are represented by symbols.

The calculated incremental change in total stress distribution slightly overestimates the
measured amount of arching. The overestimation of arching could have resulted from an
inappropriate relationship to represent the influence of density on cement-induced
strength and stiffness. This could have led to the model overestimating the actual
material stiffness, which would have promoted additional arching.

It should be noted that the bulk modulus of the material used in this experiment is at the
lower end of what would be expected in a real cemented mine backfill situation. Should
this stiffness be increased, the transfer of stress (through the cemented layer) due to the
undrained application of total stress would reduce in accordance with strain

6.14
Mechanics of Mine Backfill Matthew Helinski
Centrifuge Modelling The University of Western Australia

compatibility between the soil and water phases. Strain compatibility has been taken
into account when establishing the initial conditions in Minefill-2D to address this
aspect.

6.6 CONCLUSION
This chapter has presented a centrifuge experiment that was designed to investigate the
interaction of consolidation, cement hydration and the distribution of stress. However,
where the time-scales of consolidation and cement hydration are similar in a full-scale
stope (and hence these processed are coupled), the times-scales are very different in a
reduced-scale model, since the consolidation time is reduced (by a factor N 2), but the
hydration time is unaltered. Thus, it was not possible to devise an experiment where
these processes could be fully coupled. However, the results have experimentally
confirmed a number of key aspects relating to this thesis. These include:

The distribution of stress is heavily influenced by consolidation and largely


independent of the cement-induced bond strength.

Even in a material with an unconfined compressive strength of 300 kPa, if


saturated, the application of stress initially results in the load being supported by
the water phase, with very little arching occurring.

Even in a material with an unconfined compressive strength of 300 kPa, if


saturated, the undrained application of vertical stress results in an increase in
horizontal total stress of equal magnitude.

As consolidation takes place, the material is able to mobilise shear strength at


the soil/boundary interface and redistribute some of the vertical stress onto the
surrounding stiff medium.

Minefill-2D is capable of providing a reasonable representation of the


conventional consolidation and arching behaviour, using material parameters
measured in element testing.

6.15
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

CHAPTER 7
SENSITIVITY STUDY

7.1 INTRODUCTION

The preceding chapters have described the development of a rigorous numerical model
for simulating the mine backfill deposition process. This model is based on fundamental
material properties and, through comparison with well established analytical solutions,
was shown to provide a good representation of individual mechanisms. It was
demonstrated experimentally that the model is capable of accurately coupling the
interaction between consolidation and stress development in conditions similar to those
that are likely to be encountered in a stope filling environment. This provides
confidence that the Minefill-2D program is capable of providing a good representation
of the cemented mine backfill process and that the tool can be used with confidence.

This section uses Minefill-2D to assess the sensitivity of the overall filling response to
various characteristics. The aim of this work is to understand the backfill process and
provide strategies for managing the process that are based on sound logic. Sections 7.2
involves a comparison between hydraulic and paste fill. This is followed by two
sensitivity analyses, one on material that consolidates immediately after placement
(typically hydraulic fill) and another on material that does not consolidate immediately
after placement (typically paste fill). Throughout the sensitivity analysis, emphasis is
placed on the resulting barricade loads, but the final section (Section 7.5) considers the
effect of the application of effective stress to in situ material during curing.

7.2 COMPARISION OF HYDRAULIC FILL AND PASTE FILL

In the mining industry, slurry mine backfills are commonly divided into two main
groups, paste fill and hydraulic fill. As Minefill-2D undertakes analysis using
fundamental material properties, it provides an opportunity to investigate the
relationship between paste and hydraulic fills without introducing simplifications that
pre-empt the final outcome.

When considered from a fundamental soil mechanics point of view, hydraulic fill and
paste fill are essentially the same product. The main difference is that hydraulic fill

7.1
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

generally contains less fines than paste fill. The typical consequences of this difference
are highlighted in Table 7.1

In order to investigate the significance of these characteristics on the overall filling


response, an experimental and numerical study was undertaken. This involved testing 4
different cemented mine backfills using the (previously described) hydration test,
triaxial test and Rowe Cell test to determine both cemented and uncemented material
properties. Testing was carried out on two hydraulic fills and two paste fills. Hydraulic
Fill A (HFA) is from a zinc mine, Hydraulic Fill B (HFB) is from a copper mine, Paste
Fill A (PFA) is from a gold mine tailings and Paste Fill B (PFB) is from a nickel mine.

7.2.1 Experimental results

The experimental program commenced with particle size distribution analysis of each
tailings specimen. The results of this analysis are presented in Figure 7.1. As expected,
Figure 7.1 indicates that the hydraulic fill materials have far less fines than the paste
fills.

The aim of this investigation is to assess how the tailings characteristics influence the
filling behaviour. Therefore, in order to maintain consistency, testing was carried out
with each of the materials combined with 3% ordinary Portland cement and each mixed
with water to achieve an equivalent void ratio of 0.9, assuming fully saturated
conditions. Each mix was subject to a hydration test to determine the cement hydration
properties (as discussed in Chapter 3) followed by a triaxial test to determine the
strength properties. Rowe Cell testing was also undertaken on the tailings material
(without cement) to determine properties that represent the behaviour prior to the onset
of cement hydration.

A summary of the material properties for the various fill types is presented in Table 7.2.

From Table 7.2, it can be seen that the final unconfined compressive strengths (qu-f) are
very similar for the four samples, even though the materials have different rates of
hydration (d), efficiency of hydration (Eh), and hydraulic conductivity parameters (ck
and dk). Figures 7.2 and 7.3 show how the hydraulic conductivity (k) and cohesion (c)
evolve with time, respectively.

7.2
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

7.2.2 Modelling

In order to assess the impact of the various material properties on the filling process,
Minefill-2D was used (in plane strain mode) to simulate the filling of a plane strain
stope 20 m wide and 40 m high. A drawpoint height of 5 m was adopted with a
barricade offset distance of 5 m. A boundary condition of atmospheric pore pressure
was assigned along the boundary that represents the barricade. To allow direct
comparison between the various materials, a standard filling sequence was adopted.
This sequence consisted of filling the first 8 m over a 16-hour period (0.5 m/hr)
followed by a 14-hour rest period, and then filling the remaining 32 m over a period of
64 hours (0.5 m/hr).

In an actual stope, the drawpoint width is typically less than the side length of the stope,
whereas in the plane strain representation it occupies the full side length. This means
that the actual drawpoint represents a greater choke to outflow than the plane strain
representation. In order to account for this, the hydraulic conductivity in the drawpoint
area was halved.

Modelling was undertaken using the material properties presented in Table 7.2 to assess
the impact of tailings type on the consolidation behaviour and on the resultant barricade
stresses.

Figure 7.4 shows a plot of the total horizontal stress, developed at a point immediately
behind the barricades, for the various fill materials, using the described filling sequence.
This indicates that barricade stress reaches 108 kPa for PFA, 150 kPa for both HFB and
HFB and 240 kPa for PFB. Thus, even with fills that reach the same ultimate strength
(qu-f), stresses applied to barricades can vary significantly.

Furthermore, there is no obvious relationship between any one material property and the
resulting barricade stresses. For example, barricade loads for PFA are the lowest, even
though the hydraulic conductivities of HFA and HFB are higher, while that of PFB is
lower. Also, HFA and HFB show a faster rate of hydration (lowest d value) and a higher
efficiency of hydration (highest Eh value), when compared with the paste fill materials,
but ranks in the middle in terms of ultimate barricade stress.

7.3
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

The reasons for the stress variation may be better understood with reference to the
development of pore pressure at the opposite side of the stope to the barricade. This is
presented in Figure 7.5 along with the steady state seepage pore pressure that is
created when the water table is maintained at the fill surface and atmospheric pore
pressures are maintained at the barricade boundary. This is calculated in accordance
with the relationship presented by Helinski and Grice (2007). Pore pressures greater
than steady state indicate that the material has not fully consolidated, while pore
pressures equal to this value indicate that the material has completely consolidated.
This is a somewhat artificial situation, since ongoing water flow without further
addition of water to the stope would result in lowering of the water table (with the
possibility of further consolidation as the equilibrium situation changes) but it does
provide a reference for assessment of excess pore pressures.

Comparison between Figures 7.4 and 7.5 indicates that there is a clear relationship
between the development of pore pressure and the stresses placed on barricades. This is
logical, as higher pore pressures are associated with less consolidation and lower
effective stress. As discussed in Chapter 2, with less effective stress, less interface shear
strength is mobilised resulting in higher total vertical stresses in the stope. In addition,
the conversion from total vertical stress to total horizontal stress adjacent to the
drawpoint is dependent both on the proportion of the load being carried by the soil
skeleton (effective stress) and that carried by the pore water (pore pressure), in
accordance with Equation 7.1:

h = v .K o + u (7.1)

where h is the total horizontal stress, v is the vertical effective stress, K o is the

lateral earth pressure coefficient ( 0.3-0.5) and u is the pore pressure.

Therefore, if all of the total vertical stress is supported by the water phase ( u = v ) the
vertical and horizontal total stresses would be equal, but if the pore pressure is zero, the
horizontal total stress would approximately 30% to 50% of the total vertical stress.

The other aspect to note about Figure 7.5 is with respect to the development of pore
pressure (for the different mine backfill materials) relative to the steady state seepage
pore pressure. Comparing the pore pressures for the four cases with the steady state

7.4
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

seepage pore pressure, it is clear that both HFA and HFB hydraulic fills exhibit close to
steady state seepage pore pressures throughout the filling process, indicating immediate
consolidation, PFB develops pore pressures that are greater than steady state seepage
pore pressures indicating that consolidation is not complete, and PFA exhibits pore
pressures that are even less than steady state seepage pore pressures. This may be more
clearly depicted in Figure 7.6, which shows the pore pressure isochrones along the stope
centre line at the completion of filling. Also shown in Figure 7.6 is the line representing
the steady state seepage pore pressures for the stope.

The reason for the three different types of behaviour is as follows:

HFA and HFB


Due to the initial high value of the coefficient of consolidation (i.e. higher permeability
and stiffness compared to the paste fill materials), excess pore pressures dissipate
immediately and the pore pressures in the fill mass are the steady state seepage pore
pressures resulting from the reduced flow area in the drawpoint. It is also interesting to
note that the efficiency of hydration for the HFA hydraulic fill is double that of the HFB
hydraulic fill but the pore pressures and barricade loads are almost identical. This
suggest that this mechanism plays little role in the consolidation of hydraulic fills.

PFA
For PFA, the initial low stiffness and low hydraulic conductivity result in very little
conventional drainage-type consolidation prior to initial set. Close inspection of
Figure 7.5 indicates that during the early stages of filling, pore pressures in PFA are
higher than in HFA and HFB. This is reflected in higher barricade loads during this
period. However, this material has a high propensity for self desiccation after initial
set, where the water volume reduction from self desiccation combines with the rapidly
increasing material stiffness to reduce the pore pressures. For higher permeability
materials, this would be counteracted by an inflow of water from above that would
restore steady state seepage pore pressures (as shown for the hydraulic fills), but for
PFA, the permeability is so low that volumes being consumed by the self-desiccation
mechanism are greater than the inflow from above. Therefore, the very steep hydraulic
gradients being generated by the low pore pressures can be maintained. This suggests

7.5
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

that very low pore pressures, below the steady state line, may be produced by self
desiccation in association with low hydraulic conductivity.

PFB
Like PFA, the initial value of the coefficient of consolidation of PFB is very low (i.e.
low permeability and high stiffness), and therefore drainage-induced consolidation is
insufficient to dissipate excess pore pressures. However, unlike PFA, the prosperity for
self desiccation is too low to give a significant pore pressure reduction, with the result
being significant pore pressure development. Also, even though the material is gaining
stiffness, the low permeability is preventing any conventional dissipation of excess pore
pressures. The resulting high pore pressures (low levels of consolidation) are reflected
in high barricade loads.

This situation seems to be most prominent when the tailings being used to form the
paste have high active clay content. Clay particles reduce the coefficient of
consolidation (suppressing conventional consolidation) and have also been shown to
adversely influence the cement hydration process. This is the case with PFB.

7.2.3 Comparison of hydraulic fill and paste fill

This analysis has identified two significantly different responses depending on whether
the material undergoes immediate consolidation or if very little consolidation takes
place prior to the onset of hydration. Broadly, it may be assumed that hydraulic fills
consolidate immediately while paste fills require cement hydration to achieve
consolidation, but this outcome will be dependent on both the coefficient of
consolidation and the filling rate. Therefore, both of these factors must be taken into
account when characterising the expected behaviour. One method of characterising the
combination of material properties and filling rate is through Gibsons (1958) analytical
solution that was introduced in Section 2.2.

The following two sections describe sensitivity studies that investigate the behaviour of
firstly consolidating fills and secondly non-consolidating fills.

7.3 CONSOLIDATING FILL


This section refers to fill types that when first deposited, at the specified rate of rise,
have a coefficient of consolidation that is sufficient to dissipate any excess pore

7.6
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

pressures while filling is in progress. This characteristic is most commonly associated


with hydraulic fills, but is equally applicable to coarse paste fills that are placed with
low rates of rise. The analysis carried out to investigate the sensitivity of consolidating
fills has used the HFB parameters, which were presented in Table 7.2. In each case the
stope geometry simulated was a 40 m tall, 13 m wide plane-strain stope, with a 5m long
and 5 m high drawpoint. The 13 m wide plane-strain stope was selected as this
dimension is expected to provide a similar vertical stress distribution to a stope with
plan dimensions of 20 m x 20 m, which is considered typical, and the plane-strain
configuration allows the drawpoint to be represented. In most consolidating fill
situations, fill rates are often much slower than 0.5 m/hr, which was adopted in Section
7.2 for direct comparison with paste fills. Therefore, in order to provide more realistic
results, the analysis in this section is carried out using a constant fill rate of 3 m per day
or approximately 0.125 m/hr.

7.3.1 Influence of stope geometry

Stopes often have different configurations with regard to plan dimensions and the
number and size of drawpoints. Different configurations will lead to differences in the
restriction to flow at the base of the stope. This section is focused on investigating the
influence of the drawpoint restriction on barricade loads.

To represent different drawpoint restrictions the stope geometry has been maintained
constant but the drawpoint permeability was modified by an order of magnitude. This
modification has the equivalent impact as modifying the number of drawpoint openings
or the installation of drains through this region. In addition, to demonstrate the extreme
situation, an impermeable barricade was simulated.

The development of pore pressure at a point 2 m above the stope floor on the opposite
side of the stope has been plotted against time for each of these cases, in Figure 7.7.
Figure 7.7 indicates that the pore pressure at the opposite side of the stope to the
drawpoint is heavily influenced by the restriction created at the base of the stope. It
should be noted that the reduced pore pressures (at the base of the stope) coincide with a
significant change in pore pressure distribution throughout the stope rather than a
significantly different phreatic surface elevation. This point is demonstrated in Figure
7.8, which presents the pore pressure isochrones along the centre line of the stope for

7.7
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

the different cases. It can be seen that the phreatic surface elevation is effectively the
same for each of the cases, but the rate of pressure increase (with depth) is significantly
reduced as the restriction to flow through the drawpoint reduces (or the drawpoint
permeability increases).

The impact of the different drawpoint restrictions on barricade stress is illustrated in


Figure 7.9, which presents the total horizontal stress placed on the barricade against
time for the different cases.

Figure 7.9 indicates that the reduction in pore pressure and increase in effective stress
associated with the reduced drawpoint restriction significantly influences loads applied
to consolidating fill barricade structures. Even with the same phreatic surface elevation,
the different drawpoint restrictions can have a significant influence on the distribution
of total stress and hence on barricade loads. This is most significant in the extreme case
of the impermeable barricade. In this case there is no water flow and the resulting
hydrostatic pore pressures leads to very high barricade stresses.

This result is consistent with the analytical analysis results published by Kuganathan
(2002), who also suggested that the installation of engineered drainage systems in the
stope drawpoint can minimise barricade stresses.

7.3.2 Influence of permeability

To investigate the significance of changes in the permeability of a consolidating fill


material, the HFB permeability was modified by an order of magnitude in both a
positive (k=10kHFB) and negative (k=0.1kHFB) direction. It should be noted that,
with the change in permeability, the coefficient of consolidation remained sufficiently
large to ensure that there was no build up of excess pore pressures during filling.

Figure 7.10 indicates that, until approximately 220 hours, the material permeability has
little influence on the pore pressure that develops in a stope. At 220 hours, the high-
permeability material allows the phreatic surface to fall below the fill surface, which
leads to reduced pore pressures. But for the same phreatic surface elevation, the pore
pressures are independent of the permeability. The reason for the independence is that,
provided the permeability is sufficient to dissipate the build up of excess pore pressures,
in situ pore pressures would be dictated by the relative flow resistance between the

7.8
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

stope and drawpoint. Provided the relative flow resistance (between the stope and
drawpoint) is constant, for a given water table elevation, the pore pressure profile will
be similar.

Should the permeability be further reduced, there is potential to accumulate excess pore
pressures, which would change the deposition behaviour. This is discussed in Section
7.3.4.

The impact of fill mass permeability on barricade stress is presented in Figure 7.11,
which shows the development of barricade stresses against time for the various cases.
As with the pore pressure, it can be seen that for the same phreatic surface elevation the
barricade loads remain largely independent of permeability.

It is interesting to note that, for the k=10kHFB case, when the phreatic surface falls
below the fill surface at approximately 220 hours, there is an associated significant
decrease in barricade stress. The reduction in pore pressure is associated with an
increase in effective stress, which acts to mobilise more shear stress (within the fill
mass) and reduce the total stress transferred to the barricade.

7.3.3 Influence of cementation

The following sensitivity study investigates the influence of cementation on the


barricade loads in a consolidating fill. Again the HFB material is adopted but the
cement content is varied between 0%, 1.5%, 3% and 8%.

Figure 7.12 shows the development of pore pressure at the base of the stope (on the
opposite side of the stope to the barricade) against time for the different cement
contents. Based on the results, it appears that the development of pore pressure against
time in a consolidating fill is largely independent of cement content. This is expected to
be due to the high initial coefficient of consolidation, which rapidly creates and sustains
steady state conditions (as discussed in Section 7.2.2). Consequently cementation is
not required in the consolidation process.

Figure 7.13 presents the development of barricade stresses in a consolidating fill mass
against time with varying cement contents. It can be seen that between cement contents
of 3 and 8 %, the presence of cementation has little influence on the calculated barricade
stresses, but as the cement content drops to 1.5 %, there is an increase in barricade

7.9
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

stress. The barricade stress for the 1.5% case is similar to the 0% case. This increase in
stress is a result of the weaker material (associated with the lower cement contents)
yielding at the fill/rock interface. To demonstrate this point, Figure 7.14 presents the
calculated shear stress and cohesive strength against time for an element at the interface
between the fill and the rockmass, 7 m above the stope floor. This result indicates that
with 3% cement, the cohesive component of strength is sufficient to support the applied
shear stress, and no softening occurs. But in the 1.5% cement case, there is some strain
softening and subsequent breakdown of the cementation at the fill/rock interface, which
is indicated by the reduction in cohesion from approximately 80 hours onward. Another
illustration of this cementation breakdown is shown in Figures 7.15 (a) and (b), which
show contours of cohesion for the 3.0% and 1.5% cases, respectively. Figure 7.15 (a)
shows relatively constant cohesion horizontally across the stope while Figure 7.15 (b)
shows a dramatic reduction in cohesion at the fill/rock interface.

A consequence of this softening is that stress that was previously supported in shear by
the boundary element is redistributed, increasing the vertical stress within the fill mass,
and therefore increasing the barricade stresses. The influence of strain softening on the
vertical total stress is illustrated in Figure 7.16 (a) and (b), which show vertical total
stress contours at the completion of filling for the 3 % cement and 1.5 % cement cases,
respectively. Comparison of these figures indicates that strain softening (associated with
the 1.5% case) results in a 150 kPa increase in vertical total stress at the base of the
stope.

This increase in total vertical stress does not induce an associated increase in pore
pressure, since the coefficient of consolidation is sufficient to dissipate any excess pore
pressures that are created, but this increase in total vertical stress creates an associated
increase in horizontal effective stress, which contributes to barricade stresses.

In addition to the implications of interface strain softening on barricade stress, this


mechanism should also be considered when assessing fill exposure stability (as
discussed in Section 2.2.3). For example, the assumption of a fully cohesive fill/rock
interface (Mitchell and Wong 1982) would not be valid in the 1.5% cement example,
and instead a friction-only interface should be considered when undertaking exposure
stability analysis for this case. By reducing the filling rate, such that the development of

7.10
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

shear strength exceeded the application of shear stress, it could be possible to avoid
yield at the interface, which would result in increased fill stability at the time of
exposure.

7.3.4 Influence of filling rate

The following section presents an investigation into the influence of placement rate on
material that might be considered to be consolidating fill. Again the HFB parameters
(from Table 7.2) have been adopted, but the rate of fill rise is varied from a constant
0.06 m/hr to 4 m/hr. The development of pore pressure is presented against time for
each of the cases in Figure 7.17.

Figure 7.17 indicates that the maximum attained pore pressure for all filling rates up to
0.6 m/hr is relatively constant at 200 kPa. However, for filling rates of 2 m/hr and 4
m/hr, pore pressures reach a higher maximum before reducing to around the same
constant value. These higher pore pressures are excess pore pressures resulting from the
higher filling rates.

Figure 7.18 presents the calculated barricade stresses against time for the different
filling rates. As with the development of pore pressure, the maximum barricade load
remains relatively constant (at approximately 130-140 kPa) up to a filling rate of 0.6
m/hr. But with the filling rates of 2 m/hr and 4 m/hr the peak barricade stress is
significantly greater. The higher barricade stresses are a result of the excess pore
pressures, which change the loading mechanism.

7.3.5 Consolidating fill: discussion

The modelling results presented indicate that aspects such as drawpoint restriction,
cement content and filling rate can influence loads applied to barricade structures for
consolidating fill. But, apart from the drawpoint restriction, barricade stresses appear
relatively constant over a range of cement contents and filling rates. It was demonstrated
that if these factors vary beyond a given thresholds, there is a step-change, where the
deposition behaviour is modified. Specifically, the change in mechanism involves strain
softening at the rock/fill interface and the development of excess pore pressures. The
discrete change in behaviour results in a change to barricade stresses. Therefore, when
estimating barricade stresses with consolidating fill, it is first necessary to define if the

7.11
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

fill/rock interface is likely to yield during deposition and secondly to define the rate of
rise that would allow excess pore pressures to develop. Provided that excess pore
pressures are not generated and the calculation of vertical stress within the stope takes
account of the fill/rock interface behaviour, barricade stresses should remain relatively
independent of these factors.

Section 7.3.4 showed that steady state seepage pore pressures are dependent on the
elevation of the phreatic surface and the restriction to flow through the drawpoint.
Helinski and Grice (2007) showed that the restriction to flow through the drawpoint can
vary significantly between stopes due to the presence of macro pores, which can
dominate the drainage behaviour through the drawpoint region.

Comparison between Figures 7.8 and 7.9 indicates a trend between barricade stresses
and the pore pressure on the floor of the stope opposite the drawpoint. To examine this
relationship, the calculated pore pressure was plotted against the barricade stress in
Figure 7.19, for each of the cases analysed in Section 7.3.1. This figure indicates a
unique relationship between barricade stress and this pore pressure. The significant
dependence on pore pressure is problematic from a design point of view (as it is
difficult to accurately predict pore pressures), but since it is very easy to measure
positive water pressure accurately, these measurements can be used to efficiently
manage filling activities.

For example, based on an assumed drawpoint restriction, the required barricade capacity
can be estimated. As the most significant assumption in the design is the drawpoint flow
restriction, pore pressure measurements can be used in operation to manage filling
operations such that the actual pore pressures do not exceed those assumed in the
design. For example, consider the case presented in Section 7.31. If the design assumes
a drawpoint resistance that is equivalent to kdp = kst, an ultimate barricade stress of 118
kPa would be expected. Applying the relationship between pore pressure and barricade
stress presented in Figure 7.19, filling should be undertaken such that the pore pressure
at the opposite side of the stope floor to the barricade does not exceed 196 kPa (the
estimated peak pore pressure for this condition). If the pore pressures increased at a rate
faster than expected, filling should be suspended prior to the pore pressures reaching
196 kPa. Filling would then continue, ensuring that this target is not surpassed.

7.12
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

Alternatively, should pore pressures remain low, filling could be accelerated, ensuring
that excess pore pressures are not generated. The relationship presented in Figure 7.19
can be used along with the peak pore pressure to estimate the ultimate barricade stress.

7.3.6 Consolidating fill: conclusion

Based on the consolidating fill sensitivity study, it can be concluded that the important
aspects to consider in determining barricade stress level are:

Should the placement rate be limited to ensure that no excess pore pressures
develop, the calculated barricade loads would be largely independent of filling
rate and permeability.

A step change in barricade stress was shown to occur depending on whether or


not the material yields at the fill/rock interface during placement. For both the
yielding and non-yielding cases, the barricade stresses are largely independent
of cement content.

The restriction to flow through the drawpoint dictates the pore pressure
distribution within a consolidating fill stope. This pore pressure distribution
significantly influences the effective stress and therefore barricade stresses.

For the same stope geometry with different drawpoint restrictions, a unique
relationship exists between pore pressure and barricade stress. This unique
relationship can provide a useful means of managing filling activities.

7.4 NON-CONSOLIDATING FILL

The definition of non-consolidating fill means that, during fill deposition, very little
conventional drainage-type consolidation occurs. This scenario would most frequently
be associated with paste fill, but could be equally applicable to fine hydraulic fills that
are placed at fast rates of rise.

For the purpose of the non-consolidating fill sensitivity study, PFA material properties
(from Table 7.2) were adopted. The filling sequence involved the first 8 m of material
being placed over a 16-hour period (0.5 m/hr), followed by a 14-hour rest period, and
then filling the remaining 32 m over a period of 64 hours (0.5 m/hr filling rate). The
stope geometry adopted represents a 13 m wide, 40 m high plane strain stope with a 5m

7.13
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

long drawpoint. Through the drawpoint the fill permeability is reduced by 50% to
represent the reduction in drawpoint flow area. The aspects that were covered in the
consolidating fill sensitivity study have been addressed here also. These include the
influence of stope geometry, permeability, cementation and filling rate.

7.4.1 Influence of stope geometry

This study involved varying the permeability of the stope drawpoint area, which in turn
varied the restriction to conventional drainage-type consolidation through this region. In
this study the drawpoint permeability was increased (kdp= 10kstope) and decreased (kdp=
0.1kstope) by an order of magnitude.

Figure 7.20 presents the development of pore pressure against time during filling for a
point on the stope floor on the opposite side of the stope to the drawpoint. This figure
indicates that during the early stages of filling, the pore pressure is relatively
independent of the drawpoint permeability. The reason for this is that the dissipation of
pore pressure, at the opposite side of the stope to the drawpoint, is primarily dictated by
self desiccation rather than conventional drainage-type consolidation. At the later stages
of filling, when water migrates down to the base of the stope, the restriction to flow at
the drawpoint becomes more influential on the pore pressures.

The influence of the drawpoint permeability on barricade stress is presented in Figure


7.21, which shows the development of barricade stress with time for the different cases.
This result indicates that, even during the early stages of filling, the drawpoint
restriction can influence barricade stresses. This is also the case later in the filling cycle.
This is due to the hydraulic gradient that exists through the drawpoint region. With
higher resistance to flow through the drawpoint the hydraulic gradient is steeper, which
effectively results in lower effective stresses and less stress being transferred to the
surrounding rockmass. This is illustrated in Figures 7.22 (a) and (b), which presents
pore pressure contours at the completion of filling for the kdp=0.1kstope and
kdp=10kstope cases respectively. Figure 7.22a shows an almost linear reduction in
pore pressure when progressing across the stope floor, but Figure 7.22b shows much
higher, almost constant, pore pressures within the stope and a very steep hydraulic
gradient in through the drawpoint. It is this pore pressure profile that creates the
difference in barricade stress.

7.14
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

7.4.2 Influence of permeability

To investigate the influence of permeability on non-consolidating fill deposition,


sensitivity modelling was undertaken using PFA material properties, but with the
hydraulic conductivity increased (k=10kPFA) and decreased (k=0.1kPFA) by an order
of magnitude.

Figure 7.23 presents the evolution of pore pressure with time at a point on the stope
floor at the opposite side of the stope to the drawpoint. This result suggests that
permeability significantly influences the consolidation behaviour. Soon after placement,
a reduction in permeability creates an increase in pore pressure. This is consistent with
conventional consolidation theory. But interestingly, during the later stages of filling a
reduction in permeability is associated with lower pore pressures.

The reason for this unusual response is that, after cement hydration begins to create an
increase in material stiffness, the consolidation behaviour is dictated by the self-
desiccation mechanism. If the propensity to self desiccation is high, very large hydraulic
gradients can be created in the fill mass (due to the material being at different stages of
hydration). If the permeability is low, these hydraulic gradients can be sustained,
effectively suppressing the development of pore pressures. But if the permeability is
increased, the hydraulic gradients will cause water to flow, recharging voids that were
previously depleted due to self desiccation. This will re-establish steady state seepage
pore pressures. An example of this is demonstrated in Figure 7.23 for the k=10.kPFA
case. Here the permeability is high enough for water to flow through the fill mass, and
pore pressures are dictated by the build up that occurs at the drawpoint as discussed in
Section 7.3.

Figure 7.24 presents the calculated barricade stress against time for the different
material permeabilities. The barricade stress trends closely follows those for pore
pressure, with lower permeabilities initially creating higher barricade stresses (due to a
reduction in conventional drainage-type consolidation) but as filling continues the lower
permeability material produces lower barricade stresses.

7.15
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

7.4.3 Influence of cementation

To investigate the influence of cementation on the filling response, modelling involving


PFA with cement contents of 1.5% and 4.5% in addition to the base case of 3.0% was
undertaken.

Figure 7.25 presents the development of pore pressure with time for the different
cement contents, at point on the stope floor on the opposite side of the stope to the
barricade. This figure indicates that even minor variation in cement content can have a
significant influence on the development of pore pressure within a paste-fill stope. The
significant impact is due to the increased stiffness achieved by the higher cement
contents as well as the increase in self-desiccation volumes that come about from higher
cement contents.

Figure 7.26 presents the calculated barricade stress against time for PFA with the
various cement contents. As with pore pressures, changes in cement content
significantly influence barricade stresses for non-consolidating fill. In this particular
case, an increase in cement content from 3% to 4.5% results in a barricade stress
reduction of over 50%, while a reduction in cement content from 3% to 1.5% results in
a 150% increase in barricade stress. Again, the change in barricade stresses can be
attributed to the influence of cementation on the consolidation behaviour, which in turn
influences the total stress distribution. In addition to the influence that cementation has
on the consolidation behaviour, a reduction in cement content also increases the
likelihood of yielding at the rock/fill interface. As discussed in Section 7.3.3, yielding at
the rock/fill interface can increases the vertical total stress within the stope, which
increases barricade stresses.

To investigate the influence of the interface behaviour, an analysis was undertaken


using the base case material properties (PFA with 3% cement) but in this analysis the
cohesive bond along the rock/fill interface was set to zero. The calculated barricade
stresses and pore pressure at the opposite side of the stope to the drawpoint are
presented in Figure 7.27, along with the results for the fully bonded case. Figure 7.27
indicates that yielding at the fill/rock interface can increase barricade stresses, but
unlike the consolidating fill case, this yielding is associated with an increase in pore
pressure. This is because with non-consolidating fill material the increased vertical total

7.16
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

stress creates an increase in excess pore pressure, and due to the low coefficient of
consolidation these pressures cannot be dissipated rapidly.

7.4.4 Filling rate

During the deposition of cemented mine backfill three different time scales interact,
these being the rate of hydration, rate of consolidation and rate of placement. In order to
investigate the influence of the third timescale (rate of placement) a series of numerical
experiments were undertaken. These experiments used PFA material with filling rates of
0.2 m/hr, 2.5 m/hr as well as the base case of 0.5 m/hr. The filling sequence adopted
involved filling the first 8 m followed by a 14 hour rest period before filling the
remaining 32 m.

The development of pore pressure, at the opposite side of the stope to the barricade, is
plotted against time in Figure 7.28. As expected increasing the filling rate caused an
increase in pore pressures. Increasing the filling rate increases the rate of total stress
application but as pore pressures are being dissipated primarily as a result of self
desiccation (which is independent of the pore pressure magnitude) faster filling rates
will create an overall increases increase in pore pressure. The reverse occurs when
filling rates are reduced. In the 0.2 m/hr case the rate of application of total stress is
reduced but the rate of (self desiccation induced) pore pressure reduction remains
constant resulting in lower pore pressures. But, the other influence of slowing filling
rates is to extend the loading timescale to be comparable to the timescale associated
conventional drainage-type consolidation. In this case, drainage-type consolidation acts
to restore steady state seepage pore pressures leading to an increase in pore pressures.

The calculated barricade stress against time is presented in Figure 7.29 for the different
filling rates. Again the trend of the pore pressure and barricade stress plots are similar,
with the highest filling rates being associated with the highest barricade stresses. The
higher stresses can be attributed to reduced consolidation.

7.4.5 Non-consolidating fill: discussion

As with consolidating fills, the overall result of this study indicates that barricade
stresses are closely related to the degree of consolidation (or the pore pressures). But it
is interesting to note that material properties that appeared to have little influence on the

7.17
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

behaviour of consolidating fill, such as cement content and permeability, had a


significant impact on the behaviour of non-consolidating fills. The fundamental reason
for the significant influence of these characteristics is their influence on the
consolidation mechanism.

As with consolidating fills, the significant influence of pore pressures on barricade


stresses poses a problem for mine backfill designers, since pore pressures are very
difficult to predict numerically. However, this dependence presents the opportunity to
manage the situation using in situ pore pressure measurements. Due to the complex
interaction of mechanisms in a non-consolidating fill a unique relationship cannot be
developed between pore pressure and barricade stress. But pore pressure measurements
can provide an indication of the field situation varying from the design.

Using fully-coupled numerical modelling, and the associated laboratory experiments to


define the material characteristics (as outline in Chapter 3), a reasonable understanding
of the likely behaviour during filling can be developed. This provides a rational
approach to defining an initial filling sequence. From the analysis, barricade stresses can
be estimated as well as providing an understanding of the expected pore pressure
regime.

As illustrated in Section 7.4, the modelling results can be significantly influenced by a


number of different characteristics and it is usually necessary to assess the performance
of the model results in situ. Clayton and Bica (1993), Take and Valsangkar (2001) and
Fourie et al. (2007) suggest that the direct measurement of stress within a soil can be
problematic due to the inclusion of a loadcell that possesses a different stiffness to the
soil medium. This problem is exacerbated when attempting to measure stress in a soil,
such as a cementing minefill, in which the stiffness changes (often by an order of
magnitude) with time. Therefore, managing filling operations based on loadcell
measurements is considered unreliable.

However, as demonstrated in this thesis, with non-consolidating fills most variations


from the design assumptions (such as delayed initial set, permeability changes,
excessive drawpoint flow resistance, fill/rock interface softening or inappropriate
material stiffness development) result in higher pore pressures. Therefore, during filling
in situ pore pressure measurements may be compared with modelling results to assess if

7.18
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

the proposed sequence is suitable. Should the measured pore pressures vary from the
predicted value, the schedule can be adjusted accordingly. As barricade stresses
progressively increase during filling, the information gathered during the early stages of
filling can be used to manage filling activities in the later stages, where barricade
stresses may approach their maximum. This approach may not clearly identify the root
cause of any problem, but in most cases it can identify if the actual situation varies from
the design assumptions and hence avert barricade failure.

7.4.6 Non-consolidating fill: conclusion

The two-dimensional mine backfill consolidation program Minefill-2D was used to


undertake a sensitivity study on cemented mine backfill that would be unlikely to
consolidate during the filling process without the assistance of cement hydration. This
study highlighted a number of interesting aspects that should be considered when using
this type of fill. These include:

As with consolidating fills, the degree of consolidation (or increase in pore


pressure) has a significant influence on barricade stresses.

The resistance to flow through the drawpoint region can influence loads
applied to barricade structures. This suggested that the use of free-draining
barricades help to reduce the barricade stresses during filling.

During the initial placement of non-consolidating fill, an increase in


permeability may increase the amount of conventional drainage-type
consolidation, reducing pore pressures and barricade loads.

Contrary to conventional consolidation theory, a reduction in permeability can


actually lead to higher consolidation and lower barricade stresses after cement
hydration is initiated.

The most significant factor influencing loads applied to barricade structures


from non-consolidating fills is cement content. Non-consolidating fills are
highly dependent on cementation to achieve consolidation due to the stiffness
increase and self-desiccation characteristics that cementation imparts. Other
properties that influence these characteristics, such as placed density and

7.19
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

mineralogy, can have a comparable influence on consolidation and barricade


stresses.

A number of different factors were shown to influence stresses applied to non-


consolidating fill barricade. The most significant of these factors result in
higher pore pressures. Therefore, it is suggested that fully-coupled numerical
analysis, combined with appropriate in situ monitoring, can provide a safe and
efficient means of managing filling activities.

7.5 DEVELOPMENT OF EFFECTIVE STRESS DURING


CURING

7.5.1 Comparision between consolidating and non-consolidating fill

Coring of in situ cemented backfill has shown in situ strengths to be frequently greater
than those measured in the laboratory, for the same mix (Revell 2004, le Roux et al.
2002, Belem et al. 2002). The higher in situ strengths are expected to be a result of
different curing conditions. One aspect that differs between material cured in situ and
that cured in the laboratory, is the level of effective stress during curing. Application of
effective stress during curing has been shown to increase material strengths by
increasing the number of contact points and improving the intimacy of the contacts
(Blight 2000, Rotta et al. 2003, Consoli et al. 2000).

While the development of total stress to an element within a stope is relatively easy to
predict, it is the rate that effective stress develops relative to the hydration period that
must be appropriately understood to develop a more representative approach to sample
preparation. For example, application of effective stress at the same rate as the total
stress will result in an initial compression of the soil matrix and hydration at an
increased density, while application of the entire self-weight stress at the completion of
filling may result in crushing of weak cement bonds, and may thus have a detrimental,
rather than a positive, effect on strength.

As Minefill-2D is a fully coupled model, it can be used to estimate the rate and
magnitude that effective stress develops in material within a stope. Given the rate of
effective stress development in situ, the significance of this with respect to the final
strength of the material can be understood.

7.20
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

To investigate the impact of different fill types on the rate at which effective stress
develops, the Paste fill B (PFB) and Hydraulic Fill A (HFA) modelling results, from
Section 7.2, are plotted in Figure 7.30. This shows a plot against time of the
development of vertical effective stress (v) at a point located 12 m above the stope
floor, at the opposite side to the drawpoint, for the HFA (HFA v) and PFB (PFB v).

Also plotted in Figure 7.30 are the total self-weight stress (v) and the effective self-
weight stress (v) that would develop in the absence of any arching. Assuming the same
density, for the same filling sequence, this would be equal for both fill types. The other
information presented in Figure 7.30 is the magnitude of decrease in pore pressure for
the two materials due to self desiccation in isolation (u SD only) (using Equation
3.27).

The first point to note is that effective stress develops in HFA material immediately
after placement, while no effective stress develops in PFB until approximately 6 hours.
Also, in the first 30 hours of curing, the rate of development of effective stress in the
HFA material is almost double that in PFB. Finally, the rate of effective stress
development is relatively constant with the HFA material, while that with PFB is
initially very slow but then increases exponentially.

7.5.2 Development of effective stress in consolidating fill

Figure 7.30 suggests that the rate of development of effective stress in the HFA material
is in accordance with the effective self-weight stress. The reason is that the coarse
nature of HFA causes immediate dissipation of excess pore pressures, but due to the
restriction (to flow) created at the drawpoint a phreatic surface is established within the
fill mass. This phreatic surface creates approximately hydrostatic pore pressures within
the stope and, if arching is neglected, this would result in effective stress developing in
accordance with the effective unit weight of the material.

If arching occurs as filling progresses, the effective stress applied is less than the
effective self-weight stress. The significance of this increases as the stope plan area is
reduced, with the consequence of reduced vertical total stress and a reduced rate of
development of vertical effective stress . Another factor that could influence this
response is the restriction at the drawpoint. As illustrated in Section 7.3.1, the resistance
to flow through the drawpoint modifies the gradient of the pore pressure isochrones

7.21
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

within the stope. Using the effective unit weight and the filling rate to define the
effective self-weight stress assumes hydrostatic pore pressures. A reduction in flow
resistance through the drawpoint area would act to reduce pore pressures leading to an
increase in vertical effective stress.

To investigate the applicability of this theory for different filling rates, the effective
stress at the same point was monitored for different filling rates using HFA. Figure 7.31
presents the development of vertical effective stress against time for a point 12 m above
the base of the stope with different filling rates. Also presented in Figure 7.31 are the
self-weight stresses, for each filling rate, if arching is neglected.

In each case, the actual effective stress is very similar to the effective self-weight stress.
The rate of development of effective self-weight stress is dependent on the filling rate
and the effective unit weight of the material and independent of the rate of cementation.
Obviously, the maximum effective stress in this case depends on the amount of material
placed above the location of interest.

It is also interesting to note that, for the HFA material, the calculating the effective
stress using the dissipation of pore pressure that would occur due to self desiccation in
isolation significantly overestimates the effective stress for this material. This is because
with the high permeability of HFA, only very small hydraulic gradients are required to
recharge water volumes removed through self desiccation. These pores are recharged
and steady state seepage pore pressures are re-established.

7.5.3 Development of effective stress in non-consolidating fill

Figure 7.30 indicates that the rate of development of vertical effective stress in PFB
material is approximately equal to the rate that pore pressure is dissipated as a result of
self desiccation alone. This is because the low coefficient of consolidation allows very
little conventional (drainage-type) consolidation to take place. Furthermore, with little
conventional consolidation taking place, the application of total self-weight stress (from
fill accretion) creates an equivalent increase in pore pressure and no change in effective
stress. As the only mechanism causing dissipation of any significant proportion of pore
pressure is self desiccation, this mechanism alone can appropriately capture the rate at
which effective stress develops in the soil matrix.

7.22
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

To investigate this theory in more detail, modelling was undertaken using PFB and the
same fill/rest schedule as before, but this time with filling rates varying from 0.05 m/hr
to 1.0 m/hr. The calculated vertical effective stress is plotted against time, for a point 12
m above the base of the stope, in Figure 7.32. Also presented in Figure 7.32 is the
dissipation of pore pressure as a result of self desiccation occurring in isolation. As this
characteristic is material dependent, the relationship is consistent for all cases.

Figure 7.32 indicates that, the development of vertical effective stress is similar for all
filling rates greater than 0.05 m/hr, and that these are close to the magnitude of pore
pressure reduction from self desiccation alone. This is particularly the case during the
early stages of hydration, where the effective stress level has the greatest influence on
the cured strength.

The exception is the slowest filling rate of 0.05 m/hr. At this filling rate, the time
required to fill the stope is of the same order as the time required for conventional
drainage-type consolidation to take place. With reference to Gibsons (1958)
consolidation chart (Figure 2.4) this situation corresponds to a dimensionless time factor
(T=m2t/cv) equal to 0.5. Figure 2.4 suggests that this time factor corresponds to
(du/dx)/ equal to 0.15, indicating drained filling conditions. This essentially changes
the situation to that of a consolidating-type fill, where the effective stress is similar to
the effective self weight.

To investigate the spatial variation around a stope, the base case (PFB with a filling rate
of 0.5 m/hr) was selected and the development of vertical effective stress at different
elevations was monitored, and the results plotted in Figure 7.33. Also presented in
Figure 7.33 is the total vertical self-weight stress (from the accumulating fill mass) as
well as the reduction in pore pressure from self desiccation alone.

Figure 7.33 indicates that for non-consolidating fills, the development of vertical
effective stress throughout a stope can be appropriately represented by the self
desiccation mechanism in isolation.

7.5.4 Curing of fill: discussion and conclusion

From the above discussion, it is clear that effective stress should be applied to
laboratory control samples during curing at the same rate that they develop in situ,

7.23
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

otherwise the cured strengths of the laboratory samples will not be the same as the in
situ cured strengths. One approach to doing so is to simulate each individual filling
sequence using a tool such as Minefill-2D, and, based on the results, the laboratory
specimen could be loaded to create an equivalent stress regime throughout the curing
period. However, the results from the previous sensitivity analyses can also provide
some useful guidance in formulating an appropriate curing technique.

For material with a high uncemented coefficient of consolidation relative to the filling
rate (such as the HFA case presented above), the development of vertical effective stress
was shown to be in accordance with the effective self weight vertical stress. This logic
assumes:

Excess pore pressures are immediately dissipated

Hydrostatic pore pressures are established within the fill mass due to the
restriction at the drawpoint

There is no arching.

It is also important to note that the maximum effective stress depends on the depth of
the sample within the fill mass.

For material with a very low coefficient of consolidation relative to the filling rate (such
as the PFB case presented above), it is expected that effective stress develops due to the
drop in pore pressure that occurs from self desiccation occurring in isolation. In this
case, the setup similar to the hydration experimental setup described previously (Section
3.6) can be used to allow the effective stress to develop as hydration proceeds. In this
setup, high total stress is applied to the saturated sample enclosed in a membrane, which
generates pore pressure practically equal to the applied total stress, and thus there is
initially practically zero effective stress, just as in the stope. However, the process of
hydration causes pore pressures to reduce (and effective stresses to increase) due to the
self-desiccation mechanism, in a manner that exactly mimics what happens within the
stope. Thus, at all stages of the test, the effective stress develops at exactly the same
rate as it does within the stope assuming of course that the material is sufficiently fine-
grained to prevent any conventional consolidation occurring prior to hydration.

7.24
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

Modelling demonstrated that the rate of development of vertical effective stress is the
same throughout the fill mass in the stope, which suggests that a single experiment may
suitably define the material strengths throughout the stope. This logic assumes:

The material remains saturated in the field

No conventional drainage-type consolidation takes place.

It is important to note that an increase in the coefficient of consolidation can have the
effect of increasing or decreasing the rate of development of effective stress.
Conventional consolidation acts to restore steady state seepage pore pressures.
Therefore, if the self-desiccation mechanism is unable to reduce pore pressures below
the steady state condition (as explained in Section 7.2.2), conventional consolidation
acts to assist with the dissipation of pore pressures and thus, in this case, this increases
the rate of development of effective stress. This is the case where a filling rate of 0.5
m/hr is adopted with PFB. On the other hand, should the self-desiccation mechanism be
capable of reducing the pore pressures below steady state (as was demonstrated with
PFA in Section 7.2.2), conventional consolidation causes water to flow downwards,
recharging the pores, and re-establishing steady state seepage pore pressures. This
reduces the rate that effective stress develops.

7.6 CONCLUSION

The tools developed throughout this thesis have been used in this chapter to investigate
the behaviour of tailings-based mine backfill. This investigation addressed a comparison
between a range of tailings-based fill types, focusing on barricade stresses and the
development of effective stress during curing. At the completion of each section, a
specific conclusion section was provided. Some of the major conclusions that resulted
from this analysis are presented below.

Regardless of the fill properties, the most significant factor influencing barricade
stresses is consolidation. Reduced consolidation resulted in higher barricade stresses.

Broadly speaking, tailings-based fills can be divided into two groups: fills that
consolidate immediately after placement (consolidating fills) and those that are unlikely
to consolidate without the influence of cementation (non-consolidating fills). The
fundamental difference between these fill types is that the mobilisation of strength in

7.25
Mechanics of Mine Backfill Matthew Helinski
Sensitivity Study The University of Western Australia

consolidating fills is dependent on the rate of deposition, while the mobilisation of


strength in non-consolidating fills is dependent on the rate of cement hydration.

Pore pressures in consolidating fills are largely independent of cementation,


permeability (for a given phreatic surface elevation) and filling rates, but are influenced
by the flow restriction through the drawpoint at the base of a stope. Reducing this flow
restriction can significantly reduce pore pressures throughout the stope, thereby
increasing effective stresses and reducing barricade stresses.

Cementation, permeability, drawpoint restriction and filling rate can all have a
significant influence on the barricade stresses from non-consolidating fill. A variation in
any of these factors results in a pore pressure change, which leads to higher barricade
stresses.

7.26
Mechanics of Mine Backfill Matthew Helinski
Concluding Remarks and Recommendations of Future Work The University of Western Australia

CHAPTER 8
CONCLUDING REMARKS AND RECOMMENDATIONS
FOR FUTURE WORK

8.1 CONCLUDING REMARKS


This thesis has presented an investigation into the mine backfill deposition process that
is based on fundamental soil mechanics principles. The main feature of the work was to
investigate and numerically couple three time-dependent processes that interact during
the cemented tailings backfill process, these being: the loading rate (or filling rate), the
conventional consolidation rate and the cement hydration rate.

Based on the interaction of these different processes, a number of useful results were
obtained that are relevant to the cemented mine backfill process. These results and some
concluding remarks are contained in this chapter.

8.2 MAIN OUTCOMES

The first significant result from this work was to highlight the importance of
consolidation in estimating barricade stresses. It was demonstrated numerically that
stresses applied to mine backfill barricade structures can vary by an order of magnitude
depending on the degree of consolidation that occurs during filling. The significance of
consolidation on the total stress distribution was also demonstrated experimentally
using geotechnical centrifuge modelling.

Furthermore, it was demonstrated that, without the presence of cement, the degree of
consolidation during filling for typical tailings-based backfills can range from fully
drained to fully undrained conditions.

Investigation of the fundamental characteristics that influence the consolidation of


cemented mine backfill showed that even with the minor amounts of cement typically
added to mine backfill (2% to 10%), the material properties can change significantly.
Specifically these changes include:

An increase in stiffness, which can be greater than an order of magnitude.

8.1
Mechanics of Mine Backfill Matthew Helinski
Concluding Remarks and Recommendations of Future Work The University of Western Australia

A reduction in permeability, which can also be greater than an order of


magnitude.

A volumetric reduction, which, while only very small, can lead to a significant
drop in pore pressure when combined with the high material stiffness achieved
through cement hydration (this is the so called self-dessication effect).

By modifying the tailings consolidation program MinTaCo, the time-dependent process


of cement hydration was successfully coupled with filling and conventional
consolidation to assess the influence of cementation on the filling process. This
modified program (termed CeMinTaCo) demonstrated that:

The cement-induced increase in stiffness and self desiccation can make a


significant contribution to the mine backfill consolidation process

Because of the substantial influence of cementation on the consolidation


behaviour, cement content can have a major influence on the consolidation of
some cemented mine backfills.

Contrary to conventional consolidation theory, the combination of cementation


with low permeability material can act to generate lower pore pressures (higher
consolidation) than higher permeability material.

To investigate the influence of stress arching (onto the surrounding rockmass) during
filling, a new two-dimensional consolidation program (termed Minefill-2D) was
developed. Like CeMinTaCo, Minefill-2D coupled the three time-dependent processes,
but, unlike CeMinTaCo, Minefill-2D also takes account of the influence of the stiff
surrounding rockmass on the stress distribution. In addition, Minefill-2D allowed the
stope drawpoint geometry to be incorporated (albeit in plane strain), which allowed
stresses applied to barricade structures to be determined. This new model was compared
with results from a series of laboratory tests and shown to provide a good representation
of the process.

To investigate the ability of Minefill-2D to represent the consolidation behaviour in an


actual filling situation, modelling results were compared with in situ measurements. A
comparison was carried out between the measured pore pressure in the centre of the
stope floor and that predicted using Minefill-2D, using material properties that were

8.2
Mechanics of Mine Backfill Matthew Helinski
Concluding Remarks and Recommendations of Future Work The University of Western Australia

independently derived using the material characterisation technique described in Section


3.6. This comparison indicated that Minefill-2D can provide a very good representation
of the consolidation behaviour.

A sensitivity study to investigate barricade stresses using Minefill-2D indicated:

Broadly speaking, tailings-based fills can be divided into two groups: fills that
consolidate immediately after placement (consolidating fills) and those that
are unlikely to consolidate (during the filling period) without the influence of
cementation (non-consolidating fills). The fundamental difference between
these fill types is that the mobilisation of strength in consolidating fills is
dependent on the rate of deposition, while the mobilisation of strength in non-
consolidating fills is dependent on the rate of cement hydration.

Provided that no excess pore pressures are generated during deposition (i.e.
consolidating fills as defined in Chapter 7) pore pressure is largely independent
of cementation, permeability (for a given phreatic surface elevation) and filling
rates, but is influenced by the flow restriction through the drawpoint at the base
of the stope. Reducing this flow restriction can significantly reduce pore
pressures throughout the stope, resulting in an increase in effective stress and a
reduction in barricade stresses.

Cementation, permeability and filling rate can all have a significant influence
on the barricade stresses imposed by a non-consolidating fill. A variation in
any of these characteristics results in a pore pressure change, which leads to
changes in barricade stresses.

As consolidation is the characteristic that most significantly influences


barricade stresses (in both fill types), it is recommended that the most
appropriate means of managing the risk of barricade failure is through in situ
pore pressure monitoring strategies. The recommended management strategies
for consolidating fills differ from those for non-consolidating fills.

A sensitivity study to investigate the development of effective stress during curing using
Minefill-2D indicated:

8.3
Mechanics of Mine Backfill Matthew Helinski
Concluding Remarks and Recommendations of Future Work The University of Western Australia

The application of effective stress (at rates similar to those experienced in situ)
during curing increases the final strength of the material. This was primarily
due to an increase in density, which is consistent with the results of previous
experimental studies.

The development of effective stress during curing in a consolidating fill can


be closely related to the accumulation of effective self-weight stress from the
overlying fill mass

The development of effective stress during curing in a non-consolidating fill


can be closely related to the reduction in pore pressure from self desiccation in
isolation.

8.3 RECOMMENDATIONS FOR FUTURE WORK

The main focus of this thesis was to develop a framework for understanding the
cemented mine backfill deposition process that may be used to assess the significance of
various aspects and help with the interpretation of in situ monitoring results. In
achieving the final outcome, a series of assumptions and simplifications were made to
firstly represent the material behaviour and secondly simulate the behaviour. Much of
this work should be revisited with the view of refining some of the material
characteristics that were shown to be most critical. Specifically material modelling
should focus on:

Improved techniques for quantifying the self-desiccation process, including


directly quantifying volumetric changes that occur during the hydration process
and taking account of temperature changes during the hydration process.

An improved model to represent the influence of cement hydration and soil


compression on the permeability of the material.

Additional experimental and constitutive modelling to more appropriately


represent the variation in material strength and stiffness during the cement
hydration process.

Minefill-2D is a new finite element model that was developed specifically for the
purpose of representing the cemented mine backfill deposition process. In its current

8.4
Mechanics of Mine Backfill Matthew Helinski
Concluding Remarks and Recommendations of Future Work The University of Western Australia

state, the program is considered suitably rigorous for the applications in this thesis, but
nevertheless improvements to some of the calculation algorithms would probably result
in increased calculation speed and accuracy. Specifically these improvements might
include:

An improved time-stepping algorithm. This algorithm would calculate the most


appropriate time-stepping size based on the element size, coefficient of
consolidation and information from previous calculations in a manner similar
to that adopted by Yong at al. (1983).

Implementation of a constitutive model that takes account of yielding due to


volumetric compression as well as a more rigorous approach to numerically
accounting for strain softening. This would be useful when investigating the
behaviour of very weak material.

Implementation of interface elements to more appropriately represent the


behaviour of the interface between the fill and the surrounding rockmass.

Extension of the geometry to more appropriately represent the stope drawpoint.


This may include extension from two to three dimensions or coupling of the
axi-symmetric version of Minefill-2D (to represent the stope) with another axi-
symmetric version of Minefil-2D to represent the drawpoint.

The stress distribution around a stiff fill mass can be influenced by the
deformation behaviour of the barricade structure. Consideration should be
given to implementing beam elements to represent any flexibility in the
barricade structure.

The combination of slow filling rates and high self-desiccation rates can lead to
the development of large matrix suctions. As discussed by Grabinski and
Simms (2006) these suctions have the potential to desaturate the fill matrix
which could change the modelling response. Consideration should be given to
taking account of matrix desaturation in the numerical analysis.

This thesis demonstrated that devising a centrifuge experiment in which the three time-
dependent processes can interact is very difficult, and in order to achieve this, full-scale
field testing is required. While the repetitive nature of stoping and backfilling provides

8.5
Mechanics of Mine Backfill Matthew Helinski
Concluding Remarks and Recommendations of Future Work The University of Western Australia

opportunities to undertake full scale parametric studies, full scale field testing
introduces problems regarding the suitability of instrumentation. Gathering quality data
from cemented mine backfill stopes can be difficult and care should be taken to address
the following concerns;

It is well documented (Clayton and Bica, 1993, Take and Valsangkar 2001)
that the deformation of earth pressure cells can lead to under registration (i.e.
the cell measures less than the actual stress). The degree of under registration
depends on the stiffness of the cell relative to the surrounding soil. As
cemented mine backfill can gain very high stiffness during the hydration
process, care should be taken to ensure that the cell stiffness matches the
material stiffness appropriately.

In the centre of a large backfill mass, the cement hydration process can lead to
temperature changes that are in the order of 20C. With fluid-filled pressure
cells, this can cause the internal fluid to expand, leading to an artificial pressure
increase in the cell. Should this type of pressure cell be adopted, care must be
taken to ensure the fluid has an appropriately low coefficient of thermal
expansion to eliminate or minimise this error

The measurement of positive pore pressure can be successfully achieved using


vibrating wire piezometers, but when (and if) pore pressures become negative,
the porous disk at the face of the piezometer can desaturate, creating a capillary
block. In order to gather accurate negative pore pressure measurements, care
should be taken to select a porous disk with an appropriate pore size to avoid
desaturation. Also, to minimise the likelihood of desaturation, the porous disk
should be saturated with a low-viscosity fluid under vacuum. Grabinski at al.
(2007) adopted heat dissipation sensors in an attempt to measure pore water
suctions, but are yet to report on the success of this approach.

An obvious model verification strategy is to use in situ monitoring results to further


verify the performance of the modelling approach presented in this thesis. However,
prior to undertaking such back analysis, it is considered imperative that a number of
assumptions regarding boundary conditions should be more clearly defined. Aspects
that should be investigated include pore pressure accumulation immediately behind

8.6
Mechanics of Mine Backfill Matthew Helinski
Concluding Remarks and Recommendations of Future Work The University of Western Australia

barricade structures (assumed to be zero in this analysis) and the pore pressure
boundary condition within the stope at the fill / rock interface.

8.7
Mechanics of Mine Backfill Matthew Helinski
References The University of Western Australia

CHAPTER 9
REFERENCES

Asaoka, A. Naknao, M. Noda, T and Kaneda, K. (2000). Delayed compression /


consolidation of natural clay due to degradation of soil structure. Soils and Foundations,
Vol. 40, No. 3, 75-85.

Aubertin, M. Li, L. Arnoldi, S. Belem, T. Bussire, B. Benzaazoua M. and Simon, R.


(2003). Interaction between backfill and rock mass in narrow stopes. Soil and Rock
America (Calligan Einstein and Whittle eds.), V1, 1157-1164.

Baig, S. Picornell, M. and Nazarian, S. (1997). Low strain shear moduli of cemented
sands. Journal of Geotechnical and Geoenvironmental Engineering, ASCE, Vol. 123,
No. 6, 540-545.

Baldwin.W.F. (2004). Responses to requests to comment, Handbood on Mine fill (eds.


Potvin, Thomas, Fourie), Australian Centre for Geomechanics, ISBN 0-9756756-2-1.

Bentz, D.P. (1995). A three-dimensional cement hydration and microstructure program.


Hydration rate, heat of hydration, and chemical shrinkage. Building and Fire Research
Laboratory, National Institute of Standards and Technology, Gaithersburg, Maryland
20899, USA. Available from http://fire.nist.gov/bfrlpubs/build96/art095.html (last
accessed 19 April 2007).

Bentz. D.P, Garboczi. E.J, Lagergren. E.S (1998). Multi-scale microstructure and
transport properties of concrete. Construction and Building Materials, Vol. 10, No. 5,
293-300.

Bentz. D.P. Garboczi. E. J. and Lagergren. E. S (1998). Multi-scale microstructure


modelling of concrete diffusivity: identification of significant variables. Cement,
Concrete, and Aggregates, CCAGDP, Vol. 20, No. 1, 129-139.

Belem, T. Benzaazoua, M. Bussire, B. and Dagenais, A.M. (2002). Effects of


settlement and drainage on strength development within mine paste backfill. Tailings
and Mine Waste 02, Swets & Zeitlinger, ISBN 90 5809 353 0, 139-148.

9.1
Mechanics of Mine Backfill Matthew Helinski
References The University of Western Australia

Belem, T. H, A. Simon, R. and Aubertin, M. (2004). Measurement and prediction of


internal stresses in an underground opening during filling with cemented fill. Proc. of
the Fifth International Symposium on Ground Support, Villaescusa and Potvin (eds),
Balkema Publishers, 28-30 September, Western Australia, ISBN 9058096408, p. 619-
630.

Belem,T. El Aatar, O. Bussiere, B. Benzaazoua, M. Fall, and M.Yilmaz, E. (2006).


Characterisation of self-weight consolidated paste backfill. Paste 06, (Jewell, Lawson
& Newman eds.). Limerick, Ireland, 333-345.

Biot, M.A. (1941). General theory of three-dimensional consolidation, Journal of


Applied Physics, Vol. 12, 155-164.

Bishop, A.W. and Henkel, D.J. (1962). Measuring Soil Properties in the Triaxial Test.
2nd ed, Edward Arnold, London.

Black, D.K. and Lee, K.L. (1973). Saturating laboratory samples by back pressure.
Journal of the Soil Mechanics and Foundations Division, ASCE, Vol. 99, No. SM1, 75-
93.

Blight, G.E. (2000). The use of mine waste as a structural underground support. Geo-
eng 2000, V1: Invited paper, Technomic, ISBN No. 1-58715-067-6.

Blight, G.E. and Spearing, A.J.S. (1996). The properties of cemented silicated backfill
for use in narrow, hard-rock, tabular mines. Journal of South African Institute of Mining
and Metallurgy, January/February 1996, 17-28.

Breysse, D. and Gerard, B, (1997). Modelling of permeability in cement-based


materials: Part 1- Uncracked medium. Cement and Concrete Research, Vol.27, No. 5,
761-775

Brouwers, H.J.H. (2004). The work of Powers and Brownyard revisited: Part 1. Cement
and Concrete Research, Vol. 34, 1697-1716.

Carrier, W.D. Bromwell, L.G. and Somogyi, F. (1983). Design capacity of slurried
mineral waste ponds. Journal Geotechnical Engineering Division, ASCE, 109(GT5), p.
699-716.

9.2
Mechanics of Mine Backfill Matthew Helinski
References The University of Western Australia

Cayouette, J. (2003). Optimization of paste backfill plant at Louvicourt mine. Canadian


Institute of Mining Bulletin , 96, 1075, 51-57.

Clayton, C.R.I. and Bica, A.V.D. (1993). Design of diaphragm-type boundary total
stress cells. Gotechnique, 43 (4), 523-535.

Clough, G.W. Sitar, N. Bachus, R.C. and Nader, S.R. (1981). Cemented sands under
static loading. Journal of the Geotechnical Engineering Division, ASCE, Vol. 107,
GT6, 799-817.

Clough, G.W. Iwabuchi, J. Rad, N.S. and Kuppusamy, T, (1989). Influence of


cementation on liquefaction of sands. Journal of Geotechnical Engineering, ASCE,
Vol. 115, No. 8, 1102-1117.

Consoli, N. C. Rotta, G.V. and Prietto, P.D.M. (2000). Influence of curing under stress
on the triaxial response of cemented soils. Gotechnique 50(1), 99-105.

Consoli, N.C. and Sills, G.C. (2000). Soil formation from tailings: comparison of
predictions and field measurements. Gotechnique 50(1), 25-33.

Consoli, N. C, Foppa, D, Festugato, L. and Heineck, K. S. (2006). Key parameters for


strength control of artificially cemented soils. Journal of Geotech. and
Geoenvironmental Engrg, ASCE, vol. 133, No. 2, 197-205.

Coop, M. R. and Atkinson, J. A. (1993). Mechanics of cemented carbonate sands.


Gotechnique, 43(1), 53-67.

Coroners Report (2001) Ref No 20/01, Record of investigation into death.

Cowling R, Grice A.G. and Isaacs L.T. (1987). Simulation of hydraulic filling of large
underground mining excavations. Numerical Methods in Geomechanics, (Innsbruck
1988), Swoboda (ed.), Balkema, ISBN 90 6191 809 X, 1869-1876.

Cuccovillo, T. and Coop, M.R. (1997). Yielding and pre-failure deformation of


structured sands. Gotechnique, 47(3), 491-508.

Dyvik, R. and Olsen, T.S. (1989). Gmax measured in oedometer and DSS tests using
bender elements. Proceedings of the 12th International Conference on Soil Mechanics
and Foundation Engineering, Rio de Janeiro, Vol. 1, 3942, Balkema, Rotterdam.

9.3
Mechanics of Mine Backfill Matthew Helinski
References The University of Western Australia

Fahey, M. and Carter, J.P. (1993). A finite element study of the pressuremeter test in
sand using a non-linear elastic plastic model. Canadian Geotechnical Journal, 30, 348
362.

Fahey, M. Finnie, I. Hensley, P.J. Jewell, R.J. Randolph, M.F. Stewart, D.P. Stone,
K.J.L. Toh, S.H. and Windsor, C.S. (1990). Geotechnical centrifuge modelling at the
University of Western Australia. Australian Geomechanics, No. 19, 33-49.

Fahey, M. and Newson, T.A. (1997). Aspects of the geotechnics of mining wastes and
tailings dams. Theme Lecture, GeoEnvironment 97: Proc. of the 1st Australia - New
Zealand Conference on Environmental Geotechnics, Melbourne, Australia, 115134,
Balkema, Rotterdam.

Fahey, M. Newson, T.A. and Fujiyasu, Y. (2002). Engineering with tailings. Keynote
Lecture, Proc. 4th International Conference on Environmental Geotechnics, Rio de
Janeiro, Brazil, Vol. 2, 947-973, Balkema, Lisse.

Fernandez, A. and Santamarina, J. C. (2001). Effect of cementation on the small strain


parameters of sand. Canadian Geotechnical Journal, 38(1), 191199.

Fredlund, D.G. and Rahardjo, H. (1993). Role of unsaturated soil behaviour in


geotechnical engineering practice. Proc. Southeast Asian Geotechnical Conference, 11,
SEAGC, Singapore, May, 3749.

Fourie, A.B. Helinski, M. Fahey, M. (2007). Using effective stress theory to


characterise the behaviour of backfill. Minefill 07, Montreal, Paper # 2480.

Garboczi, E.J. and Bentz, D.P. (1996). Modelling of the microstructure and transport
properties of concrete. Construction and Building Materials, Vol. 10, No. 5, 293-300.

Gens, A. and Nova, R. (1993). Conceptual bases for a constitutive model for bonded
soils and weak rocks. Geotechnical Engineering of Hard Soils Soft Rocks, R4
Anagnostopoulos, A., Schlosser, F., Kalteziotis, N. and Frank, R. (eds), Balkema,
Rotterdam, 1: 485-494.

Gibson, R.E. (1958). The process of consolidation in a clay layer increasing in thickness
with time. Gotechnique, 8(4), 177-182.

9.4
Mechanics of Mine Backfill Matthew Helinski
References The University of Western Australia

Gibson, R.E. England G.L. and Hussey, M.J.L. (1967). The theory of one-dimensional
consolidation of saturated clays. I. Finite non-linear consolidation of thin homogeneous
layers. Gotechnique, 17(3), 261-273.

Gibson, R.E. Schiffman, R.L. and Cargill, K.W. (1981). The theory of one-dimensional
consolidation of saturated clays. II. Finite non-linear consolidation of thick
homogeneous layers. Canadian Geotechnical Journal, 18, No.2, 280-293.

Grabinski, M. and Simms, P. (2006). Self-desiccation of cemented paste backfill and


implications on mine design. Paste 06, Limerick April 3-7, Australian Centre for
Geomechanics, ISBN 0-9756756-5-6.

Grabinski, M. and Bawden, W.F. (2007). In situ measurements for geomechanical


design of cemented paste backfill systems. Canadian Institute of Mining Bulletin,
Special Edition on Mining with Backfill (August 2007).

Guo, J. (1989). Maturity of concrete: method for predicting early-stage strength, ACI
Mater J., 86 (1989). (4), 341353.

Helinski, M., Fourie, A.B. and Fahey, M. (2006). Mechanics of early age cemented
paste backfill. Paste 06, Limerick April 3-7, Australian Centre for Geomechanics,
ISBN 0-9756756-5-6.

Helinski, M. and Grice, A.G. (2007). Water management in hydraulic fill operations.
Minefill 07, Montreal, Paper # 2481.

Hua, C. Acker, P. and Ehrlacher, A. (1995). Analysis and models of autogenous


shrinkage of hardening cement paste. Cement and Concrete Research, Vol. 25, No. 7,
14571468

Illston, J.M. Dinwoodie, J.M. and Smith, A.A. (1979). Concrete, Timber and Metals.
Van Nostrand Reinhold Company, New York. ISBN 0419154701.

Ismail, M.A., Joer, H. and Randolph, M.F. (2000). Sample preparation technique for
artificially cemented soils. Geotechnical Testing Journal, ASTM. Vol. 23, No. 2, 171
177.

Jamiolkowski, M. Lancellotta, R. Lo Presti, D.C.F. and Pallara, O. (1994). Stiffness of


Toyoura sand at small and intermediate strain. Proceedings of the 13th International

9.5
Mechanics of Mine Backfill Matthew Helinski
References The University of Western Australia

Conference on Soil Mechanics and Foundation Engineering, New Delhi, India, Vol. 1,
169172, Balkema, Rotterdam.

Jewell, R.J. Andrews, D.C. and Khorsid, M.S. (eds). (1988). Engineering for
Calcareous Sediments: Proceedings of the International Conference on Calcareous
Sediments, Perth, 15-18 March. Proceedings, Vol. 1-2.

Kavvadas, M. and Amorosi, A. (2000). Constitutive model for structured soils.


Gotechnique, 50(3), 263-273.

Koenders, E.A.B. and Van Breugel, K. (1997). Numerical modelling of autogenous


shrinkage of hardening cement paste. Cement and Concrete Research, Vol. 27, No. 10,
14891499.

Kuganathan, K. (2002). A method to design efficient mine backfill drainage systems to


improve safety and stability of backfill bulkheads and fills. Proceedings of the 7th
Underground Operators Conference, Townsville, Qld., Australasian Institute of Mining
and Metallurgy, 191-188

Lagioia, R. and Nova, R. (1995). Experimental and theoretical study of the behaviour of
calcarenite in triaxial compression, Gotechnique, 45(4), 633-648.

Landriault, D. (2006). They said It will never work- 25 years of paste backfill 1981-
2006. Paste 06, Limerick, Ireland, Australian Centre for Geomechanics, ISBN 0-
9756756-5-6, 277 292.

Landriault, D. (1995). Paste backfill mix design for Canadian underground hardrock
mining. Canadian Institute of Mining Report, V22, No.1, 11-13.

Le Roux, K.A. Bawden, W.F. and Grabinski, M.W.F. (2002). Assessing the interaction
between hydration rate and fill rate for a cemented paste backfill. Proceedings on the
55th Canadian Geotechnical and 3rd joint IAH-CNC Groundwater Specialty Conference,
Niagara falls, Ontario, October 20-23, 2002, pp.427-432.

Le Roux, K. (2004). Responses to requests to comment. Handbood on Mine Fill (eds.


Potvin, Thomas, Fourie), Australian Centre for Geomechanics, ISBN 0-9756756-2-1.

9.6
Mechanics of Mine Backfill Matthew Helinski
References The University of Western Australia

Le Roux, K.A. Bawden, W.F. and Grabinski, M.W.F. (2005). Field properties of
cemented paste backfill at the Golden Giant Mine. Proceedings of the 8th International
Symposium on Mining with Backfill, Beijing, September, 233-241.

Leroueil, S. and Vaughan, P.R.(1990). The general and congruent effects of structure in
natural soils and weak rocks. Gotechnique 40(3), pp.467-488.

LI, L. and Aubertin, M. (2003). A general relationship between porosity and uniaxial
strength of engineering materials. Canadian Journal of Civil Engineering, August 03,
Vol. 30, No.4, 644-658.

Liu, M.D. Carter, J.P. Desai, C.S. and Xu, K.J. (1998). Analysis of the compression of
structured soils using the disturbed state concept. University of Sydney, Department of
Civil Engineering, Research Report R770.

Liu, M.D. and Carter, J.P. (2002). A Structured Cam-Clay model. Canadian
Geotechnical Journal, Vol ??, No 39, 1313-1332.

Liu, M.D. and Carter, M.D. (2005). Some applications of the Sydney Soil Model.
Proceedings of the 16th International Conference on Soil Mechanics and Geotechnical
Engineering, Osaka, September, 2005. Proceedings, vol. 2, 881-884.

Marston, A. (1930). The theory of external loads on closed conduits in the light of latest
experiments. Bulletion No.96, Iowa Engineering Experiment Station, Ames, Iowa.

McCarthy, P. (2007). Digging Deeper. AMC Newsletter, March 07.

Mitchell, R.J. and Wong, B.C. (1982). Behaviour of cemented sands. Canadian
Geotechnical Journal, Vol. 19, 289-295.

Mikasa, M. (1965). The consolidation of soft clay, a new consolidation theory and its
application. Japanese Society of Civil Engineers (Reprint from Civil Engineering in
Japan 1965): 21-26.

Naylor, D.J. Pande, G.N. Simpson, B. and Tabb, R. (1981) Finite elements in
geotechnical engineering, Pineridge press, Swansea, U.K. ISBN 0-906674-11-5.

Potts, M.D. and Zdravkovi, L. (1999). Finite element analysis in geotechnical


engineering (theory), Thomas Telford, ISBN 0 7277 2783 4.

9.7
Mechanics of Mine Backfill Matthew Helinski
References The University of Western Australia

Potvin, Y. Thomas, E. and Fourie, A. (eds.). (1995). Handbook on Mine Fill. Australian
Centre for Geomechanics, ISBN 0-9756756-2-1.

Powers, T.C. and Brownyard, T.L. (1947). Studies of the physical properties of
hardened Portland cement paste. Bull. 22, Res. Lab. of Portland Cement Association,
Skokie, IL, U.S.A., reprinted from J. Am. Concr. Inst. (Proc.), Vol. 43, 1901-132, 247-
336, 469-505, 549-602, 669-712, 845-880, 933-992.

Powers, T.C. (1956). Structure and physical properties of hardened Portland cement
paste. Bull. 94, Res. Lab. of Portland Cement Association, Skokie, ILK, U.S., J. Am.
Ceram. Soc. 41, 1-6

Powers, T.C. (1979). The specific surface area of hydrated cement obtained from
permeability data. Materials and Structures, Vol. 12, No. 3, 159-168, Springer.

Qiu, Y. and Sego, D.C. (2001). Laboratory properties of mine tailings. Canadian
Geotechnical Journal, Feb 2001, No.38, 183-190.

Randolph, M. F. and S. Hope (2004). Effect of cone velocity on cone resistance and
excess pore pressures. IS Osaka - Engineering Practice and Performance of Soft
Deposits, Osaka, Japan, n/a: 147-152

Rankin, R. (2004). Geotechnical characterisation and stability of paste fill stopes at


Cannington mine. PhD thesis, James Cook University.

Rastrup, E. (1956). The temperature function for heat of hydration in concrete.


Proceedings of the RILEM Symposium on Winter Concreting, Copenhagen, Danish
National Institute for Building Research, Session B11. 3-20.

Revell, M. (2002). Underground mining at Aurion gold Kanowna Belle. AusIMM


Bulletin, May/June, 37-41.

Revell, M.B. (2004). Paste How strong is it? Proceedings of the 8th International
Symposium on Mining with Backfill, September, Beijing, The Nonferrous Metals
Society of China, 286-294.

Revell, M. and Sainsbury, D (2007) Paste bulkhead failures, Minefill 07, Montreal,
Paper # 2472.

9.8
Mechanics of Mine Backfill Matthew Helinski
References The University of Western Australia

Rouainia, M. and Wood, D.M. (2000). Kinematic hardening constitutive model for
natural clays with loss of structure. Gotechnique, 50(2), 153-164

Rotta, G.V. Consoli, N.C. Prietto, P.D.M. Coop, M.R. and Graham, J. (2003). Isotropic
yielding in an artificially cemented soil cured under stress. Gotechnique 53(5), 493-
501.

Santamarina, J. C., Klein, K. A., and Fam, M. (2001). Soils and Waves Particulate
Materials: Behaviour , Characterisation and Process Monitoring. Wiley, Chichester,
England.

Schnaid, F, Prietto, P.D.M. and Consoli, N.C. (2001). Characterization of cemented


sand in triaxial compression. Journal of Geotechnical and Geoenvironmental
Engineering, ASCE, Vol. 127, No. 10, 857-868.

Schofield, A.N. (1980). Cambridge Geotechnical Centrifuge operations. Gotechnique


30(3), pp.227-268.

Seneviratne, N., Fahey, M., Newson, T.A. and Fujiyasu, Y. (1996). Numerical
modelling of consolidation and evaporation of slurried mine tailings. International
Journal for Numerical and Analytical Methods in Geomechanics, Vol. 20, No. 9, 647
671.

Sideris K. (1993). The cement hydration equation. Zem-Kalk-Gips;12:E337-55, Edition


B.

Sideris, K.K. Manita, P. and Sideris, K. (2004). Estimation of the ultimate modulus of
elasticity and Poissons ratio of normal concrete. Cement & Concrete Composites, No
(not Vol??) 26, 623-631.

Smith, I.M. and Griffiths, D.V. (1998). Programming the Finite Element Method. 3rd
edition. John Wiley and Sons, ISBN 0 471 96542 1.

Tan, T.S. and Scott, R.F. (1988). Finite strain consolidation a study of convection.
Soils and Foundations, Vol. 28, No. 2, 64-74.

Take, W.A. and Valsangkar, A.J. (2001). Earth pressures on unyielding retaining walls
of narrow backfill width. Canadian Geotechnical Journal, Vol. 38, No. 6, 1220-1230.

9.9
Mechanics of Mine Backfill Matthew Helinski
References The University of Western Australia

Terzaghi, K. (1943). Theoretical Soil Mechanics. John Wiley & Sons, New York.

Thomas, E.G. and Holtham, P.N. (1989). The basics of preparation of de-slimed mill
tailings hydraulic fill. Innovations in Mining Backfill Technology, Hassani et al. (eds),
Balkema, Rotterdam. 425-431.

Toh, S.H. (1992). Numerical and centrifuge modelling of mine tailings consolidation.
PhD thesis, The University of Western Australia.

Turcry, P. Loukili, A. Barcelo, L. and Casabonne, J.M. (2002). Can the maturity
concept be used to separate the autogenous shrinkage and thermal deformation of
cement paste at early age? Cement and Concrete Research, Vol. 32, 1443-1450.

Vermeer and Brinkgreve (1998). Plaxis Finite element code for soil and rock analysis.
Balkema, Rotterdam.

Vick, S.G. (1983). Planning, design and analysis of tailings dams. John Wiley and
Sons, New York.

Winch, C.M. (1999). Geotechnical characterisation and stability of paste backfill at


BHP Cannington Mine. Undergraduate thesis, School of Engineering, James Cook
University, Australia.

Williams, D.J. Carter, J.P. and Morris, P.H. (1989). Modelling numerically the lifecycle
of coal mine tailings. Proc. 12th Int. Conf. Soil Mech. Found. Engrg, 3, Rio de Janeiro,
1919-1923.

Yin. J. H and Fang. Z, (2006). Physical modelling of consolidation behaviour of a


composite foundation consisting of a cement-mixed soil column and untreated soft
marine clay. Gotechnique 56(1), 63-68.

Yong, R.N. Siu, S.K.H and Sheeran, D.E. (1983). On the stability and settling of
suspended solids in settling ponds. Part 1. Piecewise linear consolidation analysis of
sediment layer. Canadian Geotechnical Journal, Vol. 20, 817-826.

Zienkiewicz, O.C. and Cormeau, I.C. (1967). Viscoplasticity, plasticity and creep in
elastic solids. A unified numerical solution approach. Int. J. Num. Meth. Eng., No.8,
821-845.

9.10
Mechanics of Mine Backfill Matthew Helinski
References The University of Western Australia

Zienkiewicz, O.C. and Zhu, J.Z. (1992), The superconvergent patch recovery and a
posteriori error estimates. Part 1: The recovery technique. International Journal for
Numerical Methods in Engineering, Vol. 33, 1331-1364.

9.11
Mechanics of Mine Backfill

Figures
Mechanics of Mine Backfill Matthew Helinski
Ch 1 Introduction - Figures The University of Western Australia

Figure 1.1. Schematic of a typical mine tailings based backfill system (contributed by Cobar
Management Pty Ltd).

1.1
Mechanics of Mine Backfill Matthew Helinski
Ch 1 Introduction - Figures The University of Western Australia

Fill deposition
point

10 - 40 m Stope

20 - 100 m
Saturated fill 5-6m Containment
material barricade

5-6m

5-6m

Drawpoint

Figure 1.2. Schematic showing a typical stope filling situation (with typical dimensions).

Figure 1.3. Photograph showing a failed barricade (from Revell and Sainsbury 2007).

1.2
Mechanics of Mine Backfill Matthew Helinski
Ch 2 Background Literature Review - Figures The University of Western Australia

50

45

40
Height above base (m)

Self weight stress v


35

30
v
25

20
Drained
15

10
Undrained
5

0
0 100 200 300 400 500 600 700 800 900 1000
Vertical total stress, v (kPa)

Figure 2.1. Stress distribution down the centreline of a stope assuming drained and
undrained filling.
7

5 Undrained
Height up barricade (m)

analysis
x
4

Drained
3 analysis

0
0 100 200 300 400 500 600 700 800 900
Barricade stress, x (kPa)

Figure 2.2. The impact of drained and undrained filling on barricade stress.

F2.1
Mechanics of Mine Backfill Matthew Helinski
Ch 2 Background Literature Review - Figures The University of Western Australia

v=v+ u

h=v.K0 + u

Figure 2.3. Conversion from vertical total stress to horizontal stress.

1.2

Classified tailings Full-stream


tailings Vick
1 Vick (1983) Impermeable (1983)
base

0.8
(du/dx)/' at surface

Cu tailings, Qiu
0.6 and Sego
(2001)
Ag tailings,
0.4 Qiu and Sego
(2001)

0.2
Permeable
base

0
0.01 0.1 1 10 100 1000 10000 100000
T=m2t/cv

Figure 2.4. Gibson's(1958) consolidation chart with typical minefills.

F2.2
Mechanics of Mine Backfill Matthew Helinski
Ch 2 Background Literature Review - Figures The University of Western Australia

Void ratio, Structure


e permitted space

Compression of bonded
material

Compression of destructured
material

Mean stress, p

Figure 2.5. Comparison between structured and unstructured compression behaviour.

Deviator Critical state line


stress, q
Equivalent
unstructured yield Structured yield
surface surface

Loading surface

p0 pc ps Mean
stress, p

Figure 2.6. Comparison between structured and unstructured yield surfaces.

F2.3
Mechanics of Mine Backfill Matthew Helinski
Ch 2 Background Literature Review - Figures The University of Western Australia

Figure 2.7. Powers illustration of the Cement hydration process (from Illstron et al. 1979).

70

60
Critical
porosity
Cement content required (%)

50

40

30

20

10
Typical cemented minefill
operating range
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
Void ratio, e

Figure 2.8. Relationship between void ratio and binder content to achieve critical porosity and
typical mine backfill range.

F2.4
Mechanics of Mine Backfill Matthew Helinski
Ch 3 Behaviour of Cementing Slurries- Figures The University of Western Australia

Mean effective stress, p

Yield stress increment


py
Cemented yield
py

Cemented
Uncemented compression
yield curve

Uncemented
Void Ratio,
e
yield

Figure 3.1. Incremental yield stress as it is defined in this thesis.

2000
Lines: Eq. 3.6
Points: Data
Unconfined compression strength, qu (kPa)

1600

1200

800

Cc = 10%, 28 day
400

Cc = 5%, 7 day
0
0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95 1
Void ratio, e

Figure 3.2 (a). Relationship between void ratio and q u for CSA hydraulic fill.

F3.1
Mechanics of Mine Backfill Matthew Helinski
Ch 3 Behaviour of Cementing Slurries- Figures The University of Western Australia

2500
Lines: Eq. 3.6
Points: Data
Unconfined compression strength, qu (kPa)

2000 Cc = 6% (All at 28 days)

1500

Cc = 4%
1000

500

Cc = 2%

0
0.5 0.7 0.9 1.1 1.3
Void ratio, e

Figure 3.2 (b). Relationship between void ratio and q u for Cannington paste fill from Rankin
(2004).

1 CSA H.F. qu 0.9


exp
Cannington P.F. qu (max) t 0.16
0.8
qu/qu(max)

0.6

0.4

qu 2 .6
0.2
exp
qu (max) t 0.16

0
0 5 10 15 20 25 30
Time, t (day)

Figure 3.3. Normalised qu against time for CSA hydraulic fill and Cannington paste fill.

F3.2
Mechanics of Mine Backfill Matthew Helinski
Ch 3 Behaviour of Cementing Slurries- Figures The University of Western Australia

2000

1600
R2 = 0.9137
Incremental Go (MPa)

1200

800

400

0
0 0.5 1 1.5 2 2.5
Unconfined compression strength, qu (MPa)

Figure 3.4 (a). Incremental small strain shear stiffness against q u for CSA hydraulic fill.

250

200 R2 = 0.9809
Young's modulus, E (MPa)

150

100

50

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
Unconfined compression strength, qu (MPa)

Figure 3.4 (b). Young's modulus (at large strains) against q u for Cannington paste fill.

F3.3
Mechanics of Mine Backfill Matthew Helinski
Ch 3 Behaviour of Cementing Slurries- Figures The University of Western Australia

1.1
Cc = 5%, 1 day

1 Cc = 5%, 16 day

0.9
Void ratio, e

Cc = 5%, 5 day
0.8
Uncemented

0.7
Lines: model
Points: data
0.6
10 100 1000 10000
Vertical effective stress, 'v (kPa)

Figure 3.5. Comparison between one-dimensional compression experiments and the model
results.

8.0E-06

7.0E-06

6.0E-06
Permeability, k (m/s)

5.0E-06

4.0E-06

3.0E-06
10%
2.0E-06 5%
2%
1.0E-06 Model
0.0E+00
0.66 0.68 0.70 0.72 0.74 0.76 0.78 0.80 0.82
Effective void ratio, eeff

Figure 3.6. Comparison between e ff and permeability.

F3.4
Mechanics of Mine Backfill Matthew Helinski
Ch 3 Behaviour of Cementing Slurries- Figures The University of Western Australia

100%

90%

80%

70%
Percentage passing

60%

50% KB Paste

40%
CSA HF
30%

20%

10%

0%
1 10 100 1000
Size (microns)

Figure 3.7. Particle size distribution curves.

600
Confining
500

KB PF u
400 CSA HF '
Pressure (kPa)

300

200
CSA HF u
KB PF '
100
KB PF: Kanowna Bell Paste Fill
CSA HF: CSA mine Hydraulic Fill
0
0 20 40 60 80 100 120 140 160
Time, t (hr)

Figure 3.8. Pore water pressure (u ) and effective stress changes in triaxial samples hydrating
under constant total stress and undrained boundary conditions.

F3.5
Mechanics of Mine Backfill Matthew Helinski
Ch 3 Behaviour of Cementing Slurries- Figures The University of Western Australia

8 0.01

0.008
6
0.006
4

Received wave amplitude


0.004
Sent wave amplitude

2
0.002

0 0

-0.002
-2
-0.004
-4
-0.006
-6
-0.008

-8 -0.01
-100 0 100 200 300 400 500 600 700 800 900
Time, t ( Seconds)

Figure 3.9. Typical result from bender element test.

1000

900
Total stress
800
u
700
u backup to minimise
Pressure (kPa)

600
effective stress
500
Cumulative u
400 drop
300
p'
200

100

0
0 50 100 150 200 250
Time, t (hr)

Figure 3.10. Typical pore water pressure (u ) and effective stress changes in a triaxial sample
(CSA hydraulic fill material with 5% cement) hydrating under constant total stress and
undrained boundary conditions (with periodic re-establishment of back pressure, to minimise
effective stress change).

F3.6
Mechanics of Mine Backfill Matthew Helinski
Ch 3 Behaviour of Cementing Slurries- Figures The University of Western Australia

3000
10% Binder (Maturity relation, eqn 3.31)
10%Binder (from Go experiment)
2500 5% Binder (Maturity relation, eqn 3.31)
5% Binder (from Go experiment)
Soil bulk modulus, Ks (MPa)

2% Binder (Maturity relation, eqn 3.31)


2000 2% Binder (from Go experiment)

1500

1000

500

0
0 50 100 150 200 250
Time, t (hr)

Figure 3.11. The development of bulk stiffness Ks with time for CSA hydraulic fill:
experimental data (symbols) and Equation 3.31 (lines).

30
10% Binder
25 5% Binder
Incremental u reduction (kPa/hr)

2% Binder
20

15

10

0
0 20 40 60 80 100 120 140 160 180 200 220 240 260
Time, t (hr)

Figure 3.12. Rate of pore water pressure (u ) reduction with time after initial set for various
cement contents for CSA hydraulic fill.

F3.7
Mechanics of Mine Backfill Matthew Helinski
Ch 3 Behaviour of Cementing Slurries- Figures The University of Western Australia

0.001
10% Binder
0.0009
5% Binder
Normalised water consumption rate

0.0008
2% Binder
0.0007
(cm3/cem gram/hr)

0.0006

0.0005

0.0004

0.0003

0.0002

0.0001

0
0 20 40 60 80 100 120 140 160 180 200 220 240 260
Time, t (hr)

Figure 3.13. Normalised apparent water loss rate plotted against time for different cement
contents for CSA hydraulic fill: experimental data compared with Equation 3.32.

4000
10% Theoretical
Cumulative pore pressure reduction, u (kPa)

3500 10% Experiment


5% Theoretical
3000
5% Experiment
2500 2% Theoretical
2% Experiment
2000
0% Experiment
1500

1000

500

0
0 40 80 120 160 200 240
Time, t (hr)

Figure 3.14. Comparison of experimental reduction of pore water pressure ( u ) against time and
adjusted theoretical solution for CSA hydraulic fill.

F3.8
Mechanics of Mine Backfill Matthew Helinski
Ch 3 Behaviour of Cementing Slurries- Figures The University of Western Australia

600
2% Binder Prediction
2% Binder Experiment
Cumulative pore pressure reduction, u (kPa)

500
5% Binder Preciction
5% Binder Experiment
400

300

200

100

0
0 20 40 60 80 100 120 140
Time, t (hr)

Figure 3.15. Predicted and measured reduction in pore water pressure (u ) for KB paste
backfill.

35

30

25
Temperature (C)

tBC =20C t0 =30C


20

15
Monitoring
10 location

5
0m 5m

0
0 1 2 3 4 5
x-coordinate (m)

Figure 3.16. Temperature variation across stope half-space after 20 hours.

F3.9
Mechanics of Mine Backfill Matthew Helinski
Ch 3 Behaviour of Cementing Slurries- Figures The University of Western Australia

Cell

Sample enclosed
Bender elements
in membrane

Cell
pressure Top back-pressure
control control

Bender Element Base back-pressure


processing system control

Pore pressure transducer

Figure 3.17. Hydration test setup.

F3.10
Mechanics of Mine Backfill Matthew Helinski
Ch 4 One Dimensional Consolidation Modelling - Figures The University of Western Australia

a + da

(a+da,t)

x 1
1 x

a (a,t)

Figure 4.1. Schematic representation showing the relationship between a, and x in the
convective coordinate system.

n vw vs w

a V sh
a
t

n vw vs w n vw vs w a
a

Figure 4.2. Schematic representing pore water continuity across an element a.

F4.1
Mechanics of Mine Backfill Matthew Helinski
Ch 4 One Dimensional Consolidation Modelling - Figures The University of Western Australia

i+1

a2

a1

i-1

Figure 4.3. Schematic showing mesh used in CeMinTaCo finite difference approximation.

900

800

700

600
Pore pressure, u (kPa)

Experiment:
500 CSA HF, Cc = 5%, e = 0.7

400
CeMinTaCo
300

200

100

0
0 50 100 150 200 250 300
Time, t (hr)

Figure 4.4. Comparison between the self desiccation pore pressure reduction in a hydration
test and CeMinTaCo output.

F4.2
Mechanics of Mine Backfill Matthew Helinski
Ch 4 One Dimensional Consolidation Modelling - Figures The University of Western Australia

Typical
Typical stope Idealised
Idealised one-
geometry dimensional stope
stope one-
geometry dimensional
stope

Drainage
Drainage
Vertical drainage
through Vertical
through
drawpoint drainage
draw-point

Figure 4.5. Idealisation of the base boundary conditions used to represent a stope in
CeMinTaCo.

600

500 Total vertical stress

No cement
Pore pressure, u (kPa)

400

300
Cc = 5%, no self desiccation
200

100
Cc = 5%, with self desiccation
0
0 50 100 150 200 250 300
Time, t (hr)

Figure 4.6. CeMinTaCo output illustrating the influence of the cement induced mechanisms on
the pore pressure response.

F4.3
Mechanics of Mine Backfill Matthew Helinski
Ch 4 One Dimensional Consolidation Modelling - Figures The University of Western Australia

1.0E-06

k1
Permeability, k (m/s)

1.0E-07 k2

1.0E-08

k3

1.0E-09
0 50 100 150 200 250 300
Curing time, t (hr)

Figure 4.7. Variation in permeability against time

180

160

140 k1
Pore pressure, u (kPa)

120
k2
100

80

60

40

20 k3

0
0 50 100 150 200 250 300
Time, t (hr)

Figure 4.8. Pore pressure against time for the different cases analysed.

F4.4
Mechanics of Mine Backfill Matthew Helinski
Ch 4 One Dimensional Consolidation Modelling - Figures The University of Western Australia

40

35

30
Height above base, h (m)

25 k1 k1
SSPP
5m
20
k2

15

10 k3
Stope
5
Drawpoint
0
0 20 40 60 80 100 120 140 160 180 200
Pore pressure, u (kPa)

Figure 4.9. Pore pressure isochrones for the different permeability cases analysed.

1.30

1.25

1.20

1.15 b=0.05
1.10

1.05
e

1.00

0.95

0.90 b=3.0

0.85

0.80
0 200 400 600 800 1000 1200 1400 1600 1800 2000
Vertical effective stress, v' (kPa)

Figure 4.10. e against v for different damage parameters.

F4.5
Mechanics of Mine Backfill Matthew Helinski
Ch 4 One Dimensional Consolidation Modelling - Figures The University of Western Australia

Pore pressure, u / Effective stress, 'v / Total stress, 500

450
v

400
u
350

300
v (kPa)

250

200

150 '
v
100

50

0
0 50 100 150 200 250
Time, t (hr)

Figure 4.11. CeMinTaCo output for different damage parameters.

400

350
qu (in situ )
Stress, v / Strength, qu, vy' (kPa)

300 v
One-dimensional
250
yield stress

200

150

100 qu (unstressed )

50

0
0 50 100 150 200 250 300 350 400
Time, t (hr)

Figure 4.12. Development of material strength against time for different damage parameters.

F4.6
Mechanics of Mine Backfill Matthew Helinski
Ch 4 One Dimensional Consolidation Modelling - Figures The University of Western Australia

Pore pressure, u / Effective stress, 'v / Total stress, 500

450

400 v

350

300 u
v (kPa)

250

200

150

100 '
v
50

0
0 50 100 150 200 250
Time, t (hr)

Figure 4.13. CeMinTaCo output for different damage parameters in free draining material

400
qu (in situ)
350 One-dimensional
yield stress
Stress, v / Strength, qu, 'vy (kPa)

300 '
v
250

200

150

100 '
qu ( without v )

50

0
0 50 100 150 200 250 300
Time, t (hr)

Figure 4.14. Development of material strength for different damage parameters with free
draining fill.

F4.7
Mechanics of Mine Backfill Matthew Helinski
Ch 4 One Dimensional Consolidation Modelling - Figures The University of Western Australia

7%

Hydraulic fill,
6%
0% cement

5%
v

Paste fill,
Vertical strain,

4%
5% cement
3%

2%

1%

0%
0 50 100 150 200 250 300
Time, t (hr)

Figure 4.15. Axial strain levels for different filling scenarios.

140

120
Pore pressure, u (kPa)

100
u,
CeMinTaCo
80

60
u, In situ
measurement
40

20

0
0 10 20 30 40 50 60 70
Time, t (hr)

Figure 4.16. Comparison between CeMinTaCo and in situ pore pressure measurements.

F4.8
Mechanics of Mine Backfill Matthew Helinski
Ch 5 Two Dimensional Consolidation analysis (Minefill 2D) - Figures The University of Western Australia

Displacement only
z nodes
x
Displacement and pore
pressure nodes

Figure 5.1. Element geometry adopted for plane-strain displacement and pore pressure finite
element calculations in this thesis.

(a) (b)

Figure 5.2. 8 noded isoparametric element (taken from Potts and Zdravkovi , 1999) showing
(a) the parent element and (b) the global element.

F5.1
Mechanics of Mine Backfill Matthew Helinski
Ch 5 Two Dimensional Consolidation analysis (Minefill 2D) - Figures The University of Western Australia

Displacement only
nodes

1 radian Displacement and water


pressure nodes
r2
CL r1
r1
r2

Figure 5.3. Element geometry adopted for axi-symmetric displacement and pore pressure finite
element calculations in this thesis.

6000 6000
Peak
strength Residual
5000 strength Cemented
5000
yield surface
Shear stress, (kPa)

4000
Shear stress, (kPa)

4000

3000 3000
Uncemented
yield surface
2000 2000

Plastic shear Triaxial stress


1000 1000
strain to path
destroy
0 i 0
0 5 10 0 2000 4000 6000 8000
Axial strain, q (%) Mean stress, p (kPa)

(a) (b)

Figure 5.4. (a) shear stress against axial strain and (b) shear stress against mean stress for a
triaxial test.

F5.2
Mechanics of Mine Backfill Matthew Helinski
Ch 5 Two Dimensional Consolidation analysis (Minefill 2D) - Figures The University of Western Australia

1.0
Experimental
Model
0.8
Gt(cem) / G0(cem)

0.6

0.4

0.2

0.0
0 0.2 0.4 0.6 0.8 1 1.2
t/tmax

Figure 5.5. Tangent shear stiffness normalised by small strain shear stiffness against shear
stress normalised by the peak shear strength.

1m

u=0

E = 100 MPa, 4m
k=1e-6 m/hr

u=0

Figure 5.6. Illustration of one-dimensional consolidation problem.

F5.3
Mechanics of Mine Backfill Matthew Helinski
Ch 5 Two Dimensional Consolidation analysis (Minefill 2D) - Figures The University of Western Australia

3.5

3 Initial pressure
Analytical 30 hr
2.5
Depth (m)

Analytical 50 hr
2
Minefill-2D 30 hr
1.5
Minefill-2D 50 hr
1

0.5

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1
Excess pore pressure normalised against initial value, u ex/uex0

Figure 5.7. Comparison between Minefill-2D and the analytical solution for one-dimensional
consolidation analysis.

1m

u=0

E = 100 MPa, 4m
k=1e-6 m/hr,
=19.5 kN/m3

u=0

Figure 5.8. Illustration showing the one-dimensional self weight consolidation problem used in
the Minefill-2D verification.

F5.4
Mechanics of Mine Backfill Matthew Helinski
Ch 5 Two Dimensional Consolidation analysis (Minefill 2D) - Figures The University of Western Australia

4.5
Initial pressure 0 hr
4
Plaxis 6 hr
3.5
Plaxis 20 hr
3
Minefill-2D 6 hr
Depth (m)

2.5
Minefill-2D 20 hr
2

1.5

0.5

0
0 10 20 30 40 50 60 70 80 90
Excess pore pressure, uex (kPa)

Figure 5.9. Comparison between Minefill-2D and Plaxis for a self weight consolidation
problem.

500 kPa

500 kPa Impermeable 500 kPa

Figure 5.10. Numerical simulation undertaken to verify the performance of the self desiccation
mechanism.

F5.5
Mechanics of Mine Backfill Matthew Helinski
Ch 5 Two Dimensional Consolidation analysis (Minefill 2D) - Figures The University of Western Australia

Pore pressure, u / Mean stress, p' (kPa) 600

500

400 u

300 Minefill-2D
Analytical
200

100 p'

0
0 20 40 60 80 100 120 140
Time, t (hr)

Figure 5.11. Comparison between Minefill-2D and the analytical solution for self desiccation.

Phreatic surface
height

E = 1x1020 kPa, 4m
k=5x10-5 m/sec

u=0

Figure 5.12. Numerical geometry for comparison between Minefill 2D and Darcy's law for a
falling head permeability test.

F5.6
Mechanics of Mine Backfill Matthew Helinski
Ch 5 Two Dimensional Consolidation analysis (Minefill 2D) - Figures The University of Western Australia

0.8 Minefill-2D
Phreatic surface height (m)

0.6

0.4
Analytical
0.2

-0.2

-0.4
0 5000 10000 15000 20000 25000
Time, t (sec)

Figure 5.13. Comparison with Minefill-2D and Darcy's law for the flow through the surface
layer of the fill.

300

v, CeMinTaCo
Pore pressure, u/ Vertical total stress, v (kPa)

250

v, Minefill-2D
200

u, CeMinTaCo
150

100 4 m/day
u, Minefill-2D
50
u

0
0 20 40 60 80 100 120 140 160 180 200
Time, t (hr)

Figure 5.14. Comparison between Cemintaco and Minefill 2D.

F5.7
Mechanics of Mine Backfill Matthew Helinski
Ch 5 Two Dimensional Consolidation analysis (Minefill 2D) - Figures The University of Western Australia

Pore pressure, u/ Vertical total stress, v (kPa) 300


v, CeMinTaCo

250
v, Minefill-2D
200
u, CeMinTaCo

150

u, Minefill-2D
100 4 m/day

50
u

0
0 20 40 60 80 100 120 140
Time, t (hr)

Figure 5.15. Comparison between Cemintaco and Minefill 2D with a modified "initial set"
point.

(a) (b) (c)

Figure 5.16. Finite element mesh used to represent (a) coarse mesh, (b) medium mesh and (c) a
fine mesh.

F5.8
Mechanics of Mine Backfill Matthew Helinski
Ch 5 Two Dimensional Consolidation analysis (Minefill 2D) - Figures The University of Western Australia

180

160
Centre-line pore pressure, u-cl (kPa)

140
Medium mesh
120 u
Fine mesh
100

80

60

40
Coarse
20 mesh

0
0 20 40 60 80 100 120
Time, t (hr)

Figure 5.17. Calculated pore pressure in the centre of the stope floor for different mesh shapes.

140

120

Fine mesh
Barricade stress, x (kPa)

100 x

80 Medium mesh

60

40
Coarse
mesh
20

0
0 20 40 60 80 100 120
Time, t (hr)

Figure 5.18. Calculated barricade stress in the centre of the stope floor for different mesh
shapes.

F5.9
Mechanics of Mine Backfill Matthew Helinski
Ch 5 Two Dimensional Consolidation analysis (Minefill 2D) - Figures The University of Western Australia

(a) (b) (c)

Figure 5.19. Vertical total stress contours at the completion of filling for the (a) coarse mesh, (b)
medium mesh and (c) the fine mesh.

300
Vertical total stress, v / Centre-line pore pressure,

v, No arching
250

200
ucl (kPa)

150
ucl, In situ
measurement ucl, Minefill-2D
100

50

0
0 20 40 60 80 100 120 140 160 180
Time, t (hr)

Figure 5.20. Comparison between Minefill-2D and in situ measurements.

F5.10
Mechanics of Mine Backfill Matthew Helinski
Ch 5 Two Dimensional Consolidation analysis (Minefill 2D) - Figures The University of Western Australia

13 m 13 m

40 m

5m 5m

5m 5m

(a) (b)

Figure 5.21. Illustration showing the boundary conditions adopted for the (a) fixed-BC and (b)
free-BC case in the "arching" analysis.

1000
Vertical total stress, v/ Pore pressure, u (kPa)

900
Self weight
stress
800
v
700 v, free-BC
u
600

500

400

300
v, fixed-BC
200

100
u, fixed-BC u, free-BC
0
0 20 40 60 80 100 120
Time, t (hr)

Figure 5.22. Comparison between u and v in a stope with fixed and free vertical displacement
boundary conditions.

F5.11
Mechanics of Mine Backfill Matthew Helinski
Ch 5 Two Dimensional Consolidation analysis (Minefill 2D) - Figures The University of Western Australia

v v

(a) (b)

Figure 5.23. v contours for a stope with (a) fixed vertical displacement boundary conditions
and (b) with free vertical displacement boundary conditions.

40

35

30

25
Height (m)

20
v, fixed-BC
15
Self weight
10 stress

5 v, free-BC

0
0 100 200 300 400 500 600 700 800 900
Total vertical stress, v (kPa)

Figure 5.24. Total vertical stress along the stope centreline for the fixed and free BC.

F5.12
Mechanics of Mine Backfill Matthew Helinski
Ch 6 Centrifuge Modelling - Figures The University of Western Australia

Water overflow line 180

1
Strain gauge set
Rough 2
wall 100
3
505
620
Base plate 4

O-ring seal 5

Base load cells 6

Pore pressure transducer

Figure 6.1. Schematic showing a section through the sample container.

Figure 6.2.(a) Photograph of strain gauged cylinder that was used to represent the stope walls
and (b) the inside of the cylinder showing the rough cylinder walls.

F6.1
Mechanics of Mine Backfill Matthew Helinski
Ch 6 Centrifuge Modelling - Figures The University of Western Australia

Figure 6.3. Photograph of the false base and loadcells that were used in the experiment.

Figure 6.4. Experimental apparatus positioned in a strong box on the UWA geotechnical
centrifuge.

F6.2
Mechanics of Mine Backfill Matthew Helinski
Ch 6 Centrifuge Modelling - Figures The University of Western Australia

18 900

16 Total soil force 800

14 Base load cell force 700

Pore pressure, u (kPa)


12 600
Vertical force (kN)

10 500
Base u
8 400

6 300

4 200

2 Cylinder axial 100


force
0 0
0 1 2 3 4 5 6 7 8
Time, t (hr)

Figure 6.5. Change in pressure and stress during Stage 1 loading.

v, Layer 2
100
Pore pressure increment, u (kPa)

Base u
80

60

40
u

20 Pore pressure
transducer
0
22.3 22.8 23.3 23.8 24.3
Time, t (hr)

Figure 6.6. Incremental change in u during Stage 2 loading.

F6.3
Mechanics of Mine Backfill Matthew Helinski
Ch 6 Centrifuge Modelling - Figures The University of Western Australia

4 100

90
h #5

h
80
Vertical force increment (kN)

Horizontal total stress increment,


Applied force 70
increment #5
h
60

#6
2 50

(kPa)
Base loadcells
force 40

30
1
20
Cylinder (wall)
axial force (#6) 10

0 0
22.4 22.9 23.4 23.9 24.4 24.9 25.4
Time, t (hr)

Figure 6.7. Incremental load / stress distribution in second stage of loading.

Figure 6.6
2.5

2.4

2.3

2.2 Fitted relationship used in the modelling


2.1
Void ratio, e

1.9

1.8
Test data
1.7

1.6

1.5
0 20 40 60 80 100 120 140 160 180 200
Effective vertical stress, v (kPa)

Figure 6.8. Relationship between vertical effective stress and void ratio from the Rowe cell test.

F6.4
Mechanics of Mine Backfill Matthew Helinski
Ch 6 Centrifuge Modelling - Figures The University of Western Australia

1.0E-05
Direct measurement
From cv obtained from displacement rate
From cv obtained from pore pressure dissipation
1.0E-06
Permeability, k (m/s)

Fitted relationship used in the modelling


1.0E-07

1.0E-08

1.0E-09
1.8 1.9 2 2.1 2.2
Void ratio, e

Figure 6.9 Relationship between void ratio and permeability from Rowe cell.

700
Minefill-2D
600
Base pore pressure, u (kPa)

500

Experiment
400

300

200 u

100

0
0 1 2 3 4 5 6 7 8
Time, t (hr)

Figure 6.10. Comparison between measured and calculated pore pressure in Stage 1.

F6.5
Mechanics of Mine Backfill Matthew Helinski
Ch 6 Centrifuge Modelling - Figures The University of Western Australia

18

16 Applied force

14 Base loadcells
Vertical force (kN)

12
Base
10 Minefill -2D Base Force
8

6 Minefill-2D Axial Force


4

2 Cylinder (wall) axial


load (#6)
0
0 1 2 3 4 5 6 7 8
Time, t (hr)

Figure 6.11. Comparison between the measured and calculated load distribution in Stage 1.

350 700

300 Application 600


G o fit Unconfined compressive strength, qu
Small strain shear stiffness, Go

of layer 2
250 500

q u fit
200 400
(MPa)

(kPa)

150 300

100 200

Go measurement
50 100
qu measurement
0 0
0 10 20 30 40 50 60 70 80 90 100
Time, t (hr)

Figure 6.12. Evolution of G o and qu against time for the kaolin with 25% cement mix.

F6.6
Mechanics of Mine Backfill Matthew Helinski
Ch 6 Centrifuge Modelling - Figures The University of Western Australia

v, Layer 2
100
Pore pressure increment, u (kPa)

Minefill-2D, Base u
80
Base u

60

40
u

20
Pore pressure
transducer
0
22 23 23 24 24 25
Time, t (hr)

Figure 6.13. Comparison between measured and calculated pore pressure in Stage 2.

4 100
Minefill-2D, Incremental axial force
Minefill-2D, Incremental base force 90

Increment of horizontal total stress, h (kPa)


Minefill-2D, Incremental h 80
Vertical force Increment ( kN)

3 h
Applied force #5 h 70
increment 60
#6

2 50
Base loadcells force
40

30
1
20
Cylinder (wall) 10
axial force (#6)
0 0
22.4 22.9 23.4 23.9 24.4 24.9
Time, t (hr)

Figure 6.14. Comparison between the measured and calculated load distribution in Stage 2.

F6.7
Mechanics of Mine Backfill Matthew Helinski
Ch 7 Sensitivity Study - Figures The University of Western Australia

120
HFA
100
HFB

PFA
80
Portion passing

PFB
60

40

20

0
1 10 100 1000
Size (micron)

Figure 7.1. Particle size distribution of backfills tested.

1.0E-05 HFA

HFB
Permeability, k (m/s)

1.0E-07

PFA

1.0E-09
PFB

1.0E-11
0 50 100 150 200 250
Time, t (hr)

Figure 7.2. Evolution of permeability against time for different mine backfills.

F7.1
Mechanics of Mine Backfill Matthew Helinski
Ch 7 Sensitivity Study - Figures The University of Western Australia

120

100
PFA

80
Cohesion, c' (kPa)

HFB HFA
60
PFB

40

20

0
0 20 40 60 80 100 120 140 160 180 200
Time, t (hr)

Figure 7.3. Evolution of cohesion against time for different mine backfills.

250

PFB
200
Barricade stress, x (kPa)

150
HFB HFA

100

PFA
50

0
0 20 40 60 80 100 120 140 160 180 200
Time, t (hr)

Figure 7.4. Minefill 2D results of barricade stress against time for different backfill types.

F7.2
Mechanics of Mine Backfill Matthew Helinski
Ch 7 Sensitivity Study - Figures The University of Western Australia

350
Steady state seepage
pore pressure PFB
300

HFA
250
Pore pressure, u (kPa)

HFB
200
u

150

100

50 PFA

0
0 20 40 60 80 100 120 140 160 180 200
Time, t (hr)

Figure 7.5. Development of pore pressure against time for different mine backfills.
40

35

PFB Steady state seepage


30 pore pressure

25
Height, h (m)

20

15

10
PFA HFA
5
HFB
0
0 50 100 150 200 250 300 350 400
Pore pressure, u (kPa)

Figure 7.6. Pore pressure isochrones for different mine backfills.

F7.3
Mechanics of Mine Backfill Matthew Helinski
Ch 7 Sensitivity Study - Figures The University of Western Australia

500

450
Impermeable
400 u
barricade
350
kdp=0.1kstope
Pore pressure, u (kPa)

300

250

200 kdp=kstope

150

100 kdp=10kstope

50

0
0 100 200 300 400 500 600
Time, t (hr)

Figure 7.7. Influence of drawpoint permeability on pore pressure at the base of a stope with
consolidating fill.

45

40

35

30
Height, h (m)

25

20
Impermeable
Barricade
15

10 kdp=kstope
kdp=10kstope
kdp=0.1kstope
5

0
0 50 100 150 200 250 300 350 400 450
Pore pressure, u (kPa)

Figure 7.8. Pore pressure isochrones for consolidating fills with various drawpoint
permeabilities.

F7.4
Mechanics of Mine Backfill Matthew Helinski
Ch 7 Sensitivity Study - Figures The University of Western Australia

450
Impermeable
400
x
barricade
350
Barricade stress, x (kPa)

300

250
kdp=0.1kstope
200

150
kdp=kstope
100

50 kdp=10kstope

0
0 100 200 300 400 500 600
Time, t (hr)

Figure 7.9. Barricade stress against time for different drawpoint permeabilities with
consolidating fills.
250
k=0.1 kHFB
u
200
Pore pressure, u (kPa)

150 k=kHFB

100

k=10 kHFB
50

0
0 100 200 300 400 500 600
Time, t (hr)

Figure 7.10. Pore pressure against time for consolidating fills with different permeabilities.

F7.5
Mechanics of Mine Backfill Matthew Helinski
Ch 7 Sensitivity Study - Figures The University of Western Australia

140
k=0.1 kHFB
120 x

100
Barricade stress, x (kPa)

k=kHFB

80

60

40 k=10 kHFB

20

0
0 100 200 300 400 500 600
Time, t (hr)

Figure 7.11. Barricade stress against time for consolidating fills with different permeabilities.

250
8.0% cement 3.0% cement

200
Pore pressure, u (kPa)

1.5% cement
0.0% cement
150

100

50 u

0
0 100 200 300 400 500 600
Time, t (hr)

Figure 7.12. Pore water pressure against time for consolidating fill with different binder
contents.

F7.6
Mechanics of Mine Backfill Matthew Helinski
Ch 7 Sensitivity Study - Figures The University of Western Australia

200

180 0.0% cement


x
160 1.5% cement
Barricade stress, x (kPa)

140
3.0% cement
120

100
8.0% cement
80

60

40

20

0
0 100 200 300 400 500 600
Time, t (hr)

Figure 7.13. Barricade stress against time for consolidating fills with different binder contents.

160

140
Shear stress, /Cohesion, c' (kPa)

120 3.0% cement,


cohesion
100

3.0% cement,
80
shear stress
60
1.5% cement,
40
shear stress

20 1.5% cement,
cohesion
0
0 100 200 300 400 500 600
Time, t (hr)

Figure 7.14. Comparison between applied shear stress and cohesion for a boundary element.

F7.7
Mechanics of Mine Backfill Matthew Helinski
Ch 7 Sensitivity Study - Figures The University of Western Australia

Interface shear
induced softening

(a) (b)

Figure 7.15. Contour of cohesion at the end of filling for the (a) the 3.0% cement and (b) the
1.5% cement case.

(a) (b)
Figure 7.16 Total vertical stress calculated for the (a) 3.0% cement and (b) the 1.5% cement
case.

F7.8
Mechanics of Mine Backfill Matthew Helinski
Ch 7 Sensitivity Study - Figures The University of Western Australia

450

400
4 m/hr
350
Pore pressure, u (kPa)

300

250 2 m/hr
200

150 0.6 m/hr 0.12 m/hr


0.06 m/hr
100

50 u

0
0 100 200 300 400 500 600
Time, t (hr)

Figure 7.17. Influence of filling rate on consolidating fill pore pressures.

350

300

250
Barricade stress, x (kPa)

4 m/hr

200

2 m/hr
150
.
0.6 m/hr
100
0.12 m/hr 0.06 m/hr
50 x

0
0 100 200 300 400 500 600
Time, t (hr)

Figure 7.18. Influence of filling rate on consolidating fill barricade stress.

F7.9
Mechanics of Mine Backfill Matthew Helinski
Ch 7 Sensitivity Study - Figures The University of Western Australia

400
Kdp=Kstope
350
Kdp=0.1Kstope
300 kdp=10Kstope
Pore pressure, u (kPa)

250

200

150

100

u x
50

0
0 50 100 150 200 250
Barricade stress, x (kPa)

Figure 7.19. Relationship between pore pressure and barricade stress in a consolidating fill.
250

200 u
Pore pressure, u (kPa)

150
kdp=0.1 kstope

100
kdp=kstope

50
kdp=10 kstope

0
0 20 40 60 80 100 120 140 160 180 200
Time, t (hr)

Figure 7.20. Pore pressure against time for non-consolidating fills with different drawpoint
permeabilities.

F7.10
Mechanics of Mine Backfill Matthew Helinski
Ch 7 Sensitivity Study - Figures The University of Western Australia

200

180

160 x
Barricade stress, x (kPa)

140
kdp=0.1 kstope
120

100
kdp=kstope
80

60 kdp=10 kstope
40

20

0
0 20 40 60 80 100 120 140 160 180 200
Time, t (hr)
Figure 7.21. Barricade stress against time for non-consolidating fills with different drawpoint
permeabilities.

(a) (b)

Figure 7.22. Pore pressure profile at the end of filling for (a) kdp=10k stope and (b) kdp=0.1kstope.

F7.11
Mechanics of Mine Backfill Matthew Helinski
Ch 7 Sensitivity Study - Figures The University of Western Australia

250

200 u
k=10 kPFA
Pore pressure, u (kPa)

150

100

k=kPFA
50
k=0.1 kPFA

0
0 20 40 60 80 100 120 140 160 180 200
Time, t (hr)

Figure 7.23. Pore pressure against time for non-consolidating fills with different permeabilities.

160

140
x k=10kPFA
120
Barricade stress, x (kPa)

100

80 k=kPFA

60

40

20 k=0.1kPFA

0
0 20 40 60 80 100 120 140 160 180 200
Time, t (hr)

Figure 7.24. Barricade stress against time for non-consolidating fills with different
permeabilities.

F7.12
Mechanics of Mine Backfill Matthew Helinski
Ch 7 Sensitivity Study - Figures The University of Western Australia

400

350
u
300
Pore pressure, u (kPa)

250 1.5% cement

200

150 .

100 3.0% cement

50 4.5% cement

0
0 20 40 60 80 100 120 140 160 180 200
Time, t (hr)

Figure 7.25. Pore pressure against time for non-consolidating fills with different cement
contents.

300

250
x
Barricade stress, x (kPa)

1.5% cement
200

150

100
3.0% cement
50
4.5% cement
0
0 20 40 60 80 100 120 140 160 180 200
Time, t (hr)

Figure 7.26. Barricade stress against time for non-consolidating fills with different cement
contents.

F7.13
Mechanics of Mine Backfill Matthew Helinski
Ch 7 Sensitivity Study - Figures The University of Western Australia

200
u , Bonded
180 u , Cohesionless
Barricade stress, x / Pore pressure, u (kPa)

interface
interface
160

140
x , Cohesionless
interface
120

100

80
x , Bonded
60
interface
40

20 u x

0
0 20 40 60 80 100 120 140
Time, t (hr)

Figure 7.27. Barricade Stress and pore pressure against time for non-consolidating fill with a
bonded and unbonded interface.

500

450
u
400
Pore pressure, u (kPa)

350

300 2.5 m/hr


250

200

150 0.2 m/hr


0.5 m/hr
100

50

0
0 20 40 60 80 100 120 140 160 180 200
Time, t (hr)

Figure 7.28. Pore pressure against time for non-consolidating fills with different filling rates.

F7.14
Mechanics of Mine Backfill Matthew Helinski
Ch 7 Sensitivity Study - Figures The University of Western Australia

350

300 x
2.5 m/hr
Barricade stress, x (kPa)

250

200

150

100 0.5 m/hr


0.2 m/hr
50

0
0 20 40 60 80 100 120 140 160 180 200
Time, t (hr)

Figure 7.29. Barricade stress against time for non-consolidating fills with different filling rates.

500
Vertical effective stress, v' / Vertical total

450 v(self weight) HFA, - u


S.D. only
stress, v / Pore pressure, u (kPa)

400

350

300
v(self weight)
250

200 '
HFA, v
'
150 PFB, v

100 PFB, u
50
S.D. only

0
0 10 20 30 40 50 60 70 80 90 100
Time, t (hr)

Figure 7.30. Development of effective stress within an element of consolidating and non-
consolidating fill against time.

F7.15
Mechanics of Mine Backfill Matthew Helinski
Ch 7 Sensitivity Study - Figures The University of Western Australia

400

350
Effective self weight
Effective self weight 0.5 m/hr
1 m/hr
Vertical effective stress, v' (kPa)

300
Effective self weight
250 0.05 m/hr

200
v' 0.5 m/hr
150 (minefill-2D)
v' 1.0 m/hr v' 0.05 m/hr
100 (minefill-2D) (minefill-2D)

50

0
0 50 100 150 200 250 300 350 400 450 500
Time, t (hr)

Figure 7.31. Development of effective stress against time in a consolidating fill.

350
1 m/hr
300 Self desiccation - u
Vertical effective stress, v' (kPa)

250

200
0.5 m/hr

150

0.15 m/hr 0.05 m/hr


100

50

0
0 50 100 150 200 250 300 350 400
Time, t (hr)

Figure 7.32. Development of effective stress against time in a non-consolidating fill.

F7.16
Mechanics of Mine Backfill Matthew Helinski
Ch 7 Sensitivity Study - Figures The University of Western Australia

600

v from fill self weight


500
without arching
Vertical effective stress, v' (kPa)

400
25 m 29 m
19 m
300

2m
200

15 m
100
Self desiccation - u

0
0 20 40 60 80 100 120 140 160 180 200
Time, t (hr)

Figure 7.33. Development of effective stress against time at different elevations in a non-
consolidating fill.

F7.17
Mechanics of Mine Backfill

Tables
Mechanics of Mine Backfill Matthew Helinski
Ch 5 Two Dimensional Consolidation analysis (Minefill 2D) - Tables The University of Western Australia

Table 5.1. Material properties adopted for CeMinTaCo - Minefill-2D comparison.

Material Property Value


Uncemented compression parameter (Eqn 4.1), ac 1.2
Uncemented compression parameter (Eqn 4.1), bc -0.1
Rate of hydration parameter, d (day1/2) 1.4
Initial set, t o (day) 0.5
Efficiency of hydration, E h (cm3/g) 0.032
Permeability parameter (Eqn 4.1), c k (m/s) 5.00E-06
Permeability parameter (Eqn 4.1), d k (-) 40
Ultimate cohesion, c max (kPa) 125
Ration between 1D compressive yield and C 3
Friction angle, degrees () 30
Rigidity 6000
Cement Content (% by weight) 3
Damage Constant, b 0.1

T5.1
Mechanics of Mine Backfill Matthew Helinski
Ch 5 Two Dimensional Consolidation analysis (Minefill 2D) - Tables The University of Western Australia

Table 5.2. P.F.-A material properties (from Helinski et al. 2007) adopted for back analysis of in situ test results.

Ultimate
Rate of Friction unconfined
Initial bulk Ultimate bulk hydration Efficiency of Permeability Permeability Ultimate angle, compressive
modulus, modulus, parameter, Initial set, hydration, Eh parameter, parameter, cohesion, degrees strength, qu-f
3
K max-i (MPa) K max-f (MPa) d (day1/2) to (day) (cm /g) ck (m/s) dk (-) c-f (kPa) () (kPa)
50 750 1.4 0.3 0.032 5.0x10-6 40 125 30 433

Table 5.3. Material properties adopted in the investigation of the arching mechanism.

Rate of Friction
Initial bulk Ultimate bulk hydration Efficiency of Permeability Permeability Ultimate angle,
modulus, modulus, parameter, Initial set, hydration, Eh parameter, parameter, cohesion, degrees
3
K max-i (MPa) K max-f (MPa) d (day1/2) to (day) (cm /g) ck (m/s) dk (-) c-f (kPa) ()
35 650 0.9 0.4 0.032 5.36 15 400 35

T5.2
Mechanics of Mine Backfill Matthew Helinski
Ch 6 Centrifuge Modelling - Tables The University of Western Australia

Table 6.1. Material properties for kaolin with 25% cement content.

Initial Ultimate Ultimate


bulk bulk Rate of Uncemented Uncemented Friction unconfined
modulus, modulus, hydration Initial Efficiency of compression compression Permeability Permeability Ultimate angle, compressive
K max-i K max-f parameter, set, to hydration, parameter parameter parameter, parameter, cohesion, degrees strength, qu-f
3
(MPa) (MPa) d (day1/2) (day) Eh (cm /g) (Eqn 5.5), ac (Eqn 5.5), bc ck (m/s) dk (-) c-f (kPa) () (kPa)

40 530 0.9 0.2 0.025 -2.0x10-3 1 3.0x10-15 25 287 23 870

T6.1
Mechanics of Mine Backfill Matthew Helinski
Ch 7 Sensitivity Study - Tables The University of Western Australia

Table 7.1. Comparison of hydraulic fill and paste fill properties.

Hydraulic Fill Paste Fill


(classified tailings) (full-stream tailings)
Less than 10 % finer than 10 m Greater than 15-20 % finer than 20 m
Transported in a turbulent manner Transported in a laminar flow regime at
with high water contents lower water contents
Settles on placement, creating Non-settling on placement
surface water
High permeability Low permeability
High uncemented stiffness Low uncemented stiffness
Generally faster rates of cement Generally slower rates of cement
hydration hydration

Table 7.2. Material Properties.

Property Symbol & Units Hydraulic Hydraulic Paste Paste


Fill A Fill B Fill A Fill B
(HFA) (HFB) (PFA) (PFB)
Initial bulk K max-i (MPa) 80 60 50 31
modulus
Ultimate bulk K max-f (MPa) 630 690 750 950
modulus
Rate of d (day1/2) 1.2 0.9 1.4 2.3
hydration
parameter
Initial set t o (day) 0.3 0.3 0.3 0.3
Efficiency of E h (cm3/g) 0.064 0.032 0.032 0.018
hydration
Permeability c k (m/s) 2.5x10-1 6.2x10-4 5.0x10-6 5.0x10-8
parameter
Permeability d k (-) 80 15 40 10
parameter
Ultimate c max (kPa) 110 120 125 116
cohesion
Friction angle () 35 35 30 28
Ultimate q u-f (kPa) 422 461 433 380
unconfined
compressive
strength

T7.1

Anda mungkin juga menyukai