Vous êtes sur la page 1sur 435
|| Mineral Comminution Circuits Their Operation and Optimisation Mineral Comminution Circuits Their Operation and Optimisation TJ. Napier-Munn S. Morrell R.D. Morrison T. Kojovic JKMRC Monograph Series in Mining and Mineral Processing 2 Series Editor TJ. Napier-Munn & ‘JULIUS KRUTTSCHNITT MINERAL RESEARCH CENTRE "THE UNIVERSITY OF QUEENSLAND Published by: ? Julius Kruttschnitt Mineral Research Centre Isles Road, Indooroopilly, Queensland 4068, Australia 3 Copyright © 1996, 1999 Julius Kruttschnitt Mineral Research Centre, ‘The University of Queensland First edition 1996 Reprinted with minor corrections 1999 All Rights Reserved. No part of this publication may be reproduced, stored in a retrieval system S ‘or transmitted in any form or by any means: electronic, electrostatic, magnetic tape, mechanical, 5 photocopying, recording or otherwise, without permission in writing from the publishers. ‘National Library of Australia Cataloguing-in Publication Entry: Mineral comminution circuits: thetr operation and optimisation i ISBN 0 646 28861 X. 1, Size reduction of materials. 2, Ore-dressing. t I.Napier-Munn, TJ. I. Julius Kruttschnitt Mineral Research Centre, 1 (Geries: JKMRC monograph series in mining and mineral processing, No.2). on2r7 Printed in Australia Cover design and production by Hall & Jones Pty Ltd, Brisbane ‘This book can be ordered directly from the Publisher: email: jkmrc@mailbox.ugedu.au 7 Phone: +61 7 3365 5888 Fax: 46173365 5999 ye DEDICATION eee The authors dedicate this book to Professor Alban Lynch and Dr Bill Whiten who started it all. iii pana nas SAR AAC FOREWORD Jcomminution circuits. Slowly but surely the tools have emerged which have allowed this optimisation goal to become more attainable and successful. Many of the modem methods are based on increasingly sound process models coupled with process simulators which together form a powerful engineering tool O= have always sought to obtain better performance from their None of this has happened overnigh. The Julius Kruttschnitt Mineral Research Centre (and its forerunner group at The University of Queensland) has been involved in the modelling and simulation of comminution circuits since 1962. Professor Alban Lynch produced the first detailed account of the early work in a book published in 1977. We feel that both time and progress since then justify the preparation of a new book. “The major incentive for this book has been the 25th Anniversary of the JKMRC which occurred in 1995. The mineral industry, the JKMRC’s partner in comminution research for over 30 years, and the University of Queensland have generously provided the funding which has allowed the production of this volume. These supporters are acknowledged overleaf. Above all else, the industrial data and associated models which form the foundation of this book are the work of many JKMRC postgraduate students and staff. Twenty-two PAD theses and thirteen Masters theses will give some idea of just how much effort has been devoted to developing the concepts and techniques which are presented. To be really useful, research outcomes have to be applied. We believe the power of the optimisation methods has been unequivocally demonstrated through numerous case studies conducted both by research projects and by the JKMRC Commercial Division, JkTech. Many of the next advances in comminution performance will result from further applications of the techniques to new problems and circuits. We hope the book will assist this process. DJ. McKee Director - JKMRC ACKNOWLEDGEMENTS ‘The writing and production of this book has been made possible by the generous financial contributions of the following organisations. Gale DeBeers MINORCO Anglo American Group @BHP BHP Australia Coal Pty Lid MIM Holdings Limited Newerest Mining Limited a NORTH = North Limited Pancontinental Mining Limited Pasminco Limited Placer Pacific Limited & The University of Queenstand © Western Mining Corporation Limited vi ACKNOWLEDGEMENTS Production of this book would not have been possible without the dedication and hard work of many people. The authors would in particular like to thank the following: ‘+ DrTed Bearman, who made contributions to Chapter 4, + DeStephen Gay, who made contributions to Chapter 3. + DrDon McKee, who read much of the book in draft, made many useful suggestions for improvements, and whose support made the enterprise possible. + Dr David Sutherland, who read part of Chapter 3 and suggested improvements. + Fern Ausland, who undertook the desktop publishing, and Diane Fraser, who prepared many of the drawings. Both showed great professionalism under difficult circumstances to meet apparently impossible deadlines. Libby Hill helped with corrections. + Michael Dunglison and Godfrey Dzinomwva, postgraduate scholars of the JKMRC, who proof-read the final draft. Michael also assisted with diagrams. + Derek Jones of Hall & Jones Pty Lid, who managed the book production, design and publicity very efficiently and tolerantly. + The JKMRC Policy Committee, who gave its support to the project as a celebration of the JKMRC 25th Anniversary. ‘+ All the staff and students of the JKMRC who contributed over many years to the gathering of the knowledge embodied in this book; the companies who funded the work so generously; and the Australian Mineral Industries Research Association (AMIRA) through ‘which much of the research funding was brokered. ‘The authors would like to thank the following organisations for permission to use illustrations in the text: ANI (particularly John Hadaway and Peter O'Halloran), Figures 7-1 and 9.1; British Jeffrey Diamond, Figure 610; Krebs Engineers, Figure 12.20; Nordberg Australia, Figures 12.9 and 12.10; Niugini Mining (Australia), Figures 5.7 and 5.11; Outokumpu Mintec Oy, Figure A32; Pasminco Mining, Figure 54; Russell Mineral Equipment, Figure 82; Svedala Australia, Figures 68, 69, 8.1 and A3.1; Western Mining Corporation, Figures 55,5.8and 5°. As always, any errors and omissions remain the responsibility of the authors. vii nes porae SIR ELIS ER ER ERIN SRIRAM NOLAN ADORE DONE yey PREFACE from a textbook. Nor is it a manual of computer simulation, even though simulation is the principal optimisation tool used. It is a reference book for practising engineers concemed with getting the best out of their crushing and grinding circuits. The emphasis is on the individual process units, their characteristics and the factors influencing their performance. T= is not a textbook, even though it contains much of what would be expected ‘The philosophy followed is to provide the intelligent practitioner with the means to understand his or her circuit better, and so to optimise its performance. A pragmatic description of each unit operation and its role in the comminution circuit is therefore inseparable from an account of the mathematical models used to describe it and represent it in the computer simulator. Likewise, simulation requires data to characterise both feedstock and machine, and the book includes a practical guide to sampling a circuit, mass balancing the data, estimating model parameters and determining breakage and classification characteristics. It does not claim, however, to provide a comprehensive account of the science of rock breakage and comminution. ‘This has been attempted by others in the past. After more than 30 years of development, simulation is now a mature technology and increasingly used routinely by minerals engineers. It is our purpose to provide such engineers with the background knowledge and methodology to enable them to exploit the technology effectively. We feel it important that they understand the basis from which these techniques proceed, and we have therefore included some of the principles involved, with case studies where appropriate. For this reason, and because the approaches described are at the cutting edge of current engineering practice, students studying mineral processing will also find the book of value. ‘The extent of the modern comminution literature is so great as to preclude a comprehensive account. We have therefore included only those references appropriate to the methodologies described, or which we have found useful in their own right. ‘The perceptive reader will have noticed the omission of the word ‘design’ in the title. Design - the specification and sizing of equipment and flowsheets - is a very different business to optimisation, and deserves an independent treatment. It will not be dealt with here, even though of course the knowledge bases for design and optimisation have much in common. TJ. Napier Munn aan CONTENTS FOREWORD. ACKNOWLEDGEMENTS, PREFACE... ‘TABLE OF CONTENTS. CHAPTER 1 INTRODUCTION 11 THETECHNOLOGY OF COMMINUTION.... 1.2. SIMULATION AS AN OPTIMISATION TOOL 13. THIS BOOK AND HOW TO USEIT... CHAPTER 2 MODELS OF COMMINUTION PROCESSES 21 INTRODUCTION enw 22 ABRIEF HISTORY OF COMMINUTION MODELS... 23. CLASSES OF COMMINUTION MODELS, 24 FUNDAMENTAL MODELS 241 Model Principles . 242 Testing of Fundamental Models 243 Energy Considerations... 244 Interactions with Ore Particles 25 BLACK BOX MODELS. 25.1 Principles .. 252 The Population Balance Model - 253 The Perfect Mixing Ball Mill Model 234 —— MultiSegment Mill Model.. 25. Whiten Crusher Model 256 gbicakag Power Modeling. 26 CONCLUSIONS CHAPTER 3 MINERAL LIBERATION 31 INTRODUCTION 32 ‘THE MEASUREMENT AND DESCRIPTION OF LIBERATION .... 33 32.1 Defining Liberation wm ee 322 Measuring Liberatior 323 Describing Liberation sna 33. SIMPLE MODELS OF LIBERATION IN COMMINUTION....- 39 33.1 Weedon Approach 332 Morrell Approach... 34 CONCLUSIONS CHAPTER 4 ROCK TESTING - DETERMINING THE MATERIAL-SPECIFIC BREAKAGE FUNCTION 41 INTRODUCTION 7 42 ROCK BREAKAGE secs 421 Macro Measures of Response. 422 Micro Fracture Mechanisms... 423° Summary: Application. oftoxk Stent “Testing in Comnminution. 43 THE BONDTEST : 431 Introduction. 432 Bond Crushability Test... 433 Bond Rod Mill Grindability Tes. 434 Bond Ball Mill Grindability Test. 435 Discussion of the Bond Method... 44 BATCH GRINDING TESTS. 7 45. SINGLE PARTICLE BREAKAGE CHARACTERISATION 451 Pendulum Tests noon 452 Drop Weight Tes. - 453 JKMRC Rock Characterisation Procedures. 46 OREPARAMETERS IN MODELLING AND SIMULATION 461 AG/SAG Mill Ore Parameters. 462 Crusher Model Appearance Function £63 al Milan Rod Ml Appesrance Function, 47 CONCLUSIONS B@RSBNIseRges CHAPTER 5 SURVEYING COMMINUTION CIRCUITS 5.1 INTRODUCTIOY 52 SOME SAMPLING PRINCIPLES, 521 Introduction. 522 Size of Sample for Size Analysis. 523 Propagation of Error in Calculated Quantities mn 524 Sample Cutters... 53. DESIGN OF COMMINUTION CIRCUIT SURVEYS... 531 Objectives and Requirements.. 7 532 Typical Survey Procedure fora Grinding Circ 1105 533 Data tobe Collected. se 54 SAMPLING PRACTICE. ae 541 Equipment and Resources Required... . 542 Mill or Crusher Feed a rene HO 543 Mill Contents... ae 12 344 Mill Discharge... 19 3.45 Mill Sats or Crusher Product... 19 546 Cyclone Feed .. 347 Cyclone Underflow and Overflow 54.8 Residence Time Distributions 5S SAMPLE ANALYSIS PROCEDURES... 551 Sample Splitting 552 OreSpecific Gravity a2 553 Particle Sizing smn 11 56 DATA ANALYSIS... ~ 123 56.1 Initial Data Inspection. 123 5.62 Mass Balancing. 563 Model Fitting... 57 CONCLUSIONS... CHAPTER 6 CRUSHERS: 61 INTRODUCTION 62 PROCESS OBJECTIVES. 63. PROCESS DESCRIPTION. 136 64 KEY VARIABLES fae 137 65 MODELLING THE CRUSHER. 138 66 CRUSHER POWER PREDICTION... at 67 MODELLING THE CLASSIFICATION AND BREAKAGE FUNCTIONS. 68 CRUSHER MODEL FITTING. : : 69 — MODELLING OTHER CRUSHER TYPES. osname 145 69.1 Vertical Shaft Impact Crushers. 145 692 Horizontal Shaft Impact Crushers. 147 69.3 Toothed Roll Crushers. 148 694 High Pressure Grinding ROlIS vow mcom 148 6.10 CRUSHING OPERATION AND OPTIMISATION, a 151 (CHAPTER 7 AUTOGENOUS AND SEMI-AUTOGENOUS MILLS 71 INTRODUCTION... pales 154 72 MILL DESIGN 721 Shell Design’ 722 Discharge Mechanisms... 723 —_Liftersand Liners. 73. CIRCUITDESIGN 74 PROCESS MECHANISMS AND MODEL... 741 Overview and Model Structure. 742 Breakage .w 743 Material Transport. 78 OPERATION OF AG AND SAG MILLS... 731 Overview. 752 Feed Sizes 753° Ore Hardness on 75a Ball Charge Volume and Ball Size. 755 MillSpeed. 756 Closing a Circuit with Fine Classifier... 787 Closing a Circuit with a Recycle (Pebble) Crusher. 738 Control ee 76 OPTIMISATION - 761 General — 762 Know Your Mill nnn en 763 Define he Optimisation Ctra and Develop Programe 764 Simulation. oa 765 Optimisation Criteria. CHAPTER 8 ROD MILLS &1 INTRODUCTION 82 PROCESS DESCRIPTION wnvsnnmnnn 83. KEYVARIABLES 84 MILL LINERS 85 CIRCUIT CONFIGURATION, 86 ROD MILL MODELLING. 87 SCALING THE ROD MILL MODEL. 88 FITTING THE ROD MILL MODEL. 89 ROD MILL OPERATION. orm 810 OPTIMISATION xiv CHAPTER9 BALL MILLS 91 INTRODUCTION a 92 PROCESS DESCRIPTION. 93 KEY VARIABLES 94 MILLLINERS 9. CIRCUIT CONFIGURATIONS. we 96 THE TRADITIONAL APPROACH - BOND... 9.7 BALL MILL MODELLING... 98 SCALING THE BALL MILL MODEL... 99. FITTING THE BALL MILL MODEL TOTEST DATA.. 9.10 BALL MILL OPERATION. aa 9101 Factors influencing Operation. 9102 Slurry Density . 9.103 Slurry Viscosity « 9.104 Liner Selection 9.105 Ball Size Selection 9.10.6 Classifier Operation. 930.7 Recirculating Loads. 911 MILL POWER ae 9.12 OPTIMISATION OPTIONS... CHAPTER 10 ‘TOWER MILLS AND STIRRED MILLS 101 INTRODUCTION. ee 102 DESIGN. 1021 General 1022 — Stirrers. Z soe 232 1023 Performance Envelope wn 103. GRINDING ACTION. rns 304 MEDIA MOTION 1041 Tower Mills 1042 —_Pintype Stirred Mills. 105 MODELLING AND SCALE-UP OF TOWER MILLS. 1051 Power Draw. 1052 Size Reduction... 106 OPTIMISATION OF TOWER MILLS. CHAPTER 11 na u2 u3 4 ‘THE PREDICTION OF POWER DRAW IN CRUSHERS AND TUMBLING MILLS INTRODUCTION THE PREDICTION OF CRUSHER POWER DRAW. 1121 Introduction... 11.22 Ore-specific Energy - Size Reduction Relationships 1123 The Crusher Size Reduction Model. 11.24 The Crusher Power Modelenn: ‘THE PREDICTION OF MILL POWER DRAW 1131 Introduction. 1132 Charge Shape. 1133 Charge Position. 1134 Mathematical Description of the Variation in Toe and Shoulder Position... 1135 Charge Motion. 1136 Mathematical Description ofthe Velocity Profile. 1137 Power Draw Equations 1138 Overall Model Structure... 1139 No-Load Power 11340 Model Accuracy 11311 Worked Example. ‘CONCLUSIONS CHAPTER 12 ‘SCREENS AND HYDROCYCLONES 121 122 123 124 INTRODUCTION -THE ROLE OF SIZING IN COMMINUTION CIRCUITS ....273 EFFICIENCY CURVES ‘SCREENS AND SIEVE BENDS. : ee 1231 Introduction 1232 Vibrating Screens. 1233 Screening Surfaces : ae 1234 Screening Efficiency rennin BB 1235 Sereen Models 1236 JKMRC Vibrating Screen Model 1237 DSMScreens 1238 Optimisation of Vibr ing Screw and DBM Ser. : 308 HYDROCYCLONES i 324.1 Introduction 1242 Process Description 1243. Hyérocyclone Models. é 1244 Hydrocycione Optimisation -Eifect of Design and Operating Variables CHAPTER 13 ‘OPTIMISATION OF COMMINUTION CIRCUITS 131 INTRODUCTION... 132. APPROACHES 70 OPTIMISATION. 1321 The Optimisation Process 3322 Energy Minimisation... 7 1323 _ Balancing Comminution Stagesimm 133. OPTIMISING CIRCUIT THROUGHPUT AND PRODUCT SIZE o. 133.1. Defining Plant Optimisation Criteria... i: 1332 Generating the Base Case. 1333 Three Case Studies 0.0 Case Study 1 - Optimising a Crusher Cieult Case Study 2 - Maximising Grind at Constant Throughput. Case Study 3 - Increasing Throughput at Constant Grind ... 134 MATCHING THE SEPARATION PROCESS TO THE ‘COMMINUTION CIRCUIT... case Study 4- Masimising Value Production from a Flotation Circuit 135 _ USING AN ENERGY/PRODUCTION MODEL TO OPTIMISE GRIND... 332 332 ase Study 5 - Optimising Production Rates at Minimum Eee 356 358 186 CONCLUSIONS APPENDDXC1: CUBIC SPLINE FUNCTIONS. APPENDIX2: THE EFFECT OF SLURRY VISCOSITY ON GRINDING. 366 APPENDIX3: MEASUREMENT AND REPRESENTATION OF SIZE DISTRIBUTIONS .... ae : 372 APPENDIX4: THE SIMULATOR /KSIMMET AND MASS BALANCE PACKAGE JKMBAL... e o 387 REFERENCES 392 ‘THE AUTHORS $07 INDEX 08 SO OE CHAPTER 1 INTRODUCTION 1a ‘THE TECHNOLOGY OF COMMINUTION fhe word comminution is derived from the Latin comminuere, meaning ‘to make small’. Making small particles out of large particles is a surprisingly pervasive human technology. The breaking of rock, if not quite the oldest profession, certainly has a pedigree stretching far back into pre-history, whether in building shelters, temples or military roads, or in creating tools or weapons. ‘A modem industrial civilisation cannot exist without exploiting a wide range of comminution technologies, from the coarse crushing of mined ore and quarry rock, to very fine grinding for the production of paint, pharmaceuticals, ceramics, and other advanced materials. Rock cutting and blasting can also, without too much semantic risk, be considered the first stage of comminution in mining and quarrying operations, Indeed, there is increasing evidence that integrating the comsninution stages of mining and mineral processing in an holistic way, rather than seeing them as decoupled or even competitive elements of the production process, can produce substantial ‘economic benefits; this is an exciting field of current research. Lest there be any doubt as to the importance of comminution in modem society, a USS. ‘National Materials Advisory Board report in 1981 on approaches to improving the energy consumption of comminution processes estimated that 1.5% of all electrical energy generated in the U.S.A. was consumed in such processes (including the energy required to produce the steel media used in comminution). The report estimated that realistic improvements in the energy efficiency of comminution, including aspects of classification and process control, could result in annual energy savings in the USA. exceeding 20 billion kWh per annum, or about 15% of Australia’s entire annual consumption of electrical energy (in 1993/1994). Nearly all minesite mineral processing operations, including the beneficiation of metalliferous and industrial minerals, iron o1e, coal, precious metals and diamonds, and the preparation of quarry rock, are major users of comminution machinery (mineral sands beneficiation is a notable exception for which nature has done the job already), In the minesite context, the term ‘comminution’ encompasses the following unit : operations: Crushers ‘Tumbling mills + ~ Jaw crushers = Autogenous (AG) mills : ~ Gyratory crushers ~ Semi-autogenous (SAG) mills ~ Cone crushers ~ Rod mills ~ Rolls crushers ~ Ball mills : ~ High pressure grinding rolls ~ Impact crushers Stirred mills Sizing processes 3 = Tower mills ~ Screens 3 ~ Vertical pin mills + Sieve bends - Horizontal pin mills ~ Hydrocyclones ~ Other classifiers (Although sizing processes are not in themselves size reduction devices, they are an integral part of any comminution circuit, and contribute directly to circuit performance and energy utilisation efficiency.) ‘Comminution forms a correspondingly large proportion of any mineral processing plant's capital and operating cost. Cohen (1983) estimated that 30-50% of total plant ; power draw, and up to 70% for hard ores, is consumed by comminution. The proportion of total plant operating cost attributable to comminution (power plus steel plus labour) is variable, depending as it does on the nature of the plant and the ore being treated. However, for a ‘typical’ metalliferous concentrator quoted by Wills (1992) it was exactly 50%, and a similar figure can be inferred from the operating data for a range of metalliferous concentrators given by Weiss (1985). For those operations in which comminution is the predominant unit operation, such as quarries, or iron ore crushing and screening plants, the figure will clearly be much higher. Capital cost figures also vary, but lie in the range 20-50% for most mixed-process plants, The corollary of these statistics is that there is much to be gained from improving the Practice of comminution. Improvements can be of two kinds: 5 © Fundamental changes in the technology, or the introduction of novel technology. ‘Chapter 1: Introduction SEES eC eee ‘+ Incremental improvements in the technology, its application and operating practice. ‘The latter essentially implies optimising the performance of comminution machines, that is, ensuring that the installed capital asset is exploited as efficiently as possible in an economic sense, The benefits of optimisation may be captured as: ‘+ reduced unit operating costs (¢/t treated), ‘+ increased throughput, and thus value production, + improved downstream process performance as a result of an improved feed size specification, ‘or some combination of these. This monograph provides the information to assist the process engineer to realise the benefits of optimisation, through a methodical and technically sound approach to understanding and analysing his or her comminution circuit. Optimisation here implies the engineering process of adjusting machine and cieuit variables to attain some improved operating condition. The book does not discuss mathematical process ‘optimisation procedures such as evolutionary operation (EVOP) and simplex search techniques; these are covered elsewhere (e.g. Bacon 1967, Mular 1972). 12 SIMULATION AS AN OPTIMISATION TOOL Inevitably, in view of its pedigree, there is an emphasis in the book on computer simulation as the principal optimising tool. Simulation here implies the prediction of the steady state performance of a circuit, in terms of stream properties such as mass flow, solids concentration and size distribution, as a function of material properties, machine specifications and operating conditions. Dynamic simulation explores time dependencies for use in plant design and process control system design, and is not considered here. ‘The great power of simulation as an optimisation, and indeed design, tool is its ability to explore many different scenarios quickly and efficiently - the “what if?” questions. ‘This enables the engineer to prescribe with confidence the condition for optimum performance, in terms of maximising throughput or minimising product size for example, without the need for expensive, difficult and often inconclusive plant-scale testwork. At the very least, simulation permits confirmatory plant work to be efficiently designed, leading to reduced costs (e.g. minimising lost production) and better confidence in the final result. Chapter 1: Introduction Circuit optimisation by simulation is not a trivial business. It requires engineering skill Which, like any other skill, needs to be learned through study and experience. Computer simulation is simply the vehicle for the exercise of engineering judgement As is discussed in detail elsewhere in the book, the process model structure which provides the platform for the simulation methodology seeks to decouple and separately estimate material (ore) properties and machine characteristics. Each are described by parameters which must be estimated from real life. The practice of simulator-based optimisation therefore comprises the following steps: 1. Characterising the feed material in laboratory tests. 2, Estimating machine parameters, from plant surveys (‘calibrating’ the models). Running simulations to explore ways of meeting the optimisation criteria through changes in flowsheet, machine or operating conditions. 4. Testing and/or implementing the chosen conditions. Figures 1.1 and 1.2 illustrate the process in more detail. Figure 1.1 shows the procedure for estimating model parameters from plant surveys, and Figure 12 the simulation procedure which uses these parameters for plant optimisation. Note carefully the decision loops involved in judging the quality of data collected from the plant and in estimating model parameters, before committing to the simulations. ‘This monograph provides most of the information required to plan and implement the approach shown in Figure 1.2, The parameter estimation step (Figure 1.1) using a particular simulator will usually be covered in detail in the simulator manual. However the treatment of each process model in the following chapters includes discussion of the interpretation of the model parameters, with some comments on parameter estimation in Chapter 5. In practice, parameters are sometimes chosen from a ‘library’ or from previous experience, rather than fitted to a particular dataset. Chapter 1: Introduction amp alect Suvey | _Semple | matera-specie a ata (ab test) cut y Fitmodsl a No / cata es ime 1 gatstato (machine) b- = Resormuate | 7 aremetars ect | oO [check model chock check | pradiations data model parameters SIMULATOR Use model perameters for simutation Seo Figure 1.2 Figure 1.1: Parameter estimation (calibration’ for models to be used in simulation; data taken {rom plant surveys. Chapter 1: Introduction ‘Setmechine arametors 1 Modis (in Simulator) t ‘Chango fowsheet simutare je} Change fowene ‘Product charactonstcs {and crcui performance Yor SIMULATOR Figure 1.2: Using a computer simulator to seek the conditions for optimum circuit performance Chapter 1: Introduction 13 THIS BOOK AND HOW TO USEIT ‘The book can be read on three levels: ‘+ Asa general introduction to the ideas and methodology underlying the practice of comminution circuit optimisation, including practical ideas which do not require the use of a computer simulation package. + As amore detailed review of specific unit processes, their principles, models, ‘operational features and optimisation options. + Asa companion to the more advanced optimisation methods using computer simulation packages. ‘The book is the outcome of over 30 years of research and consulting in comminution processes by the staff and research students at the JKMRC. It therefore unashamedly reflects the methodologies (and no doubt prejudices) developed in the research, and honed by application in a very large number of case studies in Australia and in many other parts of the world. An important element of JKMRC work in that period has been the encapsulation of much of this knowledge in what was the first commercial user-friendly, PC-based, dedicated mineral processing computer simulator, JKSimMet (Wiseman et al 1991). JKSimMet is supported and marketed internationally by the JKMRC Commercial Division, JKTech, and in 1996 celebrated 10 years as a commercial product. Its main features are outlined in Appendix 4. In some senses this monograph updates the book by the founding Director of the JKMRC, Alban Lynch (1977). When Lynch and his co-workers published their work, ‘personal computers were still only a dream, and optimising comminution circuits by modeling and simulating them on a computer was an exotic activity for the privileged few, though Lynch's book did much to introduce the methodology to the practising engineer. ‘Twenty years later, PCs have invaded our offices and homes, and circuit optimisation by process simulation is a mature (though still developing) technology, widely if not universally practiced. ‘The present monograph reflects the advances of that generation, in terms of models, process understanding, and the entire methodology of circuit optimisation. It also reflects changes in circuit design and operating practice, particularly the widespread use today of autogenous and semi- autogenous grinding, and the introduction of new technologies such as tower mills and new forms of impact crusher. Unlike most conventional textbooks, this book encourages the reader to browse. There is some logic to the ordering of the chapters - the basics, followed by a detailed description of the unit processes, and ending with some ideas and examples of the (Chapter 1: Introduction methodology. However the approach to the book will depend on the reader's experience and objectives. + Someone new to the ideas of mathematical modelling as a prerequisite to ‘optimisation should read Chapter 2 before going further. + Those who simply need a deeper understanding of a particular process should read the appropriate chapter, Nos 6-10 or 12, which have been designed to stand alone as far as possible. + Anyone seeking advice on surveying a comminution circuit is directed to Chapter 5, and if only a reminder of Bond’s test or the JKMRC single particle breakage test procedure is required, the details are given in Chapter 4. + There is even some justification for starting with a review of the optimising methodology in Chapter 13 (the last), since this gives examples of how to go about the task, and so provides a basis on which to think about the particular problem in hand; it can then be re-visited once the reader has familiarised himself or herself with the principles presented in the earlier chapters. ‘There is no assumption in the book that the reader has a computer simulation package available. However the fact is that simulation is increasingly seen as a standard approach, and the book has been written with that in mind. It will not be long before the casual browser is confronted with its value as a routine tool, and the inevitability of its wider use. A brief discussion of the contents of each chapter and appendix follows. CHAPTER 2 gives a general background to the way comminution processes are modelled mathematically. This chapter can be omitted in a first reading, or if simulation is not to be used in the optimisation process. CHAPTER 3 discusses some aspects of the measurement and description of mineral liberation, and its practical use in the prediction of grinding performance. Again, it can be omitted in a first reading, or if liberation is not an issue for the circuit under ‘consideration. CHAPTER 4 is a detailed treatment of methods of assessing the breakage characteristics of rocks in the context of comminution, including Bond’s methods and the JKMRC single particle breakage testing procedures. Characterising the crushability or grindability of the feed material is an essential clement of any optimisation exercise, ‘CHAPTER 5 is a practical guide to surveying comminution circuits, including some discussion of sources of error, sample size, the data that need to be collected, and 8 (Chapter 1: Introduction appropriate sampling procedures for particular streams. It includes a limited discussion of issues relating to data analysis, including mass balancing and parameter estimation (these aspects are handled more fully elsewhere, including in the manuals for the appropriate software). CHAPTERS 6 - 10 deal in depth with the specific unit operations of crushing, autogenous and semi-autogenous grinding, rod milling, ball milling and stirred milling respectively. Each chapter includes a process description, discussion of how the process is modelled, operational features, and comments on optimisation options. CHAPTER 11 describes new and powerful methods developed at the JKMRC for predicting the power draw of crushers and tumbling mills. The management of power draw is an essential element of any optimisation exercise, since as noted earlier it comprises a significant component of plant operating cost. CHAPTER 12's a detailed description of sizing devices such as screens, hydrocyclones and cone classifiers. Again, it includes a process description, models, operational features and optimisation options. CHAPTER 13 describes some strategies for approaching the optimisation problem, with some practical examples to illustrate some of the issues and methods involved. APPENDIX 1 gives a short account of spline functions and how they are used to represent data in JKSimMet. APPENDIX 2 is a discussion of recent JKMRC research which has identified some useful cozzelations between grinding performance and slurry rheology. These trends should be borne in mind when interpreting grinding performance data and considering optimisation strategies. However they are not yet sufficiently quantified to enjoy a formal place in the process models of grinding, APPENDIX 3 reviews the more important laboratory techniques for particle size analysis, with comments on the features of each, and the problems that can be encountered in obtaining reliable results. It is important to emphasise that careful size analysis forms the basis of all comminution circuit surveys and optimisation studies. APPENDIX 4 is a short specification for the steady state comminution circuit simulator JKSimMet, and its associated mass balancing routine, JKMBal. ‘The REFERENCES have been chosen principally to support the statements in the text. Taken together they form a useful bibliography in mineral processing comminution and its modelling and optimisation. However the literature on these topics is very large, and this selection is not intended to be comprehensive. CHAPTER 2 MODELS OF COMMINUTION PROCESSES 24 INTRODUCTION action of water and wind. However for the most part these processes are too slow to be of interest to mineral processors. Speeding up the breakage process requires an intense application of energy. Typically, several kilowatt hours of ‘energy are applied to each tonne of material in mineral comminution processes. This is a great deal of energy. Dropping an ore particle 10 m generates only 1/37 of a kilowatt hour per tonne. A single kilowatt hour per tonne of potential energy requires a lft to a height of 367m. T he mechanisms which cause most rock breakage are those of nature - the ‘The main reason for this large energy requirement is that the particle must be heavily stressed before any substantial breakage occurs. This stress is mostly stored as elastic ‘energy and is lost when the particle fractures. Industrial crushers are about 75% efficient in energy utilisation as breaking rocks one at a time in a laboratory single particle breakage device, such as a pendulum or drop- weight tester (Morrell et al 1992), But even such ideal devices do not use energy ‘efficiently’ in a fundamental sense. Calculations based on theoretical considerations suggest that most industrial breakage processes, especially grinding, are at best only a few percent efficient in terms of the theoretical energy needed to create new surface (Austin ef al 1984). However, nobody has yet developed an industrial scale method for breaking rocks at high throughput which does not require hitting or squeezing them. Perhaps the most practical approach to date for fine size reduction is the high pressure rolls crusher (Section 6.9.4) which retains elastic energy to some extent, although localised breakage tends to relieve stress in adjacent particles. ‘The efficiency of comminution is important because the cost of breakage will be one of the factors determining whether low grade mineralisation constitutes an orebody, For example, almost none of the porphyry deposits (which provide most of the world’s 10 Chapter 2: Models of Comminution Processes LECH C eee SEES ER eee CEP a ee eee copper production) would be economic without the low cost comminution technology which has evolved in this century. Useful models of comminution processes must therefore find a way of representing the application of energy by a breakage machine (such as a crusher or ball mill) fo an ore. “Useful! in this context means a model which can be used in simulation to solve practical problems of comminution circuit optimisation - the subject of this book. ‘The model therefore has to describe two elements of the problem: ‘+ The breakage properties of the rock - essentially the breakage which occurs as @ result of the application of a given amount of specific energy. © The features of the comminution machine - the amount and nature of energy applied, and the transport of the rock through the machine. 22 A BRIEF HISTORY OF COMMINUTION MODELS ‘The modelling of comminution has historically been dependent on the computational power available to perform the necessary calculations. Before computers, all models related energy input to the degree of size reduction expressed as a percent passing size - typically 50, 80 or 90% - or to the proportion of final product generated. In mathematical terms, consider the incremental energy dE required to produce an incremental change, dx, in size (say P80). The following discussion follows Lynch (2977). Tt was always clear from even simple experiments that more energy was required to achieve a similar relative degree of size reduction as the product became finer. Therefore energy and breakage were related by dE = -K.dx/x™ (2a) Researchers in the second half of the nineteenth century applied some ideas from physics to estimate n: + constant energy per unit mass for similar relative reduction: E = Kinja /x) (Kick 1883) 22) + constant energy per unit of surface area generated (Rittinger 1867) @3) "1 Chapter 2: Models of Comminution Processes Bond (1952) based the following intermediate relationship on an extensive experimental investigation into rod and ball milling, the so-called Third Law’ ete BE = K - (24) Fo 4 In all cases x} is the feed size and xz denotes product size - usually the 80% passing size; K is some constant Hukki (1961) reviewed a wide range of industrial devices and concluded that no single relationship was adequate. The regions of applicability are shown in Figure 2.1. At crushing sizes, Kick’s relationship was appropriate. Bond’s ‘law’ worked well for rod and ball mills as might be expected. At finer sizes, Rittinger’s ideas about surface were more plausible, 108 108 10% Conventional Grinding Range 10° 10? 107 10° Energy Consumed (kWh/tonne) 1o+ 10% 109 102 104 108 ‘Size (um) Figure 2.1; Relationship between energy input and particle size in comminution {after Hukki 1961) It is strongly suspected that all these underlying ideas are incorrect. Materials science provides compelling evidence that cracks initiate at points of weakness or flaws in the atomic structure of the material. It is assumed that most rocks contain a distribution of flaws of various sizes - from geological faulting or jointing down to dislocations in crystal structures on the atomic scale. Hence, at large particle sizes, there would be plenty of flaws available to initiate breakage. Indeed any structural imperfections will tend to amplify local stresses and initiate breakage. A glass cutter scoring a pane of glass, o a diamond cutter using a flaw to help cleave a diamond, are examples of initiating breakage in a controlied manner. 2 Chapter 2: Models of Comminution Processes [As particle size reduces, the larger flaws will tend to become external particle surfaces along with many of the smaller ones. When an excess of flaws is available, constant size reduction at constant energy input per unit mass is reasonable, ‘The single particle breakage tests described in Chapter 4 confirm this effect for most ore particles in the range 3-100mm, although some material becomes distinctly ‘softer’ at larger particle sizes, The overall effect of breakage is to reduce the notional intemal area of flaws relative to particle volume. Hence the energy required to achieve a certain degree of size reduction will increase, as suggested by the Bond and Rittinger equations which do NOT depend on geometric reduction but on product fineness. Indeed the definition of the Bond Work Index is the energy per unit mass to reduce a particle from ‘infinite’ size to 80% passing 100 microns. This is reasonable, as the last term in equation 2.4 simply disappears as x, becomes infinite. ‘One other piece of evidence strongly suggests that the flaws control breakage. As the particles become finer stil there must be a size at which they will contain no flaws at all and fracture under stress will be replaced by plastic deformation. Some remarkable experimental work by Schénert (1979) demonstrated this process for several materials at particles sizes less than 10 microns. As might be expected, this brittle/plastic transformation also exhibits some loading rate dependence; thus, brittle fracture is more likely at high loading rates (Inoue and Okya 1994). The chief shortcoming of this ‘ideal’ fracture model is that we have no satisfactory method for quantifying the flaw distribution at present, although scanning electron microscope technology offers some possibilities. Pure energy models provide a useful gross description of total breakage. However, they do not consider particle transport, or the expenditure of energy which does not result in breakage. Further, the underlying assumption of all single point size measures is that the shape of the size distribution remains relatively constant, regardless of breakage history. This is usually true for rod and ball mills but is often in serious error for crushers, autogenous mills and SAG mills. To try to overcome these deficiencies, researchers have considered both breakage and transport at ever increasing levels of complexity. 23 CLASSES OF COMMINUTION MODELS It is fair to say that the development (and certainly the use) of comminution models results from the evolution of the digital computer. ‘The inversion of a 30 x 30 matrix - 13 ‘Chapter 2: Models of Comminution Processes even a symmetric matrix - requires a huge investment in time and concentration without a digital computer. The early modellers, such as Austin, Lynch and Whiten, were severely limited by available computational power. This challenge did result in some elegant and simple models which were also very useful. However, as computational power per unit cost has doubled about every 18 months since perhaps 1960, computational cost has largely ceased to be an issue, except for discrete element models (DEM) and computational fluid dynamics approaches (CFD). An unfortunate side effect is a tendency to increase ‘model complexity. This complexity reduces model utility as tools for understanding. ‘Comminution models can be divided into two main classes: * Those which consider a comminution device as a transform between a feed and product size distribution, and © Those which consider each element within the process. ‘The former are now in common use. The latter require huge computational resources but will become practical as computer power per unit cost continues to rise. For want of better terminology these classes are referred to as Black Box and Fundamental respectively. A black box model aims to predict the product size distribution from an ore feed size distribution, breakage characterisation and experience with similar devices, ie. a data base, encapsulated in an appropriate algorithm. It is phenomenological in the sense that it seeks to represent the phenomenon of breakage, rather than the underlying physical principles. The population balance model is the most widely used example of this class. ‘A fundamental model considers directly the interactions of ore particles and elements within the machine, largely on the basis of Newtonian mechanics; they are also referred to as mechanistic. Adequate computer power for fundamental modelling has only become affordable on the desk top since about 1990, and such models are much less developed than the black box variety. 24 FUNDAMENTAL MODELS: 241 Model Principles ‘The objective of a fundamental model is to generate a relationship between detailed physical conditions within a machine and its process outcome. In practice this means considering a substantial number of elements within a grinding mill, or flows within a classifier. “4 Chapter 2: Models of Comminution Processes a ‘The constraint for this type of modelling is computational power. The main centres for this endeavour initially were the Comminution Centre at the University of Utah under the direction of J. A. Herbst and later RP. King, and the work of P, Radziszewski at the University of Quebec. Inoue and Okya (1994, 1995) have subsequently contributed in this area. To make the computation more manageable, these researchers considered selected zones of each problem. Mishra and Rajamani at Utah (1992, 1994a, 1994b) considered a ball mill as a two dimensional slice of circles. However, the ‘circles’ were provided with the mass of equivalent spheres. Radziszewski et al (1989) reduced computational demand by dividing the mill into zones of impact, abrasion/attrition and little action, and then characterising each. For either approach, the simple application of Newton's Laws of Motion very quickly becomes quite complex. While steel balls (or rods) are approximately perfectly elastic, the ore particles in between them are definitely not - if they were, the mill would not produce any product. Mishra and Rajamani (1994a) approximate ball behaviour using, a spring and dashpot model as shown in Figure 2.2. There is a considerable range of opinion amongst DEM researchers about appropriate methods for modelling elastic/damped interactions. Inoue and Okya (1994), for example, use a non-linear spring with friction and hysteresis effects. shear : + Assembly of balls normal Figure 2.2: Spring-dashpot representation of a contact (after Mishra and Rejamani, 1994a) ‘This model considers the motion of each ball in each dimension i (Le. x, y in 2D or x, y, zn 3D) as a set of vectors. 6 Chapter 2: Models of Comminution Processes IM) + (C] | Sil—> Pe y Breakage out Figure 2.6: Mass balance for a single size fraction inside a mil For a continuous mill, what goes in must equal what goes out (at steady state), and solutions of this mass balance can be derived based on assumptions about the nature of mixing and breakage. For non steady state mills or batch mills, a net accumulation or loss within the size fraction can be assumed (or measured) and somewhat more complicated solutions can be derived. Most practical dynamic solutions consider the balance within a time slice. For ball mill modelling, two variations on this approach dominate the application of simulation techniques. These are the population balance model, and a related model, the Whiten perfect mixing ball mill model, described in the next two sections. A variation on the perfect mixing ball mill model is the Whiten crusher model outlined in section 255. For a long period there was vigorous academic rivalry between these two models. However, Whiten (19722) clearly demonstrated that they were more similar than different. A more recent paper confirms that under certain assumptions both approaches reduce to a common model form (Morrell, Sterns and Weller, 1993). Both models remain in common use and have been used to model many types of comminution devices. They differ essentially in the assumptions made about mixing and residence time, Chapter 2; Models of Comminution Processes enn 252 The Population Balance Model ‘The population balance model was introduced by Epstein (1947) with further development by many, including Kelsall et a! (1969), Whiten (1974), Herbst and Fuerstenau (1968, 1972), and Austin et al (1983). The model concepts are widely used in modem process simulation. ‘This model is sometimes called the ‘first order rate model’ because it effectively assumes that the production of ground material per unit time within the mill depends only on the mass of that size fraction which is present in the mill contents, Le. there is rate constant k; for each size fraction, which characterises its rate of disappearance, ‘Thus rate of breakage = kj 5 (210) where kj breakage rates (min) 5 mass of i size fraction in the mill kj can be back calculated as described below or estimated from batch tests. “While this assumption has the advantage of simplicity itis clearly not justifiable over a wide range of operation, What does remain approximately constant is the number of available impacts in each energy range. If there are too many or too few particles to match the available number of impacts, this key assumption is clearly invalid. ‘The Whiten perfect mixing model (described in the next section) has a similar assumption and similar shortcomings. However, in an industrial context with only relatively small changes in operating conditions, these models are very useful and have certainly contributed to increased grinding efficiency world wide. Because of their inherently short ranges of prediction they must be used as part of an iterative optimisation process, as explained in Chapter 13 To derive the actual model, consider the balance around a single size fraction (Figure 25) as an equation: Feed in + Breakage in = Product out + Breakage out au) In addition to ki, the rate of breakage, a breakage function, bij, describes the fraction of, size range j which reports to size range i after breakage. Now the balance equation can be written. ia fie D by yy = pith: 2) ra 23 Chapter 2: Models of Comminution Processes To estimate the product, this equation is re-arranged to: iat Peeks + Db Sy 213) it Given measurements of pj, {and si, and a suitable function (or matrix) for bj a set of kj values can be calculated in a straightforward manner starting from the coarsest size fraction, The more usual format of this equation uses size distributions for each pulp description designated by a dash ’. id Pie ff = kids + AF by ky sf 214) jt where 2 (the ‘mean’ solids residence time) = Es, / Xf, ‘This mean is a somewhat artificial concept which is not valid at feed sizes of more than 1 or 2mm. Actual use requires an experimental method for estimating 4. The following two paragraphs provide a succinct description and are drawn from Morrell, Stems and Weller (1993). One approach is to measure the response of a mill (and classifier circuit when present) to an impulse of soluble tracer added to the mill feed. Weller (1981) has described how the overall mean residence time of the tracer in the mill (or circuit) is calculated from the concentration-time data gathered from the mill discharge, corrected when necessary by the returns from a classifier. Tt is then assumed that the residence time of the tracer, water and all solids size fractions in the mill (not the circuit) are the same. ‘This assumption allows the mean solids residence time 2. and the total mass of solids in the mill Es; to be estimated, assuming that the density of pulp held in the mill is the same as that in the mill discharge (see also Section 5.4.8). ‘The problem now is to estimate s{. Kelsall etal (1969) and Weller (1981) have shown that the residence time distribution, measured from the impulse tracer data, can be closely fitted by one large and several equal small perfect mixers in series. The fitting yields two parameters; the fraction of hold-up in the big mixer (M) and the number of small mixers (N). As mixing is perfect, then by definition s{q = py, where the subscript A refers to the first mixer. Using the product from one mixer as feed to the next effectively eliminates s{ from equation 2.14 and enables k; to be calculated. Once 24 3 ae PrOnOn Chapter 2: Models of Comminution Processes EEE eget S dT PREECE eRe eeeeePeeceeceS ee jis known, 5 of the total hold-up can be calculated from the pin (i.e. the product of ‘each mixer), N and the total solids hold up. Melvor et al (1990) proposed an heuristic approach. He assumed that the average mill ‘contents were intermediate between mill feed and discharge. Combining this idea with reasonable assumptions about charge voidage and pulp density allows k; to be calculated directly. 2.53 The Perfect Mixing Ball Mill Model Although derived independently, the perfect mixing model (Whiten 1976) is quite similar to the general population balance model, and can be considered a special case Most of the complexities in the general population balance model arise from @ consideration of mixing. The assumption of a perfectly mixed mill removes these complications. Start with the same material balance, ie. Feed in + Breakage in = Product out + Breakage out a) However, the Whiten terminology uses 1; for the rate of breakage and ay for an appearance function in which some of the original particle may remain in that size fraction after breakage. AAs the mill is perfectly mixed, mill contents are related to mill product with a discharge rate, dj, for each size fraction. Peds or sp /d (215) ‘The balance equation around each size fraction is +S ay qgertnss 216) it Now substitute for the experimentally troublesome mill contents: + & [EPL |e py + SPL Qn mal 4 4 This means that the ratio rj /dj can be calculated for each size fraction from a set of actual feed and product measurements, subject to a reasonable form of the appearance function. 2B ‘Chapter 2: Models of Comminution Processes To provide a simple correction for variations in residence time, dj is scaled in terms of ‘mill volume and volumetric feed rate, Q, to the term dj. (Bs ans) where D and L are the diameter and length of the mill. Calculating (t/d*) starting at the coarsest size is very simple, The computation can even be carried out using a pocket calculator (Whiten 1976) and is very easy to program for a spreadsheet or almost any computer language. A full set of (r/d*) values provides an exact transform detween feed and product. However, these rates are poorly determined as individual values. Each value is the result of a set of difference equations. Hence, individual values are less well determined than measured inputs. To overcome this, the (t/d*) function is represented by a cubic spline function as shown in Figure 2.6. us, 8, We, Ln Wd, Ln (Particle size) Figure 2.6: Grinding rates variation with particle size ‘The spline function (Whiten 1972c) provides a smooth curve which is defined by 3 or4 values at user-selected size knots (x; -x,). Three or four values are appropriate for ball mills with five used for AG/SAG mill modelling (splines are discussed further in Appendix). FOUL reson ites: sa 6 Chapter 2: Models of Comminution Processes HEREC eee EE SERRE ‘Touse this model, {" , p{ and mill conditions are measured. A breakage function ais then measured (Chapter 4) or assumed from previous results on similar ores. Suitable values of x -x, are selected, and the best values of r,~ rq are sought using a suitable search technique such as non linear least squares, which chooses values of r) -™4 which ‘minimise the sum of the squares of the differences between the observed and model- predicted product size distribution. The procedure is implemented in a user-friendly form in the simulator JKSimMet. This model has proved to be very robust and adaptable in practice. Where mixing effects cannot be ignored, a mill may be considered as a number of perfectly mixed segments - as in the multi-segment mill model (Kavetsky and Whiten 1982) - see Section 2.5.4 below. Jf the mill content, 5; , is retained and a holdup function added, it becomes a useful dynamic model (Lees 1973). Adding a more sophisticated breakage function leads to the Leung AG/SAG model (Chapter 7). Multi mineral grinding may be considered by dividing a mill into segments which are proportional to feed mineral content and applying the model to each component. Where a suitable breakage function can be defined, the perfect mixing mill model can be used to describe almost any grinding device. If the breakage characterisation method can relate breakage energy to breakage function, then the model can be used to estimate effective comminution energy. A range of additional scale up factors have also been added for ball mills and these are described in Chapter 9. 254 Multi-Segment Mill Model For ball mills with discharges containing a few percent of +2mm material, the perfect mixing model does not predict variations in operation very well. In practice, it is found that there is a dependence of both the discharge rate and breakage rate curves on increasing amounts of coarse material in the ball mill feed, as shown in Figures 2.7 and 28. To improve prediction accuracy, Kavetsky and Whiten (1982) considered a mill ag a series of perfectly mixed segments with some forward and backward mixing allowed between segments. Each segment used constant breakage rates (r). However, while the shape of the discharge function (d) was kept constant, it could be systematically scaled as shown in Figure 2.7. By considering the contents in each segment, some dependence on feed sizing is added which provides reasonably realistic predictions of coarse mill discharge with changes in feed rate and sizing. a Chapter 2: Models of Comminution Processes Discharge rate Coarse feed Ln (Particle size) Figure 2.7: Typical variation of discharge rate in the mult-segment model with increasing feed coarseness Coarser feed Ln(ed) Ln (Particle size) Figure 2.8: Typical variation of breakage rate inthe mult-segment model with increasing feed coarseness wt Chapter 2; Models of Comminution Processes oe nee Oo OoO?$ Figure 2.9 shows the conceptual design where M is the mixing factor between segments. This mode! is complex to implement and is fully described by Kavetsky and Whiten (1982). r 5 Breakage 5 Feed Product yt) | (2) Flow. 4j(h-t) f Pp >> > ate acts Mi Ll et Segments 1 2 Mt M Figure 2.9: Schematic of the mul-segment model structure (Kavelsky and Whiten, 1982) A reasonable approximation of this type of dependence can be made by considering the mill as two halves, with an intermediate product which is the average of the actual feed and product. JKSimMet allows r/d" values for each segment to be constrained to the same values. For the future, a model which considers different degrees of mixing at different sizes may well provide a more realistic response for coarse feed mills. For fine feed mills, perfect mixing is a good approximation. 2.5.5 Whiten Crusher Model ‘The crusher model is explained in more detail in Chapter 6. However, it is outlined here for completeness in the same notation as for the ball mill models. 2 Discharge —_y Function [> f P Appearance <——} ‘A+ C+ S| Function os Figure 2.10: Schematic of the crusher model 2 Chapter 2: Models of Comminution Processes For each size fraction i i+ Dd aijej sy 219) F SEPT 5 220) BaH-si + ajejsy @21) 7 ‘The reader will note that this expression is essentially identical to the perfect mixing ball mill model except that the classification function (cj + dj = 1) takes a different form (Figure 6.7) a=00 fr qK2 a=10- (Hx) Kreg em 2) ft K2-K1, ‘i ‘This model therefore also provides a transform between feed and product given a suitable breakage and discharge function. It is also very flexible and has been widely ‘used. ‘The crusher model reduces an operational data set to K1, K2 and tyg: three numbers instead of two sets of size fractions (K3 is usually held constant at 2.3). ‘As for the ball mill model, a crusher might be considered as a number of segments. Altematively a device such as a vertical shaft impactor which has a discharge function ‘where ¢; = 1 (ie. ital falls through) will only require a breakage function. 25.6 Breakage Power Modelling The techniques detailed in Chapter 4 relate input energy to resulting particle breakage. A careful examination of the perfect mixing ball mill model reveals that 7; is the fraction of each size range in the load which is broken in unit time. If the load weight in each size range and the energy required to cause the desired degree of breakage is, known, the effective breakage energy which is being achieved by the comminution device can be calculated. This is of limited use for the mill models but is very useful for the crusher model (Andersen and Napier-Munn 1988), providing an excellent estimate of required power when scaled by an efficiency factor in the range 1.25 - 1.5 for cone and jaw crushers (see Section 11.2). 30 od we Go jhe Chapter 2: Models of Comminution Processes SEH sve eres seeeecee suc eseeeeeeL eee ee See 26 CONCLUSIONS ‘The ‘fundamental’ models described are the way of the future, but are limited at present by computational power and the understanding of mechanical force and particle interactions. However these models are likely to become more common over the next decade, Black Box models are mature technology. With today’s computers they are easy to use. ‘Their strongest feature is data reduction - that is, an ability to reduce a complex operation to a few numbers or parameters, These parameters can be made independent of ore type and operational factors to some degree, which helps make real world data easier to interpret. The numbers provide both guidance to improved performance (via simulation) and a better basis for decision making, because the effects of ore type changes and operational factors (e.g. tonnage, overflow density, etc.) are reduced, Provided the iterative nature of the optimisation process is understood (see Chapter 13) there is no real alternative to using process models for analysis and optimisation. 4 CHAPTER 3 MINERAL LIBERATION 3a INTRODUCTION components of the ore from the worthless parts, and to concentrate them to make a saleable product. The concentration steps are usually preceded by a comminution stage whose purpose is to break the ore to a size small enough to free or liberate the valuable components from the gangue. T: purpose of a mineral beneficiation plant is to separate the valuable ‘One of the most critical design criteria for mineral processing plants is the choice of size to which the commination step must reduce the host rock (grind size) to ensure an ‘economic level of liberation. Size reduction and liberation are inextricably linked. If the size to which the rock is reduced is insufficient, then a relatively large proportion of the valuable constituents will not bé extracted, leading to loss of potential revenue. If the size chosen is too small, an oversized and overcostly plant with unnecessarily high energy costs will result. To ensure an efficient process, therefore, mineral liberation and its association with size reduction should be well characterised. However, the understanding of liberation, and its coupling to the prediction of size reduction, has not kept pace with the modelling of other operations in mineral processing, for three main reasons: 1. The lack, until recently, of accurate and convenient means to measure the degree of liberation in process streams on a size-by-size basis, 2. The lack of a validated and manageable description of the process of liberation during comminution. 3. The complexity of the problem of describing liberation. Advances in the measurement and modelling of mineral liberation are now beginning to have a positive impact, though much remains to be done. This chapter provides a brief overview of the subject, and describes some simple and practical approaches that ads Chapter 3: Mineral Liberation have been applied recently in coupling liberation information to the prediction of size reduction. 3.2 THE MEASUREMENT AND DESCRIPTION OF LIBERATION 321 Defining Liberation ‘The term liberation, like the term ore, fs essentially an economic one. In practical terms, a particle is liberated when its composition is such that the downstream process can selectively recover it into a concentrate (or conversely, if it is gangue material, reject i). ‘An extreme example is a particle of a single phase (value or gangue) whose composition is then no longer a limiting factor in its recovery (Figure 3.1a). In reality, comminution circuits create a distribution of particle compositions which result in a corresponding distribution of recovery probabilities (rates, in the case of flotation). Sutherland (1989) has shown that flotation rates can be high even when there is a large proportion of multiphase particles which are therefore poorly liberated by any simple definition. In such a case, recovery of value may be high but there is a penalty in reduced grade due to the other minerals present in the composite particles recovered. In flotation, therefore, a particle may be classed as liberated in roughing but not so in cleaning; regrinding is then necessary to increase liberation to attain the required concentrate grade. Figure 3.1 shows an imaginary multiphase section of rock before and after breakage, illustrating the different forms of particle which are the products of comminution. It is clear that the size of each phase and its relationship to every other phase, both before and after breakage, is critical to understanding and describing the liberation process. ‘These relationships are collectively embodied in a property called the texture of the rock. A description of texture is an essential component of any liberation model, as, discussed in Section 3.23. 3.2.2 Measuring Liberation ‘The classical method of measuring the composition of rocks or broken particles is by manual mineral identification, fractionation, and counting under a binocular microscope. Jones (1987) has given a comprehensive account of the procedures involved in such methods, collectively referred to as point-counting, ‘The problems with these techniques include the need for some skill in mineral identification, and the tedious and time-consuming nature of the analysis. A more automated approach became possible by coupling modern image analysis techniques to the use of the scanning electron microscope. Rock and particle images can be produced and analysed in which phase (mineral) identification is achieved by a combination of electron backscatter and x-ray fluorescence, the former identifying 33 Chapter 3: Mineral Liberation phases from their mean atomic number, and the latter particular elements. Jones (1987) describes the general principles. O ® (tested parties & & © eta Progeny alter breakege © come CD var GD varer Figure 3.1: Varying mineral associations in the progeny particles {from breakage of a multiphase composite ‘The development of such automated systems for identifying mineral phases, and determining the degree of liberation in samples taken from mineral processing plants, hhas had a major impact on the study of liberation in comminution. Its principal advantage over the traditional technique of point-counting is its accuracy and speed. QEM‘SEM, developed by the CSIRO Division of Minerals, is one such system which hhas gained wide acceptance. QEM"SEMis an acronym for "Quantitative Evaluation of ‘Materials using Scanning Electron Microscopy” ‘The basic system is described by Sutherland and Gottlieb (1991). ‘The hardware comprises a computer-controlled scanning electron microscope fitted with backscattered electron (BSE) and x-ray detectors. QEM'SEM uses the BSE signals to tocate the particles rapidly and to provide a mean atomic number for the phase being ‘measured. Simultaneously, an x-ray spectrum is collected so that chemical elements can be identified and their relative proportions estimated. This information is used to select, in real time, the target mineral from a list of possible species stored in the computer, ‘The sample to be analysed is set in an epoxy resin to form a briquette which is then carefully polished to present a surface suitable for analysis, The briquette is placed on. a motorised stage which can be driven to scan automatically a number of briquettes. QEM*SEM constructs a digital image of the sample being investigated using a grid patter. As each pixel within the grid is scanned the number of the identified mineral is stored. Subsequent processing of this raw image defines the phase boundaries, and groups the minerals into metallurgically important types, e.g, sulphides, non-sulphide 4 Chapter 3: Mineral Liberation gangue etc., depending on the application. Once complete, the pixel data are then analysed to produce estimates of mineral volumes and degree of liberation. In standard QEM*SEM useage, the data are used to classify particles into 12 compositional classes based on their volumetric content determined by measurement of the exposed surface. At one extreme (0%), particles are classified as barren, whilst at the other (100%), particles are classified as fully liberated. In between, classes are constructed in 10% increments of degree of liberation. Thus, for example, particles in the 40-50% class have 40-50% of their volume exposed in section (Liberated), whilst 50- 60% of their volume is locked with other minerals. The 90-100% class is often chosen to represent the liberated material. ‘The data are presented in tabular format on a size-by-size and mineral-by-mineral basis. They are also described pictorally in false colour maps in which the outline of each particle and the phases present are shown. Each phase has a different colour to aid in identification. Such maps are helpful in obtaining a qualitative indication of the degree of liberation and nature of locking for particular process streams. An example from a gold ore grinding circuit is shown in Figure 3.2. 7 TC] eactground I Carbonatos Siicetes BB Prite Onides Figure 3.2: Example of a QEM"SEM false colour map showing liberated and composite particles. (fepatierned for black and white reproduction) 35 Chapter 3: Mineral Liberation (Other useful measures include the phase-specific surface area, PSSA (Miller eta! 1982), and the liberation indices (Gottlieb ef al 1994). The PSSA is the surface area of a particular phase per unit volume of that phase in the total sample; it varies strongly with particle size, being relatively constant at the coarser (unliberated) sizes, increasing rapidly at the finer sizes as the phase in question liberates. Stereology Most measurement techniques rely on analysing planar images, in two dimensions. ‘What is required from the analysis, however, is information regarding the liberation of discrete particles which are of course 3-dimensional. Stereology is the process by which the 3-dimensional geometric features of an object are inferred from a 1- or 2- dimensional measurement. Without appropriate correction, the transformation of 1- or 2-dimensional measurements into estimates of 3-dimensional properties leads to error which is generally known as stereological bias. To illustrate the problem, Figure 3.3 shows an imaginary particle whose degree of liberation is required to be known. The particle is a binary sphere comprising a 50% white and 50% black phase. If the particle is prepared for analysis by scanning a polished section of the particle, then sections marked a, b, cin Figure 3.3 are all equally possible if the sectioning is carried out at random. The result from measurements taken from section a and c will both indicate, incorrectly, that the particle is comprised ‘of 100% white or black phase respectively. Only section b will give the correct result that the particle is in fact a composite. The net result is that plane section data will always tend to overestimate the number of totally liberated particles in a sample. Figure 3.3; Sectioning of a binary sphere Until recently the correction of stereological bias in liberation measurement has remained an unsolved problem. However, recent research has provided a solution (Gay and Lyman 1995) and it is now possible to determine the true distribution of valuable minerals from sectional data (Chapter 3: Mineral Liberation 323 Describing Liberation Modelling liberation using the new measurement techniques has been an active research field in recent years. The following discussion identifies the main issues. A key assumption at the basis of most liberation models is that of norepreferential breakage. At its simplest this assumption implies that the fracturing of a particle is independent of its texture, or that mineral phases within the particles have the same hardiness or breakage rate. Although this assumption may not always be true, it does allow a simple framework with which to commence liberation modelling. Natural extensions of the non- preferential breakage assumption are that: * there is no grain boundary fracture, + the rate of breakage of particles is independent of texture, + liberation distributions can be predicted from the texture, and + liberation can be modelled separately to comiinution. It is the idea of separating liberation from comminution which leads to a major simplification in the mathematical modelling of liberation, and this simplification is why the non-preferential breakage assumption is so popular. The idea is explained with the following example. Suppose that one has a set of similarly-sized particles, called the parent particles. The particles are ground and the liberation distribution as a function of size is determined by sampling the progeny particles. These are then ground further and a second liberation distribution as a function of size is determined. If the nonpreferential breakage assumption is valid, for the size range in which the two liberation distributions correspond, the two liberation distributions should be the same. Hence the liberation distribution is a function of size only. Thus, liberation is modelled separately to size reduction. If the non-preferential breakage assumption is not vali modelling of liberation cannot be separated from size reduction, greatly increasing the complexity of the system to be modelled. |, the two liberation distributions will not be the same and this means that Barbery (1991) reviewed a number of experiments used to test the assumption of non- preferential breakage and it is clear from this review that the assumption is not universally valid. However, to date, most liberation models are based on the non- preferential breakage assumption and this is likely to remain the case until an acceptable method is developed of modelling liberation when breakage is preferential. Even with the assumption of non-preferential breakage, liberation modelling is still quite complicated and no single model has yet become popular. There are however 37 Chapter 3; Mineral Liberation two approaches, The first approach is to model the process of liberation, based on the knowledge that the liberation process is related to the texture of the ore. For example, if the average grain size of valuable mineral is small, the size at which the mineral is ‘"iberated’ will also be small. Thus the first approach is generally based on using Uberation data to describe the texture of an ore, and from the ore texture predict how the liberation will change if the size is reduced. This approach is called texture modelling. ‘The second approach is based on representing observed patterns in liberation data and using these together with size reduction models to predict changes in liberation that occur as a result of changes in size reduction. To date the JKMRC has been following the second approach, with methods developed by Weedon (1992) and Morrell (Morrell, Dunne and Finch 1993). These approaches are discussed in Section 3.3. Texture modelling ‘The first model that can be regarded as a texture model is that of Gaudin (1939) which ‘was subsequently improved by Wiegel and Li (1967). In Wiegel and Li's model a cubic lattice is constructed, built from cubic cells with each cell having a width a. The texture is made by defining each cell as belonging to one of two phases. The assigning process is random. The size réduction process is simulated by specifying that particles are also cubic with size B. The texture of these particles is constructed by placing these cubic particles randomly within the original lattice but oriented in the same direction as the original lattice. The mathematics for the simulation is achieved by analytical methods. ‘One can see that the process of modelling involves two main steps: use available information to determine the three dimensional texture, and superimpose particles over the texture to simulate size reduction. A number of other liberation models use similar steps. These include the models of Barbery (1991) and Meloy (1984). They differ mainly in what texture is used. Barbery uses a number of different textures of Which the most common is the Boolean texture based on Poisson polyhedra. Although Meloy and his co-workers also use a number of textures, the texture which provides the conceptual understanding of Meloy’s method is a block consisting of equidistant parallel rectangular blocks. The information used for both Barbery’s and Meloy’s models is the average grade of the valuable mineral phase and the interfacial surface area, Both of these parameters can be determined from examining sections. Note that the interfacial surface area is directly related to the valuable mineral grain size. That is, fa small average mineral grain size will tend to lead to a large value of interfacial surface area, Barbery’s model is generally regarded as more realistic than Meloy’s, but it is mathematically very complex and therefore difficult to understand. Meloy’s texture is Chapter 3: Mineral Liberation SSE EECEECEC eee gage ECE eee eee regarded as too simplistic but its mathematical description is quite understandable. Unfortunately, the challenge for a practical model is for it to be both realistic and also simple to understand. King’s model (1979) provides an interesting alternative. Whereas the texture modellers discussed so far use information to create the texture in three dimensions, King’s approach is to model the texture in one dimension, and then superimpose on this texture linear intercepts that would correspond to particles of a particular size. In this way the liberation distribution for the linear intercepts is simulated. This approach «greatly simplifies the computational requirements of the model. The key then to obtain the volumetric iberation distribution is to apply a stereological correction to the linear intercept distribution (see above). King’s model has been subsequently improved by King and Schneider (1993). Like Barbery’s and Meloy’s models, King and Schneider's model is also based on using the interfacial surface area. ‘The relationships between texture, liberation modelling and stereology were also demonstrated by Davy (1984), who derived a direct relationship between the covariance of the texture of an ore and the variance of the liberation distribution as a function of size. This covariance function is measured by using linear intercepts or planai’sections, and was actuelly the basis of Barbery’s Boolean texture model (1991). Use in Simulation Packages “Although none of these liberation models have yet been widely applied, some of them have been included in computer simulation packages. Principally, these include Wiegel’s model which is included in SIMBAL (Hejjden 1991), and King and Schneider's model which is incorporated in MODSIM (Schneider and King 1995). The implementation of the models is still at an early stage. However it is expected that liberation models will in due course become an important component of mineral processing simulation packages. 33 ‘SIMPLE MODELS OF LIBERATION IN COMMINUTION To date, two approaches to describing liberation during size reduction have been developed at the JKMRC. Neither can yet be regarded as constituting a true model of ‘mineral liberation in that they describe the process of liberation. Rather they are ways of representing observed patterns in liberation data, and using these to predict the changes in liberation that occur as a result of changes in size reduction in comminution. 39 Chapter 3: Mineral Liberation 33.1 Weedon Approach Lelines Weedon (1992) developed a means of representing liberation data in a manner analagous to the t-curves used by Narayanan and Whiten (1988) in their breakage characterisation work (Section 4.5.3). Weedon took a number of rock samples and broke them using the twin pendulum device described in Section 4.5.3 He then sized the products and subjected them to QEM'SEM analysis. The result was a series of data sets which contained information on the degree of liberation of selected minerals in ‘each product size class following breakage of a parent particle, A typical result is shown in Table 3.1. In this particular test, 100 galena-bearing particles from Pasminco Broken Hill rod mill feed in the size range -5.6 + 4.75mm were broken at the same energy level. The total amount of galena in each product size fraction, as well as the amount of fully liberated galena, was then determined by QEM"SEM. ‘The data show that in this case 54.15% of the galena in the original particle appeared as berated galena after breakage in the -300jum size fractions; +300um was not analysed by QEM'SEM, but can sometimes be inferred by extrapolation (Weedon et al 1990). In ‘cumulative form these data are represented in Figure 3.4. This curve is analagous to the familiar cumulative percent passing size plots. 60 50 40 30 20 Cumulative % berated 10 ° 0 50 100 150 200 250 300 350 Size (um) Figure 2.4: Cumulative percent liberation curve for galena Such curves can then be used to represent the cumulative proportion of total mineral present in the original sample which is liberated below some size of interest, after a particular breakage event. 40 Chapter 3: Mineral Liberation ee eee eee ‘Table 3.1: Example of liberation data from a pendulum breakage test (after Weedon, 1992) ge EeceeCee eee ees ecEeeeeeeeeee ese CeCCececeee ee Cece BEFORE PENDULUM BREAKAGE Feed Sample Size 4.75 - 5.6mm ‘Weight of sample (100 particles) 8.659 Percent galena in sample (chemical assay) | 4.9% ‘Weight of galena in sample 0.429 AFTER PENDULUM BREAKAGE 7 2 3 4 5 6 7 8 Size | Weight | %galena| Weight | %galena| Weight | Percent | Cum, % Gm) (@ | insize | gelena | Iiberated | galena | liberated | liberated (%) | insize | insize | tiberated | (%) ~ (a) (%)__| in size (g) +300 | 5.26 nla : nla - : 5 3004212) 0.81 8.68 0070 | 808 | 0.056 | 1357 | 64.15 -2124150) 0.68 913 | 0062 | 823 | 0051 | 1221 | 40.58 -150+106] 0.60 5.79 | 0.028 | 821 0.023 5.68 28.37 “106+75) 0.41 972 | 0039 | ats 0.032 7.76 | 22.69 -75+53 | 0.32 7.32 0023 | 88.8 0.02 497 14.93 -53+38 | 0.24 7.84 oots | e69 | oo | 391 9.96 -38 043 658 | 0028 | eas | 0025 | 5.95 5.95 TOTALS| 865 - 027 - 0.22 5415 | = Notes Col. 2is weight measured after breakage Cols. 3 and 5 obtained directly from QEM‘SEM Col. 4= Col, 2x Col. 3/100 Col. 6 = Col. 5 x Col. 4/100 Col. 7 = Col. 6 x 100/total weight of galena in sample (0.428) Col. 8 = Col. 7 cumulated Using the tyo parameter analogy (Chapter 4), an Lyq9 parameter can then be defined as the cumulative percentage of mineral liberated at a size smaller than 1/100th of the original rock size. It was found that this parameter could be related to other members of the L-parameter family (Lo, Lgo, Lizs. L250, Liooo) by a series of L-lines. An ‘example of such a series is given in Figure 2.5. a (Chapter 3: Mineral Liberation ‘Thus given the family of L-lines in Figure 35, then for a given value of Lg the entire ‘cumulative liberation curve of Figure 3.4 can be generated. 50 40 30 20 Cumulative % liverated 10 o 2 4 6 8 10 12 14 “L' Parameter Figure 3.6: Example of a family of Lines Predicting Mineral Liberation in Ball Mills ‘The breakage parameter r/d! is used in the perfect mixing ball mill model (Chapter 9) to describe the breakage of particles in each size fraction. Weedon applied the model on a mineral-by-mineral basis and generated a mineral-specific breakage parameter (tm/d"%,). He assumed that the individual minerals broke in a similar manner to the host rock and therefore used the same breakage distribution (appearance) function for ‘each species. This structure provided a prediction of the size distribution of a mineral phase in the product. To these data Weedon then applied his L-parameter concept to edict the size-by-size degree of liberation of each mineral phase. Just as with the tm/ dy, distribution, which had to be fitted to feed and product data, the L-parameter distribution also required feed and product liberation data to be determined. Using data from Pasminco's primary No. 1 and secondary No. 1 ball mills the L-parameter ‘was fitted. The results for galena are shown in Figure 3.6. A similar shape of distribution is seen to result, though the values for the secondary mill are significantly higher over most of the size range. To test his model, Weedon used the L-parameter distribution from the No. 1 primary ball mill to predict the performance of the No. 2 and No. 3 mills which were run under different conditions, The results showed that the overall degree of liberation was predicted well (Figure 3.7). However the size-by-size predictions showed considerable overestimation of liberation in the finer size fractions, and the opposite in the coarser 42 Chapter 3: Mineral Liberation ‘Secondary Ball Mil parameter Primary Ball Mill ° 1000 2000 Size (um) Figure 3.6: Example of the L-parameter distribution for a primary and secondary ball mill 100 80 60 40 Predicted 20 0 20 40 60 80 100 ‘Observed Figure 3,7: Observed vs predicted degree of beration using the Weedon model 33.2 Morrell Approach In Morrel's approach (Morrell, Dunne and Finch 1993) no attempt was made to ‘mathematically reduce liberation data. Instead the size-by-size degree of liberation was used in its original form, obtained from QEM'SEM, coupled to a population balance size reduction model. The basis for using this approach was the observation that, from feed and product samples which had very different size distributions, it appeared that the size-by-size degree of liberation varied by amounts which were 4a Chapter 3: Minoral Liberation within their expected measurement errors, i.e. for a given mineral species the relationship between degree of liberation and size is the same, regardless of the conintinution treatment of the ore. This is essentially the random or non-preferential breakage referred to earlier in Section 3.23. An example is shown in Figure 38. bee Tay Tate) & 80 b = A feed § cE 3 Bb RrECT CS ia eas B69 Fenton 4 £ 9f jbl. Qg- 4 cl é fone fneoet ep ea ete ees ee o 1 2 3 40 50 ‘Size (microns) Figure 3.8: Degree of sphalerte iberation in a tower mill feed and product stream Here the y-axis is the percentage of mineral (in this case sphalerite) in a particular size interval which is liberated, a ‘liberated’ particle being defined as at least 90% mineral; this is sometimes termed the CLY90 (Miller et al 1982, Gottlieb et al 1994). This curve was subsequently referred to by Gottlieb ef al (1994) as the liberation profile. The liberation modelling approach of Morrell was originally applied to pyrite liberation in a ball milling circuit. This approach was subsequently modified and extended for application in tower milling. The conceptual model is shown in Figure 3.9. The feed is split into gangue, locked, and liberated valuable streams. For simplicity, only a 2-component system is shown, although multi-component systems are possible. Each stream is modelled in terms of size reduction, using a population balance model (Morrell, Sterns and Weller 1993). Within the mill all streams are reduced in size. However, in the case of the locked stream, liberation of valuable and gangue also occurs. The locked fraction, therefore, divides into three streams, two of Which join those of the liberated valuable and gangue streams. The third comprises the locked valuable remaining after breakage. Within the model this process is, described using two relationships (A and B). The combined effect of these is to reproduce the relationship typical of that shown in Figure 3.8, At the same time they enable the generation of the gangue, liberated and locked stream data necessary to maintain a mass balance across the mill. Chapter 3: Mineral Liberation Gang Locked valgable Liberated Lookea! vaisabie + valuable Lberates valuable Figure 3.8: Schematic diagram of conceptual iveration model for a comminution machine (in this case a tower mil If itis assumed that the data in Figure 3.8 can indeed be represented by a single line then a reasonable prediction of the data is obtained (Figure 3.10). 100 80 Observed (%) 20 40 60 80 400 Prodicted (%) Figure 3.10: Observed vs predicted degree of sphalerite liberation using the Morrell model 45 Chapter 8: Mineral Liberation Relationship A is represented by the example curve shown in Figure 3.11. It describes the relationship between size and the fraction of locked mineral which becomes liberated during size reduction. Relationship B (Figure 3.12) is effectively a size-by- size grade curve relating to the fraction of the locked rock within each size class which contains valuable mineral. Data from QEM’SEM analysis of one of the tower mill streams were used to generate both relationships, which were represented in the model as tabulated size-by-size functions. Using a representative ore sample, curves A and B should be determinable from QEM*SEM analysis of the products from controlled breakage tests, laboratory batch grinding tests or an existing grinding installation. 06 05 4 oat oab oa Fraction berated oat 4 ° 10 ; 100 size um) Figure 3.11: Relationship between size and the fraction of composite material which sphalerte during size reduction (back calculated) iberates 08 os F 4 2 x anos, eae 4 2 ~ x & os} ie B oo2f 4 & ott 4 ieee EEE ee eres 10 100 ze (um) Figure 3.12: Relationship between size and the fraction of locked material which is sphalerite (back calculated) %6 Chapter 3: Mineral Liberation fsa SEE ese Ecce eee To test the model, data from two surveys of an industrial tower mill were used. The conditions under which the surveys were conducted are given in Table 3.2. Table 3.2: Tower mill survey conditions Variable Suvey 1 ‘Survey 3 Feed Rate (uh) liberated ZnS 327 191 locked ZnS 9.90, aot gangue 3.03 1.84 total 162 766 Feed Size (dgo} (rm) liberated ZnS 473 328 locked ZnS a7 486 gangue 203 24 total 53.6 349 Product Size (dgo) (mm) liberated ZnS 165 104 locked ZnS 289 13.9 gangue 124 90 to te wos | Product Pulp Density (% solids) 63.0 622 Power Draw (kW) 164, 205: ‘There are three notable differences between the two surveys: + an increase of 25% in the tower mill power draw in survey 3 © 253% lower feedrate in survey 3 ‘+ amuch finer feed and product size distribution in survey 3 Using survey 1 data, the A and B relationships were determined (Figures 3.11 and 3.12) and the breakage rates for liberated sphalerite, locked sphalerite and gangue subsequently calculated from the perfect mixing ball mill model (Figure 3.13). ‘The performance of the tower mill under the conditions of survey 3 was then predicted using the model derived from the survey 1 data, The results are shown in Table 33. Tt a ‘can be seen that the overall increases in liberated sphalerite and gangue, as well as the ‘consequent reduction in gangue flowrate, are well predicted. In general so are the size distributions of each of the streams, although there was a tendency for the model to predict a somewhat coarser locked sphalerite product size distribution than was observed. ar Chapter 3: Mineral Liberation 100 3 T n Breakage Rate (hr') T 1 o locked Zns * liberated 2ns = ganguo 0.4 10 100 size (um) Figure 3.13: Breakage rates of gangue, locked and liberated sphalerite ‘Table 3.3: Comparison between observed and predicted sphalerite liberation size Mill Discharge - Cumulative % Passing (am Liberated ZnS Locked ZnS Gangue Total (Observed] Prediced | Observed] Predeled | Observed] Predicted | Observed] Prediaed 38 | 986 | 054 | 959 | a8 | 982 | s72 | o79 | a6 188 | 79 | 771 | 592 | 572 | 923 | 39 | 731 | 739 ga_| 479 | 491 | 365 | 377 | 557 | 565 | 45.0 | 405 wr | 280 | 267 | 225 | 221 | 202 | 208 | 766 | 765 34 CONCLUSIONS Modern techniques in liberation measurement have enabled accurate data to be generated relatively easily (though at a cost). Research into modelling liberation in comminution machines has already benefitted from the availability of such data, though much work still needs to be done. Research programmes are currently in place which should result in advances in this field as more data become available. This chapter has described two simple methods of integrating liberation information from QEM'SEM into the ball mill model, in order to predict liberation as a function of size reduction, 48 CHAPTER 4 ROCK TESTING - DETERMINING THE MATERIAL- SPECIFIC BREAKAGE FUNCTION 41 INTRODUCTION in the mining industry, and the literature on the subject is correspondingly large. In the work related to comminution, the main objective has been to derive reliable ways of assessing how a particular material breaks in a comminution machine such as a crusher or mill. Clearly some ores, coals o quarry rocks are harder than others, and the appropriate machine, its size and its consumption of energy will depend not only upon the duty (feedrate and desired product size) but also upon the ‘hardness’ of the material to be broken. Hardness itself isa difficult property to define, and its relationship with the machine will not necessarily be straightforward: an autogenous mill treating a soft ore may have a low capacity because of the lack of hard ‘ore media to promote grinding; a cone crusher of a given set will crush a wide range of ores to a similar product size, but will draw different powers in doing so. T= is a long history of research into rock and particle breakage characterisation Standard rock mechanics tests of strength, such as fracture toughness and uniaxial compressive strength, are not usually seen as appropriate in comminution studies, ‘They provide information on the stress required to cause failure under particular modes of loading, in the form of a single hardness or strength parameter. In comiminution, itis also important to identify the product size distribution resulting from applying a particular breakage mechanism to a given feed size, and the energy required to generate that product size. In particular it is the enezgy-size reduction relationship which is the main focus of the laboratory tests developed to assist in comminution equipment specification, and circuit design and optimisation. Such tests include the Bond tests, batch grinding, and single particle testing. However evidence is accumulating that some rock mechanics fracture tests can also be usefully interpreted in cormminution terms, as discussed below. ‘This chapter describes the commonly available laboratory breakage characterisation tests, with particular respect to comminution energy and size reduction, and the use of these tests in predicting the performance of industrial comminution machines. It 49 Chapter 4: Rock Testing - Determining the Material Specific Breakage Function commences with a brief review of the fundamentals of rock breakage, and the measurement of rock properties. It then discusses the industry standard tests such as the widely used Bond tests, The chapter concludes with a detailed description of the single particle breakage tests developed at the JKMRC, and their application to the determination of material-specific parameters in the comminution process models described elsewhere in the monograph. 42 ROCK BREAKAGE ‘The breakage of rock in comminution depends upon how the rock material behaves under applied load. The form of the load application is dependent upon the particular ‘unit process involved (crusher, mill, et.) Rock, as any other engineering material, will exhibit: macro measures of response, such as ultimate strength (both in compression and tension), and properties that describe its response to loading (Young’s Modulus, Poisson's Ratio). * microfracture mechanisms: crack initiation and propagation. 4.2.1 Macro Measures of Response Compressive Loading Figure 4.1 shows a stress versus strain plot for the uniaxial compression of cylindrical rock samples (after Jaeger and Cook 1979). The region OA is the elastic zone, and it is the ratio of stress (force per unit area) to strain (relative deformation) in this region that in its simplest form gives the Young's modulus. At point A the material is said to yield; A is known as the yield point and the corresponding value of applied stress is referred to as the yield stress (00). Between points A and B the rock continues to deform without losing its ability to resist load; this region is known as the ductile region. Once past point B, the ability of the material to resist load decreases with increasing deformation and the rock enters the brittle region. The point B therefore denotes the transition from ductile to brittle behaviour and it is the stress at this point that defines the Uniaxial Compressive Strength (UCS) of the material, denoted as Cg. Sudden failure of the material will then ‘occur somewhere in the BC region of the graph. In many cases, however, sudden catastrophic failure of the sample will occur very near to the point B, due to the inability of the machine-rock system to cope with the conditions Chapter 4: Rock Testing - Determining the Material-Specttic Breakage Function In terms of macro measures of response, therefore, the most comunonly quoted values relating to the compressional properties of rock are Uniaxial Compressive Strength (UCS), Young's modulus of Elasticity (E) and Poisson’s Ratio (v). tress (MPa) Figure 4.1: (a) idealised stress-strain response for a cylindrical sample of rock under Uniaxial compression, () Typical stress-strain curve for quartzite for comparison with the idealised example. Only a limited post yield region is identified, and neither ductile nor a noninear zone is tecorded, Failure is catastrophic. ‘All these parameters are obtained from the slow compression test of a cylindrical rock sample, using a stiff testing machine, in accordance with the recommendations of the International Society for Rock Mechanics (ISRM). The UCS is simply the maximum force recorded at the point of ductile-brittle transition, divided by the cross sectional area of the sample. It is a measure of strength that is important in designing structures in mining and civil engineering. Typical values for UCS range from weak limestone at < 50 MPa to highly competent microcrystalline hornfels at 450 MPa. ‘Young's modulus can be quoted in a variety of ways depending upon the degree of linearity of the response; however it is most easily defined as the slope of the elastic region of the stress versus strain graph (Figure 4.1, region OA). Young’s modulus values range from <10 GPa for marble and weak sedimentary rock types to >70 GPa for certain quartzites and igneous rocks capable of sustaining high stresses with restricted longitudinal strains. To obtain the Poisson’s Ratio of a rock material, strain gauges must be applied to the surface of the sample both laterally and longitudinally to the direction of force. Poisson's Ratio is then determined as the ratio of the lateral strain to the longitudinal strain. Poisson's Ratio is far more variable in its values than other rock properties. 54 Chapter 4: Rock Testing - Determining the Materi However, values between 0.05 and 0.15 can be regarded as low and associated with weak, poorly cemented rock types. Values of 0.25 and above generally indicate an ability to deform under load whilst maintaining integrity and are associated with competent, linear elastic rock types. Tensile Loading Although behaviour under compressive loading is helpful in understanding rock breakage, itis important to realise that it is the tensile strength of rock that controls its failure. In rock the tensile strength is only about 10% of the compressive strength. ‘The measurement of tensile strength in rock is accomplished via an indirect measure known as the Brazilian test (Hondros 1959), which relies upon the diametral compression of a rock disc. The centre of the disk is put into tension and as a result a crack initiates at the centre and propagates towards the loading platens (Figure 4.2). Figure 4.2: Loading cycle for a Brazilian test showing plots of principat stress difference, from ‘numerical simulation (Clark, 1992) Given that the tensile strength of rock controls failure of rock in comminution, it is clear that this mode of failure is critical to an understanding of comminution behaviour. Clark (1992) has investigated the Brazilian style of test using numerical modelling techniques, and has considered how these techniques can be applied to comminution. oO Chapter 4: Rook Testing - Determining the Material-Speofic Breakage Function EEE 4b EgPEEUEEeLULt319egeeSEEEELEL-OS On Eoeeseeee ecco eseeenEESee 4.22 Micro Fracture Mechanisms ‘The microscale features control the mechanism by which the rock responds to load, and thus what macro sale mechanical properties the rock possesses. In terms of modem thinking the basis for examining the response of rock to loading land crack propagation is the work of Griffiths (1921). He examined the low tensile strength of glass using the assumption that small penny-shaped cracks existed throughout the fabric of the material. Subsequent workers accepted this assumption and eventually this led to the development of the discipline known as fracture mechanics, and more specifically Linear Elastic Fracture Mechanics (LEFM). LEFM assumes that at the tip of all cracks within a loaded material there is a stress intensity factor (Ki). Just as a material has a critical tensile stress capacity, it wil also have a critical stress intensity (Kjc), also known as the fracture toughness, For the crack to advance the condition must be that Ki2 Kjc. Fracture Toughness Fracture toughness, which can be defined as the resistance of rock to catastrophic crack extension, is thus an intrinsic material property of the rock, and is indicative of how rock behaves under load. The application of fracture toughness to the prediction of ‘comminution potential is thus a logical extension. Inits simplest form the fracture toughness refers to a tensile failure, known as a mode 1 fracture. For the purposes of assessing comminution potential, mode I fracture toughness testing is accepted as the most appropriate test method. Bearman et al (1991), for example, have shown that it can be correlated with cone crusher product size and power draw, and subsequently found that it is also correlated with the energy-size relationship derived from the JKMRC single particle breakage tests {equation 4.20, Section 46.1) ‘The International Society of Rock Mechanics (ISRM) has standardised two methods of model fracture toughness testing, namely the chevron bend and the short rod pull test. Bearman et al (1991) used the chevron bend test in their study. Figure 4.3 shows the experimental layout of this apparatus. The bending of the core induces a tensile failure ‘with the crack propagating from the tip of the notch to the opposite edge of the core. Fracture toughness is calculated from the force at fracture and specimen geometry from the following equation (Figure 43): ves fmsrafgon(g « 53 Chapter 4: Rock Testing - Determining the Material Specific Breakage Function | F- signals to EX - racorder Ot eo be Flock core ‘Suppor roller ip gauge ~~ Supper = diameter of rock core specimen support span {maximum} notch apn reduced cross-sectional roa rack front ongth toad on specimen |88 = creckmouth opening displacement Figure 4.3: Experimental configuration forthe chevron bend fracture toughness test. The Hopkinson Pressure Bar To study the breakage properties of rock materials, a horizontal Hopkinson Pressure Bar (HPB) has been constructed at the JKMRC, as part of its collaborative research within the Centre for Mining Technology and Equipment, The HPB has strong similarities in its fracture principles to the Brazilian Test. ‘The bar is similar to the vertically-mounted Ultra-Fast Load Cell (UFLC) that has been successfully used at the Comminution Center, University of Utah to analyse breakage functions for ball mill modelling (Hioffler and Herbst 1990; Bourgeois et al 1992 - see Section 4.5.2). The purpose of the apparatus is to measure both the force and the ‘energy required to initiate catastrophic failure in rock specimens. The bar constructed at the JKMRC was originally designed to simulate the energy levels commonly found in industrial cone crushers. ‘The HPB consists of a horizontally suspended steel bar 6.4m long, to which a rock specimen is attached at one end. The rock sample is then impacted and compressed by the collision of an impact bar (1.2m long). Figure 4.4 shows the HPB test apparatus. ‘The impact bar is mounted on linear bearings and is attached to a linear spring mounted at the opposite end of the bar which is compressed fo a known distance to control the impact velocity. ‘The input energy is then determined by measuring the speed of the impact bar at the point of impact, using an optical sensor. Chapter 4: Rock Testing - Determining the Material-Specitic Breakage Function cee tS Es PeSEPEE EES es eeecececeece ett eeeeseeee creer eeeeeereereece 4.2m 4m strain gauges a aa > Impact Bar [] ‘rack Hopkinson Bar loptcal lsensor Figure 4.4: Experimental configuration of Hopkinson Pressure Bar (plan view) ‘The force experienced by the rock is equal to the forces exerted by the ends of the two bars during the collision. This force propagates as a longitudinal strain wave down the impact and Hopkinson bars, which is measured by a strain gauge bridge arranged on ‘both bars. The force on the rock can therefore be resolved, and the particular force at which failure occurs is measured as a sudden release of force in the strain signals. Rock failure is defined by the point on the force-time signal at which a sudden drop in force is observed, which is associated with fracture of the zock into at least two pieces through the middle of the sample. Figure 45 shows an example for granite. The integration of the force-time profiles from both the Hopkinson and impact bars allows lost strain energy to be subtracted from the input energy, so that the actual energy responsible for breakage can be calculated. ‘The Hopkinson Pressure Bar has the potential to test the strength of individual rocks, providing a convenient means of assessing non-destructive damage in rocks, witich can contribute to fracture probability in comminution (Briggs and Bearman 1995). In order to investigate damage in a pre-failure regime, discs of granite were impacted at increasing energy levels until failure occurred. This energy was reduced in stages, and the number of successive hits required to cause failure was recorded and used as a ‘measure of the damage induced in a material prior to breakage. The results are shown in Figure 46. Figure 4.6 suggests that above 0.06 kWh/t a particle will break on the first impact, a regime in which breakage energy is utilised efficiently, despite the possibility of breakage at lower energies. Hence the HPB can be used not only to give an inherent measure of rock strength, but also to quantify the optimum energy for breakage. Briggs and Bearman (1995) found that the HPB results correlate well with other measures of rock competency, for example, Brazilian Tensile Strength, and the JKMRC ty parameter (Section 4.52). 55 Chapter 4: Rock Testing - Determining the Material Specific Breakage Function 3 25x10 20x 10° 3 15x10 10x e Force (kN) ° 0.0008 0.001 0.0015 0.02 Time (ms) Figure 4.5: Typical breakage profile in Hopkinson Bar, showing force to fist fracture Number of Hits 0.06 0.05 008 007 0.08 0.09 Spaeific Energy (kWhit) Figure 4.6: Number of repeated impacts to achieve breakage, for granite Being able to use the Hopkinson Bar to determine the energy required to guarantee breakage, or alternatively if no breakage then the ‘damage’ energy, suggests potential applications in assessing critical size problems in SAG/AG milling and blast induced damage. It may also aid crusher design. Chapter 4: Rock Testing - Determining the Material-Specitic Breakage Function 4.23 Summary: Application of Rock Strength Testing in Comminution Most standard rock mechanics testing procedures still do not provide information that can be directly used in the design and optimisation of comminution processes. However there is now accumulating evidence that the results of certain rock strength tests are correlated in a useful way with the performance of comminution machinery; the Hopkinson Bar has indeed already been used to determine breakage functions for the population balance model (Bourgeois ¢t al 1992). ‘These techniques are also providing an insight into aspects of comminution. For example «Fracture toughness is clearly correlated to crusher performance (Beaman ad 1991). + Numerical modelling of standard breakage events such as the Brazilian Test provides valuable information on the mode of fracture in comminution, particularly for heterogenous materials (Clark 1992). «The Hopkinson Pressure Bar is being used to identify critical energy levels for the design of comminution equipment, and the role of damage in comminution (Briggs and Bearman 1995). ‘The role of these tests in understanding and therefore optimising comminution is likely to increase in the future. For the present, however, the Bond and single particle tests described below represent standard industry practice. 43 ‘THE BOND TEST 43.1 Introduction In order to size crushing and grinding machinery, and also to specify motor sizes, a method is required for determining the energy requirements of a comminution process. Rittinger’s and Kick’s 19th century theories, reviewed in Chapter 2, had severe practical limitations and were consequently of little use in real applications. It was not ‘until 1952 that Fred Bond published the approach that has continued to be the major comminution design tool used by industry, particularly the equipment manufacturers, to the present day. It has also served as a useful operating tool to evaluate and optimise crushing and grinding circuits. ‘According to Bond's ‘third theory’ of comminution, the work input is proportional to the new crack tip length produced in particle breakage, and equals the work represented by the product minus that represented by the feed. He expressed this relationship as follows: Chapter Flock Testing - Determining the Material-Specitic Breakage Function qed w = towr{—--_1_ @2) (es) and Pos oTeW 3) where W_ = work input (kWh/t) work index - a material-specific constant (kWh/t) size at which 80% of the product passes (im) size at which 80% of the feed passes (jim) = throughput of new feed (t/h) = power draw (kW). ‘The work index is the comminution parameter which expresses the resistance of the material to crushing and grinding. Numerically, it is the kWh per tonne required to reduce the material from theoretically infinite feed size to 80% passing 100 microns. Grinding power calculated using work indices obtained from Bond grindability tests (see below) applies reasonably well in the range of conventional rod mill and ball mill gtinding conditions, and can be cotrected for other conditions, Equation 4.2 is also useful for determining an ‘operating work index’ for an existing comminution ‘operation, which can be used to compare {eed ore types, or assess the crushing or grinding performance in relative terms (see Sections 4.35 and 13.22). Bond derived equations for finding the work index from several types of laboratory tests. These are briefly introduced below. 43.2 Bond Crushability Test Particles of broken rock passing a 75 mm square grid and retained on a 50 mm square are mounted between two opposing equal 13.6 kg weights which swing on bicycle wheels, When the wheels are released the weights impact simultaneously on opposite sides of the measured smallest dimension of the rock. The height of fall is progressively increased until the rock breaks. The work index is obtained from the average of ten breaks and is calculated as Ics Wi = 53.49 a) where ICS is the impact crushing strength in Joules per mm of rock thickness, SG is the specific gravity of the rock, and WI is the work index in kWh/t. 43.3 Bond Rod Mill Grindability Test ‘The feed is crushed to -12.7 mm, and 1250 cc packed in a graduated cylinder is weighed, screen analysed, and ground dry in closed circuit with 100% circulating load 58 Chapter 4: Rock Testing - Determining the Materia-Spectic Breakage Function (see below) in a 0.305 m diameter by 0.610 m long tilting rod mill with a wave-type lining and revolution counter, running at 46 rpm. The grinding charge comprises six 31.8 mm diameter and two 44.5 mm diameter steel rods 0.53 m long and weighing in all 33.380 kg. To equalise segregation at the mill ends, the mill is rotated level for eight revolutions, then tilted up 5 degrees for one revolution, down 5 degrees for another revolution, and returmed to level for eight revolutions continuously throughout each grinding period. Tests are made at all sieve sizes from 4.75 to 0.212 mm, At the end of each grinding period the mill is discharged by tilting downward at 45 degrees for 30 revolutions, and the product is screened on test sieves of the required size. The sieve undersize is weighed, and fresh unsegregated feed is added to the oversize to make its total weight ‘equal to that of the 1250 cc originally charged into the mill. This is returned to the mill and ground for the number of revolutions calculated to give a circulating load equal to the weight of the new feed added (see ball mill example in Section 4.3.4 below). The cycles of the grinding period are continued until the net grams of sieve undersize produced per revolution reach equilibrium and reverse its direction of increase or decrease, Then the undersize product and circulating load are screen analysed, and the average of the last three net grams per revolution (Grp) is the rod mill grindability. ‘The rod mill work index (kWh/t) is then given by WI = 68.4 a (45) HP x a x( 2 - x ) Pe0 YFo where P; is the opening in microns of the sieve size tested. This WI value should conform with the motor output power (at the pinion) to an average overflow rod mill of 2.44 m internal diameter grinding wet in open circuit. On average, the rod mill WI is hhigher than the ball mill WI. Note that Bond’s original paper included an error in this equation and the following. equation, 4.6. Note also that in this monograph, all Bond’s kWh/t terms are expressed in metric tonnes. Confusing mass units is a common error in applying Bond's methods (he originally expressed his results in short tons). 434 Bond Ball Mill Grindability Test Basic Principles ‘The objective of the Bond ball mill test is to determine the so-called standard work index, which is defined as the specific power required (kWh/t) to reduce a material 69 Chapter 4: Rock Testing - Determining the Material-Specific Breakage Function from a notional infinite size to a P80 size of 100 um. The test involves a series of consecutive batch grinds in a laboratory mill, After each grind the contents are screened to remove undersize which is replenished with an equal mass of new feed. The length of time of each batch grind is adjusted until the mass of the oversize fraction is consistently 25 times greater than the undersize. Under these conditions the test approximates the performance of a closed circuit continuous mill with a recycle Toad of 250%. (Once the standard work index is known, Bond’s ‘Third Law’ (equation 4.2) is used to adjust this value to give the specific power required to reduce a given F80 to the required P80 in an 8 ft diameter wet overflow ball mill. For a given throughput (t/h) the specific power (kWh/t) is converted to a power draw (kW). Mill dimensions for an industrial unit are then chosen which are predicted to draw the required power, using an appropriate mill size-power correlation; several such correlations are available (see for example Section 11.3). Where these dimensions and/or proposed operating conditions deviate from Bond’s standard, correction factors recommended by Bond are applied to the work index. Conducting and Analysing the Test ‘The standard feed is prepared by stage crushing to pass a 3.35 mm sieve, but finer feed can be used when necessary. It is screen analysed and packed by shaking in a 1000ce graduated cylinder, and the weight of 700cc is placed in the mill and ground dry at 250% circulating load. ‘The standard Bond mill is 0.305 m by 0.305 m with rounded corners and a smooth lining, except for a 100 mm by 200 mm door for charging. It has a revolution counter and runs at 70rpm. The grinding charge consists of 285 balls, the total weighing 20.125 kg. The commercial test typically uses 25x38.1 mm (1.5 inch) balls, 39x31.75 mum (1.25 inch) balls, 60x25.4 mm (1 inch) balls, 68x22.23 mm (7/8 inch) balls and 93x19.1 min 8/4inch) balls. Tests can be made at all sieve sizes below 600 microns. After the first grinding period of 100 revolutions, the mill is dumped, the ball charge is screened out, and the 700cc of material is screened on the test sieve of the required closing size, with coarser protecting sieves if necessary. The closing size chosen will depend upon the application (e.g. expected liberation size). 75 um and 150 jim are ‘commonly used closing sizes. ‘The undersize is weighed, and fresh unsegregated feed is added to the oversize to bring its weight back to that of the original charge, Then it is returned on to the balls in the mill and ground for the number of revolutions calculated to produce a 250% circulating load, dumped and rescreened. The number of revolutions needed is 60 Dia sd tes Chapter 4: Flock Testing - Determining the Materlal-Specfic Breakage Function fp feaeae see TTAetsseceecece see seeeeeeeSeS OEE determined from the results of the previous period to produce sieve undersize equal to 1/35 of the total charge in the mill. ‘The grinding period cycles are continued until the net grams of sieve undersize produced per mill revolution reach equilibrium and reverse its direction of increase or Gecrease, Then the undersize product and circulating load are screen analysed, and the average of the last three net grams of final product size generated per revolution (Gbp) is defined as the ball mill grindability. Table 4.1 illustrates the typical results and calculations in performing a Bond grindability test The ball mill work index (kWh/t) is calculated from the following revised equation (Bond 1961): 49.1 Based on historical data, the average value of Pgg for a 150 um Pyclosing sieve size is “114 microns; at 106 jim it is 76 microns, at 75 um it is 50 microns, and at 45 jim it is 26.7 microns. These relations are used in equation 4.6 when Pgq cannot be found from size distribution analyses. ‘The WI value from equation 4.6 should agree with the motor output power per tonne of ore treated in an average overflow ball mill of 244m internal diameter grinding wet in closed circuit. wi = 6) ‘The Bond work index is a basic measure of rock hardness and, as such, has been found to be broadly related to the UCS (Section 4.2). Typical classifications are shown in ‘Table 4.2, which indicate large variations in results. For dry grinding in both rod and bail mills, the work input should normally be multiplied by 1.30. There are six different efficiency factors that are applied to the reference energy requirement to translate it to a particular commercial circuit installation (Rowland 1982). These are related to: dry grinding, open circuit ball milling, oversized feed, diameter efficiency, fine grinding in ball mills to product sizes finer than 80% passing 75 microns, and high or low ratio of reduction ball milling, ‘Thus, on the basis of three variables - the work index, the feed size and product size = the Bond model can be used to calculate the specific energy requirement for a given grinding duty. 6 Chapter Fock Testing - Determining the Material-Spocific Breakage Function Table 4.1 and Figure 4.7: An example of @ Bond grindabilty test Bond Ball Mill Work Indox Ore Source ‘Sample No Closing sieve size Py (um) | 75 Mass of orginal feed, X(g}| 1156.4 Food F0 (um) Product PO (um) % -Téum in teed, ¥ Ideal Product, Z=X/3.5 (@) T Moga Mino ABCI23 ne Seow | ove | stsim | sin | tem | mom | rat | thee | ate | tiie | meow | en feet | ns | pctot | rg | Rent | com | vom | tom | ane cee ing a ew fare sua eae al age eet cerca [se v"] aso | sous | ass | oo | ante | seea | ose | cea | sae | atnze | am 2 | a | aes] Sar | os | Sos | aso | ee | ares | tema | mm | amo 8 | ao | rar | ses | oo | ama | aos | as | dee | tim | som | ase «| 2 | ams | see | oo | see] saa | zs | ates | tas | aun | oe s | ae | ows | sess | oo | sss | se | zor | ome | tz | mer | 20 TACT we Carer | oy | Ge | Fine | a [Maan mass 75im per rev (g) | 1.22 Bond Work index 124 (KWhishortton) 187. (kWhitonne) 00 Feed Final Product Sieve | Cum%| Sieve | Cum% ia Sizoum| Pass’ | sizeum| Pass 3360 [100.00 | 75 | 10000 2000 | 8960 | 63 | a8.10 60 zeco | oat | 53 | 7509) & : z uct | roo | sito | as | seco fe | Pe 850 | 2857 ee 425 | 1953 2t2 | 1364 20. 160 | 11:30 7 | 7.40 : H wo 10 1000” 10000 ‘ize um) ‘Chapter 4: Rock Testing - Determining the Material-Specific Breakage Function “Table 4.2: Relationship between UCS and Bond WI Property Soft Medium Hard Very Hard UCS (MPa) 60 -100 400-150 150-250 >250 Bond Wi (Wht) 19 9-14 14-20 220 For a given ore work index WI (kWh/t) and required new feed capacity T (tph), the required power P (kW) to grind the feed F80 (ym) to product P80 (1am) is given by: P= 10x Tx WIx (BF, x BE, x EF, an xf Pao Pao where EF, ... are Bonds efficiency factors. 43.5 Discussion of the Bond Method Limitations of the Bond Method Austin and Brame (1983) and Herbst and Puerstenau (1980) have reviewed the limitations of the Bond test. Perhaps the most serious in a fundamental sense is Bond’s claim that equation 4.2 is a ‘universal law’, whereas it is known that the specific grinding energy needed to take a feed with a certain 80% passing size to a product with a certain 80% passing size cannot be the same for a batch test, the standard Bond locked-cycle test, or a steady-state continuous mill with 2 real residence time distribution, The method ignores the fact that the shape of the product size distribution and the associated specific grinding energy is different for these three Criticism of Bond’s method as a ‘model’ is not really warranted since it is not intended to be a model in any comprehensive sense. Its real value is as a method of distinguishing the grindability of different ores in design (though better tests may be available), and as a yardstick to check the energy utilisation of current operations. Experience in its use has demonstrated the following limitations: + It tends to be poor predictor of what happens in a real closed circuit when throughput is increased, unless classifier performance is ‘adjusted’ to cope with the new conditions, + As implied earlier it is not appropriate in systems where size distribution slopes change. 63 Chapter 4: Rock Testing - Determining the Material Specific Breakage Function * Itis not a good predictor of the grinding of large rocks, and specifically for AG/SAG mill behaviour (though it has been adapted to such circumstances - see below). ‘+ It is unreliable for particles with unusual screening characteristics (.e. of unusual shape), though this is a general problem for any technique relying on sieving for the determination of particle size. + The square root relationship assumed in equation 4.2 (arising from Bond's assumption that work is proportional to new crack tip Iength) does not always hold. See also Figure 2.1 ‘There is also a perception that Bond’s method failed to predict the performance of the very large ball mills introduced in the early 1970s, though this was probably more to do with the early learning curve associated with operating these mills and their associated classification processes. Certainly there was never a dedicated Bond test for the AG and SAG mills now in common use. Other problems and inconveniences of the Bond standard procedure include the need to use a Bond mill (which may not be readily available), the requirement of about 10 kilograms of feed sample, which needs special preparation, and the fact that conducting the test is time consuming, A number of attempts have been made in the past to determine the Bond work index by means of simplified procedures (Yap et al 1982). Empirical methods which allow for a direct calculation of the work index using a straight forward procedure are often more practical and better suited for plant application. For example the Anaconda simplified method (Yap et al 1982) which uses an ordinary laboratory mill and only 2 kilograms of sample, is not time consuming and has proven to be very practical and flexible in a number of different applications. ‘However, the full Bond test gives the maximum confidence in predicti ‘The Bond method attempts to partially compensate for idealisations in the test procedure and shortcomings of the model by making use of correlations between grindability test results and results obtained in a ‘standard’ 2.44 m diameter milling circuit. Complete compensation is not possible because this procedure cannot, for instance, take into account inevitable deviations in transport and classification characteristics of the commercial circuit from those of the standard’ circuit. Since the method does not incorporate the above second-order effects, it cannot be used as a guide to optimise a given system, from either the operating point of view or AG Chapter 4: Rock Testing - Determining the Material-Speciic Breakage Function the economic point of view. It is valid only as a gross method for mills operating under normal closed circuit conditions. In spite of these limitations, the Bond method has two main advantages: it is very simple, and experience has shown that it works for many circumstances to a reasonable degree of accuracy, since the corrections are based on a large industrial database. ‘The Operating Bond Work Index Work indices obtained from operating data on any mill can be compared to grindability test results. The operating work index, Wio, can be obtained using ‘equation 4.2, by defining W as the specific energy being used (power draw + new feedrate), F80 and P80 as the actual feed and product 80% passing sizes, and Wlas the operating work index, Wlo. Once corrected for the particular application and equipment-related factors, Wloc, can be compared on the same basis as grindability test results. This allows a direct comparison of grinding efficiency. Ideally WI should bbe equal to Wiog and grinding efficiency should be unity. Power draw is determined using equation 47, assuming grinding efficiency is unity. Application of Bond Theory to Autogenous Mills For the case of autogenous and semi-autogenous milling it has been determined that the work indices obtained from standard Bond grindability tests cannot be used (ection 9.6). It is necessary, therefore, to calculate a work index from operating data (eg. from a pilot plant) so that power draw and mill sizing can be calculated. In some cases the feed is not used in the calculation as the factor 10/-/F80 becomes insignificant because of the large size of the ROM feed. For open circuit AG/SAG. mills, the discharge can be hypothetically classified at the desired closing screen size, to provide a more realistic P80 value for the work index calculation. Macpherson has developed a test procedure for the design of AG/SAG mill circuits. It includes the conduct of standard Bond rod mill and ball mill grindability tests, together with a dry autogenous test in a 45 cm air-swept Aerofall mill. A 250 kg sample of -3.175 mmm ore is required. The autogenous test is used to compute an autogenous work index, which is then corrected for a full scale installation using ‘empirical scaling procedures (based on historical data) and by comparison with the rod and ball mill Wls. The autogenous WI and the other data are then used to determine AG/SAG milling suitability, power requirement and circuit configuration. Further details can be found in Knight et a (1989), Chapter 4: Rock Testing - Determining the Material-Specitic Breakage Function Using the Bond Work Index in a Process Model Although Bond chose to use the results of his laboratory ball mill test to determine an operating work index, much more information about grinding of an ore can in principle be extracted. His testis a physical model which approximates a closed circuit ball mill, and so product sizings and flowrates of each stream can be obtained from it. Research currently underway at the JKMRC is testing the use of these data to determine the relevant parameters for a ball mill model. The model is then used to predict the performance of a full scale ball mill. This approach has the advantage of simultaneously generating @ Bond ball mill work index and a set of model parameters, which can be used in a complementary fashion in design and optimisation. 44 BATCH GRINDING TESTS Another approach to grinding, much favoured by chemical engineers, is to treat it as a rate process. The kinetics of a grinding process can be studied with batch grinding tests on monosize samples. The typical procedure involves batch grinding the material within a certain size fraction, from which the breakage rate of that size fraction is determined, together with the breakage function of the material. The breakage function is essentially a size distribution resulting from breakage, usually expressed in matrix form. The results of the test are interpreted in terms of the population balance model (Chapter 2), usually by estimating the values of the selection and breakage functions in the model. ‘The test material is classified into a specific size fraction, usually a V2 size interval, and loaded into the mill. The selection of the other grinding conditions depends on the objectives of the test and should be as close as possible to the conditions envisaged for the final application. The material is discharged at milling times decided prior to the test. After the product is sieved, the material and the grinding media are reloaded into the mill. The test is repeated until a grinding time appropriate to the expected residence time in the industrial mill, or the desired product grind, is achieved. For the -monosize fraction tested, if the specific rate of breakage of the particles in the feed is constant, the following equation is valid (Austin and Weller 1982): eT) 4s) ‘where my(t) is the mass percentage of the feed size at grinding time t and 5, is the specific rate of breakage of the feed size (minute). The integrated form of the ‘equation may be written as: Jog mi(t) = log my(0)- 5;t/In 10 49) ou oe CChapler 4: Rock Testing - Determining the Material-Specitic Breakage Function ‘Thus, by plotting the logarithm of the mass percentage remaining in the top size fraction, m,(t), versus the grinding time, t, the breakage rate of that size fraction, sj, can be determined from the slope of the curve, which is usually linear for the ball grinding process, After conducting the monosize grinding test for a number of size fractions, the selection function based on the breakage rates from tested size fractions can be estimated. It can be fitted using the following equation (Austin and Brame 1983): (4.10) x; is the upper size of interval i, and x, is a standard size. A, and m are parameters related to the grinding conditions, and otis a characteristic of the material. ‘The breakage function of the material can be calculated by the methods of either Herbst and Fuerstenau (1968) or Austin and Luckie (1971). The monosize grinding test is closely related to the population balance model described in Chapter 2, and a review of this model is appropriate here. ‘The population balance model can be described by the equation k SOL gmt) + bysjmj) fori=123..k (4.11) a at ‘where by is the size discretised breakage function. The development of this model was initially based on the statistical approach to the breakage of solids established by Epstein (1947). In particular, he introduced the two basic comminution functions: (a) the probability of breakage of material of a given size, and (©) the size distribution of the products of a breakage event. For particular cases of these probability and distribution functions he was able to show that the products of a multiple breakage process would tend to a log-normal size distribution, as is frequentiy observed in practice. Reid (1965) later developed a solution to the fundamental integro-differential equation for continuous grinding systems. This solution calculates the product from a continuous mill by taking the solution of the batch mill equation and applying a residence time distribution #(t). That is ifa fraction ¢(t) dt of the feed stays in the mill for a time t, and leaves between t and t+dt, this fraction will be broken as if it were batch ground for time t Its size distribution upon discharge from the mill will be mt), obtained by the solution of equation (4.11) for each value of i. At the same time, 67 Chapter 4: Rock Testing - Determining the Material-Specific Breakage Function material that has been in the mill for different durations will also be discharging, and at steady state, the total product in size fraction, i, which leaves the mill will be the ‘weighted sum of all fractions of size in the product: t P, =/milt) ot) dt (4.12) a This is the basic equation of steady-state, continuous, first-order grinding, The integral is in practice evaluated for at least three mean residence times. Note that equation 4.12 assumes all particle sizes in the mill have the same residence time distribution, and there is no classification of mill discharge. The solution of equation 4.12 usually assumes the RTD to be of the form: ()=ke* (19) Herbst and Fuerstenau (1973) introduced the specific grinding energy into the population balance model and linked the grinding energy to the entire size distribution as one mathematical system. This model has been successfully applied in mill design and optimisation in recent years by researchers such as Herbst and Fuerstenau (1980), Kelsall et al (1973), Weller etal (1988) and Austin et al (1976). ‘The size distribution can be described without difficulty using very simple numerical methods, and the Kelsall breakage function (Austin and Weller 1982) has been applied in a wide range of circumstances (e.g. Morrell, Sterns and Weller 1993). Despite the wide use of the population balance method by researchers, a standard test procedure is yet to be established and accepted. Often, the simulation of the process has to resort to the back-calculation of the model function, which is basically a data fitting process. Furthermore, the method is basically a descriptive model and has proved to be rather weak as a predictive model. It is not yet possible for example to predict with confidence the variations of the selection and breakage functions. This is a serious weakness because although the model describes existing operations well, it cannot be used confidently for the selection of equipment and the design of flowsheets. Experience has shown that the assumption of first order grinding implicit in the batch grinding approach (equation 4.11) works quite well for ball milling and fine grinding in general. A problem is the decoupling of breakage function and breakage rates, which requires tedious laboratory experimentation, but again the assumption of a constant breakage function (over a range of ore types) has been shown to work in many cases. wOGu BG Ou ‘Chapter 4: Rock Testing - Datermining the Material-Specific Breakage Function SERGE Ae eed zee EEE Reece eee eeeeee eee However evidence is accumulating that for more complex grinding systems, and in particular AG and SAG milling, the assumption of first order breakage and constant rates is not correct (Morrellet ai 1994). Indeed there is no particular a priori reason to suppose the process to be first order. For example, breakage mechanisms may change 1s grinding proceeds and the properties of daughter particles ‘evolve’, and in many cases the availability of impacts (and thus the behaviour of the charge, and transport through the mill) may be process-determining, rather than simply the mass of material present of a particular size. In summary, the batch grinding method can be a useful approximation to real systems where its assumptions are not tested too far, particularly for comparative work, It does not seem to be appropriate for AG/SAG milling, and it does not have the flexibility to bea consistent predictor for large industrial ball mills. 45 SINGLE PARTICLE BREAKAGE CHARACTERISATION A disadvantage of the batch grinding test and population balance mociel is that it is difficult to separate the influence of the material-specific properties (the breakage function) and the machine-specific properties (the breakage rates or selection function) in a way which is meaningful in the context of industrial grinding. It is common practice, for example, when applying the population balance model to the simulation of industrial grinding to use some standard breakage function, and then fit the size- specific rates to industrial grinding data (Austin and Weller 1982). This ignores possible changes in the breakage characteristics of the ore, and requires any observed variation to be taken up in the rates, which can be misleading, It has been an objective of the research at the JKMRC in recent years to improve the decoupling of material and machine in the simulation of comminution. The early work by Whiten and his students Awachie (1983) and Narayanan (1985) led to the development of procedures for the separate estimate of the breakage (appearance) function by the controlled breakage of single particles. Single particle breakage tests have been conducted by a number of researchers to investigate various important features of the complex comminution process. Basically three types of breakage systems ate distinguished: © impact, ‘slow compression, and © shear. This section focuses on single particle breakage tests in which the breakage mechanism is by impact. In these tests, the particle(s) is crushed between two hard surfaces. One Cy Chapter 4: Rock Testing - Determining the Material-Specific Breakage Function such arrangement is the twin pendulum in which the input pendulum is released from a known height to swing down and break a particle attached to the rebound pendulum. Another is the drop weight apparatus in which a particle(s) resting on a hard surface is struck by a falling weight. 45.4 Pendulum Tests The first experiments using a twin pendulum device seem to have been carried out to study the effect of velocity on single particle breakage and to emphasise the need for such a device for conducting breakage studies under controlled conditions (Pahernwald et al 1938). Fahernwald found that an increase in impact velocity increased fines production resulting from breakage of a particle. Bond (1946) directed a research group that developed a twin pendulum apparatus to determine the impact crushing strengths of 76 x 50mm rock particles under controlled ‘crushing conditions. This became known as the crushability test. Gaudin and Hukki (1946) carried out single particle crushing tests using a pendulum device and determined the size distribution and new surface of the comminuted products. Other features of comminution that have been studied using pendulum tests include correlating the specific fracture energy with the new surface area created. This relationship had a fineness limit to account for the fact that material of infinite fineness cannot be produced by infinitely increasing the specific fracture energy (Yashima et al 1981, 1982). It was also shown that the crushing efficiency of irregular specimens was better than that of spherical particles for a wide range of input energy. ‘Awachie (1983) investigated several single particle breakage tests as a means of determining the breakage function for crusher modelling, The products from slow compression, drop weight and pendulum tests were compared and found to be similar. The compression test was chosen by Awachie because it was considered that this breakage mechanism most closely resembled the breakage mode in an operating crusher. In the compression test, a single particle was loaded in a confined volume between two platens. The platens were constrained for each test by an insert which corresponded to aselected degree of reduction. The loading rate was low at 0.021mm per second. A study on energy losses during the breakage of regular shaped materials using a twin pendulum and a small calorimeter has shown that most of the input energy is dissipated as heat (Zeleny and Piret 1972). 70 aL ‘Chapter 4: Rock Testing - Determining the Material-Specific Breakage Function Hates ever Cee vec teececeeeeee Se eeeteeNeeeeeeeeeee ere 452 Drop Weight Tests "Drop shatter’ tests are commonly used for assessing the breakage characteristics of soft materials (e.g. coal; Berenbaum 1961), or for simulating the degradation of coal and ores in handling (Norgate etal 1986). The principle is simple and well known. The particle is dropped from a known height and the size distribution of the broken product is measured. The energy applied is simply the potential energy, mgh, in Joules, where m = mass of particle (kg), g = 9.81 m/s and h = drop height (m). The specific input energy is then mgh/m = gh in Joules /kg. Since 1 kWh/t = 3600 J/kg, it requires a drop height of 367 m to apply 1 kwh/t (a typical crushing energy), clearly impractical for simulating breakage of ores in crushers and mills, in which these and higher energies are believed to dominate. As pointed out by Morrell and Morrison (1989), a more convenient (and less hazardous) approach is rather to drop a suitable weight on the particle, for which heights of @ metre or less are adequate to achieve the necessary energies. Such drop weight tests are discussed below. Drop weight testers of various kinds have been described in the literature. An early form was reported by Gross (1938) who used the device to establish a relationship between surface area produced and input energy. Piet (1953) modified Gross’ original simple apparatus and used it to crush brittle materials, from which he developed a relationship between energy input and product size distribution. Fairs (1954) made use of the drop weight apparatus developed by Gross and modified by Piret to conduct particle bed breakage tests, and established a relationship between the net input energy and the new surface area produced. The size distributions resulting from these tests could be described by the Schuhmann equation. He established an energy-size reduction linear relationship of the form: jg = AKO? (4.14) where Eis is the specific energy input, A and n are constants, and kis the size modulus in Schuhmann's equation. Schénert (1972) employed a drop weight apparatus to determine the energy utilisation in single particle crushing, and used it as a basis with which to evaluate the performance of an actual grinding mill. He defined the energy utilisation as a ratio of the new surface to the energy needed to create the new surface area. Jowett and Van Der Waedern (1982) further modified the drop weight originally developed by Protodyakanov (1950) and adapted it to testing iron ores and associated m Chapter 4: lock Testing - Detormining the Material-Specilic Breakage Function rocks to determine hardness relevant to drilling, blasting and crushing in the field. ‘They concluded that the hardness test they developed might be useful for a rough estimate of rock hardness in the field but did not recommend the use of such field data for accurate prediction of rock fragmentation involving crushing and grinding. They did not establish a relationship between energy and size reduction. Pauw and Maré (1988) criticised the correlation of the surface area created during breakage with the energy input as a method for determining the efficiency of energy usage in a single breakage event. They stated that the method did not demonstrate how significant the results obtained for the breakage of a single particle were in the optimisation of breakage events in a grinding mill. They used a drop weight device to break single particles of a gold ore of various sizes at varying levels of impact energy. It was shown that an optimum impact energy can be determined for the breakage of an core particle of a given size to a specified product size, and that the optimum usage of impact energy is often obtained by the use of a number of breakage steps for the size reduction of a particle. The optimum level of the impact energy was shown to be highly dependent on the size of the particle to be broken, and to be influenced by the size of the final product required and the efficiency of secondary breakage. Hotfler and Herbst (1990) employed a form of drop weight apparatus, the Ultrafast Load Cell (UFLC), to conduct single particle and particle bed breakage tests for application to ball mill modelling. The UFLC consists of a 5m long vertical steel rod (essentially a Hopkinson Bar - see Section 4.2.2) which is equipped with an anvil to allow round and fiat tops to simulate ball-bal, ball-flat and flat-lat impact geometries. ‘The rod was instrumented with solid state strain gauges to capture the impact event. ‘The results from the particle bed drop weight tests were used in ball mill modelling based on the population balance approach. Bourgeois et al (1992) ater used the UFLC to study the fundamental processes associated with the impact fracture of particulate material. They measured the force-time histories of individual fracture events, and reported that the resulting breakage function overestimated the absolute breakage function because the ball continues to stress the breakage fragments after primary breakage, producing secondary breakage. 453 JKMRC Rock Characterisation Procedures ‘As noted earlier, the JKMRC crushing and grinding simulation models aspire to a clear goal - that of separating ore characteristics from those of the processing machine, in a form which can be exploited by the simulation models of crushing and grinding. This is achieved through the use of separate ore and machine parameters. ‘The ore parameters can be determined from ore characterisation tests on representative samples. The machine parameters (eg. breakage rates) are calculated from plant survey R eet Chapter 4: Rock Testing - Determining the Material-Spectic Breakage Function age Eee CECE e eee data. Once the models have been customised to an existing circuit, the behaviour of that circuit over a wide range of operational conditions can be accurately predicted. ‘The ore parameters are derived through the use of single particle impact breakage testing, which aims to characterise rocks in the context of comminution. The JKMRC has tested a very wide range of ores, quarry materials and coal samples over many years. These data have been used extensively in modelling and simulation research activities and in consulting work. The ore characterisation tests essentially measure the ore-specific energy/size reduction behaviour. This can be conveniently expressed as a degree of breakage at a specific comminution energy level, Ecs (kWh/t), or as breakage or appearance functions; these measures are discussed in detail later. Details of the twin pendulum and drop weight devices developed at the JKMRC are presented below. The abrasion test used to provide information on low energy breakage in AG/SAG mills is discussed in Section 4.6.1. Twin Pendulum Test ‘The twin pendulum (Narayanan 1985; Narayanan and Whiten 1988) was the first single particle breakage device to be developed at the JKMRC. It is used for conducting single particle breakage tests from which the comminution energy, defined as the energy available for the breakage of a particle, can be determined, together with the resultant product size distribution (Narayanan 1985). The single particle breakage test resulls are used to generate ore-specific breakage functions for each of the associated comminution models. The breakage rate is assumed to be machine dependent, The incorporation of the single particle breakage test results in the JKMRC comminution models is described in Section 4.6. Two pendulums of different sizes are used for breaking different particle size ranges at different energy ranges. The ‘large pendulum’ is used to relate the energy consumption in single particle breakage to industrial crushers and semi-autogenous /autogenous mills (-31.5+11.2 mm particle size), while the ‘small pendulum’ is used to relate energy consumption in single particle breakage to industrial rod and ball mills (-11.2+4.75 mm particle size). Each device consists of an input and a rebound pendulum suspended from a rigid frame, as illustrated in Figure 4.8, The rebound pendulum swings between a laser source and a detector which are mounted on an optical bench at right angles to the plane of swing of the pendulum. The motion of the rebound pendulum is monitored with a computer, which records the time taken for a multiple fin arrangement (attached to the pendulum) to pass through the laser beam. 25 swings of the zebound 73 Chapter 4: Rock Testing - Detormining the Material-Specttic Breakage Function pendulum are monitored to determine the period. The particle selected for testing is fixed to the rebound pendulum by 2 piece of tape, or a similar arrangement, and the input pendulum is released from a known height to swing down and collide with the particle. oe pendulum Rebound pendulum Rock Colection box Figure 4.8: Schematic ofa twin pendulum device ‘The energy transmitted to the rebound pendulum, Bt, is determined from the computer-logged timing signals (Narayanan and Whiten 1988). This is calculated from Et = Mr (L-L cose) (4.15) where @ is the angle of displacement of the rebound pendulum from its equilibrium position, Mr is the mass of the rebound pendulum and L is the length of the pendulum. ‘The angle of displacement is computed from the corrected period P, and the calibration constants, a and b, using the expression P = a+bee (4.16) a and b are determined by regression on a set of measured p and values. The principles of conservation of momentum allow the velocity of the input pendulum after impact and therefore its residual energy, Er, to be computed. The energy balance during a collision of the input pendulum with a particle attached to the rebound pendulum can be written as: Bi = Br+Et+Ec (417) 74 Chapter 4: Rock Testing - Determining the Material-Specific Breakage Function Esse ELEC o ese EeEEE Eee eeeeeeeeeee see SereeEeeeeeECee where Ei is the input energy, and Ec is the total energy loss, representing the energy consumed by the specimen for breakage and other losses such as acoustic, thermal anc strain energy losses which cannot be separately determined. This energy loss term, Ec, is defined as the comminution energy and is a lumped representation of all losses ‘occurring in single particle breakage. Ei is defined as Mi g h, where Mi is the mass of the input pendulum and h is the vertical height of the centre of gravity of the pendulum above the impact point. “The specific comminution energy was defined by Narayanan (1985) as the difference in the energy of the input pendulum before impact and the energies of the rebound and input pendulum after impact. It therefore includes the comminution energy for size reduction plus any acoustic and heat energy dissipated during the breakage process. ‘The specific comminution energy Ecs (kWh/t), defined as the comminution energy per ‘unit mass, can be determined from these results using equation (4.17), which may also be expressed as: (4.18) where ‘mass of the rebound pendulum (small: 6.364 kg; large: 40.35 kg) ‘mass of the input pendulum (small: 4.£41 kg; large: 19.86 kg) coefficient of restitution (obtained from measurement of the input pendulum velocity and the initial rebound pendulum velocity) specific input energy, kWh/t = Ei/m mass of the particle. ‘The particles are closely sized before breakage, usually using a J2sieve series. Standard values of Bis are generally selected, to give an appropriate range of product sizes ‘Observation during single particle breakage tests using the twin pendulum device indicates that the collision tends to follow the rebound pendulum, implying that an inelastic collision has occurred. Therefore, the coefficient of restitution is close to zero for the lower input energies where no compaction of the particles between the pendulums is observed. There is some evidence that the coefficient of restitution will increase with the specific comminution energy, but the increase is negligible. ‘The mass tatio coefficient in equation 4.18 is approximately 0.58, and for low coefficients of restitution may be used to calculate the specific comminution energy 26 accurately as can be determined from the pendulum monitoring. A value of the coefficient of restitution between 0.0 and 0.2 results in a maximum error of less than 5% % Chapter 4: lock Testing - Determining the Material-Specitic Breakage Function in the calculation of the specific comminution energy. This assumption may therefore simplify the experimental procedure with a significant saving in time (Andersen 1988). Although the twin pendulum is a simple device for estimating the energy used in particle breakage, its operation and the results obtained have limitations, The device is restricted in its energy and particle size range and additionally is time-consuming in its ‘operation. Also calculation of the breakage energy is sometimes imprecise due to ‘secondary motion of the rebound pendulum. In view of these limitations, the drop ‘weight tester was developed. as an alternative to the twin pendulum, Drop Weight Test ‘The drop weight tester was introduced at the JKMRC to replace the twin pendulum apparatus for assessing impact breakage characteristics of ores (Brown 1992). It consists of a steel drop weight mounted on two guide rails and enclosed in perspex, as shown in Figure 4.9. An electric winch is used to raise or lower the weight to a known height. The weight is released by a pneumatic switch and falls under gravity to crush a single particie placed on a steel anvil. The device is built on a heavy steel frame which is mounted and bolted on to a concrete block. By changing the release height as well as the mass of the drop weight, a wide range of input energy can be produced. The energy range generated is wider than that of both pendulum devices, as shown in Table 4.3. Perspex encour [| — ska tead weights Gui ra. Adjustable height (energy Fock ‘Stee! anu east Large concrete base Figure 4.9: The JKMRC drop weight testing device Chapter 4: Rock Testing - Determining the Material-Specific Breakage Function ‘The standard drop weight device is fitted with a 20kg mass, which can be extended to 50kg. The effective range of drop heights is 0.05 to 1.0m, which represents a wide operating energy range, from 0.01 to 50 kWh/t (based on 10 to 50mm particles). These masses were designed for testing hard rock ores, whose specific gravity ranges from 2.8 to over 4 g/cc. In coal, the breakage energies of interest are much lower, which necessitated the re-design of the device to include a Light Impact Breakage Head of 2kg. This effectively lowered the base breakage energy to 0.001 kWh/t, which is equivalent to a 0.35m drop under gravity. Typical patticle sizes and input energies used for AG/SAG mills are given in Section 4. Table 4.3: Comparison of pendulum and drop weight energy operating range Particle Size Pendulum Drop Weight (rum) cewht) cxWht) -46.0+38.0 0.01 - 0.09 0.004 - 1.07 A822. ot - 9.30 0.081 = 41.2. Normal practice for coarse particle assessment (e.g. crushers, AG/SAG mills) is for 20- ‘50 particles to be tested at each size/energy combination, requiring typically 500-1500 particles in all, or 50-100 kg of material. Following sample preparation (screening into narrow size ranges), the mean mass ii of each set of particles to be broken is calculated. Based on the required specific input energy for each test, the height from which the drop weight is to be released is determined using the relationship below: im Eis ae 9) 0.0272Ma eee where hi is the initial height (cm) from which the drop weight is released (or the initial height of drop weight above the anvil), and Md is the mass of the drop weight (kg). ‘Typically 10 mm is added to the calculated drop height for each test. This ensures that the correct final specific comminution energy is obtained, to allow for the fact that after breaking a particle, the drop weight is brought to rest at some height above the anvil, due to the presence of the crushed particle, The average offset hf can be measured for each sample of particles broken, in which case the actual applied energy is 0.0272 Md(hi-hf) Eis (4.20) Note that the specific comminution energy Ecs (kWh/) is equal to the specific input energy Eis (kWh/t) as long as the drop weight does not rebound after impact, It has been occasionally observed that at energies above 3 kWh/t, and with elastic materials, 7 Chapter 4; Rock Testing - Determining the Material-Spectfic Breakage Function the drop weight rebounds. Although this rebound energy is not currently measured, it is known to be small relative to the input energy. ‘The drop weight tester provides exactly analagous data to the twin pendulum. It has several advantages, however: Extended input energy range Shorter test duration Extended particle size range Greater precision Possibility to conduct particle bed breakage studies. grepe ‘The drop weight device is suitable for the impact breakage assessment required for ball mill, SAG/AG and crusher models. It can also be used to detect the effects of particle size (up to 100mm) or rock damage (blast or crushing) on comminution strength. Data Reduction A key concept in analysing data from the pendulum and drop weight tests, and in ‘establishing ore breakage functions, is that the product size distributions are a function of the size reduction or specific comminution energy, Ecs (kWh/t). To model this breakage process, a simple way of relating energy to geometric size reduction is employed. The method developed at the JKMRC is to use a set of curves (cubic splines, Whiten 1972c- see Appendix 1) to describe the size distribution produced by breakage events of increasing size reduction or energy input. Ifa single particle of known size (eg. lying in a narrow sieve range) is broken, one might consider the resulting size distribution as cumulative percent passing a ~/2 series of screen apertures (Figure 4.10a). Figure 4.10a can be replotted in terms of the original particle size by dividing the x axis by the original particle size (Figure 4.105). Next some marker points on the size distribution are selected, defined as the percentage passing, t, a fraction, y, of the original particle size. Thus t2 is the percentage passing an aperture of half the size of the original particle size, ty is the percentage passing one quarter of the original aperture and to the percentage passing cone tenth of the original particle size. tyo is employed as a characteristic size reduction. For the original particle, all of the t values are zero. For a slightly broken particle, tig is a few percent. In crusher breakage ty9 is typically 10 to 20%. Tumbling mills probably operate over a broad range of ty values. However JKMRC models of tumbling mills tend to use tyo values in the range 20-50%. 78 Chapter 4: Rock Testing - Determining the Material-Specific Breakage Function ee ee NN SO ‘To make use of this description of ore breakage, marker points ty, ty tas gg. and tys are stored in matrix form against ty9. These same data can be represented graphically as shown in Figure 4.11, which covers a wide range of ore types. : | erence 2 a, a & § 3 g i i i“ icing (ean Zn 3 cy Particle size mm mow on ul Particle size mm Figure 4.100: Product size distbuton Figure 4.100: Relave size cstibuton, showing t markers. 100 e a te eo i 10 = «0 fb es o ‘80 ts 0 T 1 [aerate soca Broakage Index, tg (8) ' Figure 4.11: tvs degree of breakage, tyo (afler Narayanan 1985, for ¢ range of ore types) 79 Chapter 4: Rock Testing - Determining the Material-Specific Breakage Function Figure 4.11 is a useful graph. Each vertical line (or value of t10) represents a complete size distribution, expressed as cumulative weight percent passing. Therefore if the data represented in Figure 4.11 can be measured for a particular ore type, it can be used to predict the size distribution which will result at any known degree of breakage, ot tho value. This convenient representation of breakage data is commonly referred to as the one-parameter family of curves (Narayanan and Whiten 1988). The parameter, tyg, is defined as the cumulative percent passing Y/10, where Y is the geometric mean of the size interval for the test particles. Knowing the curves for a particular material, and given a ty (from a given Ecs, or from a model), the full product size distribution can then be reconstructed. Extensive measurements have shown that the same family of curves describes the breakage of a wide range of ores , with “harder or “softer” ores having lower and higher values of typ respectively for a given comminution energy. Such information forms the essential ore-specific information necessary for accurately ‘modelling and simulating mineral comminution processes, viz. rod and ball milling, AG/SAG mills, and crushers. These breakage product size distributions, also known as appearance functions, therefore constitute an integral part of the mathematical models of industrial comminution equipment described in this book. To ensure that the more recently developed drop weight device consistently provides accurate results, a study was initiated to compare the results of the twin pendulum and drop weight tester over a wide range of energies and ore types. Comparative single particle breakage tests conducted on five ore samples showed that the devices produce similar results, as defined by the respective energy /product size relationships. The drop weight tester, however, has the added advantage that it provides less data scatter resulting in better precision in the determination of 9, as shown in Table 4.4, based ‘on six replicate tests conducted on Broken Hill ore. ‘Table 4.4: Error bar values for the determination of tio Device 95% Confidence Interval ‘Twin Pendulum 29.1% Drop Weight £3.6% Not unexpectedly, the to shows some similarity to the traditional fracture mechanims tests, Bearman et ai (1991), for example, have shown that the ty is strongly inversely correlated with the mode I fracture toughness Kjc (Section 4.2.2), based on ‘measurements with five quarry rocks. ‘The early work on the pendulum test indicated that the energy-size reduction relationship was linear and independent of original particle size. However, over the 80 Chapter 4: Rock Testing - Determining the Material-Specific Breakage Function flat geste PEPE eeeeere rece last 10 yeats,it has become apparent that this relationship varies with factors such as particle size, energy and type of industrial comminution equipment, Consequently, the way the breakage test results are used in the mathematical models differs depending con conditions. Examples of how the breakage data are used in JKMRC comminution models are presented below. 46 ‘ORE PARAMETERS IN MODELLING AND SIMULATION 4.61 AG/SAG Mill Ore Parameters ‘As noted in Chapter 2, the mill models are based on the assumption that the contents of the mill are perfectly mixed. At steady state, the following material balance can be determined for each size fraction: feed + appearance via breakage = breakage tosmaller + product (4.21) of larger particles particles ‘The AG/SAG mill appearance function for a given material is determined for at least three size fractions, and defined in terms of three impact breakage and abrasion parameters (A, b, and t,). The appearance functions are size-energy dependent, and the prevailing energy levels in the mill are estimated from the mill load (determined iteratively) and mill size. The amenability of an ore to impact breakage is measured by the drop-weight or pendulum test, described earlier. ‘The current JKMRC practice is to use five narrow size fractions and three input specific energy levels, as defined in Table 45: ‘Table 4.5: Size intervals and nominal input energy levels (kWh!) used in JKMAC AG/SAG drop weight test test_| -63+5omm_|_ -45437.5mm_| -31.5+26.5mm | .22.4+19mm | -16+13.2mm 1 0.10 0.0 025 0.25 0.25 2 0.25 0.26 100 | 4.00 1.00 Ls 0.50. 1.00 250 | 2.50 2.50 ‘Typical data from one energy input level is shown in Table 4.6, a computer printout. Included in the printout are the test conditions (mass of material before and after testing, input energy, particle size and number of particles), the product size distribution and the interpolated t-values. et Chapter 4: Rock Testing - Determining the Material-Specific Breakage Function ‘The amount of breakage, or breakage index, tio, is related to the specific comminution energy as follows: : Alt -e6b-Ea) (422) bo where tio is the percent passing 1/10th of the initial mean particle size, es is the specific comminution energy (kWh/t), and A and b are the ore impact breakage parameters. Graphically this relationship is shown in Figure 4.12 for a primary gold bearing ore. Table 4.6: Example of JKMRC ore impact breakage test results, -31.5+26.5 mm [Sk Pam Sta: 31.5 +26. Sem BELT ‘StEom) wrigrs) WES CMA BASS 22.400 14.870 © 1.911 98.088 aMTTIAL wercen(gr=) "78.3000 is:c00 48060 6.308 91.786 FAL WexGHT(owe) 728,270 31/200 279128028455 @2.351 «EU? BE lao) 19.435 ‘000 336.240 17.808 aa-e25 ROUT SREY Uni/tomma) 433 : 5.600 53.810 12.056 32.771 ROTIAL PAIGE Siz tm) | 287890 ieee 2.000 52.470 6742251030 «—MIMBER OF PARICLES TESTO 20 see 2.900 44.300 5.692 20.337 2000 28:20 3.762 36.575 000 329.000 16.575 000 CURR & FASSRD ¥/n (fF INITIML PARTICLE S125 (mm) & n = 20,2,4,25,50 end 75) tomate PE} _ 0005S i Specific Comminution Energy Eee (kWhit) J Figure 4.12: Effect of specific comminution energy on the breakage index, io ‘Chapter 4: Rock Testing - Determining the Material-Spectic Breakage Function steep tlant edie itilasbisfeeseeebeses sire snianiabioueensnam ese “Acsteeper gradient of the tg ~ Ecs curve indicates @ softer ore. For a constant value of ‘A, this translates into a higher value of b. The way A and b are currently fitted, A remains relatively constant at around 50 for most hard-rock ores. The tyo can be interpreted as a “fineness index” with larger values of typ indicating a finer product size distribution. The value of parameter A is the limiting value of tyo. This limit indicates that at higher energies little additional size reduction occurs as the Ecs is increased, ie, the size reduction process becomes less efficient. A*b is the slope of the curve of ‘zero’ input energy. Because there is some interaction between A and b, it is not as easy to characterise ‘typical’ values of these parameters for particular ore types, as can sometimes be done with a Bond WI. However they can be used for simple comparative purposes. In. AG/SAG milling, size reduction is believed to occur both by impact (crushing) and abrasion or chipping (Section 7.4). Abrasion and chipping breakage events leave the original particle largely intact, the products being relatively fine particles. Impact, however, typically breaks an entire particle into fragments of a range of sizes. ‘The pendulum or drop-weight device provides an effective means for ore characterisation for high energy breakage (crushing). Although these devices can at low energy produce breakage products similar to those typical of chipping/ abrasion, it is extremely time consuming to generate sufficient material for analysis. The generation of chipping/abrasion appearance functions is more suited to a tumbling test, in which a quantity of ore is tested autogenously under a standard set of conditions. This test is outlined below. “The original modelling approach assumed the mean specific energy for an AG/SAG mill fo be a function of mill diameter. Later a modification was introduced to allow the mean specific comminution energy to be calculated separately for each particle size. Given this specific comminution energy, the model calculates the tyg for crushing breakage, from which the crushing appearance function is interpolated from @ standard breakage map (Leung 1987). Table 4.7 shows the standard appearance function (or size distribution) developed at the JKMRC and used in the AG/SAG ‘snodel, This is based on the observation (over many cases) that the t-curves for most brittle ores behave similarly, which greatly simplifies the computations. ‘Table 4.7: Standard appearance function data (24) used in JKMRC AG/SAG mode! [0 (4) 75 ‘so [ts w | 2 10, 2.33 3.06 4.98 23.33 50.53 30 6.88 24 1562 6158 92.49 50 1032 1471 25.88 82.86 9647 83 ‘Chapter 4: Rock Testing - Determining the Material Specific Breakage Function Note that the size-by-size crushing (or high energy) appearance function is combined. in the model with the abrasion (or low energy) appearance function to provide a combined appearance function, a ac Nestle the - ahe ay te * the where a combined appearance function tle t parameter for chipping /abrasion the t parameter for crushing breakage ale chipping/abrasion appearance function ahe = crushing appearance function (The appearance functions are lower triangular matrices). Ore Specific Gravity In AG and SAG miills in particular, the ore specific gravity has a significant effect on charge density and hence power draw. However where blends of ore are being ‘treated, a harder component may be present which has a different SG to that of the rest ‘of the ore, in which case the SG of the ore in the mill will be different to that of the feed ore. To determine if this may be an issue, as part of the drop-weight test procedure, the JKMRC measures the SG of 100 individual particles of approximately 25mm size using a He gas pycnometer, The SG of a sub-sample of the -6.7mum fraction is also measured to identify variations in coarse liberation. Ore Abrasion Test ‘The abrasion testis performed in a 300mm diameter x 300mm long tumbling mill with four 10mm lifter bars. In the standard JKMRC test a 3kg sample of -55+38mm ore is ground for 10 minutes at 70% of critical speed, ie 53 rpm. ‘Typically a bulk sample is riffled down to approximately 3kg, from which rocks are randomly selected to meet the 3kg specification as close as possible (#5g). After each. test is completed, the mill contents are removed for size analysis. For each test, the actual RPM and charge mass are recorded. ‘The product from the mill is dry sieved to -38um on a V2 series of sieves. From the ‘aw particle size data the typ values are generated using a program which converts the actual mass retained data into percent mass retained and cumulative percent mass passing size distributions, from which the to values are interpolated using cubic spline techniques. Table 4.8 shows an example of a program output, which illustrates the bimodal size distribution typically obtained in abrasion breakage. a Chapter 4: Rock Testing - Determining the Material-Specitic Breakage Function Heda eaecaeeet ee ESTsST eect eee Note that the abrasion ore parameter, t, used in the SAG/AG model (tye in equation 4.23), is defined to be 1/10th of tio, which for the example shown in Table 4.8 is 0.8. The ty value is as low as 0.2 for very hard ores, to above 2 for very soft ores; a value of 0.88 is indicative of a medium resistance to abrasion. ‘The JKMRC abrasion test should not be confused with the Allis-Chalmers abrasion test, which measures metal abrasion. In this test a single paddle impactor, requiring 16 kg of -19+13.2 mm dry material, is used to determine the wear characteristics of the ‘ore. During the course of the test, the paddle abrades to such an extent that its weight {oss is readily determined by weighing on an analytical balance, this being expressed as the Abrasion Index (AD). The abrasion index varies between 0 and 1, with typical values for lead zinc ore being 0.21, gold ore 0.48, and bauxite 0.02. Correlations have been established relating the weight loss of the paddle to metal wear in comminution equipment, but is very sensitive to the prevailing conditions in the plant. ‘Table 4.8: Example of JKMRC ore abrasion test results ore a STR cacy tina) art coarse epnesgrarinn) 2-6 fi Re RE REE Pinad aa | Oo ig oe oe Bs ig COATES DSB ann n+ 2.245,59 nd) Correlations with work index ‘There is some correlation between the ore impact (A*b) and abrasion (fa) parameters and the Bond ball mill work index, as shown in Figure 4.13. However, the work index measures the reaction of a -3.35mum feed to the mix of abrasion and impact present in a dry ball mill. The JKMRC breakage tests try to separate impact and abrasion breakage and, as such, would not be expected to correlate strongly, as supported by the data on 47 different ore types presented in Figure 4.13. 8 Chapter 4: Rock Testing - Determining the Material-Specitic Breakage Function 1.34 = 19700 Impact Parameter A°D ‘Abrasion Parameter Bond Werk Index, WI (kWh) ‘Bond Work Index, WI (kWh) Figure 4.13: Correlations between A*b and t, parameters, and Bond work index 46.2 Crusher Model Appearance Function ee The current crusher model (Andersen 1988), which is a development of Whiten’s original crusher model (1974), uses the energy-size reduction results derived from the Grop-weight or large pendulum test on coarse particles to predict both breakage and crusher power consumption. In this case the drop weight test results are used to determine two sets of crusher ore-specific parameters. ‘The first part is an appearance function which relates the degree of breakage, tio, to the remainder of the size distribution (Table 4.9). Unlike the AG/SAG model, the crusher model uses actual ore appearance data, not the standard mapping in Table 4.7 above. Table 4.9: Example of ore parameters used in JKMRC crusher model (@) Appearance Function Data, tn (2) to) 75 ‘60 25 Py @ ae 10 292 3.89 397 20.78 8 20 554 7.48 1173 40.16 7.04 ; 30 B24 10.97 1728 58.35, 88.35 (b) Size reduction/specitic comminution energy, Ecs (kWh) 1G tr (%) Initial Particio Size (mm) 122 175 23.0 70 a2 027 0.25 20 085 0.53 0.50 30 0.98. 079 076 ‘Chapter 4: Rock Testing - Determining the Material-Spectfic Breakage Function SHSs cence gees amines anaes peers esse acter eeeeeeeeesesaeeaeceeeeeemey ‘The second part of the crusher ore breakage description relates the size-specific energy requited to achieve a particular degree of breakage (Table 4.9b). The crusher model uses splines (or mathematical curves) to relate tyo to Ecs, instead of the exponential relationship (equation 4.22). These splines are described at nominal yg levels of 10, 20 and 30%, for each of three particle sizes used in the drop weight or pendulum test. ‘The latter information is also used to determine how much power (per unit feed mass) would be required to crush feed size to product size if all the crushing was carried out in the pendulum or drop weight tester. This ‘pendulum power’ is closely correlated with net crusher power consumption. This method of power correlation is much more flexible than the traditional F80/P80 approach (Morrell ef al 1992 - see also Section 1124). In applying the JKMRC crusher model when the feed is known and the product is the desired output, the problem resolves itself into obtaining estimates of the breakage function and classification parameters for a particular machine and feed material ‘Whilst the classification parameters are directly related to the crusher closed and open side settings, the ore-specific energy-size-teduction relationship for the crusher is ‘usually unknown. This is summarised in a parameter T10 which is related to the size distribution of the breakage products. Typically this parameter needs to be fitted to survey data from the crusher in question. However, in vertical shaft impact crushers it may be possible to determine the applied specific comminution energy and hence T10 parameter from first principles. The basis of this information comes from the drop- weight or pendulum tests on the rock sample, rotor speed and diameter. Further details are given in Section 6.9.1. ‘The drop-weight or pendulum test should be conducted on representative ore particles over the range of the crusher feed size. Where a specific mathematical performance model has been developed from plant surveys, the power draw may then be predicted for different operating conditions. Ina design situation, given the feed and the destred product size distributions, the tio- size-Ecs relationship (equation 4.22) for the ore to be processed must be obtained from the drop-weight or pendulum test and this information used in the model to calculate the total comminution energy required. The crushing power requirements can then be determined for a similar crusher from a power correlation obtained from another site. ‘When calculating power in this way, large particles appear to be softer in the impact tests, The current large pendulum test is not suitable for particles of diameter larger than 55mm. Hence, using specific comminution energies derived on smaller particles a7 Chapter 4: Rock Testing - Determining the Material-Specitic Breakage Function may overestimate required pendulum power in coarse crushing. The drop weight tester can however process larger particles. ‘The effect of particle size on the tyg-Ecs relationship is usually negligible in hard ores, but the A and b parameters may both be dependent on particle size, shape and ore type. Figure 4.14 illustrates an example where there was a significant difference in the ‘energy required to break fine and coarse particles. Each of the four curves in the plot is, described by equation 4.20. 40 = = Initial Avo, Particle Size $ Do 71.0mm a» 3 © 589mm 6 3 ° 28.1 mm. 2 io > 473mm ° ° 0.05 on os, 02 ‘Specitic Comminution Energy, Es (kWh) Figute 4.14; Ettect of patcle size on the t,p-Ecs relationship fo a pisoltc ore Effect of particle shape The way in which shape affects breakage is clearly shown in Figure 4.15, which suggests that distinct breakage distributions exist when two different shapes are broken. The shape definition used here is based on the aggregate industry's Flakiness Index, which relates the particle's smallest dimension to its nominal size. A flaky particle will have a least dimension (thickness) less than 0.6 of its mean dimension. Comprehensive data from pendulum tests on basalt (Kojovic 1994) showed clearly that flaky and non-flaky particles produce different product size distributions due to their different geometry; for a given size range the flaky particles will appear to be ‘smaller’ to the crusher than non-flaky particles, and will also exhibit lower strength when. 8 ae oat Chapter 4: Rock Testing - Determining the Material-Spectic Breakage Function eee eee stressed across the length because of their relatively thin cross-section. Hence flaky rocks produce more fines when broken at the same specific energy. ‘The classification of flaky particles in jaw, gyratory and cone crushers is expected to be different to non-flaky particles, because the flaky particles tend to align themselves in the crusher cavity in such a way that they behave as particles of smaller size, presenting their longest dimension parallel to the crushing faces. The crusher model which accounts for shape of the feed therefore requires shape specific breakage and classification functions. The above results have been used in the development of a new crusher simulation model which predicts product size and shape distributions (Kojovic 1994). 30 20 10 (%) 10 © Flaky © Nonflaky 00 05 10 15 20 Eos (kWh) Figure 4.18; Effect of particle shape on the tyg-Ecs relationship 463 Ball Mill and Rod Mill Appearance Function As the JKMRC ball and rod mill models assume that the contents of the mill are perfectly mixed, at steady state the material balance can be described by equation 4.21, If the contents of the mill are known, the breakage rates can be computed from equation 4.21 knowing the feed and product size distributions and the appearance function (Chapter 2). In practice, this is rarely if ever the case and the equation is modified to replace the mill contents with the term product/discharge rate. Since the ore-specific appearance function can be independently determined from the pendulum breakage test, the parameters of the model then become the breakage rate/discharge rate (r/d) terms, which are determined by non-linear least squares fitting of the modified equation to operating data, Normally x/d is found to be a regular function of particle size, and this dependency is represented in the model by spline functions, the estimation of the 89 Chapter 4: Rock Testing - Determining the Material-Specitic Breakage Function spline knots being part of the model fitting procedure. In practice r/d terms are also normalised through mill dimension which allows the model to scale (Chapters 2 & 9) ‘The bal! mill and rod mill models use an ore-specific appearance function determined using a specific comminution energy for a standard input energy level (Ei) of 41.788 kg.cm, applied to the -5.6+4.75 mm size fraction (Narayanan 1985). This has been traditionally determined using a smaller pendulum device than that used for the SAG/crusher tests, and more recently the lower drop weight tester energy range. Narayanan (1985) found that the relationship between the specific comminution ‘energy (Ecs) and the standard input energy (Ei) is of the linear form: Kes = a+ bE: 428) while the appearance function (to) is related to the specific comminution energy by the following equation: tyo= a+ bin(Ees) (4.25) where a and b are constants. The particle size effects are normalised in this relationship. With a standard input energy level itis possible to compare the relative characteristics of different ores. However, different mill diameters generate different energy levels for breakage and therefore this use of a fixed input energy level may not be appropriate. It follows that the specific energy should be a function of mill dimensions, as in the AG/SAG model. Current research is aimed at unifying the ball mill and AG/SAG mill models, which will address this question. ‘As noted earlier, it has been found that the relationship between typ and the specific ‘comminution energy Ecs (kWh/t) is best described by the exponential equation: tox afte PE] (422) ‘where A and b are dependent on the ore type ‘Hence the ball or rod mill appearance function is determined from the ore-specific tio calculated using either equation 4.25 or 4.22, and the standard 't' family of curves (equation 4.22 is preferred as it defines a maximum value of tig). As noted above, the to is based on the Ecs obtained using a standard input energy level (Bi) of 41.788 kg.cm, applied to a-5.6 + 4.75 mm particle. Hence the test Ecs will depend largely on. 90 Chapter 4: Rock Testing - Determining the Material-Specific Breakage Function SRE EEE EEE Heeb EEE the ore sg and particle shape, that is the average particle mass in the -5.6 + 4.75 mm size fraction. ‘Once the ore-specific typ -Ecs relationship and standard Ecs have been determined, the appearance function values are obtained using a computer program which seconstitutes the product size distribution from the predicted tyo and the given t- curves. The results of this program represent a relative size distribution (mass fraction retained), as illustrated in Table 4.10. ‘The appearance function starts with the value 0 since the model assumes that every particle once broken will no longer belong to the initial size fraction. Hence the daughter particles will be distributed in a 2 sieve series, starting with the next fraction down. To illustrate the range of values seen between hard and soft ores, Figure 4.16 shows the appearance functions or breakage distributions for a primary and oxide gold bearing ores. Note the shift in the peak fo the left with the softer oxide ore. Table 4.10: Typical ball or rod mil appearance function internal | Appearance Sizing (mm) | Values 16.0 0 13, 0.043 8.00 0.084 5.68 0.114 4.00 0.119 2.83 ttt 2.00 0.100 141 0.088 1.00 0.074 0.707 0.061 0.500 0.050 0.354 0.041 0.260 0.082 0.177 0.026 0.125 0.019 0.088 0.014 0.063 0.010 0.044 0.007 0.031 0.004 0.022 0.003_| 1 ‘Chapter 4: Rock Tasting - Determining the Material-Specific Broakage Function For ball mill modelling, and fine grinding devices like tower or stirred mills, the single particle approach as well as the range of particle sizes may in some applications be inappropriate. The single impact of a ball on to a particle bed situated on a flat or round surface probably describes the micro event in a ball mill better than a single particle impact. The particle sizes for which pendulum data are generated range from 5 to 55mm (though particles as small as 2 mun have been broken). Smaller sizes which represent the majority of material in a ball mill have not been examined critically in these experiments, Accordingly attention is being given to altemative methods of obtaining the breakage function for fine particles. Work is also being done on determining the breakage function for the comminution of particle beds, such as occurs in the high pressure grinding rolls (Section 6.9.4). 15 3 | = | é * Primary © Oxi x Z Oxide : | : 01 1 1 10 100 Particle Size (mm) Figure 4.6: Example of ball mill appearance function fore hard (primar) and sof (oxide) gos ore 47 CONcr.USIONs Various laboratory rock breakage characterisation tests have been developed for studying breakage phenomena and for designing and predicting the performance of industrial comminution equipment, These tests fall into three main classes: conventional rock and fracture mechanics measurements, standard grindability tests, and the single particle tests. Fracture toughness has been empirically correlated with crusher performance and, as such, is a rock property which may be useful in crusher design. The Hopkinson Bar is, being used to identify critical energy levels for the design of comminution machines, 2 Chapter 4: Rock Testing - Determining the Material-Specific Breakage Function sr epeseEEeEEeCHLs-64zes ESEEEECEE Up ee aecsesreeeceeeee eee seeseeeeee and the role of damage in comminution. This may have potential applications in assessing the strength of critical size material in AG/SAG mills, and blast damage in ROM material. Numerical modelling of standard breakage events has provided valuable insight into the mode of fracture in comminution, particularly for heterogenous materials. ‘The JKMRC simulation models aspire to a clear goal, that of separating ore characteristics from those of the processing machine. In this context, neither of these approaches provide information on the breakage function, which is necessary in the prediction of the product size distributions from crushers and tumbling mills. ‘The conventional grindability test is based on Bond's ‘third theory’ of comminution (4952). This method continues to be the main tool for designing comminution machines, and its long use has engendered considerable confidence in its application. ‘The particular value of the Bond grinding test is its simulation of the recycle elements of closed circuit grinding systems, including the behaviour of components of different ‘nardness’. However, the Bond test also has some disadvantages. For example itis not a good predictor of AG/SAG mill behaviour (for which it was not designed), nor of situations in which it is desired to investigate changes (usually increases) in feedrate. ‘The monosize grinding test for the population balance model based on Reid's equation (3965) estimates the breakage rate and the breakage distribution function. The method has provided a useful framework for the description of the operating behaviour of ball mills. In spite of this, a standard test procedure is yet to be established and accepted. In addition it has proved to be a weak predictive model because the breakage rate and the breakage distribution function for particles are not related to the physical micro- processes involved in comminution. This has encouraged research into breakage processes using single particle tests to characterise breakage properties. Single particle breakage tests are mainly of two types - pendulum tests and drop weight tests. Various types of pendulum and drop weight devices have been developed and used in the past. This chapter has described the methodology developed and now in routine use at the JKMRC and in a number of licenced testing laboratories around the world. ‘The breakage size distributions resulting from breakage of hard-rock ores in a pendulum or drop weight device have been shown to form a one-parameter ‘t’ family of curves independent of ore type and normalised for particle size. In addition an empirical equation with ore-specific parameters is used to relate the tyo parameter to the specific comminution energy, Ecs, measured in the pendulum or drop weight test. This approach covers the range of shapes of breakage distribution functions commonly Chapter 4: Rock Testing - Determining the Material-Specitic Breakage Function used in crusher and tumbling mill modelling, depending on the Ecs level of the single particle breakage event. In summary, the energy-size reduction results from pendulum and drop-weight breakage tests are now routinely used to provide ore parameters for use in JKSimMet simulation models of ball and rod mills, autogenous/semi-autogenous mills and crushers. The unit parameters (eg. breakage rates) are calculated from plant survey data. Once the models have been customised to an existing circuit, the behaviour of the comminution circuit over a wide range of operational conditions can be accurately predicted, as evidenced in numerous optimisation studies carried out to date. In design ‘of mew plants, the extensive JKMRC database on ore types and related circuit performance is used to select appropriate ore and unit parameters for the simulation studies. Current experience suggests that the single particle test, when combined with process models, is superior to other methods for evaluating coarse breakage machines such as crushers and AG/SAG mills. Although itis also used successfully in describing ball mills, itis likely that the breakage characteristics of relatively large rocks (say down to Smm) may not describe well the same material at say 100um. Further work is therefore required to improve the independent measurement of the breakage characteristics of fine particles, and this research is underway. oii CHAPTER 5 SURVEYING COMMINUTION CIRCUITS 5a INTRODUCTION he ability to assess the performance of comminution circuits depends largely on the quality and nature of information collected from the circuit. This is particularly true of any modelling and simulation study. A survey consists of collecting data and samples from the circuit over a particular operating period, which are representative of the operation of the circuit during that period. The representative character of the survey is important, as circuit analysis and simulation is based on the ability to build a model that is representative of a real system. There are several issues which need to be addressed, including the nature of the data to be collected, sampling points, the method of sampling, the equipment used, and the sample processing which follows. Many types of comminution machine and circuit are discussed in this monograph. In broad terms, however, three main types of circuit are encountered in practice: + Crushing and screening - e.g. iron ore processing, coal processing, quarry processing, diamond recovery, and preparation for grinding. = Coarse grinding, using AG/SAG, rod and/or ball mills - e.g. metalliferous processing, as preparation for beneficiation processes such as flotation or leaching; an example of this kind of circuit is shown in Figure 5.1, including typical sampling points. + Fine grinding, using ball, tower or stirred mills - e.g. regrind in metalliferous processing, or industrial mineral processing. Each of them has particular features which must be addressed in any sampling or data collection exercise; sampling a crusher feed is very different in a practical sense from sampling a hydrocyclone overflow. However, there are some guiding principles and techniques which apply to all such campaigns, and which must be adhered to for best results 96 Chapter 5: Surveying Comminution Circuits ‘Sampling point Final product eg eee New food VS dive AGISAG Mit Figure §.1: Schematic diagram of typical comminution ctcutt, showing some altemative eub-circuits This chapter presents a general practical methodology for sampling and data collection in conventional crusher, SAG/AG and/or bali mill-cyclone circuits. Additional information is included on data mass balancing techniques, sample analysis (particularly size analysis), the selection of sample size, examples of the types of errors that occur in the sampling of pulp streams within a grinding circuit, and some notes on model parameter estimation. 52 ‘SOME SAMPLING PRINCIPLES 521 Introduction ‘The objective of all sampling is to obtain a representative sample. This is an ideal concept which is rarely realised in practice. A formal definition of the term is that “all particles present in the stream have an equal chance of being selected in the sample by the sampling device.” a3 Chapter 5; Surveying Comminution Circuits OS ‘The problem is really a statistical one, since it depends on the nature and magnitude of the errors which accumulate when collecting and processing a sample. Some of the errors and disturbances which can contribute to overall error in determining some quantity such as a solids concentration or particle size distribution are: 1. Plant dynamics. Processes are rarely in steady state, and sampling policy must be determined accordingly. A ‘snapshot’ sample may be appropriate for a single device with essentially no dynamics, such as a crusher or hydrocyclone, but a circuit is usually sampled by accumulating a number of incremental samples taken over 1-2 hours to ‘smooth out’ disturbances in the process. 2. Sample cutter design - see Section 5.2.4 3. Sub-sampling a primary sample, 4 Analytical errors, e.g. weighing, screens with worm or incorrect apertures, inadequate screening time, incorrect calibration or selection of constants, etc. Such errors are more common than generally admnitted (see also Appendix 3). 5, The propagation of error when calculating quantities - see Section 5.2.3. 6 The fundamental statistical uncertainty (error) involved in choosing a small, finite sample to represent the properties of a large (effectively infinitely large) population - see Section 5.2.2. ‘The surveyor has some control over the first five of these, but essentially no control over the fundamental error (FE) which isa statistical property of the particulate system being studied. The object of any sampling and analysis exercise should therefore be to minimise the effects of items 1 - 5 so that their contribution to overall error is small relative to FE. The size of sample is then chosen to achieve the required confidence in the light of the prevailing FE. : Sampling statistics is a complex topic. Gy (e.g. 1976) has developed a widely-used theory of particulate sampling, which has been adapted and described by Pitard (1993), who also discusses many other aspects of sampling such as cutter design. 522 Size of Sample for Size Analysis How large a sample of solids from a process stream should be presented for size analysis? One answer is: as large as possible, since this will maximise the reliability of the final result. In practice, however, the selection of sample size is a pragmatic compromise between economics (the cost of collecting and processing the sample) and the confidence needed in the answer. 7 Chapter 5: Surveying Comminution Circuits Chapter 18 of Pitard’s book deals specifically with sampling for size analysis. However, to give a simple estimate of the order of sample size required to accommodate the FE, Barbery (1972) derived an expression based on Gy’s theory, which is easy to use: fpdm ep 64) mass of sample required (g) shape factor for material (0 < f <1) density of material (g/cm?) mean size in size range of interest (cm) expected proportion of material in size range of interest (to be smeasured) @ = _ standard deviation of the number of particles in that size range. (The variance 8? is essentially the FE referred to earlier) ‘These terms require some explanation. The shape factor of a single particle is f = m/pd®, £ = 0.1 for flat, plate-like particles and approaches 1 for spheroidal particles. For most natural ores and coal, 0.3 < f < 0.7, and 0.6 is not a bad guess in many cases. : 3, can be calculated as (Barbery 1972): 2 62) where di and dz are the limiting sizes of the size range of interest. The size range of interest is that which is likely to have the least number of particles in it, which is nearly always the coarsest size interval. This ensures that the error on the proportions estimated for the other size intervals will always be less than this, thus ensuring a conservative choice, However, itis pointless to choose an interval with practically no ‘material in it. A good rule of thumb is that the coarsest size interval should be chosen togive P= 5%, @ is determined from the precision of estimation and confidence required: 2 z 63) chosen precision (telative proportion) = normal ordinate at the chosen confidence level. 8 where z 98 Chapter &: Surveying Comminution Circuits ‘Table 5.1 gives values of 2 for different confidence levels. Table 5.1: Normal ordinates (rom the normal distribution) Confidence Level (%) z 60 0.6745 80 1.2816 90 1.6449 95 1.9600 9 2.5758 20.9 3.2905 ‘A confidence level of 90% (z = 1.64) is usually adequate. The definition of @ is interpreted as follows: if the proportion of material expected to lie in the coarsest size range is 5% (P = 0.05), and we want to estimate this to a precision of 10% relative with 90% confidence, then e {e9/100) = 0061, and P = 5% + 0.5%, with 90% confidence. Example 51; A SAG mill trommel oversize stream in a base metal concentrator is to be sampled in a plant survey to determine its size distribution, It is expected that all the material will be less than 50mm, and about 10% will lie in the (top) screen size range - 50 + 25mm. How much sample should be screened to ensure 90% confidence of determining this proportion to a relative precision of 20%? (ie. P= 10 £ 2%). 3 4258 B - 2425 3 703 cm ae 2 0 = 02/164 = 0.122 ‘Trommel oversize from SAG mills is often quite rounded, so assume f = 0.7. Also p= 3.0 and P = 0.10. Substituting these values in equation 5.1 gives M = 99kg. This is a large sample and demonstrates the uncertainties in sizing coarse material. If the size interval had been -5 + 2.5mm, M = 99g, because of the much larger numbers of particles per unit mass at this finer size. The calculated mass is also conservative, because the coarsest fraction and smallest mass proportion (10%) is very much the ‘worst case; the other size intervals would have much better precision. Clearly equation 5.1 can be re-arranged to calculate the precision or confidence limits on an actual result: Chapter 5: Surveying Comminution Circuits 6.4) Example 5.2: In the preceding example, only 10kg was screened, and the amount in the -50 + 25 mm fraction was found to be 14.5%. What are the 90% confidence limits on this result? M = 104g, and P=0.145. From equation 5.4, = 0.319 and thus the 90% confidence limits are + 0.319 x 1.64 x 14.5 = 7.6%, ie. the 90% confidence interval is (14.5 - 7.6)% to (145 +7.6)%. The relative precision is therefore 7.6 x 100/14.5 = 52% - very poor! Clearly the FE for size analysis is really only an issue in the coarsest sizes, ie. in crushing, AG/SAG milling and coarse rod and ball milling. For flotation or fine gravity concentration streams and in fine grinding, the FE can be accommodated with only a few grams of material, or less, and the size of sample is then determined more by the other sources of error than by the FE. For very coarse materials, the required sample size will be larger than it is practicable to process in a reasonable time-frame (typically 5 tonnes for a 200mm top size), and some compromise is generally sought ‘Means by which an acceptable estimate of run-of-mine (ROM) sizing can be obtained are discussed in Section 5.4. 523 Propagation of Error in Calculated Quantities Some quantities in a survey will be calculated rather than measured. A good example is solids concentration in a slurry. Although it is advisable to ‘measure’ it directly by dewatering and drying a sample of slurry, it is sometimes inferred from a slurry density measured with a Marcy scale, The solids concentration by weight, C,, is then given by _Pslp-9 ae pps—D where p, and p, are the solids and slurry densities respectively, with water as the fluid. Assuming the errors in measuring or ‘knowing’ p, and p, are small, the law of partial differentiation states that Wy ay Rey = Fi Bost Fp 66) 100 Chapter 5: Surveying Comminution Circuits Ege HELE where 6 signifies ‘maximum error in’. 14 Wy Pe and 2m = = 6n Ps pps -1) Pp pps) Suppose p, = 2.65 and pp= 1.15. Then from equation 55, C, = 0.2095, or 20.95% solids, land from equations 5.6 and 5.7: BCy = (0.048 + Bp) +(1.214 + Bp) 68) Note that signs are ignored, because the objective is to estimate the maximum possible error in the answer, ie. when everything that can go wrong does go wrong, The &p, and Bp, ‘errors’ are usually taken tobe the precision in reading an instrument (eg; the Marcy scale) or the maximum error in measuring the quantity (e.g. the assumed solids density). If we assume in this case that 8p, = 69, = 0.01, then &C,, = 0.0125, or 1.3%. ‘The solids concentration is then 21% 1.3%. ‘Two points are worth noting: 1. The calculated error can be quite large, even though its components might be regarded as ‘reasonable’ in terms of normal practice. 2. Errors offen propagate in unexpected or counter-intuitive ways. Inspection of ‘equation 5.8 shows that in this example, measuring the slurry density contributes 96% of the total error, and errors in assuming the solids density are relatively ‘unimportant (assuming the two errors are of the same absolute magnitude). ‘A similar exercise should be conducted for all such calculated quantities, so that the surveyor can identify those quantities which need to be measured with particular care. 5.24 Sample Cutters Sampling coarse material on a conveyor belt is usually done by stopping the belt and removing an appropriate length of material (Section 5.4.2). Proprietary automatic belt and transfer point samplers are also available. ‘The usual sampling device used in grinding surveys is the manual sample cutter. “Unfortunately in many cases a Marcy can or bucket is used. This pract appropriate. A design of the type shown in Figure 5.2 is suitable for fine slurries. For coarser mill discharge samples a cutter design of the type illustrated in Figure 5.3 2 is not 101 Chapter §: Surveying Comminution Circults should be considered. Sometimes a 50-75mm diameter pipe may be inserted into a stream to divert some of the material to the floor, where a sample can then be taken. Ingenuity is often necessary due to difficult local conditions. The major problem with using sample cutters is their bulk and weight. Sample point accessibility in grinding circuits can be atrocious, particularly SAG/AG and ball mill discharge samples (from underneath trommels). Tyee Typo? 250 10, + tie I eND VIEW SIDE View So ost ss must be staight and arate! (Hands 1o be added as required, 9, Pip threaded at ore and} NOTES: Typo #1 : Cycione OFF ‘Type #2: Cyclone UF Dimensions ae in mm Matai shoud be 1.6mm mis steal sheet Figure 6.2: Recommended cutter design for cyclone underflow and overflow sampling ‘Another major problem is the sampling of large pulp flowrate streams (>100 tph). The large flowrate can cause particles to become ejected from the cutter even when the cutter is passed quickly through a pulp stream. Obviously if this happens the sample should be discarded and the sampling process repeated or reviewed. Chapter §: Surveying Comminution Circuits Feectangutar section asic pipe ‘Opening height > stream width Closed ena Figure 5.3: Sampling a coarse particle high volume flow (e.g. SAG mil discharge) Sampling pulp streams over 250tph should not be done with hand held cutters. Considerable practice in cutting stream samples should be accumulated before any serious sampling exercise is attempted. If possible, always stand above the stream being sampled to improve control over the cutter. ‘The cutter edges should be horizontal and parallel. For streams where the largest particle size d > 3mm, the minimum width of the cutter should be at least 3d. For smaller particles (4 <3 mm), the width should not be less than 10mum. The depth of the cutter should be large enough to prevent the sample splashing out or overflowing. ‘The cutter should be passed through the total stream cross-section at a uniform speed. If this is not feasible for large flowing streams, a number of successive cuts moving across the total stream cross-section is recommended. Care should be taken that ‘material does not fall into the cutter between cuts (eg. dust or slurry). ‘The cutter should be ‘washed’ in the slurry stream and tapped to ensure that it is empty before sample cuts are taken. Dilution with water would alter the percent solids obtained from the sample. Chapter §: Surveying Comminution Circuits 33 DESIGN OF COMMINUTION CIRCUIT SURVEYS 53.1 Objectives and Requirements ‘The objectives of the survey should be clearly established before planning and undertaking it (see also Section 13.3.1). These may include the establishment of the performance of a single unit (before or after optimisation), the audit of an entire section or circuit, or the collection of data for model parameter estimation (Section 5.633). ‘The survey may involve a simple size analysis of a crusher product, a mass balance around a cyclone, or a full circuit sampling campaign. A full campaign is defined as a survey from which the mass flowrate of solids, water and particle size in every stream of the circuit can be determined. In this discussion, the survey is assumed to take place in steady state, ie. mass flows and stream characteristics remain approximately constant during the period of a survey. The mass balancing algorithms used to smooth the data then assume that input flow = output flow for each process node, for both solids and water. ‘The design of control systems often requires step tests to be conducted to characterise the dyriamics of the circuit, In these cases, a disturbance such as a change in feedrate is deliberately introduced, and the circuit behaviour monitored over time to determine the effect, Such surveys are not discussed here. Methods of conducting comparative experiments on plants in which feed or other conditions are subject to uncontrolled variations with time are discussed by Napier-Munn (1995). Note that it is often not practicable to measure the flowrate and size distribution of every circuit stream. Providing that a minimum number of stream samples and at least one reference dry mass flowrate are taken, the remaining streams can usually be calculated. Such a survey is referred to as a minimum sampling campaign. Any excess of data above this minimum is not superfluous, but highly advantageous because it provides cross-checks within the data set (data redundancy), allowing more effective mass balancing of the data An example of the data requirements for a minimum sampling campaign for a grinding circuit of the type shown in Figure 5.1 is: 1. New feed dry mass flowrate. Trommel screen oversize flowrate (if pebble crusher running). 3. Percent solids of new feed, AG/SAG mill and ball mill discharge, cyclone feed, underflow and overflows. vos Chapter §: Surveying Comminution Circuits 4. Sizings of new feed, AG/SAG mill and ball mill discharge, trommel screen oversize (and pebble crusher product if running), cyclone underflow and overflow. Clearly if one element of data is measured inaccurately (eg. cyclone feed percent solids) problems will arise in the processing of the data. Therefore the general rule is never to conduct a minimum sampling campaign unless no other option exists. ‘The ideal full circuit sampling campaign for grinding circuits of the type shown in Figure 5.1 requires the following data: 1. New feed dry mass flowrate. 2 Trommel screen oversize flowrate (if pebble crusher running). 3. Percent solids of new feed, AG/SAG and ball mill discharge, cyclone feed, underflow and overfiow. 4. Sizing of new feed, AG/SAG and ball mill discharge, trommel screen oversize, cyclone feed, underflow and overflow. 5. Online flowmeter and slurry SG gauge on cyclone feed. 6. Online water flowrate measurements. Items 1 to 4 are usually possible on most plants although cyclone feed percent solids and sizings are sometimes omitted. This omission can cause serious problems. Item 5 is valuable to have as it provides an independent cross check of both the cyclone feed petcent solids and mass flowrate to the cyclone, and often simplifies the mass balance. Unfortunately the flowmeter readings can often be in error by as much as 10% due to calibration problems. Item 6 is even more valuable, as water flowrates can be measured to a high degree of accuracy (unlike slurry flowrates). These values can therefore be used to cross check the mass flowrates at various points in the circuit 5332 Typical Survey Procedure for a Grinding Circuit ‘The set point tonnage should be held as constant as possible during the sampling campaign. To achieve this in grinding circuits, the grinding control loops may have to be de-activated. Once set up, the grinding circuit illustrated in Figure 5.1 will run at reasonably steady state conditions providing feed sizing, flowrate and ore hardness do not change significantly. Once a set of operating conditions has been chosen, at least an hour (ie, several residence times) should be allowed for steady state conditions to occur. It should be noted that steady state condition is an ideal state and is never fully realised in practice; however an indication of the degree of steady state can usually be obtained from the consistency of the cyclone feed pump amperage draw and speed (for VS drives). Chapter §: Surveying Comminution Circuits FEES ses seseeeeeeeceeCeSese eae seeCeeL SS ueeeesesceeeeeeeee eee eee With the advent of DCS monitoring software, it is relatively straight forward to prepare and view a group of relevant operating variables prior to commencing the survey. The same data, including the period of the survey, are then stored on disk which can be easily accessed once the survey is complete. An example of the circuit data monitored during such a survey period is shown in Table 5.2 Table 5.2: Example of steady state circuit concitions TIME: SAG Mil BALL MILL. Bali Mil CYCLONE | (am) | teed | water | power speed] power | water | pressure | feed pump] density wn) | cm | caw | (rom | ce _| (Pm) | (KPa) | output (6) feed (26), 700 | 105 | se | eas | 15 | 1864 | 395) 123 | 759 | 315 710 | 115) 58 | 867 | 15 ) 18608) 396 | 118 | 738 | 318 720 | 116] 59 | 950 | 15 | 1874 | 398 | 119 | 734 | 313 730 | 116| 58 | a7 | 15 | 1863 | 395 | 110 | 702 | 304 740 | 117] 58 | a40 | 15 | 1962 | 398 | 115 | 729 | 309 zso_| 118 | so | a73 | 15 | 167 | 400 | 115 | 731_| 310 ‘Note that in this example the cycione is run at a low feed density to achieve a fine cut; cyclone ‘overflow is thickened prior to leaching. ‘The circuit should be sampled over a period of 1-2 hours (either 8 cuts every fifteen minutes, or 4 cuts every thirty minute period). Multiple samplers (people) make the sampling more efficient, as more time is made available for the cutting of the stream samples. It also allows the samples to be taken simultaneously (or in a logical sequence appropriate to the residence times in each section), which is highly advantageous. If this is not possible the order of samples should be varied. Ensure that sample containers are clearly labelled; where multiple containers are used for the single sample, each container should be marked with container number and total number of containers. Ifa significant disturbance occurs during a survey, either terminate sampling if close to the end of campaign, or otherwise abandon sampling and return another time. Back- up samples should if possible be collected in case of mishaps in the lab during sample processing, It is useful to know the ball charge filling of the mills after each sampling campaign. Therefore, after collecting the slurry samples, the mills should be crash stopped to allow feed belt samples to be collected and the mill charge level to be measured via physical inspection of mill contents. This key step in the survey is discussed in more detail in the sections which follow. Feed size analysis requires special consideration. It is important to obtain an accurate measurement of the feed size distribution particularly at the coarse end. It is therefore necessary to collect a belt cut sample and record the belt speed and length of belt over 406 wuved Chapter §: Surveying Comminution Circuits GREECE eee eee which this sample was collected. Feed ‘samples should also be taken for ore characterisation tests (Chapter 4). As snapshot samples are the usual method for sampling crushers and screens, it is important to stop all the necessary conveyors at the correct time. To ensure that the samples taken from product conveyors are representative of the measured operating conditions, the feed belt can be marked at the start of shutdown by dropping an appropriate marker onto the belt or processing unit (e.g. flour, paper, etc), and the product conveyors stopped when the marker appears. ‘The samples collected need to be processed on site to determine size distribution and percent solids. Where possible, consideration should be given to collecting a separate sample solely for percent solids determination. The procedure to be used to analyse the slurry and feed samples is given in Section 5.5. The plant dry ore feed rate (and other belt flowrates where possible) should be determined from the weightometer readings and feed sample moisture measurements. ‘The weightometer calibration should be checked using measured belt speeds and a suitable belt sample, as described in Section 5.4 [Readings of the fresh feed rate to the circuit, power draws, cyclone or screen feed rate and density, and water additions to the circuit should be recorded periodically, from which averages over the survey period can be determined. Cyclone pressures should be recorded manually. If possible power readings should be cross-checked using local power meters or kWh meters (see below). A full list of data to be collected follows. 53.3 Datato be Collected Apart from the stream samples and circuit operating condition, the information required from a full comminution circuit survey is listed in Table 5.3 below. Note that all internal cyclone dimensions should be measured, not assumed. Normal wear, and maintenance errors, can lead to significant discrepancies in cyclone dimensions. For cone crushers, lead the crusher (to determine closed-side setting) in at least three positions around the mantle immediately before or after the survey. Lead thickness should preferably be measured with a micrometer or calipers, not a ruler. Take the average of the measured readings. For crushers with automatically adjusted gaps, record gap reading and calibrate against measurements with leads. 107 Chapter 5: Surveying Comminution Circuits ‘Table 6.2: Data to be collected trom particular unit operations. Crushers Hydrocyciones ‘Type ‘Type Closed side setting Number of eyclones Eccentric throw Cyclone pressure (kPa) No-Load power draw (kW) Feed flowrate and density Operating power draw (kW) Diameter Wear condition (liner age) Inlet diameter Feedrate weightometer reading (tph) Cylinder length Vortex finder diameter Mills Spigot diameter ‘Type Cone angle Number of mills Diameter (inside liners) ‘Vibrating Screens Cylinder length (inside liners) Type Cone length (inside liners) Number of decks Speed (spm, or % critical) Length x width Discharge mechanism. ‘Aperture (length and width) Grate aperture /layout dimensions Hole type (square, round, slot etc) Pan lifter depth (grate mills only) open area ‘Trommel dimensions and aperture Angle of inclination Lifter height Deck material ‘Trunnion Diameter Ball Filling (% mill volume) Conveyors ‘Total charge filling (% mil! volume) Belt speed Ore SG Ore type No-Load power (kW) Operating power draw (kW) Feedrate weightometer reading (tph) Bearing load and pressure (kPa) Power Measurement Historical data (e.g. WI, UCS, etc) Experience shows that crusher or mill power draw is often one of the least reliable of the quantities collected during a circuit survey. The surveyor should ensure that all power records are legitimate and correct. Power draw is measured in several ways. The best way is to use an in-line power transducer to measure in-phase power, which does not therefore have to be corrected for power factor. Most modern plants will incorporate such measurement, with a readout in the plant control room or motor control centre (MCC). Even here, signal transmission or calibration problems can lead to an erroneous readout in the control room, so the readings should be cross-checked wherever possible. 108 Chapter 5: Surveying Comminution Circuits In older plants, one sometimes has to rely on an analogue ammeter (often with a wildly swinging needle). In such a case, for AC 3-phase motors, total power delivered to the load, Pz, is given by Pr = VilLv3 cos@ | KW 69) where V__ = _ line voltage (rms value, V; veries with site and machine) 7 line current (ems value, A; read from the ammeter) 4 phase angle between the phase voltage and phase current. Cos ¢ is the power factor for the circuit, which will vary with the motor’s inductive load. For normal loads, cos = 08 - 0.95. For ‘no-load’ power draw, the pf. will be lower. The plant electrical engineer should be consulted as to the correct power factor to use in the calculation. plant (average) p¥. is not necessarily an appropriate value, as the individual machine's load may differ significently from others. ‘The golden rule is to cross-check all power readings and calculations. If possible take a portable power meter to site and make independent readings. Where more than one >plant reading is available, all should be taken and cross-checked. Finally, the crusher and mill power prediction procedures described in Chapter 11 are now good enough to provide an independent check of plant data. 34 SAMPLING PRACTICE 541 Equipment and Resources Required Note that in all sampling surveys, the proper safety procedures must be followed at all times. These may include, where appropriate, the electrical isolation of equipment, and the use of safety harness when: taking awkward samples. For major sampling ‘campaigns, a written procedure should be prepared for eack participant. For coarse material, due to the mass of samples required, itis recommended that 200 litre drums be used to hold the original samples and the various size fractions after screening. A front end loader or fork-lift which is capable of lifting and up-ending the drums will also be necessary to manoeuvre the samples during sizing, Drums will have to be moved by way of chains or drum lifters in the case of using a forkclift. The following resources /equipment will be required on site: + atleast 4 persons (including front end loader /forklit driver) * 200 litre drums or similar * buckets and shovels + 0.500 kg platform balance 109 Chapter 5: Surveying Comminution Circuits Peete seseeeeeee eee tates ECE Ce EEE A eae eee + front end loader or forklift truck + 450mm (106,75, 53, 375,265, 19, and 13.2mm) and 200mm sieves (9.5, 6.7, 4.75, 3.35, 2.36, 1.18mm, and 850, 600, 425, 300, 212, 150, 106, 75,53, and 364m) + ‘one hole’ flat bar screens (100, 125, 150, 175, and 200mm) For slurry sampling good buckets are essential survey tools, Use 10 or 20 litre buckets with good seals on the lids. ‘Tare the buckets (with or without the bucket lid). Aim to take about 10-15kg of pulp from each of the sample points. Any more than this and the filtering process can start to become even more tedious than it normally i. $4.2 Millor Crusher Feed ‘As noted earlier, the size of sample required from statistical considerations for coarse streams (Section 5.2.2) can be impractical. To tackle this problem, the JKMRC has adopted an approach based on a combination of coarse and fine sampling, as outlined below. After crash-stopping the crusher or mill, and isolating the feed conveyor drive, the feed sample can be taken. This should comprise a complete 2-5m belt cut plus at least 50 475mm rocks. This typically is a total sample of at least 500-800 kg: If people are available to assist, a typical SAG mill ROM sampling run will require 4-6 persons and will take approximately 30 minutes to complete. The procedure should be as follows: 1. Check the speed of the conveyor belt using a tachometer fitted with a linear velocity measurement wheel, or a stop watch by timing how long a marker on the belt takes to pass between two points a known distance apart, say 50m. 2. Designate a section of the belt from which the feed sample will be removed, and position accordingly the 2001 drums, shovels, brush and dustpan in preparation for the sampling. 3. Once the conveyor is safe to access, walk along the conveyor belt and remove at least 50 coarse rocks visibly larger than 75mm. Determine the length of belt from which the coarse rocks were collected. Assuming the feed size distribution can be represented by the binomial distribution, this should give a reasonable estimate of the coarse end. 4, After removing the coarse racks, select a 2-5m section of the belt which represents a typical fines loading and carefully remove all the solids from within the measured section. Figure 5.4 illustrates this process at the Pasminco Elura Mine, sampling the AG mill ROM feed. 410 Chapter 5: Surveying Comminution Circuits Figure 5.4: Example of 7.14 x 3.78 m AG Mil ROM fine feed sampling at Pasminco Elura Mine 5, The coarse rocks should be sized using the “one hole” flat bar screens (+75, +100, #125, +150, +175, and +200mm), the number of rocks in each size fraction recorded, and each size fraction then weighed. Figure 5.5 shows the end result of such a procedure at WMC Leinster Nickel Operations. Figure 5.5: Example of 9.4 x 5.6 m AG Mill ROM coarse feed sizing at WMC LNO A platform balance positioned on the conveyor gallery floor is recommended, as the coarse rocks can be tossed back onto the belt once sized and weighed. This size distribution (mass per m in each coarse size fraction) is then combined with the feed size distribution determined from the 2-5m belt cut of fines (mass per m in each -75mm size fraction) to give a complete feed size distribution, as illustrated in Figure 5.6. 1 Chapter 5: Surveying Comminution Circuits 100 F (cum, % Passing a | "Lo aaleaseederoe™| oot an + 1 ae 1000 Particle Size (mm) Figure 6.6: Example of SAG mill ROM feed size distribution determined using the UKMRC belt ‘sampling procedure at the Cominco Red Dog Mine ‘The best method for a ball mill is to stop the feed belt and remove about 1-2 metres (about 20-25kg). Based on a top size of 12mm, and removing 25kg, this should take 5- 10 minutes. (Crushers fed directly from screen oversize are usually a major challenge. A crane skip positioned on top of the crusher allows a good sample to be taken. Similarly, screen discharge samples may require lip samplers of a suitable design to cut either the ‘complete stream or part thereof. Accurately timed lip samples can provide good sizing and flow rate data, A long feed conveyor is useful for obtaining a feed which can be sub-sampled as discussed above, fed through the crusher and the resulting product identified and sampled using the matker scheme mentioned in Section 5.3.2. It is almost impossible to make screens or crushers operating in parallel behave in an identical manner. Therefore it is much better to test one unit over a range of operating conditions. Mass flow measurements are available immediately after weighing. A relative difference between the feed and product flow of less than 20% is usually satisfactory. Less than 10% variation is good for sampling around crushing and screening circuits $4.3 Mill Contents Charge Distribution ‘The JKMRC AG/SAG mill model relies on the calculation of mill load mass and its size distribution (Chapter 7). This information is often made available from pilot plant studies. Should the ball and charge size distribution for a given mill not be available, it is back calculated from the surveyed discharge size distribution and estimate of load 112 Chapter §: Surveying Comminution Circuits Reef seer ee See eae eee eeEe Eee eters cree mass and filling. In this instance, the mill breakage rates may not be unique in the sense that a range of breakage regimes may produce a similar discharge and mill filling, yet show quite a difference in load size distribution. To provide the necessary data to test the model on full scale mills, the JKMRC has undertaken the arduous task of sizing the complete mill contents on three separate occasions, namely at Warrego, Alcoa Pinjazra and Red Dome. Dealing with hundreds of tonnes of charge, which comprises solids and slurry, requires the task to be split into two phases, the slurry being pumped and the solid being moved using mechanical shovels and trucks or front-end loaders. This requires the mill to have an opening in the shell to allow the contents to discharge from the mill ‘The slurry should be drained to the mill floor sump from which it may be pumped to another tank or circuit. Whilst pumping, the volume flowrate should be measured and samples of slurry taken for density and size analysis. “The solids fraction should be loaded using front-end loaders into available trucks and after weighing at the weighbridge, dumped in a suitable clear area. The total lot should then be riffled to an appropriate sample sized and hand sieved down to 13.2mm. The treatment of such samples is outlined in Section 5.6. Figure 5.7 shows a portion of the Red Dome SAG mill load resting on the mill floor, prior to being trucked. Figure 6.7: Exampie of §.0 x 6.6 m SAG mill load at the Nuigini Mining Red Dome Mine ‘An alternative procedure that may provide a reasonable estimate of the coarse end of the load size distribution is to remove a few tonnes from the mill after inching the mill for several revolutions, The inching is required to mix the contents of the mill and avoid segregation. un Chapter 5: Surveying Comminution Circuits Figures 5.8 and 5.9 show the WMC Mt. Keith SAG mill in the process of being inched around, after which a channel sample of approximately two tonnes, shovelled into six drums, was removed with the lining machine for analysis. i Figure 6.9: Sampling of SAG mill load at WMC Mt. Keith Chapter §; Surveying Comminution Circuits SREP eevee eden Sen Mill Filling ‘The mill filling is the volume of charge in the mill which is the principal determinant of power draw. in the case of ball mills the filling remains fairly steady over time, as it comprises mostly steel balls. In AG/SAG mills, however, the feed ore contributes significant quantities of rock to the grinding media, As such, variations in feed ore hardness and size distribution will affect the quantity of ore in the mill and hence the power draw. Knowledge of the mill filling is therefore a key parameter in a grinding survey, as it quantifies the response of a given mill to the prevailing feed ore characteristics. ‘A direct measurement of the load entails the crash-stopping of the AG/SAG mill under Joad whilst the mill is running under steady state. Once access to the mill is gained, typically after the feed chute is removed via the feed end trunnion, the width of the charge in three places should be taken, together with inside-liner dimensions. From these measurements the load volume may be calculated using simple geometry, as shown in Figure 5.10. Figure 5.10: Example of mill oad geometry parameters used to calculate volume Se2esin5 6.10) Gay 62) 6.13) 614) Chapter §: Surveying Comminution Circuits 6.15) where S = chord length (m) ‘mill radius (m, inside liners) depth of charge (m) = angle that chord subtends at mill centre (°) ‘cross-sectional area of mill charge (m*) = cross-sectional area of mill (m?) = estimate of mill filling (%)

Vous aimerez peut-être aussi