Wa0007.
Wa0007.
OPTIMISATION
SECOND EDITION
Spectrum 25
Edited by Christopher J Greet
Flotation Plant
Optimisation
SECOND EDITION
Edited by Christopher J Greet
Published by:
The Australasian Institute of Mining and Metallurgy
Ground Floor, 204 Lygon Street, Carlton Victoria 3053, Australia
© The Australasian Institute of Mining and Metallurgy 2019
The Institute is not responsible as a body for the facts and opinions advanced in any of its
publications.
‘The great tragedy of science – the slaying of a beautiful hypothesis by an ugly fact.’
Anonymous
1. Blainey, G, 1963. The Rush that Never Ended: A History of Australian Mining (Melbourne University Press: Melbourne).
My career-long journey of discovery of the technical literature of the art of flotation started as a new graduate in Applied
Chemistry without any specific training in mineral processing, attempting to read the seminal works in the subject then
available in the excellent technical library of Mount Isa Mines Limited. As a notorious bibliophile, I have encompassed most
of the important publications though biased towards the ‘practical’ side.
What is missing in the English language literature of flotation is a single up-to-date volume covering all aspects of the
‘practical art’ as required by a mineral processing engineer either operating or supporting a flotation concentrator. Books
written to date have usually focused on the ‘science’ or chemistry of the process to the detriment of ‘practical’ aspects with
much useful information on the latter being either in monographs or scattered through many technical papers written
for journals and symposia. Even in the age of the internet, going through this large volume of material is a daunting
prospect. In my opinion the only comparable volume to this book is the venerable Froth Flotation 50th Anniversary Volume
published in the USA nearly 50 years ago.2 Even so it was not as oriented to practical plant problem solving as Flotation
Plant Optimisation.
This book is essentially a ‘how to’ manual and will be particularly useful as a conversion course for mineral processors such
as those whose undergraduate training has been primarily in chemical engineering. Its approach is based on the relevant
design and execution of a plant survey followed by the steps required to analyse the results to a suitable level of detail.
Chris Greet has assembled a great cast of contributors for this book, all authorities in their fields representing the leading
edge of accumulated current knowledge on relevant topics such as applied mineralogy, geometallurgy, flotation chemistry,
process analysis, mass balancing, experimental design and characterisation of industrial flotation machines. I know most
of the individual authors and can attest to their mastery of their particular areas of expertise. It has been a privilege to
have worked with them over the years and participated in the occasional ‘technical highs’ on the science and application of
flotation that have made working in the field so rewarding.
I commend the book to readers and can assure them that understanding the contents and applying the knowledge
gained to real mineral processing situations will result in improved metallurgical outcomes.
Peter Munro
Senior Principal Consulting Engineer, Mineralurgy Pty Ltd
2. Fuerstenau, D W (ed), 1962. Froth Flotation 50th Anniversary Volume (American Institute of Mining, Metallurgical and Petroleum Engineers, Inc:
New York).
ACKNOWLEDGEMENTS
Well, the cattle prod alluded to in the acknowledements to the first edition of this book appears to have worked! Kristy
Burt first broached the subject of a second edition of Flotation Plant Optimisation way back in late 2012, and I thought this
would be a walk in the park! Well. How wrong was I? Once again the weak link in the publishing chain was me. However,
after six years of following up, pushing, prodding and nagging we have finally pulled the second edition together due in
large part to Kristy’s determination and tenacity. I owe a great debt to Kristy and the AusIMM publishing team.
Again, the contributing authors (some old and some new) have my sincere thanks for offering your words of wisdom
and contributing to this work. Thank you for your patience, advice, contributions and your willingness to pass on your
knowledge. I know that your contributions do assist the reader in their endeavors to improve their operation.
It is incumbent on me to acknowledge not only Alan Buckley’s contribution to this work, but also his life’s work exploring
the surface chemistry of sulfide minerals and how this effects flotation. Alan’s passing in September 2017 represents a
significant loss to our industry. However, I would encourage young mineral scientists and engineers to use Alan’s legacy in
their work.
Magotteaux must be thanked for once again loosening the purse strings and sponsoring this work. In particular, Patrick
Viseur (my long suffering manager) must be acknowledged for approving the funding and seeing the value this volume has
and will continue to bring to the mining industry.
To the readers/users of this work I implore you to apply rigour to your work, gather the right data and analyse it correctly.
Proper data collection and analysis will bring rewards.
Finally, to my kids Erica, Jessica and Matthew: yes I may be dull, colorless and boring but I still love you and appreciate
your support.
Dr Christopher J Greet
Magotteaux is the world leader in process optimisation solutions for high abrasion extracting industries.
Starting with a cross-section insight of our customers’ value chain, combining expert advice, services and resources,
products, equipment and systems, Magotteaux uses all relevant tools to help customers optimise their operations, recover
more valuable minerals, and deliver their final products or services at the best total cost of ownership with minimal
environmental impact.
Magotteaux employs more than 3200 talented individuals and has a market footprint of above 700 000 tons of grinding
media and castings through its specialised production units, sales offices and technical centres worldwide.
Our team of 200+ technical and sales experts located close to your operations helps you choose the perfect solution for
your applications.
Magotteaux is part of Sigdo Koppers, listed in Santiago, Chile.
IT TAKES MORE
THAN JUST RELIABLE
GRINDING MEDIA
TO OPTIMIZE PERFORMANCE!
PROCESS OPTIMIZATION PRODUCTS AND SERVICES FOR HIGH ABRASION EXTRACTING INDUSTRIES.
[email protected] · WWW.MAGOTTEAUX.COM
Geometallurgy
(Chapter 13: Dean David)
Mass balancing
(Chapter 3: Rob Morrison)
Data acquisition
(Chapter 2: Bill Johnson) Mineralogy
(Chapter 4: Alan Butcher)
Troublesome minerals
Questions
(NEW Chapter 5: Bill Johnson and Peter Munro)
(What and where?)
Cell characterisation
(Chapter 6: Greg Harbort, Sarah Schwarz and Thu Nguyen)
Problem definition
Machine
Liberation
Chemistry Pulp chemistry
(Chapter 7: Stephen Grano)
Laboratory testing
(Chapter 10: Kym Runge)
Electrochemistry
(Chapter 8: Ron Woods)
NEW Appendix by
Trials and statistical analysis Bill Johnson and Peter Munro
(Chapter 11: Tim Napier-Munn)
Surface analysis
(Chapter 9: Alan Buckley)
Communication NEW Appendix by
(Chapter 12: Joe Pease) Chris Greet
CHAPTER 2 55
Existing methods for process analysis
Bill Johnson
Appendix – Two aspects of sampling 99
CHAPTER 3 105
Mass balancing flotation data
Rob Morrison
CHAPTER 4 137
A practical guide to some aspects of mineralogy that affect flotation
Alan R Butcher
CHAPTER 5 157
Some troublesome sulfide and non-sulfide minerals
Bill Johnson and Peter Munro
CHAPTER 6 171
Characterisation measurements in industrial flotation cells
Greg Harbort, Sarah Schwarz and Thu Nguyen
CHAPTER 7 199
Chemical measurements during plant surveys and their interpretation
Stephen Grano
CHAPTER 8 227
Electrochemical aspects of sulfide mineral flotation
Ronald Woods OAM
Appendix – General description of the oxidation of sulfides and the consequences for the flotation process 256
Bill Johnson and Peter Munro
CHAPTER 9 263
Surface chemical characterisation for flotation plant optimisation
Alan Buckley
Appendix – Practical examples of using surface chemistry to identify problems in flotation plants 294
Christopher Greet
CHAPTER 10 321
Laboratory flotation testing – an essential tool for ore characterisation
Kym Runge
Appendix – Worked examples 343
CHAPTER 11 361
Designing and analysing plant trials
Tim Napier-Munn
CHAPTER 12 391
Project evaluation and communication
Joe Pease
CHAPTER 13 407
Operational geometallurgy
Dean David
Return to contents
CHAPTER 1
The Eureka Mine – an example of how to
identify and solve problems in a flotation plant
Christopher J Greet
ABSTRACT
This chapter will provide a methodology for identifying and solving problems in a flotation
plant. To do this, a ‘mythical’ concentrator (The Eureka Mine) will be described and used to
demonstrate how to go about the process of identifying where the losses of valuable mineral
occur and what gangue species are diluting the concentrate. It is intended that this chapter
be used in parallel with subsequent chapters to guide the reader through the steps involved
in optimising the plant.
INTRODUCTION
Many of us have, at one time or another, flicked through the newspaper or searched online
for a job, and stumbled upon an ad not unlike the one that appears below. Whether we
are jaded with our current role, looking for a step up to the next level, or wanting a new
challenge, we prepare our resume and send it off in the hope that we may be the successful
candidate.
With the interview process out of the way, the waiting and self-doubt start. After what
seems an eternity you receive a phone call or letter telling you that you have got the job.
Congratulations! Now what? You’re new, you have ambition, you have drive and you want
to make your mark! But, there’s a right way and a wrong way to do this. The first thing to
remember is that this place has a history, and the people you are going to work with have been
here much longer than you have, so communication and respect are key to your success. You
need to discover the history of the concentrator and discuss its operation with other members
of staff (operators, metallurgical technicians, shift foremen, plant metallurgists, the chemist,
mechanical and electrical maintenance, the mine (ie geologists and mining engineers) and
the supply department). They will all give you their perspective and you need to respect
their point of view. You will also work out who among these people hold the knowledge, the
history; the real story of your concentrator.
Employment conditions
Eureka Mining Limited is a wholly owned subsidiary of Stockade Resources Limited, an Australian-based mining and exploration
company with interests in Australia, New Guinea, Zambia and Peru.
Reporting to the Metallurgy Manager, the successful candidate will work closely with production to maintain and improve plant
performance. This is a residential position, with an attractive renumeration package commensurate with your qualifications and
experience.
To apply
Please submit your application including a cover letter and current copy of your resume via email to ...
While you are establishing relationships you need to determine what data exists (and what
is missing) to help you develop a technical perspective of how your concentrator performs.
The focusing questions in any process improvement strategy are:
•• Where and how do the losses of valuable mineral occur?
•• What gangue minerals are diluting the concentrate and how did they get there?
The intention of this book is to provide you with a sequence of logical steps to collect the
necessary data to define the problem(s) in your operation. Once the problems are defined you
can prioritise them and develop experimental strategies that may lead to solutions that can be
implemented in the plant.
To illustrate this journey we will consider the performance of the Eureka Mine. The Eureka
Mine treats a complex polymetallic sulfide ore supplied from an underground mine and
produces three saleable concentrates. This chapter provides a description of the Eureka
Concentrator and its metallurgical performance since commissioning. It demonstrates
the methodology to collect plant data, analyse and interpret it to determine where the
metallurgical problems – and potential solutions – lie.
Geology
The Eureka volcanogenic massive sulfide deposit occurs within the Laylor-Eureka Volcanic
sequence of the Mount Rush Volcanics. The deposit was formed when hot mineralised
solutions spewed on to the ocean floor and were rapidly quenched by the surrounding
seawater. Hence, the sulfide minerals precipitated from solution formed very fine crystals and
intricate mineral textures. Subsequent geological changes to the orebody were few; therefore,
many of the original fine-grain textures remained intact.
Deposit mineralogy
To fully appreciate the complexity of flotation at Eureka requires a rudimentary understanding
of the mineralogy of the orebody. Eureka is unusually sulfide-rich and contains a relatively
simple mineral suite: 58 per cent pyrite, 20 per cent sphalerite, four per cent galena,
two per cent arsenopyrite and one per cent chalcopyrite, with minor amounts of tetrahedrite.
The remaining 15 per cent of the ore consists of: quartz, barite, calcite, chlorite, sericite and
siderite.
Macroscopically, the mineral textures are diverse; however, the orebody can be divided
into two distinct metal zones. The demarcation between the two zones is set, arbitrarily, at
100 g/t of silver and represents a continuous horizon across the deposit. Above this level is
the hanging wall enrichment zone characterised by higher lead, zinc, silver, gold and arsenic
grades. Macroscopically, the sulfides within the enrichment zone tend to be banded and very
fine grained. The footwall-depleted zone occurs below the 100 g/t silver horizon. Pyrite and
chalcopyrite are the dominant minerals in this part of the orebody, with reduced lead, zinc,
silver, gold and arsenic grades. The footwall-depleted zone is highly recrystallised, therefore
the grain structure is comparatively coarse compared with those observed in the hanging
wall enrichment zone.
The Eureka orebody is relatively free of non-sulfide gangue mineralisation. Pyrite is the
dominant gangue mineral and is associated with all other minerals within the deposit.
Therefore, the properties of pyrite will influence greatly the behaviour of all other minerals
during processing.
The pyrite textures vary markedly across the deposit from coarse recrystallised grains in the
footwall-depleted zone to compact microcrystalline masses, spongy and colliform clots, such
as melnokovite (an amorphous pyrite of colloidal origin). Ultra-fine intergrowths of pyrite
with other sulfides are common, particularly with galena and arsenopyrite (Figure 1). The
association of auriferous arsenopyrite with pyrite is also of significance.
A B
C D
FIG 1 – Photomicrographs of various galena ore textures: (A) galena replacement in pyrite matrix, like
melnokovite (magnification ×10); (B) galena blebs in pyrite matrix (magnification ×20); (C) galena in
crystal voids around pyrite (magnification ×5); and (D) galena replacement in melnokovite colloform
(magnification ×40). (Note: the blue/grey areas are galena and the golden areas are pyrite.)
Process description
Laboratory testing of the Eureka ore clearly demonstrated that it was possible to produce
saleable copper, lead and zinc concentrates. The flow sheet developed in the laboratory was
tested at pilot scale to prove the selected process route was robust and to produce sufficient
quantities of concentrate for smelter testing. The final Eureka process flow sheet is presented
in Figure 2.
Mine geologists classify the ore into broad types based on texture: enrichment zone ore
(banded), or footwall depleted zone ore (coarse grained) and estimated pyrite content. Each
ore type is crushed in batches to nominally 100 per cent passing 100 mm, in the underground
jaw crusher before being trucked to the surface in 50 t dump trucks. Upon delivery to the
Grinding st
Zn 1 Cl
P80 = 75 microns
nd Zinc Regrinding
Pb 2 Cl Lead Regrinding
P80 = 38 microns
P80 = 38 microns
Pb 3rd Cl
run-of-mine (ROM) pad each ore type is stockpiled separately. The ore is fed onto a conveyor
belt that leads to an open stockpile in specific ratios of each ore type. Apron feeders under the
open stockpile feed the blended ore onto the primary mill feed conveyor at nominally 120 t/h.
The primary mill is a low aspect ratio semi-autogenous grinding (SAG) mill in open circuit.
The SAG mill product discharges into a common sump shared with the ball mill. The pulp
is pumped from the mill discharge sump to cyclones in closed circuit with a secondary ball
mill. The cyclone underflow feeds the secondary ball mill, and cyclone overflow reports to
flotation feed. The mill and cyclone configuration is designed to produce flotation feed P80
of 75 µm.
Secondary cyclone overflow feeds a sequential copper/lead/zinc flotation circuit. Each
flotation section consists of a rougher/scavenger, with the rougher concentrate reporting
to the cleaner circuit. Copper cleaning is achieved without regrinding and with only one
stage of cleaning. Lead rougher concentrate feeds the first of three stages of cleaning. Lead
scavenger concentrate and lead first cleaner tailings are reground and recycled to the head of
the lead rougher. Lead scavenger tailings report to zinc circuit feed. Zinc rougher concentrate
reports to two stages of cleaning. Zinc scavenger concentrate and zinc first cleaner tailings are
reground and recycled to zinc rougher feed. The concentrates produced from the copper, lead
and zinc flotation circuits are pumped to thickeners. Thickened concentrates are filtered for
stockpiling before loading into rail cars for shipment to the smelter.
Flotation tailings are dewatered and used as paste backfill in the underground workings.
Metallurgical performance
The Eureka Concentrator was commissioned in December 1995 and reached name plate
throughput by July 1996. A further 18 months were required to achieve the design concentrate
grades and recoveries. Typical metallurgical performance since 1998 is summarised in Table 1.
DATA ACQUISITION
Measurement is the first step to understanding how your plant is performing. In a modern
concentrator, the process can be monitored using a multitude of sensors; however, the data
collected from inventory samples on a shift, daily, weekly and monthly basis, coupled with
well-executed metallurgical surveys, is invaluable to define where valuable mineral losses
occur and what gangue minerals are diluting the concentrate.
As the new metallurgist, the first step is to acquaint yourself with the existing plant data.
That is, review the shift mass balance data, interrogate the monthly composite data and
examine the results of any previous plant surveys. This should provide some indication
TABLE 1
Typical metallurgical performance of the Eureka Concentrator since 1998.
Stream Wt % Grade (%) Recovery (%)
Ag (ppm) Cu Pb Zn Ag Cu Pb Zn
Flotation feed 100.0 130 0.4 3.1 12.4 100.0 100.0 100.0 100.0
Cu concentrate 0.9 3190 25.4 5.2 6.2 21.9 53.2 1.5 0.4
Pb concentrate 3.6 1356 1.2 61.5 10.8 37.6 10.1 70.4 3.1
Zn concentrate 20.0 102 0.3 2.0 53.4 15.9 16.4 13.0 86.9
Final tailing 75.3 42 0.1 0.6 1.6 24.6 20.3 15.1 9.6
of where the metallurgical weaknesses lie. Unfortunately, often the only readily available
up-to-date information are shift mass balances. Therefore, it is good practice to organise a
plant survey as soon as possible to quantify the metallurgical performance of each section
of the plant.
The shift data for the copper, lead and zinc circuits of the Eureka Concentrator for the first
half of 1998 appear in Figure 3.
These data suggest that, with the exception of lead and zinc concentrate grades, the plant
performance is somewhat unstable. The copper concentrate grade and recovery fluctuate
widely, and lead recoveries are highly variable with greater than ten per cent zinc grade in
the lead concentrate, and zinc recoveries are usually in the low to middle 80s.
Your ‘gut feeling’ probably tells you that copper, lead and zinc recoveries are lower than
expected, so it should be possible to improve performance. But you don’t know what the
limiting factors are. You need a comprehensive plant survey to identify the targets.
Now you need to decide what level of detail the survey needs. This will obviously depend
on what work has been completed previously. In this example it will be assumed that the
available data is scattered and incomplete. Therefore, the survey objective will be to collect
as much data as possible to provide sufficient information to describe the pulp chemistry
and metallurgical performance. Ideally, both sets of data can be collected together so they
complement each other. To add further value to your survey, collecting gas hold-up, superficial
gas velocity and bubble size data will provide information about the hydrodynamics of the
flotation cells.
90.0 30.0
A
75.0 27.0
Cu recovery, %
Cu grade, %
60.0 24.0
45.0 21.0
30.0 18.0
15.0 15.0
0.0 12.0
98
98
98
98
98
98
98
98
98
98
98
98
98
/
/
01
01
01
02
02
03
03
04
04
05
05
06
06
1/
9/
/
7/
4/
/
15
29
12
26
12
26
23
21
18
Time, days
Cu recovery Cu grade
90.0 90.0
B
75.0 75.0
Pb and Zn grade, %
Pb recovery, %
60.0 60.0
45.0 45.0
30.0 30.0
15.0 15.0
0.0 0.0
8
8
/9
/9
/9
/9
/9
/9
/9
/9
/9
/9
/9
/9
/9
1
01
05
04
06
6
/0
/0
/0
/0
/0
/0
/0
/0
/0
1/
9/
7/
4/
15
29
12
26
12
26
23
21
18
Time, days
Pb recovery Pb grade Zn grade in Pb con
100.0 60.0
C
90.0 50.0
Zn and Fe grade, %
Zn recovery, %
80.0 40.0
70.0 30.0
60.0 20.0
50.0 10.0
40.0 0.0
8
8
/9
/9
/9
/9
/9
/9
/9
/9
/9
/9
/9
/9
/9
1
01
05
04
06
6
/0
/0
/0
/0
/0
/0
/0
/0
/0
1/
9/
7/
4/
15
29
12
26
12
26
23
21
18
Time, days
Zn recovery Zn grade Fe grade in Zn con
FIG 3 – Time series data for: (A) copper; (B) lead and (C) zinc circuits.
TABLE 2
The sampling points and sample type for the metallurgical survey of the Eureka flotation circuit.
Sample Process stream Sample type Sample Process stream Sample type
number number
1 Flotation feed (ball mill Half moon cutter 18 Lead second cleaner tailing A Dip sample
cyclone overflow)
2 Copper rougher concentrate Lip sample 19 Lead second cleaner tailing B Dip sample
3 Copper rougher tailing A Dip sample 20 Lead third cleaner concentrate Lip sample or OSA
4 Copper rougher tailing B Dip sample 21 Lead third cleaner tailing A Dip sample
5 Copper cleaner concentrate Lip sample or OSA 22 Lead third cleaner tailing B Dip sample
6 Copper cleaner tailing A Dip sample 23 Zinc rougher concentrate Lip sample
7 Copper cleaner tailing B Dip sample 24 Zinc rougher tailing A Dip sample
8 Lead rougher concentrate Lip sample 25 Zinc rougher tailing B Dip sample
9 Lead rougher tailing A Dip sample 26 Zinc scavenger concentrate Lip sample
10 Lead rougher tailing B Dip sample 27 Zinc scavenger tailing A Dip sample or OSA
11 Lead scavenger concentrate Lip sample 28 Zinc scavenger tailing B Dip sample or OSA
12 Lead scavenger tailing A Dip sample 29 Zinc first cleaner concentrate Lip sample
13 Lead scavenger tailing B Dip sample 30 Zinc first cleaner tailing A Dip sample
14 Lead first cleaner concentrate Lip sample 31 Zinc first cleaner tailing B Dip sample
15 Lead first cleaner tailing A Dip sample 32 Zinc second cleaner concentrate Lip sample or OSA
16 Lead first cleaner tailing B Dip sample 33 Zinc second cleaner tailing A Dip sample
17 Lead second cleaner concentrate Lip sample 34 Zinc second cleaner tailing B Dip sample
Other samples that may be redundant in the mass balance still provide valuable data to
confirm the mass balance ‘makes sense’ – for example, copper rougher feed, lead rougher
feed and zinc rougher feed.
With the sample list decided, the next step is to brief the survey team, prepare the equipment
and ensure that the plant is running in a representative and steady state condition for the
survey. Set the ground rules for the survey (ie number of rounds over what time interval) and
delegate tasks to each survey member so that:
•• sufficient sample buckets with lids are cleaned, weighed and labelled
•• the sampling equipment is checked, clean and ready for use
•• everyone knows what is expected of them.
You need to inform operations that you intend to conduct a plant survey(s) and what the
objectives of the work are. Check that the plant will be available and at steady-state at the
planned survey time. You must inform the chemical laboratory of the expected sample load
and assay requirements.
The day before the survey you should conduct a full plant ‘dress rehearsal’ to ensure everyone
knows their role and sample points. All sample points should be checked and cleaned,
including flotation cell lips. Sample buckets should be pre-labelled. Most importantly, check
that the work area, access to sample points and sampling methods are safe.
On the day of the survey, attend the production meeting to establish the plant condition and
whether it is expected to operate at steady state. Inform operations that the survey will proceed.
Set out the buckets and samplers in the correct positions, check and clean the sample points and
flotation cell lips again, have a safety and procedures pre-start with the survey team.
When you are ready for ‘kick off’, do a final check that the plant is at steady state – feed
tonnage, head grades and performance. If so, you are ready to start the survey.
A survey of this type should be conducted over three residence times, with multiple rounds
of samples collected to form a composite. In this example, the survey was conducted over
three hours, during which four ‘cuts’ from each sampling point were collected randomly.
Once completed, gather the samples and take them to the laboratory.
TABLE 3
Raw data from the metallurgical survey of the Eureka Concentrator.
Sample Process stream Sample weight (g) % Solids
number
Gross wet Bucket Cutter Net wet Dry
1 Flotation feed 2083.5 202.6 1880.9 909.8 48.4
2 Copper rougher concentrate 1961.2 229.2 347.8 1384.2 708.2 51.2
3 Copper rougher tailing A 2509.5 466.3 2043.2 949.7 46.5
4 Copper rougher tailing B 2480.9 466.0 2014.9 936.1 46.5
5 Copper cleaner concentrate 1386.7 225.0 324.8 836.9 536.3 64.1
6 Copper cleaner tailing A 1622.0 202.6 1419.4 144.8 10.2
7 Copper cleaner tailing B 1668.4 202.6 1465.8 141.8 9.7
8 Lead rougher concentrate 4842.2 225.3 341.0 4275.9 2695.1 63.0
9 Lead rougher tailing A 2395.6 464.3 1931.3 812.3 42.1
10 Lead rougher tailing B 2374.1 477.7 1896.4 772.0 40.7
11 Lead scavenger concentrate 1317.1 206.9 340.9 769.3 482.4 62.7
12 Lead scavenger tailing A 2416.1 465.0 1951.1 823.7 42.2
13 Lead scavenger tailing B 2357.1 446.3 1910.8 782.0 40.9
14 Lead first cleaner concentrate 10638.1 190.4 341.1 10106.6 6660.8 65.9
15 Lead first cleaner tailing A 2646.2 234.5 2411.7 1347.1 55.9
16 Lead first cleaner tailing B 2640.5 224.8 2415.7 1343.7 55.6
17 Lead second cleaner concentrate 5832.8 214.4 325.3 5293.1 3547.9 67.0
18 Lead second cleaner tailing A 2814.2 233.6 2580.6 1608.0 62.3
19 Lead second cleaner tailing B 2828.9 228.2 2600.7 1424.7 54.8
20 Lead third cleaner concentrate 2721.8 200.6 341.0 2180.2 1395.7 64.0
21 Lead third cleaner tailing A 2905.2 238.1 2667.1 1635.1 61.3
22 Lead third cleaner tailing B 2881.6 225.8 2655.8 1650.7 62.2
23 Zinc rougher concentrate 2779.3 226.4 341.2 2211.7 1287.8 58.2
24 Zinc rougher tailing A 2090.7 224.9 1865.8 681.5 36.5
25 Zinc rougher tailing B 1992.0 234.7 1757.3 638.8 36.4
26 Zinc scavenger concentrate 1565.6 200.4 324.9 1040.3 500.5 48.1
27 Zinc scavenger tailing A 1983.0 234.9 1748.1 629.1 36.0
28 Zinc scavenger tailing B 1753.0 225.5 1527.5 536.5 35.1
29 Zinc first cleaner concentrate 3945.3 202.4 341.2 3401.7 1947.7 57.3
30 Zinc first cleaner tailing A 2316.8 225.6 2091.2 1004.9 48.1
31 Zinc first cleaner tailing B 2321.4 225.4 2096.0 1005.5 48.0
32 Zinc second cleaner concentrate 4209.1 211.9 341.2 3656.0 2020.2 55.3
33 Zinc second cleaner tailing A 2466.7 225.5 2241.2 1250.5 55.8
34 Zinc second cleaner tailing B 2474.3 234.0 2240.3 1248.2 55.7
TABLE 4
Raw assay data from the metallurgical survey of the Eureka Concentrator.
Sample Process stream % Solids Raw assays
number Ag (ppm) Cu (%) Pb (%) Zn (%) Fe (%)
1 Flotation feed 48.4 135 0.43 4.02 15.80 14.30
2 Copper rougher concentrate 51.2 588 16.50 15.20 12.30 14.20
3 Copper rougher tailing A 46.5 92 0.19 4.04 15.40 13.40
4 Copper rougher tailing B 46.5 98 0.21 4.04 15.60 13.60
5 Copper cleaner concentrate 64.1 229 22.40 4.07 7.80 16.10
6 Copper cleaner tailing A 10.2 524 15.45 16.30 11.30 13.80
7 Copper cleaner tailing B 9.7 424 5.50 26.40 18.50 9.00
8 Lead rougher concentrate 63.0 184 0.32 21.42 29.97 8.97
9 Lead rougher tailing A 42.1 43 0.16 1.22 17.90 13.60
10 Lead rougher tailing B 40.7 34 0.16 1.25 16.60 13.50
11 Lead scavenger concentrate 62.7 179 0.36 8.16 39.84 11.57
12 Lead scavenger tailing A 42.2 38 0.14 0.67 15.20 13.90
13 Lead scavenger tailing B 40.9 18 0.15 0.50 14.60 13.70
14 Lead first cleaner concentrate 65.9 600 0.30 44.68 14.10 4.10
15 Lead first cleaner tailing A 55.9 184 0.28 18.50 30.65 8.80
16 Lead first cleaner tailing B 55.6 176 0.28 20.50 30.75 9.00
17 Lead second cleaner concentrate 67.0 960 0.51 50.90 10.10 2.50
18 Lead second cleaner tailing A 62.3 814 0.23 47.30 13.40 3.85
19 Lead second cleaner tailing B 54.8 800 0.27 47.40 13.60 3.95
20 Lead third cleaner concentrate 64.0 1060 1.08 60.00 9.40 1.70
21 Lead third cleaner tailing A 61.3 890 0.32 47.00 11.26 2.40
22 Lead third cleaner tailing B 62.2 900 0.40 49.00 11.34 2.80
23 Zinc rougher concentrate 58.2 53 0.33 1.63 45.10 8.40
24 Zinc rougher tailing A 36.5 16 0.13 0.45 1.72 15.35
25 Zinc rougher tailing B 36.4 15 0.13 0.55 1.74 15.65
26 Zinc scavenger concentrate 48.1 72 0.47 2.23 15.20 14.30
27 Zinc scavenger tailing A 36.0 12 0.09 0.32 0.85 15.32
28 Zinc scavenger tailing B 35.1 13 0.11 0.36 0.87 15.48
29 Zinc first cleaner concentrate 57.3 43 0.33 1.30 55.70 6.00
30 Zinc first cleaner tailing A 48.1 63 0.38 2.02 19.40 12.10
31 Zinc first cleaner tailing B 48.0 60 0.42 1.94 19.60 12.30
32 Zinc second cleaner concentrate 55.3 39 0.29 1.17 58.40 5.30
33 Zinc second cleaner tailing A 55.8 62 0.48 1.60 42.70 8.50
34 Zinc second cleaner tailing B 55.7 64 0.44 1.64 44.70 8.50
Step 1
Determine the plant throughput when the survey was completed. If the feed tonnage is not
available, assign the fresh flotation feed a value of 100 per cent.
Step 2
Balance the ‘outer’ circuit. That is, complete a mass balance of the feed, final concentrate
and tailing for the copper, lead and zinc circuits (Figure 4). This will provide estimates of
the tonnes of copper, lead and zinc concentrates. These values can be used in subsequent
balances to estimate tonnages of the internal process streams within rougher, scavenger and
cleaner circuits.
Assign each process stream a number and identify each of the nodes. The process streams
are:
1. flotation feed
2. copper cleaner concentrate
3. copper rougher tailing
4. lead third cleaner concentrate
5. lead scavenger tailing
6. zinc second cleaner concentrate
7. zinc scavenger tailing.
The nodes are:
1. flotation feed = Cu cleaner concentrate + Cu rougher tailing (or, 1 = 2 + 3)
2. Cu rougher tailing = Pb third cleaner concentrate + Pb scavenger tailing (or, 3 = 4 + 5)
3. Pb scavenger tailing = Zn second cleaner concentrate + Zn scavenger tailing (or, 5 = 6 + 7).
In this case, at the time of surveying the plant, throughput was 120 t/h.
The tailing assays are reasonably similar and in the expected range, indicating good
sampling practice. Therefore, taking an average of the two tailing sample assays is acceptable.
If one assay was considerably higher than the other (for example, the copper cleaner tailing A, in
Table 4), it suggests that this dip sample has been contaminated with froth during collection,
and this assay may have to be rejected from the data set. However, caution is needed when
rejecting assays – this sample has a similar assay to the copper rougher concentrate, so it is
possible that samples were wrongly labelled.
Having sorted the data and identified the process streams, nodes and feed tonnage, the
numbers can be plugged into your mass balancing program. In this case MATBAL, using
a Monte Carlo simulation, was employed. The resultant mass balance is shown in Table 5.
This includes a sigma value. If sigma is less than five per cent then the data are considered
to be good.
Step 3
The next step is to balance the internal or ‘inner’ circuits within the copper, lead and zinc
circuits. These internal balances are broken into ‘bite size’ pieces to simplify the balancing
process and generate tonnage estimates for subsequent ‘inner’ circuit balances.
The copper circuit (Figure 5) is comparatively easy to balance using the tonnage estimates
for the flotation feed and copper cleaner concentrate generated in the ‘outer’ circuit mass
balance. The process streams are:
1. flotation feed (120 t/h)
2. copper rougher concentrate
3. copper rougher tailings
4. copper cleaner concentrate (1.21 t/h)
5. copper cleaner tailings.
TABLE 5
Adjusted assays for the ‘outer’ circuit mass balance.
No Stream t/h ∑ Adjusted assays (%)
Ag (ppm) Cu Pb Zn Fe
1 Flotation feed 120.00 0.09 135 0.43 3.99 15.04 13.17
2 Cu cleaner concentrate 1.21 3.31 229 22.40 4.07 7.80 16.11
3 Cu rougher tailing 118.79 0.10 95 0.21 3.99 15.12 13.14
4 Pb 3rd cleaner concentrate 6.76 2.19 1060 1.07 61.01 9.42 1.70
5 Pb scavenger tailing 112.03 0.16 28 0.15 0.55 15.46 13.83
6 Zn 2nd cleaner concentrate 27.70 2.14 39 0.29 1.18 59.92 5.35
7 Zn scavenger tailing 84.33 0.73 12 0.11 0.34 0.86 16.62
2. Cu rougher concentrate
5. Cu cleaner tailing
Cu cleaner (2)
4. Cu cleaner concentrate
(Final Cu Con)
2. Pb rougher concentrate
4. Pb scavenger concentrate
Balance 2
Pb 1st cleaner feed Pb 1st cleaner 7. Pb 1st cleaner
(3) tailing
6. Pb 1st cleaner concentrate
Balance 1
Pb 3rd cleaner feed Pb 3rd cleaner 11. Pb 3rd cleaner tailing
(5)
While this arrangement may appear to be counter intuitive, the reasons for moving
backwards through the lead circuit are driven by the need to estimate tonnage figures for the
recycle streams.
In Balance 1 (the lead second and third cleaners), by using the third cleaner concentrate
estimated tonnage generated in the ‘outer’ mass balance in Step 2 it is possible to estimate
the tonnage for the lead first cleaner concentrate and the lead second cleaner tailing. Using
these two values, it is then possible to mass balance the lead first cleaner. This gives tonnage
estimates for lead rougher concentrate and lead first cleaner tailings.
Balance 3 uses the tonnage estimate for the lead scavenger tailings from the ‘outer’ balance
of Step 2. The mass balance of the lead scavengers estimates the tonnage of lead scavenger
concentrate.
Once these three ‘inner’ mass balances are complete it is possible to fix certain tonnages
(ie copper rougher tailings, lead scavenger tailings, lead first, second and third cleaner
concentrates) and complete a mass balance of the lead circuit. The process streams used for
this balance are:
1. copper rougher tailings
2. lead rougher concentrate
3. lead rougher tailings
4. lead scavenger concentrate
5. lead scavenger tailings
6. lead first cleaner concentrate
7. lead first cleaner tailings
8. lead second cleaner concentrate
9. lead second cleaner tailings
10. lead third cleaner concentrate
11. lead third cleaner tailings.
The nodes are:
1. Cu rougher tailings + Pb scavenger concentrate + Pb first cleaner tailings = Pb rougher
concentrate + Pb rougher tailings (or, 1 + 4 + 7 = 3 + 3)
2. Pb rougher tailings = Pb scavenger concentrate + Pb scavenger tailings (or, 3 = 4 + 5)
3. Pb rougher concentrate + Pb second cleaner tailings = Pb first cleaner concentrate + Pb first
cleaner tailings (or, 2 + 9 = 6+ 7)
4. Pb first cleaner concentrate + Pb third cleaner tailings = Pb second cleaner concentrate + Pb
second cleaner tailings (or, 6 + 11 = 8 + 9)
5. Pb second cleaner concentrate = Pb third cleaner concentrate + Pb third cleaner tailing (or,
8 = 10 + 11).
The mass balanced assays for the lead circuit are shown in Table 7.
The same approach is used for mass balancing the inner circuits for zinc flotation (Figure 7).
The first balance (Balance 1) is for the zinc first and second cleaner, while Balance 2 is for the
zinc scavenger circuit. When these two internal mass balances are complete, it is possible to
mass balance the zinc circuit while fixing certain flows. The process streams used are:
1. lead scavenger tailings
2. zinc rougher concentrate
3. zinc rougher tailings
TABLE 7
Adjusted assays for the ‘inner’ lead circuit mass balance.
No Stream t/h ∑ Adjusted assays (%)
Ag Cu Pb Zn Fe
1 Cu rougher tailing 118.79 0.05 87 0.20 4.00 14.86 13.17
2 Pb rougher con 134.72 1.66 208 0.32 21.53 29.91 8.81
3 Pb rougher tailing 122.16 0.09 40 0.16 1.23 17.25 13.68
4 Pb scavenger con 10.14 1.34 179 0.36 8.20 39.79 11.55
5 Pb scavenger tailing 112.02 0.09 27 0.15 0.59 15.20 13.86
6 Pb 1st cleaner con 110.85 0.10 631 0.30 44.67 13.63 3.91
7 Pb 1st cleaner tailing 127.95 1.17 162 0.28 19.48 30.99 9.18
8 Pb 2nd cleaner con 31.98 0.09 940 0.51 50.72 10.54 2.45
9 Pb 2nd cleaner tailing 104.08 0.11 602 0.25 43.65 13.91 4.06
10 Pb 3rd cleaner con 6.75 0.09 1077 1.08 60.39 9.35 1.71
2. Zn rougher concentrate
4. Zn scavenger concentrate
Balance 1
Step 4
With the inner mass balances complete, it is now possible to balance the whole circuit
(Table 9). When reviewing your mass balance, you should check that the internal workings of
the circuit are balanced. For example, in the copper circuit the feed to the copper rougher is
fresh flotation feed plus copper cleaner tailings. The sum of these recovery values should be
the same as the copper rougher concentrate plus the tailing. That is, for copper:
Rflotation feed + RCu cleaner tailings = RCu rougher con + RCu rougher tailings
TABLE 8
Adjusted assays for the ‘inner’ zinc circuit mass balance.
No Stream t/h ∑ Adjusted assays (%)
Ag Cu Pb Zn Fe
1 Pb scavenger tailing 112.02 0.07 28 0.15 0.57 15.15 13.20
2 Zn rougher con 43.74 0.07 53 0.33 1.52 44.32 7.98
3 Zn rougher tailing 89.83 0.18 16 0.13 0.47 1.74 15.69
4 Zn scavenger con 5.51 2.56 72 0.47 2.26 15.17 14.29
5 Zn scavenger tailing 84.32 0.09 12 0.11 0.36 0.86 15.78
6 Zn 1st cleaner con 37.44 0.09 43 0.33 1.32 54.80 6.15
7 Zn 1st cleaner tailing 16.04 0.10 63 0.40 2.04 19.55 12.53
8 Zn 2nd cleaner con 27.70 0.09 39 0.29 1.22 58.66 5.34
9 Zn 2nd cleaner tailing 9.74 0.47 64 0.46 1.61 43.84 8.42
25.0
Cu cleaner concentrate
20.0
Cu grade, %
15.0
Cu rougher concentrate
10.0
Cu rougher feed
5.0
Flotation feed
0.0
0.0 20.0 40.0 60.0 80.0 100.0 120.0
Cu recovery, %
FIG 8 – The copper grade/recovery curve for the copper circuit within the Eureka Concentrator.
TABLE 9
Mass balance of copper, lead and zinc circuits of the Eureka Concentrator.
No Stream t/h ∑ Cu (%) Pb (%) Zn (%) Fe (%)
Grade Recovery Grade Recovery Grade Recovery Grade Recovery
1 Flotation feed 120.00 0.05 0.43 100.00 4.00 100.00 14.92 100 13.06 100.00
2 Cu rougher con 2.39 0.41 14.29 66.15 15.10 7.50 12.76 1.70 13.10 1.99
3 Cu rougher tailing 118.79 0.05 0.20 46.39 4.00 98.97 14.99 99.48 13.02 98.69
4 Cu cleaner con 1.21 0.11 22.74 53.61 4.07 1.03 7.71 0.52 16.84 1.31
5 Cu cleaner tailing 1.18 0.84 5.52 12.54 26.54 6.47 17.99 1.18 9.22 0.69
6 Pb rougher con 134.72 3.12 0.32 85.07 21.44 611.75 29.86 228.56 8.84 77.31
7 Pb rougher tailing 122.16 0.23 0.17 39.52 1.24 31.39 17.40 119.05 13.52 105.65
8 Pb scav con 10.14 3.00 0.36 7.25 8.17 17.72 39.71 23.11 11.58 7.70
9 Pb scav tailing 112.02 0.10 0.14 32.27 0.58 13.67 15.33 95.94 13.70 97.95
10 Pb 1st cleaner con 110.85 0.10 0.30 64.56 44.68 1030.87 13.63 84.37 3.91 27.69
11 Pb 1st cleaner 127.95 2.27 0.28 70.95 19.41 526.44 30.92 225.02 9.21 76.58
tailing
12 Pb 2nd cleaner con 31.98 0.10 0.51 31.72 50.75 337.91 10.54 18.84 2.45 5.00
13 Pb 2nd cleaner 104.08 0.11 0.25 50.44 43.64 945.57 13.91 80.84 4.06 26.95
tailing
14 Pb 3rd cleaner con 6.75 0.09 1.08 14.12 60.64 85.30 9.36 3.53 1.71 0.74
15 Pb 3rd cleaner 25.23 0.13 0.36 17.60 48.10 252.61 10.86 15.30 2.65 4.26
tailing
16 Zn rougher con 43.74 0.39 0.33 27.91 1.48 13.51 44.77 109.42 8.00 22.33
17 Zn rougher tailing 89.93 0.24 0.13 21.96 0.50 9.38 1.74 8.75 16.30 93.53
18 Zn scav con 5.51 3.59 0.47 5.12 2.23 2.58 15.14 4.71 14.26 5.07
19 Zn scav tailing 84.32 0.07 0.10 16.85 0.39 6.79 0.86 4.04 16.44 88.46
20 Zn 1st cleaner con 37.44 0.09 0.33 24.19 1.30 10.16 55.33 115.74 6.17 14.74
21 Zn 1st cleaner 16.04 1.07 0.40 12.49 1.99 6.64 19.52 17.52 12.51 12.83
tailing
22 Zn 2nd cleaner con 27.70 0.09 0.29 15.42 1.19 6.88 59.41 91.91 5.38 9.50
23 Zn 2nd cleaner 9.74 0.38 0.46 8.76 1.62 3.28 43.76 23.83 8.41 5.23
tailing
70.0
st
Pb 1 cleaner concentrate
40.0
Pb rougher concentrate
20.0
Cu rougher tailing
10.0 Pb rougher feed
Flotation feed
0.0
0.0 200.0 400.0 600.0 800.0 1000.0 1200.0 1400.0 1600.0 1800.0
Pb recovery, %
FIG 9 – The lead grade/recovery curve for the lead circuit within the Eureka Concentrator.
70.0
50.0
Zn rougher concentrate
Zn grade, %
30.0
20.0
Zn rougher feed
Pb rougher tailing
10.0
numbers 1 to 34 correspond to the sample numbers in Table 2. The next survey would be
S11, and so on. Flotation tests could be identified according to the operator’s initials, for
example, CG351/1 to CG351/5, which would be read as flotation test 351 completed by CG
with five test products.
Once a numbering system is in place it is easy to develop a storage system. Your storage
system must set rules regarding the length of time to keep samples, and allow the laboratory
technicians time to periodically maintain the storage facility. For example, it is wise to keep
flotation test products for three months and plant surveys for up to 12 months. It is also
wise to have a clean-up every three months to dispose of samples that have passed their
storage date.
When investigating the pulp chemistry in your circuit you might be tempted to sample
every stream in the concentrator. While this may be a good idea initially to understand the
variation between streams, it is soon apparent that some of this data is redundant. For example,
concentrate streams usually have high dissolved oxygen contents and very oxidising pulp
potentials because froth contains significant air. Therefore, a pulp chemistry survey of the
primary grinding/rougher flotation circuit might be skewed by measuring the pulp chemistry
of the rougher concentrate. However, if you are examining the pulp chemistry of a regrind/
cleaner circuit, you should collect a sample of the fresh concentrate feed to regrinding. If you
are comparing the surface analysis of streams (eg rougher concentrate and rougher tailing)
then you should collect the pulp chemistry of both streams.
12.0
10.0
8.0
pH
6.0
4.0
2.0
0.0
w
g
e
d
rg
rg
ee
ee
ee
lin
flo
flo
ha
ha
ai
rf
rf
rf
r
er
de
rt
he
sc
sc
he
he
ov
un
he
di
di
ug
ug
ug
e
ug
on
e
ill
i ll
ro
ro
ro
on
m
ro
l
yc
Pb
Zn
l
ll
G
yc
Zn
Ba
C
SA
Circuit position
FIG 11 – The pH profile through the grinding and flotation circuits of the Eureka Concentrator.
200.0
160.0
Eh, mV (SHE)
120.0
80.0
40.0
0.0
ow
ow
g
e
d
ed
ed
g
ilin
rg
ee
ar
fe
fe
r fl
rfl
ha
rf
ta
ch
de
r
he
he
sc
he
ov
r
s
un
he
di
di
ug
ug
ug
ne
ug
ne
ill
ill
ro
ro
ro
lo
m
m
ro
lo
yc
Pb
Zn
ll
G
yc
Zn
Ba
C
SA
Circuit position
FIG 12 – The Eh profile through the grinding and flotation circuits of the Eureka Concentrator.
6.0
5.0
4.0
DO, ppm
3.0
2.0
1.0
0.0
ow
ow
g
ge
ed
ed
d
g
ilin
ee
ar
fe
fe
ar
r fl
rfl
rf
ta
ch
ch
de
r
he
he
he
ov
r
s
s
un
he
di
di
ug
ug
ug
e
ug
on
ne
ill
ill
ro
ro
ro
lm
m
ro
l
lo
yc
Pb
Zn
G
yc
C
l
Zn
Ba
C
SA
Circuit position
FIG 13 – Dissolved oxygen profile through the grinding and flotation circuits of the Eureka Concentrator.
The pulp temperature ranged from 33°C to 35°C in most of the circuit (Figure 14). The
spike in pulp temperature in the ball mill discharge is due to part of grinding energy being
dissipated as heat.
The Eh-pH data for the grinding and flotation circuits are plotted in Figure 15 to determine
where the reactions are occurring. From the Nernst Equation 2 there is a dependence of redox
potential on pH:
a
E = E 0 + 0.059 log10 f Reactants p (2)
n a Products
Applying the Nernst equation to water results in a Pourbaix diagram that describes three
domains separated by lines of equilibria. The uppermost is the water-oxygen line (Equation 3),
above which water decomposes and oxygen is evolved and below which water is stable:
45.0
40.0
35.0
Temperature, oC
30.0
25.0
20.0
15.0
10.0
5.0
0.0
w
g
e
d
rg
rg
ee
ee
ee
lin
flo
flo
ha
ha
ai
rf
rf
rf
r
er
de
rt
he
sc
sc
he
he
ov
un
he
di
di
ug
ug
ug
e
ug
on
e
ill
i ll
ro
ro
ro
on
m
ro
l
yc
Pb
Zn
l
ll
G
yc
Zn
Ba
C
SA
Circuit position
FIG 14 – Temperature profile through the grinding and flotation circuits of the Eureka Concentrator.
250
200 6
8
150
Eh, mV (SHE)
4
5
100 1
1. SAG mill discharge
2
2. Cyclone underflow
50 3. Ball mill discharge
4. Cyclone overflow 7
5. Cu rougher feed 3
0 6. Pb rougher feed
7. Zn rougher feed
8. Final tailing
-50
4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0 12.0
pH
FIG 15 – The Eh-pH curve for the grinding and flotation circuits of the Eureka Concentrator.
This can be simplified further (Johnson, 1988; Natarajan and Iwasaki, 1973) for an oxygenated
aqueous solution with no well-defined redox couples to:
Broadly, if the changes in Eh and pH result in a line parallel to the water-oxygen line
this means that water equilibria is being maintained. That is, any change in Eh is directly
proportional to a change in pH with a relationship similar to Equation 4. If the changes in
Eh and pH result in a line that is perpendicular to the water-oxygen line, this suggests that
oxidative reactions are occurring.
Figure 15 shows that the Eh-pH lines between points 1 and 4 are perpendicular to the water-
oxygen line. This indicates that oxidative reactions are occurring in the grinding circuit. It is
likely that these oxidative reactions are corrosion of the grinding media and oxidation of
the sulfide minerals. As the pH is decreased across the ball mill, pyrite oxidation (an acidic
reaction) may be one of the dominant reactions. The changes in the Eh-pH curve between
points 4 and 5, 5 and 6, and 6 and 7 can be attributed to the addition of reagents that alter pH:
SMBS to copper flotation, lime to the lead and zinc circuits.
The EDTA extractable copper, lead, zinc and iron data are shown in Figure 16. The EDTA
extractable copper profile (Figure 9) suggests that the percentage of oxidised chalcopyrite
remained negligible throughout the grinding circuit. The EDTA extractable copper increased
during copper flotation to approximately 2.5 per cent and increases to 17 per cent in the
zinc rougher feed (which can be attributed to the addition of copper sulfate for sphalerite
activation).
Notably, the percentage of EDTA extractable lead is at least an order of magnitude greater
than the values for zinc and iron. This is because galena is a reactive mineral when in contact
with other sulfide minerals, particularly pyrite. The EDTA extractable lead profile shows
the percentage of oxidised galena remained approximately constant through the grinding
circuit, with values of around 1.0 per cent. It increased in the cyclone overflow and remained
reasonably constant through the copper circuit. After lead flotation, the lead scavenger
tailing had 17 per cent EDTA extractable lead, increasing to almost 40 per cent after zinc
flotation. This suggests that the lead species remaining in lead scavenger tailing are more
heavily oxidised.
The EDTA extractable zinc profile exhibits a similar trend to that observed for lead.
Through the grinding circuit the percentage EDTA extractable zinc was approximately
constant at around 0.1 per cent. This indicates that sphalerite oxidation was largely
unchanged through grinding. It then increased during the copper and lead flotation circuits
from 0.13 to 0.37 per cent and then significantly increased after zinc flotation to 1.0 per cent.
This indicates that the zinc species that remained in zinc scavenger tailings were more
heavily oxidised.
42.0 1.2
EDTA extractable Zn and Fe, %
EDTA extractable Cu and Pb, %
35.0 1.0
28.0 0.8
21.0 0.6
14.0 0.4
7.0 0.2
0.0 0.0
w
w
g
e
ed
ge
ee
ee
ilin
rg
flo
o
fe
ar
rfl
ha
rf
rf
er
ta
h
de
r
he
sc
he
he
sc
ov
l
na
un
di
di
ug
ug
ug
ne
Fi
ne
il l
ill
ro
ro
ro
lo
m
m
lo
yc
Zn
Pb
G
ll
yc
C
Ba
C
SA
Circuit position
Cu Pb Zn Fe
FIG 16 – The EDTA extractable copper, lead, zinc and iron profiles
through the grinding and flotation circuits of the Eureka Concentrator.
The EDTA extractable iron profile gives the best indication of the impact of grinding
media corrosion on the system. In primary grinding it ranged from 0.05 to 0.16 per cent, and
then increased to 0.30 per cent in cyclone overflow. That is, the pulp entering the flotation
circuit contains elevated levels of EDTA extractable iron, presumably after contact with the
forged grinding media. Interestingly, the levels decreased through the flotation circuit. This
is probably due to the continuing oxidation of the iron species to higher-level iron oxides,
which are not soluble in EDTA.
PROBLEM DEFINITION
From the metallurgical survey the circulating load in the lead rougher/lead first cleaner
circuit was extremely high (Table 10) for both lead and zinc. In fact, copper scavenger tailings
contributed only 15.4 and 28.6 per cent of lead and zinc units in lead rougher feed. The recycle
from the lead first cleaner tailing contributed by far the most lead and zinc to the lead rougher
feed (81.9 and 64.7 per cent, respectively).
TABLE 10
Mass balanced data for the lead rougher feed, with respect to flotation feed. Note: both the lead scavenger
concentrate and the lead first cleaner tailing are recycled back to the head of the lead rougher.
Stream t/h Grade (%) Recovery (%)
Cu Pb Zn Fe Cu Pb Zn Fe
Cu scav tailing 118.8 0.2 4.0 15.0 13.0 46.4 99.0 99.5 98.7
Pb 1st cleaner tailing 128.0 0.3 19.4 30.9 9.2 71.0 526.4 225 76.6
Pb scav concentrate 10.1 0.4 8.2 39.7 11.6 7.2 17.7 23.1 7.7
Pb rougher feed 256.9 0.3 11.8 23.9 11.1 124.6 643.1 347.6 183
You should correct for each mineral in the system as each mineral has a different specific
gravity (SG). The cyclone ‘cut’ sizes for each of the cyclosizer cones are in Table 11. When
dealing with elements that are present in parts per million (eg platinum-group minerals,
gold and silver) an understanding of their mineral associations is required. For example, if
the silver is associated predominantly with galena the cyclosizer cyclone cut sizes should be
calculated using a fSG based on the SG of galena. The same applies to gold that is associated
with pyrite; the fSG should be based on the SG of pyrite.
Note that the pre-cyclone, cyclosizer and/or the centrifuge size on hydraulic characteristics
rather than particle dimensions as obtained by screen sizing above 38 µm. Screen sizing is not
affected by particle SG; hydraulic sizing is. When the two sizing methods are used together
the transition point from one technique to the other can lead to some unusual shaped curves
due purely to the change in the sizing method. To overcome this, it is usual to combine the
-53 µm to +C2 size fractions to smooth the curves and remove the transition effect.
If you are considering a mineralogical analysis, you now need to ‘split’ out a small
representative portion from the samples of interest before pulverising the samples for assay.
Once you have reviewed the weights in each size fraction for each sample, decided which size
fractions to combine and ensured that there is enough sample for assay in each size fraction,
the samples can be prepared and submitted for assay.
TABLE 11
Cyclosizer cyclone ‘cut’ sizes for quartz, chalcopyrite, galena, sphalerite and pyrite.
Mineral
Correction factor Quartz Chalcopyrite Galena Sphalerite Pyrite
fT 0.9815 0.9815 0.9815 0.9815
fSG 0.7181 0.5038 0.7416 0.6423
fFM 0.9519 0.9519 0.9519 0.9519
fET 1.0063 1.0063 1.0063 1.0063
f 0.6751 0.4737 0.6972 0.6038
Cyclone ‘cut’ size
C1 44 30 21 31 27
C2 33 22 16 23 20
C3 23 16 11 16 14
C4 15 10 7 11 9
C5 11 7 5 8 7
100.0
90.0
80.0
Pb distribution, %
70.0
60.0
50.0
40.0
30.0
20.0 Pb rougher feed
FIG 17 – Lead distribution by size in the process streams contributing to the lead rougher feed.
the plant survey in Table 9 and for this section in Table 10. Using the mass balanced tonnages
for each process stream in Table 10, the size distribution for each sample point was applied
to calculate the solids flow in each size fraction (Table 12). The weight per cent and the assay
for each size fraction are used to calculate the head grade:
n (mi # ai)
Head grade = / /m (6)
i=1 i
where:
m is the weight per cent in size fraction i
a is the assay for that size fraction
This calculated head grade is then compared with the assay of the head sample to determine
how well the sizing and assay procedures have been completed. Unfortunately, in this
instance the mass and assay of the -5.6 µm fraction was calculated by difference. That is:
n-1
m-5.6nm = mH - /m i
(7)
i=1
and
n-1
e mH # a H - e / mi # ai oo
i-1
a-5.6nm = (8)
m-5.6nm
where:
mH and aH are the weight per cent and assay for the head
By calculating weight per cent and assay for the finest size fraction, all errors associated with
sizing, assaying and sampling accumulate to this size fraction. In an ideal world a sample of
this material would be collected.
TABLE 12
The lead recovery-by-size data for the lead rougher feed calculated from the copper
rougher tailing, lead first cleaner tailing and the lead scavenger concentrate.
Stream/size Wt Wt Pb parameter (%)
% (t/h) Grade Distribution Recovery to Pb Ro feed
Cu scavenger tailing
+22.2 µm 48.79 57.96 4.06 49.49 7.73
+16.7 µm 6.66 7.91 3.00 5.00 0.78
+11.6 µm 10.18 12.09 1.88 4.78 0.75
+7.6 µm 5.32 6.32 2.54 3.38 0.53
+5.6 µm 3.77 4.48 2.24 2.11 0.33
-5.6 µm 25.28 30.03 5.58 35.24 5.51
Head 100.00 118.79 4.00 100.00 15.62
Pb first cleaner tailing
+22.2 µm 82.44 105.48 22.12 93.96 76.72
+16.7 µm 6.13 7.84 8.84 2.79 2.28
+11.6 µm 3.19 4.08 8.64 1.42 1.16
+7.6 µm 1.17 1.50 5.10 0.31 0.25
+5.6 µm 0.81 1.04 4.10 0.17 0.14
-5.6 µm 6.26 8.01 4.18 1.35 1.10
Head 100.00 127.95 19.41 100.00 81.65
Pb scavenger con
+22.2 µm 70.32 7.13 9.66 83.14 2.26
+16.7 µm 10.48 1.06 8.20 10.52 0.29
+11.6 µm 6.69 0.68 2.62 2.15 0.06
+7.6 µm 2.15 0.22 3.20 0.84 0.02
+5.6 µm 1.37 0.14 3.68 0.62 0.02
-5.6 µm 8.99 0.91 2.49 2.74 0.07
Head 100.00 10.14 8.17 100.00 2.72
Pb rougher feed
+22.2 µm 66.40 170.57 15.46 86.72 86.72
+16.7 µm 6.55 16.82 6.05 2.25 2.25
+11.6 µm 6.56 16.85 3.55 1.97 1.97
+7.6 µm 3.13 8.03 3.03 0.80 0.80
+5.6 µm 2.20 5.65 2.62 0.49 0.49
-5.6 µm 15.16 38.95 5.22 6.68 6.68
Head 100.00 256.88 11.84 100.00 100.00
From the tonnes and assay of each size fraction you can calculate the distribution-by-size
and recovery-by-size for each stream. The distribution within a process stream is given by:
(mi # ai)
Distributiona = f p 100 (9)
(mH # aH) #
The distribution of lead by stream is shown in Table 12. In this case the recovery-by-size
is calculated for lead rougher feed because the copper scavenger tailings, lead first cleaner
tailings and lead scavenger concentrate combine to form the lead rougher feed. So, the
recovery for each size fraction, i, is given by:
J n N
/
K (m P # aP) O
Recoveryi,a = K P = 1 O # 100 (10)
K (m # a ) O
L H H P
where:
P are the process streams of interest
A similar approach is adopted for the more traditional recovery-by-size curves in Figure 18
and Figure 20.
90.0
80.0
70.0
60.0
Recovery, %
50.0
40.0
30.0
20.0
10.0
0.0
1.0 10.0 100.0 1000.0
Geometric mean particle size, microns
Lead Zinc Copper Iron Silica
FIG 18 – Recovery-by-size data for the lead final concentrate, with respect to flotation feed for the Eureka Concentrator.
6.0
to flotation feed), %
4.0
3.0
2.0
1.0
0.0 2.3
6.0
7.9
11.7
17.6
28.0
41.0
48.8
63.0
89.2
126.1
178.3
Geometric mean particle size, microns
FIG 19 – The galena distribution by size, with respect to flotation feed, in the lead scavenger tailing for the Eureka Concentrator.
100.0
90.0
80.0
70.0
Recovery, %
60.0
50.0
40.0
30.0
20.0
10.0
0.0
1.0 10.0 100.0 1000.0
Geometric mean particle size, microns
Lead Zinc Copper Iron Silica
FIG 20 – Recovery-by-size data for the zinc final concentrate, with respect to flotation feed for the Eureka Concentrator.
Figure 20 shows recovery-by-size for the zinc final concentrate. It shows a peculiar behaviour
of coarse (+20 µm) galena with recovery as high as 60 per cent. This is presumed to also be
related to galena-bearing/sphalerite-rich composite particles).
Mineralogy (Chapter 4)
Once the recovery-by-size data has been analysed and theories developed, select streams
should be examined mineralogically. This is a relatively expensive process and should only
be conducted after considering the work already discussed. In this case the recovery-by-size
analysis shows a significant recycle of lead first cleaner tailings to lead rougher feed, and
that most of this is in the +20 µm size fraction. It is hypothesised that this is due to galena/
sphalerite composites. Mineralogical examination will test this hypothesis.
TABLE 13
Galena liberation by size and mineral class for the lead rougher feed of the Eureka Concentrator.
Size (µm) Liberation class (%) Total (%)
Liberated Binary with … Ternary
Sphalerite Pyrite Gangue
+89 26 54 5 4 11 100
+28 66 28 3 0 2 100
+10 69 25 5 1 1 100
+6 93 6 0 0 0 100
-6 94 5 0 0 0 100
Head 66 28 3 0 2 100
100
90
80
70
60
Distribution
50
40
Liberated
30
Ga-Sp
20
Ga-Py
10 Liberation class
Ga-Gn
0
Head Ternary
2.3
6.0
10.1
28.2
Size, microns 89.2
FIG 21 – Size by liberation class data for the lead rougher feed. Ga – galena, Sp – sphalerite, Py – pyrite, Gn – gangue.
60.0
55.0
50.0
Pb grade, %
45.0
40.0
35.0
30.0
25.0
10.0 20.0 30.0 40.0 50.0 60.0 70.0 80.0 90.0 100.0
Pb recovery, %
No regrinding 5 minutes regrinding 10 minutes regrinding
15 minutes regrinding 20 minutes regrinding
FIG 22 – Lead grade/recovery curves for lead rougher flotation tests conducted
on lead first cleaner tailing, demonstrating the effect of regrinding.
flotation time of 12 minutes. The seven concentrate and tailing samples from each test were
prepared and submitted for assay.
To assist interpretation the kinetic data and the selectivity data can be tabulated. For example,
Table 14 shows lead concentrate grade and zinc and iron sulfide recoveries at 80 per cent lead
recovery. Selectivity curves can also be drawn.
TABLE 14
Lead grade and diluent recoveries, at 80 per cent lead recovery, for lead rougher flotation
tests conducted on lead first cleaner tailing demonstrating the impact of regrinding.
Test description Pb grade (%) Diluent recovery (%)
Zn IS
No regrinding 30.8 71.2 51.6
5 minutes regrinding 34.3 61.9 37.2
10 minutes regrinding 37.6 50.8 30.4
15 minutes regrinding 42.0 42.6 24.3
20 minutes regrinding 42.2 42.2 25.8
The improved selectivity for galena against sphalerite improves lead metallurgy; it is
expected that zinc metallurgy would also improve.
90.0
80.0
MF4 used as
Average % Rec.
50.0
13-May-98
26-May-99
5-May-99
18-Nov-98
10-Nov-99
18-Feb-98
11-Mar-98
10-Feb-99
24-Mar-99
16-Sep-98
29-Sep-99
28-Jan-98
2-Feb-00
20-Jan-99
3-Mar-99
8-Sep-99
12-Jan-00
7-Jan-98
15-Jul-98
26-Aug-98
30-Dec-98
28-Jul-99
18-Aug-99
22-Dec-99
22-Apr-98
24-Jun-98
5-Aug-98
9-Dec-98
14-Apr-99
16-Jun-99
1-Dec-99
1-Apr-98
3-Jun-98
7-Jul-99
28-Oct-98
20-Oct-99
7-Oct-98
W/E
Average Average CL UCL LCL
FIG 23 – Lead recovery versus time data from June 1998 to February 2000. Some milestones are highlighted.
These changes to the lead circuit decreased the amount of sphalerite reporting to lead
concentrate and made this sphalerite available for recovery in the zinc circuit. This improved
zinc metallurgy, with:
•• zinc recovery increased by 1.5 ± 0.4 per cent with >99 per cent confidence
•• the zinc concentrate grade increased by 2.1 ± 0.5 per cent with >99 per cent confidence.
Other statistical methods (for example, comparison of regression lines and multiple linear
regression modelling) produced similar values.
More surveys
It is necessary to continue monitoring the process through:
•• analysis of shift mass balances
•• analysis of monthly composite samples (by size and liberation class)
•• routine metallurgical and pulp chemical surveys
•• flotation cell characterisation (Chapter 6)
•• the introduction of a future ores testing program (Geometallurgy, Chapter 13).
After all, to measure is to know!
Chemistry
With the improvements in the lead circuit through better regrinding, it is time to study the
chemistry of the system.
Pulp chemistry
The pulp chemistry indicated that the lead rougher operated in oxidising conditions with
Eh values between 150 and 200 mV (SHE). Figure 15 shows the Eh-pH curve for the circuit.
The curve is essentially parallel to the water-oxygen line, which suggests that changes in Eh
or pH were due to maintenance of water equilibria. This indicates the system is operating in
oxidising conditions.
Solution chemistry
The EDTA extractable lead increased from approximately five per cent in copper rougher
tailings to 17 per cent in lead scavenger tailing. EDTA extractable zinc and iron values were
an order of magnitude lower than those for lead. This suggests the galena in lead circuit
feed is moderately oxidised, while galena remaining in the lead scavenger tailing is heavily
oxidised. A significant quantity of galena oxidation products are present in the pulp which,
under certain conditions, could cause inadvertent activation of sphalerite by lead.
Surface chemistry
To test the theory you developed from pulp chemistry observations, it is wise to complete
surface analysis on selected samples. Chapter 9 provides details of types of surface analysis
and how they may be used to provide more information. A practical example is provided in
he appendix for Chapter 9.
ion analysis on the process streams where the surface analysis samples were collected. This
information should be supplied to the surface scientist to help them understand the system
chemistry and explain their observations.
0.3
0.25 Pb cleaner con
0.2 Pb cleaner tail
0.15
0.1
0.05
0
Si
Pb
g
Ag
Zn
H
Fe
a
H
M
O
O
Pb
Fe
(+) Fragment
0.5
Pb cleaner con
0.4
Pb cleaner tail
0.3
0.2
0.1
0
C CH O OH S SO3 SO4 IEX
(-) Fragment
FIG 24 – Positive and negative mass spectra for sphalerite particles in the lead cleaner
concentrate and tailing process streams – confidence intervals calculated for 95 per cent.
•• The surface analysis confirmed the suspicion that the liberated sphalerite was activated by
lead ions. Copper and silver were also observed on the surfaces of sphalerite particles, but
in lower concentrations.
This suggests that some of the sphalerite is being recovered to lead concentrate because it is
activated by lead ions. You should propose that laboratory test work be completed examining
the effect of various reagents to depress activated sphalerite. Should these tests prove positive,
and the solution is economically sound, then it should be tested in the plant using statistically
rigorous trial methodology.
COMMUNICATION
The technical aspects of concentrator performance are very important in achieving optimum
metallurgy. However, the best technical expertise is of little consequence if people and
maintenance issues are not adequately addressed.
Process variations are probably the hardest to monitor and control and can result in operations
personnel treating symptoms rather than causes (ie ‘firefighting’). In most instances, process
variations such as ore hardness, mineralogical changes and head grade variation are very
difficult to measure online but can be partly managed with an effective ore characterisation
program (geometallurgy) and good communication between the mine and the mill.
Behavioural variations are often easily identified but difficult to correct because of the
human factor. One way to minimise behavioural variations in a concentrator is to document
and standardise each task so that each operator performs the same task the same way.
Unfortunately, this is easier said than done.
An important strategy to maximise metallurgical performance is to develop a proactive,
hands-on metallurgical team to establish and maintain metallurgical targets and procedures.
The metallurgist should be regarded as a member of the team with specialist knowledge to
call on when problems arise.
CONCLUSIONS
At the Eureka Concentrator, the use of classical metallurgical techniques (recovery- and
liberation-by-size analysis) has proved highly effective at determining the source of the
metallurgical problem. Initially, galena/sphalerite liberation was identified as a significant
impediment to better lead and zinc metallurgy. With the successful application of more
regrinding capacity, the focus shifted from liberation to chemistry. The use of pulp and
solution chemistry indicated that inadvertent activation of sphalerite by lead, copper and
silver ions was probably the reason for high recovery of liberated sphalerite into the final lead
concentrate. The judicious application of appropriate surface chemical techniques provided
confirmatory evidence. The Eureka metallurgists have now identified the reasons for the
losses of sphalerite from the circuit and can now develop solutions to this problem for testing
in the laboratory before plant trials.
REFERENCES
Johnson, N W, 1988. Application of electrochemical concepts to four sulphide flotation separations, in
Proceedings Electrochemistry in Mineral and Metal Processing II, pp 131–149.
Natarajan, K A and Iwasaki, I, 1973. Practical implications of Eh measurements in sulphide flotation
circuits, in AIME Transactions, 256:323–328.
Rumball, J A and Richmond, G D, 1996. Measurement of oxidation in a base metal flotation circuit by
selective leaching with EDTA, International Journal of Mineral Processing, 48:1–20.
Warman International Limited, 1991. Cyclosizer Instruction Manual – Particle Size Analysis in the Sub-Sieve
Range, Bulleting WCS/2 (Warman International Limited: Sydney).
FIG A1 – A bank of 12 flotation cells, and the cell grouping for a down-the-bank survey.
TABLE A1
The sample list for a down-the-bank survey conducted
on the bank of 12 flotation cells pictured in Figure A1.
Sample number Sample name Sample type
1 T bank feed Dip sample
2 T bank combined concentrate Lip or OSA sample
3 Con 1 – T bank cell 1 Timed lip sample
4 Con 2 – T bank cells 2 to 4 Timed lip sample
5 Con 3 – T bank cells 5 to 7 Timed lip sample
6 Con 4 – T bank cells 8 to 10 Timed lip sample
7 Con 5 – T bank cells 11 and 12 Timed lip sample
8 T bank tailing A Dip sample
9 T bank tailing B Dip sample
With the sampling points identified, you need to communicate your intentions to various
people involved and prepare your equipment. The same actions discussed above in the
section ‘The Metallurgical Survey, What do I have to do?’ apply to conducting a down-the-bank
survey. Communication and organisation are the keys to success.
In terms of sampling equipment, you will need:
•• dip and lip samplers that are clean and in good working order
•• sufficient clean buckets, with lids, that have been tared
•• a stop watch for timing the lip sample collection
•• a notebook to record the times and any other observations that may be of use later.
Conducting a down-the-bank survey is usually a job for two people. One person collects
the sample, and the other measures and records the time taken to collect each of the timed lip
samples. Prior to conducting the survey, it is wise to inspect the flotation bank to be sampled,
clean the cell lips to ensure that the froth flows freely, make the work area safe and free
of tripping hazards and establish how the lip sampling is going to occur. Next, perform a
dry run to practise how the timed samples will be collected. One method of sampling that
works effectively is for the sampler to yell ‘GO’ and ‘STOP’ as he/she starts and finishes the
sample collection. On these commands the time keeper starts and stops the stop watch. It is
imperative that the two people collecting the timed lip samples agree on the methodology,
and work in concert. So, with all the preparations complete and the plant operating correctly,
it is time to conduct the survey.
As with other surveys, you would have decided beforehand how many ‘cuts’ from each
process stream will be taken over a specified sampling time. The two people detailed to
conduct the down-the-bank survey will then collect the nine samples over the prescribed
sampling period. Once the sampling is finished, the samples are gathered together and taken
to the laboratory.
TABLE A2
Summary of the down-the-bank survey data collected.
No Sample name Weights (g) % solids
Gross Tare Wet Dry
Bucket Lip
1 T bank feed 1585.3 350.0 1235.3 471.9 38.2
2 T bank combined con 1891.1 350.0 245.0 1296.1 331.8 25.6
3 Con 1 – Cell 1 1789.7 350.0 245.0 1194.7 630.8 52.8
4 Con 2 – Cells 2 to 4 2217.0 350.0 245.0 1622.0 361.7 22.3
5 Con 3 – Cells 5 to 7 2049.4 350.0 245.0 1454.4 315.6 21.7
6 Con 4 – Cells 8 to 10 2142.3 350.0 245.0 1547.3 379.1 24.5
7 Con 5 – Cells 11 and 12 1795.6 350.0 245.0 1200.6 187.3 15.6
8 T bank tailing A 1428.8 350.0 1078.8 372.2 34.5
9 T bank tailing B 1610.0 350.0 1260.0 459.9 36.5
TABLE A3
Lip sample details.
No Sample name Sample time (s) Length (mm)
Cell lip Cutter lip
3 Con 1 – Cell 1 31.6 1250 × 2 105
4 Con 2 – Cells 2 to 4 28.9 1250 × 6 105
5 Con 3 – Cells 5 to 7 28.4 1250 × 6 105
6 Con 4 – Cells 8 to 10 30.8 1250 × 6 105
7 Con 5 – Cells 11 and 12 39.7 1250 × 4 105
Thus, by substituting the dry weights for each of the timed concentrate samples from
Table A2, and the sample times, number of cell lips, cell lip length and cutter lip length
from Table A3 in Equation A1, you can calculate the tonnes per hour recovered into each
concentrate. For example, for Con 1 – cell 1, Equation A1 becomes:
TABLE A4
Elemental assays for the down-the-bank survey.
No Sample name Assay (%)
Ag (ppm) Cu Pb Zn Fe SiO2
1 T bank feed 98 0.05 2.84 10.7 15.5 21.9
2 T bank combined con 444 0.24 14.20 11.1 23.0 9.2
3 Con 1 – Cell 1 644 0.47 21.40 10.7 21.2 7.6
4 Con 2 – Cells 2 to 4 490 0.28 15.70 11.3 22.9 8.5
5 Con 3 – Cells 5 to 7 406 0.21 12.60 11.3 23.8 9.1
6 Con 4 – Cells 8 to 10 346 0.18 10.60 11.6 24.2 10.0
7 Con 5 – Cells 11 and 12 306 0.18 9.15 11.7 24.3 10.9
8 T bank tailing A 70 0.02 2.04 10.5 14.8 22.6
9 T bank tailing B 72 0.02 2.13 10.6 14.9 22.6
So, from the timed lip sample collected from Cell 1, the tonnes per hour recovered are 1.71.
The same calculation was completed for each of the other concentrates. The values obtained
are given in Table A5. The tonnage distribution over the five concentrate samples was
calculated and also appears in Table A5.
The next step is to mass balance the ‘outer’ circuit. The process streams are:
1. T bank feed
2. T bank combined concentrate
3. T bank tailing.
There is only one node in this mass balance:
T bank feed = T bank combined concentrate
+ T bank tailing (1 = 2 + 8).
TABLE A5
The tonnes per hour and distribution of tonnes for the five
concentrates collected in the down-the-bank survey.
No Sample name Tonnage data
Tonnes per hour Distribution
3 Con 1 – Cell 1 1.71 14.55
4 Con 2 – Cells 2 to 4 3.22 27.36
5 Con 3 – Cells 5 to 7 2.86 24.30
6 Con 4 – Cells 8 to 10 3.17 26.91
7 Con 5 – Cells 11 and 12 0.81 6.88
Total 11.76 100.00
In completing this mass balance, it was assumed that the feed tonnage was 100 per cent.
The mass balanced data are provided in Table A6. The mass balance has calculated that the
tonnes of concentrate produced is 6.96 t/h, compared to 11.76 determined from the timed lip
tonnages. The difference is reconciled by applying the tonnage distribution from Table A5 to
the mass balanced tonnage given in Table A6. Thus, using the ‘outer’ circuit mass balanced
tonnage new values are calculated for each of the concentrates. These new values are
presented in Table A7 and are used in the second mass balance incorporating the down-the-
bank concentrates.
The down-the-bank mass balance has eight process streams:
1. T bank feed
2. T bank combined concentrate
3. Con 1 – Cell 1
4. Con 2 – Cells 2 to 4
5. Con 3 – Cells 5 to 7
6. Con 4 – Cells 8 to 10
7. Con 5 – Cells 11 and 12
8. T bank tailing.
There are two nodes:
1. T bank feed = Con 1 + Con 2 + Con 3 + Con 4 + Con 5 + T bank tailing (ie 1 = 3 + 4 + 5 + 6 +
7 + 8)
2. Con 1 + Con 2 + Con 3 + Con 4 + Con 5 = T bank combined concentrate (ie 3 + 4 + 5 + 6 + 7
= 2).
TABLE A6
The mass balanced data for the ‘outer’ balance.
No Sample time Wt Adjusted assay (%)
(%) Ag (ppm) Cu Pb Zn Fe SiO2
1 T bank feed 100.00 98 0.04 2.91 10.67 15.48 21.78
2 T bank combined con 6.96 444 0.24 14.20 11.10 23.00 9.23
3 T bank tailing 93.04 72 0.03 2.07 10.63 14.92 22.72
TABLE A7
The ‘new’ tonnes per hour for the five concentrates collected in the down-the-bank survey calculated from
the tonnage distribution and the ‘outer’ mass balanced tonnage recovered into the combined concentrate.
No Sample name New tonnes per hour
3 Con 1 – Cell 1 1.01
4 Con 2 – Cells 2 to 4 1.90
5 Con 3 – Cells 5 to 7 1.69
6 Con 4 – Cells 8 to 10 1.87
7 Con 5 – Cells 11 and 12 0.48
Total 6.96
The mass balance is accomplished by fixing the tonnage values for the T bank feed, and the
five down-the-bank concentrates, then using a mass balancing software fitting the data. The
down-the-bank mass balance is given in Table A8.
The mass balanced down-the-bank data can now be used to construct a lead grade/recovery
curve for this bank of flotation cells, as shown in Figure A2. This data can also be used to
generate kinetic data, examining the flotation rate constant and maximum recovery of the
various species.
TABLE A8
The down-the-bank mass balance.
No Sample name Wt (%) Weight (%) Recovery (%)
Ag Cu Pb Zn Fe SiO2 Ag Cu Pb Zn Fe SiO2
1 T bank feed 100.00 98 0.04 2.9 10.6 15.5 21.8 100.0 100.0 100.0 100.0 100.0 100.0
2 T bank 7.00 440 0.26 14.0 11.3 23.3 9.1 31.5 42.0 33.6 7.4 10.5 2.9
combined con
3 Con 1 – Cell 1 1.02 644 0.47 21.4 10.7 21.2 7.6 6.7 11.2 7.5 1.0 1.4 0.4
4 Con 2 – Cells 2 to 4 1.91 490 0.28 15.7 11.3 22.9 8.5 9.6 12.5 10.3 2.0 2.8 0.7
5 Con 3 – Cells 5 to 7 1.70 406 0.21 12.6 11.3 23.8 9.1 7.0 8.3 7.4 1.8 2.6 0.7
6 Con 4 – 1.88 346 0.18 10.6 11.6 24.2 10.0 6.6 7.9 6.9 2.0 2.9 0.9
Cells 8 to 10
7 Con 5 – 0.49 306 0.18 9.2 11.7 24.3 10.9 1.6 2.1 1.5 0.6 0.8 0.2
Cells 11 and 12
8 T bank tailing 93.00 72 0.03 2.1 10.6 14.9 22.7 68.5 58.0 66.4 92.6 89.5 97.1
25.0
20.0
Pb grade, %
15.0
10.0
5.0
0.0
0.0 10.0 20.0 30.0 40.0 50.0 60.0
Pb recovery, %
OUTER balance data Down-the-bank balanced data
FIG A2 – The lead grade/recovery curve constructed from the mass balanced down-the-bank survey data.
ASSUMPTIONS
The first set of assumptions relate to the minerals themselves. That is, the dominant sulfide
minerals are: galena, sphalerite, chalcopyrite and pyrite. The lead occurs as galena, the zinc
as sphalerite, the copper as chalcopyrite, and the iron occurs in sphalerite, chalcopyrite and
pyrite.
The mineral conversions are based on the assumption that each mineral is ‘pure’; for
example, galena is PbS, pyrite is FeS2, etc.
Eureka sphalerites are known to contain, in solid solution, moderate iron levels. In this
exercise 3.0 per cent was chosen. This value is an estimate of the average value.
A calculated assay of the non-sulfide gangue (NSG) is made by assuming that everything
that is not galena, sphalerite, chalcopyrite, or pyrite is non-sulfide gangue.
MINERAL CONVERSIONS
Each conversion is based on the atomic mass of the elemental components of the mineral.
Galena
Galena contains lead (207.2 amu) and sulfur (32.06 amu); therefore, the atomic mass of galena
(PbS) is:
PbS = Pb + S
PbS = 207.2 + 32.06 amu
Thus, ‘pure’ galena contains 86.6 per cent lead and 13.4 per cent sulfur. The conversion
factor (fGa) to convert the lead assay to galena is given by:
fGa = 239.26
207.2
fGa = 1.155
So, an assay of ten per cent lead is equivalent to 11.6 per cent galena.
% sulfur in PbS
fSGa =
% lead in PbS
fSGa = 13.4
86.6
fSGa = 0.155
Sphalerite
‘Pure’ sphalerite (contains no iron) is made up of zinc (65.38 amu) and sulfur (32.06 amu),
therefore the atomic mass of ‘pure’ sphalerite (ZnS) is:
ZnS = Zn + S
Thus, ‘pure’ sphalerite contains 67.1 per cent zinc and 32.9 per cent sulfur. Unfortunately,
Eureka sphalerite contains an average of 3.0 per cent iron in solid solution. Further, it is
assumed that the iron replaced zinc in the sphalerite lattice. Thus, the composition of Eureka
sphalerite is 64.1 per cent zinc, 3.0 per cent iron and 32.9 per cent sulfur. Hence, the conversion
factor (fSp) to convert the zinc assay to sphalerite is given by:
fSp = 100.00
% Zn in sphalerite
fSp = 100.00
64.1
fSp = 1.560
So, an assay of ten per cent zinc is equivalent to 15.6 per cent sphalerite. To determine the
amount of iron present in sphalerite, it is simply a matter of multiplying the zinc assay by the
ratio of the per cent iron in the sphalerite and the per cent zinc in the sphalerite. So, the iron
in the sphalerite conversion factor (fFeSp) is given by:
% iron in sphalerite
fFeSp =
% zinc in sphalerite
fFeSp = 3.0
64.1
fFeSp = 0.047
Thus, for a sample containing ten per cent zinc, the per cent iron associated with the
sphalerite is 0.47 per cent (ie 10 × 0.047). To check this, convert the zinc assay to per cent
sphalerite (ie 10 × 1.560), and multiply by the amount of iron in solid solution in sphalerite
(ie 3.0 per cent), and the per cent iron associated with sphalerite is 0.47 per cent.
Similarly, the sulfur in the sphalerite conversion factor (fSSp) is given by:
% sulfur in sphalerite
fSSp =
% zinc in sphalerite
fSSp = 32.9
64.1
fSSp = 0.513
Thus, for a sample containing ten per cent zinc, the per cent sulfur associated with the
sphalerite is 5.13 per cent (ie 10 × 0.513). To check this, convert the zinc assay to per cent
sphalerite (ie 10 × 1.560), and multiply by the amount of sulfur in sphalerite (ie 32.9 per cent),
and the per cent sulfur associated with sphalerite is 5.13 per cent.
Chalcopyrite
Chalcopyrite is made up of copper (63.55 amu), iron (55.85 amu) and sulfur (32.06 amu);
therefore, the atomic mass of chalcopyrite (CuFeS2) is:
CuFeS2 = Cu + Fe + S
Thus, stoichiometric chalcopyrite contains 34.63 per cent copper, 30.43 per cent iron
and 34.94 per cent sulfur. Hence, the conversion factor (fCh) to convert the copper assay to
chalcopyrite is given by:
CuFeS2 amu
fCh =
Cu amu
fCh = 183.52
63.55
fCh = 2.888
So, an assay of ten per cent copper is equivalent to 28.9 per cent chalcopyrite. To determine
the amount of iron present in chalcopyrite, it is simply a matter of multiplying the copper
assay by the ratio of the per cent iron in the chalcopyrite and the per cent copper in the
chalcopyrite. So, the iron in chalcopyrite conversion factor (fFeCh) is given by:
% iron in chalcopyrite
fFeCh =
% copper in chalcopyrite
fFeCh = 30.43
34.63
fFeCh = 0.879
Thus, for a sample containing ten per cent copper, the per cent iron associated with the
chalcopyrite is 8.8 per cent (ie 10 × 0.879). To check this, convert the copper assay to chalcopyrite
(ie 10 × 2.888), and multiply this by the amount of iron in solid solution in chalcopyrite
(ie 30.43 per cent), and the per cent iron associated with chalcopyrite is 8.8 per cent.
Similarly, the sulfur in chalcopyrite conversion factor (fSCh) is given by:
% sulfur in chalcopyrite
fSCh =
% copper in chalcopyrite
fSCh = 34.94
34.63
fSCh = 1.009
Thus, for a sample containing ten per cent copper the per cent sulfur associated with
the chalcopyrite is 10.1 per cent (ie 10 × 1.009). To check this, convert the copper assay to
chalcopyrite (ie 10 × 2.888), and multiply this by the amount of sulfur in solid solution
in chalcopyrite (ie 34.94 per cent), and the per cent sulfur associated with chalcopyrite is
10.1 per cent.
Pyrite
Assuming that pyrite is the dominant iron sulfide mineral present (ie no pyrrhotite), it is
relatively easy to calculate the per cent pyrite in a sample. Pyrite contains iron (55.85 amu)
and sulfur (32.06 amu); therefore, the atomic mass of pyrite (FeS2) is:
FeS2 = Fe + S
Thus, ‘pure’ pyrite contains 46.6 per cent iron and 53.4 per cent sulfur. Depending on the
assay data, the pyrite content of the ore can be calculated using either the iron or the sulfur
assay. Generally speaking, the estimation of the pyrite concentration is more accurate using
the sulfur assay because it is assumed that the vast majority of the sulfur in the ore is
associated with the sulfide minerals. Iron, on the other hand, can be present in the sulfides
as well as the non-sulfide gangue; for example, as iron oxides such as hematite or magnetite,
and feldspars to name but a few. Thus, the pyrite conversion factor (fFePy) based on the iron
assay is given by:
FeS2 amu
fFePy =
Fe amu
fFePy = 119.97
55.85
fFePy = 2.148
So, based on the conversion factor developed from the iron assay, an assay of ten per cent
iron is equivalent to 21.5 per cent pyrite.
Using sulfur, the pyrite conversion factor (fSPy) is given by:
FeS2 amu
fSPy =
S2 amu
fSPy = 119.97
64.12
fSPy = 1.871
So, based on the conversion factor developed from the sulfur assay, an assay of ten per cent
iron is equivalent to 18.7 per cent pyrite.
Unfortunately, iron is also associated with sphalerite and chalcopyrite in the ore, so these
contributions must be subtracted from the iron assay before determining the pyrite content.
Thus, the pyrite conversion factor (fFePy) based on the iron assay becomes:
From the discussion above the conversion factor for iron in sphalerite from the zinc assay
was determined to be 0.047, and a similar calculation revealed that the conversion factor for
iron in chalcopyrite is 0.879. So, the per cent pyrite is given by:
Thus, if an ore sample contained ten per cent iron, ten per cent zinc, and 0.1 per cent copper,
the per cent pyrite is 20.28 per cent. However, if the pyrite conversion factor (fSPy) is based on
the sulfur assay then:
From the discussion above the conversion factor for sulfur in galena from the lead assay
is 0.155; sulfur in sphalerite from the zinc assay was determined to be 0.513, and a similar
calculation revealed that the conversion factor for sulfur in chalcopyrite is 1.009. So, the
per cent pyrite is given by:
Thus, if an ore sample contained ten per cent sulfur, ten per cent lead, ten per cent zinc, and
0.1 per cent copper, the per cent pyrite is 6.0 per cent.
Non-sulfide gangue
If it is assumed that galena, sphalerite, chalcopyrite and pyrite are the only significant sulfide
minerals present in the orebody, it is possible to estimate the percentage of non-sulfide
gangue present. This is achieved by subtracting the per cent galena, sphalerite, chalcopyrite
and pyrrhotite from 100 per cent. That is:
Thus, if a sample assayed ten per cent lead, zinc and iron and 0.1 per cent copper, the
per cent non-sulfide gangue is 57.8 per cent.
FINAL COMMENT
These element to mineral conversions have been completed based on a lead/zinc ore;
however, a similar approach can be applied to other ore types provided the stoichiometry
of the minerals in the system is known, and the assumptions made in the calculations are
clearly explained.
CHAPTER 2
Existing methods for process analysis
Bill Johnson
ABSTRACT
This introductory chapter describes options for surveying flotation banks and calculation of
their solid and water balances, including mineral recoveries. The calculation and interpretation
of mineral recovery-size data and the more advanced mineral recovery-size-liberation data
are then covered. The related role of surface analysis for elucidating unexplained process
weaknesses in the mineral recovery-size-liberation data is described. This chapter is aimed
at recognition of the location and magnitude of process weaknesses in industrial plants and
identification of the mechanism involved in each, allowing technical solutions for some or
all weaknesses to be proposed and evaluated. Later chapters expand on topics covered here.
INTRODUCTION
The first step in analysis of the flotation process is to obtain a mass or solid balance, which
may be accompanied by a water balance, from surveying of the relevant portions of the
plant. The key benefit of obtaining the corresponding water balance is that water flow rates,
and hence pulp flow rates, can be calculated and hence residence times can be calculated, ie
kinetic data can be obtained.
Decisions must be made in the design of an experiment or survey on the level of sampling
detail that is appropriate for each portion of the plant. In general, the following options for
sampling detail exist and different levels of detail may be selected for various parts of the
plant:
•• block level, eg grouping of all cleaner banks or rougher banks
•• bank level, eg obtaining of information on feed, combined concentrate and tailing assays
•• down-the-bank data (where data on each concentrate or on groupings of concentrates are
obtained).
Of course, the whole plant may be treated as a block and this treatment is applied commonly
for obtaining shift or daily composite samples from automatic inventory samplers from which
shift or daily metallurgical performance is often calculated routinely. Such data can also be
is recovered, the grade of the concentrate will be detrimentally affected due to the adhering
gangue mineral.
For convenience, hand-held samplers will be considered in three categories: conventional
hand-held or manual samplers for sampling streams at weirs or leaving launders or pipes,
dip samplers for sampling the pulp in the pulp zone of a flotation cell, and lip sampling for
providing a sample of the concentrate from one or more flotation machines, which usually
also involves obtaining a value for the flow rate of solid by a timing procedure. Manual
samplers for sampling streams at weirs, leaving launders or pipes may be used in conjunction
with control samplers for online analysis systems and with well-formed sampling points in
automatic inventory samplers in special circumstances. For large tank cells, suitable portable
pumps or siphons may be used to obtain pulp or tailing samples.
There were major changes in the type and scale of cells employed in the last two decades
of the 20th century; this trend has continued in the first two decades of the 21st century.
The introduction of flotation cells with washed froth technology has changed the nature of
sampling problems. Further, the installation of very large tank cells has increased sampling
difficulties in many new plants due to lack of access to the large machines and decisions made
during the plant layout phase of design, which tend to increase sampling difficulties (see ‘Lip
Sampling’ in the Appendix).
From the work involved in the listed points 1 to 7, the planning and execution of a plant
survey to obtain a solid balance is a major task. In contrast, a solid balance is easily obtained
from a laboratory batch testing procedure from the readily obtained weights of dry solid in
the various concentrates and tailings, along with their assay values.
There are a limited number of methods for surveying a plant bank to enable the calculation
of a solid balance and related metallurgical information.
To provide clarity on the key aspect of obtaining a solid balance, it is valuable to state what the
generation of a solid balance does not involve. It does not involve obtaining estimates of the flow
rates of solid in the relevant streams by independent and quite different means, and assembling these
values from disparate sources. An example of such poor practice in a rougher would be obtaining
the solid balance by use of the feed solid flow from the weightometer, from timed lip sampling for the
concentrate and from instrumentation (volumetric flow and pulp density) for the tailing. In contrast,
it does involve using principally or solely the assay values for the relevant streams and a known reliable
tonnage (eg from a weightometer) as the scaling factor.
A survey of a total plant (many banks) or part of a plant (one bank or a limited number
of banks) is obtained by matching the limited number of methods to the type of equipment
in each part of the plant and the availability of sampling points, and from consideration of
the purpose of the survey, ie the type of metallurgical data required from the survey. The
limited number of methods will be described in following sections. The methods represent
the building blocks from which surveys are designed and conducted.
Water balance
The water flow rate in a stream can be calculated from the value for the solid flow rate that has
been determined as part of the solid balance. The following relationship holds:
Flow rate of solid (100 - per cent solids by weight)
Flow rate of water = (1)
Per cent solids by weight
The per cent solid value (by weight) for flotation concentrates requires some discussion for
the concentrate stream. The value of interest for analysis of flotation data is the per cent solid
value for the concentrate as it traverses the cell lip. The value for the per cent solid in the
concentrate should not reflect the water that has been added to the cell launder, to the pump
box or to the gland seal system for the concentrate pump. Taking the concentrate sample at
the cell lip provides the required per cent solid value.
Similarly, in laboratory batch tests, the water in the concentrate (for process analysis purposes)
should not include water that is added to wash the cell lip or the scraper during the collection of
a given concentrate. This can be arranged by adding such additional water from a wash bottle
for each concentrate and weighing the wash bottle before and after each concentrate is gathered.
Conceptually, the five-product and ‘higher’ cases could be used but these cases are almost
without exception impractical. In fact, the three- and four-product cases can only be utilised
under appropriate circumstances, which are now discussed.
To apply the two-product design in Box 1, the following form of the two-product equation
is used to calculate the fractional recovery of solid (ie the solid split):
a (f - t)
Solid recovery (fractional) = F = - (2)
(c t)
The concentration of at least one mineral (eg chalcopyrite or the copper assay) in the feed
needs to be significantly different from the value for the tailing, ie there must be a clear
difference between feed and tailing assays in the numerator, otherwise f - t will be a very low
Definitions
Let f, c and t be the assays for a particular element in the feed, concentrate and tailings
Let F = t/h of solid in feed
Let α, β and γ = t/h of solid in concentrate(s)
α α β α β γ
BOX 1 – Three traditional designs for acquisition of metallurgical data, including the commonly used two-product formula.
value (and possibly negative), and very susceptible to considerable error from sampling and
assaying. Therefore, the calculated fractional solid recovery will be prone to unacceptable
errors. One contributor to the overall sampling error, the fundamental error, is very dependent
on the sample mass and this issue is described in the appendix.
This separation of the mineral in question would then normally result in a significant
difference for the denominator (c - t). The assay values (feed, concentrate and tailing) for the
key element (eg copper) can then be used as shown in Box 1 to calculate the solid recovery
(α/F) or the absolute value for α if F is known. In addition to a large separation occurring for
the element used in the calculation, it is desirable that the selected element can be assayed
accurately and that sampling errors are not unusually high. Low-grade samples containing
some free gold can introduce large sampling errors.
To apply the three-product case, the concentration of at least one mineral in the feed would
need to be significantly elevated in the first concentrate while the concentration of a second
mineral would need to be significantly elevated in the second concentrate. A typical example
is the separation of galena (PbS) and sphalerite (ZnS) into two separate concentrates and the
two equations in the three-product case in Box 1 are solved by use of lead and zinc assays to
obtain the solid balance.
The four-product case can be thought of as an extension of the three-product case; however,
three different minerals require concentrating into their respective concentrates. An example
is the concentration of chalcopyrite, galena and sphalerite into their respective concentrates
and the use of copper, lead and zinc assays to solve the three equations in the four-product
case, to obtain the solid balance.
The three- and four-product cases can usually only be applied to the feed and products of
a concentrator (ie inventory sampler products and shift or daily performance calculations) of
a concentrator with multiple single mineral concentrates, as the special conditions for their
successful application exist for the products from such concentrators and do not usually exist
along individual banks. Therefore, the three- and four-product cases and ‘higher’ cases can
usually be expected to play no role or only a supporting role in designing a survey for the
banks in a concentrator. In contrast, the two-product case has often an important role in data
acquisition from individual banks within a flotation plant.
◦
Note: for the last cell in the bank, the tailing sample can be taken from the tailing stream (□) or from the pulp in the last cell ( )
(if the mixing state is known for the last cell).
M = exchange of pulp due to intermixing between adjacent cells (zero if cells are isolated) in units of M t/h of solid
Assays for calculating of solid balance for each cell using two product equation (ie ignoring the solid flows (M) from intermixing):
Feed assay Concentrate assay Tailing assay
Cell 1 f c1 p1
Cell 2 p1 c2 p2
¦ ¦ ¦ ¦
Cell n pn-1 cn pn
For the first cell in the bank: let C = t/h of solid in concentrate, F = t/h of solid in feed, and T = t/h of solid in tailing.
Stating the two-product equation and inserting the symbols for cell 1:
Solid recovery * = C = f - t
F c-t
C = f - p1
F c1 - p1
f - p1
C = F#
c1 - p1
T=F-C
*expressed as a fraction
This calculation is reapplied for cell 2, cell 3, etc.
BOX 2 – Bank survey design involving repeated use of the two-product equation.
The approach depicted in Box 2 represents repeated application of the two-product equation
to the data for each machine to obtain the solid balance. It must be recognised that the method
is based on some simplifying assumptions:
•• the pulp zone of each machine is perfectly mixed (to allow ready and reliable sampling via
a dip sampler)
•• the effect of the intermixing flow rates (M) is disregarded in the mass balance for each
machine (see Box 2).
There are some other properties of this type of sampling:
•• A relatively large number of samples are required (compared to some other options) as a
pulp sample from each cell is required.
•• Experimental error in sampling, preparation and assaying of the series of pulp samples
can result in irregularities in the data, particularly towards the end of the bank where the
additional recovery in each cell is low. For example, the recorded assay for the valuable
element in the pulp region of cell n can exceed its assay in cell n-1. It should be noted that
data smoothing techniques available since approximately 1970 may handle this issue but,
for manual calculations, this property remains a difficulty.
Use of the approach in Box 2 declined during the period from 1970 to approximately 1990
because a more effective method emerged in terms of the quantity of samples required and
its theoretical basis. This more effective method will be discussed shortly in conjunction with
Box 4. However, since approximately 1990, a change in the type and volume of flotation
machines being installed (ie the use of large tank cells) has resulted in increased use of a
closely related variant of the method in Box 2. This variant is illustrated in Box 3.
Assays for calculating of solid balance for each cell using two product equation:
Feed assay Concentrate assay Tailing assay
Cell 1 f c1 t1
Cell 2 t1 c2 t2
¦ ¦ ¦ ¦
Cell n tn-1 cn tn
For the first cell in the bank: let C = t/h of solid in concentrate, F = t/h of solid in feed, and T = t/h of solid in tailing.
Stating the two-product equation and inserting the symbols for cell 1:
Solid recovery * = C = f - t
F c-t
C = f - t1
F c1 - t1
f - t1
C = F#
c1 - t1
T=F-C
* expressed as a fraction.
BOX 3 – Bank survey design involving repeated use of the two-product equation for a series of tank cells.
The design of the tank cells precludes the possibility of any intermixing of pulp zones and
removes any objections on this basis. However, for taking dip samples from the pulp zone of
each cell, the degree of mixing in the pulp zone must be considered; the degree of mixing in
the pulp zone is also relevant for the correct use of portable pumps or siphons for obtaining
a pulp sample. However, it is also possible for valid sampling points to exist in the pipework
between the cells, eliminating the use of dip samples, which is time consuming and presents
difficulties if the pulp zone is not perfectly mixed. Banks of unit tank cells contain typically six
to ten cells and the moderate number of cells makes the workload of the method of sampling
manageable. When there is adequate access to the cell lips, a viable method to sample tank
cells down the bank is the timed concentrate sampling method, which is now discussed.
A competing method for sampling a bank emerged from approximately 1970 (Box 4).
It required samples of the feed, tailing and the various concentrates for which some grouping
was possible. For each concentrate or concentrate grouping, the flow rate of solid discharging
from the various cell lips also had to be determined (Restarick, 1976).
From Box 4, no assumptions on pulp intermixing between machines or the level of mixing
within machines were required, the only possible exception being the collection of the tailing
sample from the pulp zone (for which an alternative point, sampling the combined tailing
stream after exiting the last cell, exists in some cases). The elimination of all or virtually
all the pulp sampling reduced the quantity of work. However, a more advanced method
of concentrate sampling, preferably requiring two people, was required to obtain values
for the solid flow rate for each machine or grouping of machines. Details of this aspect of
the sampling are provided in the appendix, where its application to large circular cells is
discussed in addition to its original application to small rectangular cells.
In Box 4, the two-product equation is demonstrated to be a special case of a general equation
for removal of many concentrates (designated as n). Box 4 presents the general equation and
the method for initial examination of the data from the survey. Concentrate assays and flow
rates by lip sampling must be obtained for each concentrate and a feed and tailing assay for
the bank must also be obtained.
It is important to discuss the reliability that can be attributed to the measured values for the
solid flow rate (C1, C2, C3, … Cn) in each concentrate (see Box 4) and hence the appropriate
means for processing the data from this type of sampling.
The calculated value for F is compared with its known value from the weightometer. The observed lip tonnages for each concentrate
(C1, C2 … Cn) are then rescaled upwards or downwards to ensure the value for F based on the rescaled values for C1, C2, … Cn equals
the true value from the weightometer.
For the general case above applied to a bank with seven concentrates:
Rewriting the two-product equation where F is calculated from the concentrate flow rate C and the feed (f) and the product assays
(c and t):
F = C (c - t) and T = F - C
f-t
It can be observed that the two-product equation for F is a special case of the above equations for F with 7 and n cells in the bank.
BOX 4 – Bank survey requiring assays for the feed, tailing and concentrates and the relative flow rates for each concentrate
(no knowledge of the degree of intermixing (M) is needed for legitimate use of this method – see also Box 2).
tonnages are available. In some cases, extra metallurgical samples may have to be taken to
link the plant section being surveyed in detail to the source of an absolute solid tonnage, eg
weightometer, thereby removing reliance on the correctness of concentrate tonnages on an
absolute basis.
CASE 1
Grade
Recovery
CASE 2
Grade
Recovery
CASE 3
Grade
Recovery
FIG 1 – Illustration of the reliability of calculated metallurgical performance for a valuable mineral/element using the method
with measured lip tonnages (see also Box 4). Case 1: Lip tonnages correct on absolute basis, ie compatible with reference
tonnage and correct relative to each other. Residence times are measured correctly. Case 2: Lip tonnages correct relative to
each other but all ten per cent too high in comparison to reference tonnage (lip tonnages not rescaled as described in Box 4).
Calculated residence times are lower by ten per cent than correct value. Position of grade-recovery curves for cases 1 and 2 is
the same. Case 3: Lip tonnages correct relative to each other but all 15 per cent too low in comparison to reference tonnage
(lip tonnages not rescaled as described in Box 4). Calculated residence times are higher by 15 per cent than correct value.
Position of grade-recovery curves for cases 1, 2 and 3 is the same. Case 4: Lip tonnages not correct relative to each other.
Position of grade recovery curve becomes subject to additional errors and is no longer the same as for cases 1, 2 and 3.
meter and a density gauge (both of which are calibrated properly and checked regularly) may
be used to provide a scaling factor for the tonnes per hour of solid for a key stream in a survey
of some peripheral part of a circuit.
2. the use of point counting or related methods involving an optical microscope, image
analyser or automated electron microscope for mounted samples, often size fractions and
sometimes on the original sample containing all sizes
3. the use of approximate methods for conversion of elemental assays to mineral assays
where the approximations are specific to a given ore or site.
Methods 1 and 2 provide mineral assays but some understanding of their limitations is
required. Quantitative X-ray diffraction is suitable for assaying total samples or individual size
fractions from samples; the method is suited for minerals with high assays and is particularly
suited for non-sulfide gangue minerals. Point counting or related methods as practised via
an optical microscope are usually applied to size fractions and the finest size fraction from a
given sample cannot be measured by this method. In this situation, assumptions for the finest
size fraction have been needed to sum the data across the size fractions to obtain the mineral
head grade. Methods for obtaining mineral assays with an automated electron microscope
are also usually applied to size fractions; the ability to measure the finest size fraction has to
be determined on a case by case basis.
In general, these mineral assay methods are less accurate than traditional assaying methods
for elements and they have increased technical difficulty and/or become much more costly
if minerals are being assayed in the low concentration region of less than one per cent by
weight. In practice, methods 1 and 2 are often used in conjunction with elemental assays.
A very powerful application of method 2, particularly with an automated electron
microscope, is for an ore with multiple minerals reporting to one concentrate, eg an ore
with several copper sulfide minerals. Mineral assays can be obtained for each copper sulfide
mineral and the performance of each can be calculated.
Further, in many cases, the use of all the methods in combination is warranted, allowing a
method that draws on the strengths of the data from various methods for obtaining mineral
assays. Suitable programs are needed for execution of the necessary calculations to produce
a combined set of mineral assays.
It has been observed that operating sites with both traditional assaying facilities for elements and
recently acquired assaying facilities for minerals generally do not merge the two sources of assay data
well to maximise the benefit to the site.
Method 3 relies on the elemental assays, which are very reliable but require assumptions
for conversion to mineral assays, which are of various levels of reliability. As an example,
consider a simple ore containing lead (only as galena), pyrite, a small amount of copper (only
as chalcopyrite) and various non-sulfide gangue minerals including talc.
Clearly, the lead and copper assays can be converted to galena and chalcopyrite assays,
preferably by using the measured stoichiometry for galena and chalcopyrite in that ore.
If sulfur only existed in the galena, chalcopyrite and pyrite, the sulfur assay could be corrected
for sulfur in the galena and chalcopyrite, and the corrected sulfur assay converted to a pyrite
assay, preferably using the measured stoichiometry for pyrite in that ore. By subtracting the
mineral assays for galena, pyrite and chalcopyrite from 100 per cent, a useful estimate of the
non-sulfide gangue assay would be obtained. If a substantial and variable proportion of the
magnesium in the ore existed in non-talc minerals, conversion of magnesium assays to talc
assays may provide unreliable information on the talc. In such circumstances, a combination
of elemental assays that are converted to mineral assays and directly measured mineral assays
for the talc (eg by X-ray diffraction) could be devised for that particular ore.
As a result, a mineral balance must also exist for mineral A (see the same flow rates for
mineral A in the actual feed and recalculated feed (from summation of flows of mineral A in
the products) in Table 1b).
However, for minerals B and C, it is very unlikely that an exact mineral balance will
exist using the observed assays for minerals B and C and the solid balance calculated from
mineral A. In Table 1b, the different flow rates of minerals B and C are shown for the actual
feed and for the recalculated feed (defined as the sum of the flow rates in the concentrate
and tailing). The recovery values for minerals B and C can be expressed with respect to the
TABLE 1A
Raw mineral assays for a separator with two products.
Feed Concentrate Tailing
Mineral A 10.2 81.3 1.2
Mineral B 19.7 11.6 21.6
Mineral C 70.1 7.1 77.2
Total 100 100 100
Feed Tailing
F F –
t/h solid t/h solid
Concentrate
t/h solid
FIG 2 – Illustration of the symbols for solid flow for a separator producing two products.
TABLE 1B
Summary of flow rates and recoveries.
Feed Recalculated Concentrate Tailing Recoveryb Recoveryc
(t/h) feeda (t/h) (t/h) (t/h) (%) (%)
Mineral A 102 102 91 11 89.2 89.2
Mineral B 197 205 13 192 6.6 6.3
Mineral C 701 693 8 685 1.1 1.2
Solid 1000 1000 112 888 11.2 11.2
a. Sum of mineral flow rates in concentrate and tailing.
b. With respect to actual feed.
c. With respect to recalculated feed.
recalculated or the actual feed. It is recommended that the recovery value is calculated with
respect to the recalculated feed because:
•• the resulting recovery value is less affected by sampling and assay errors existing in the
data, particularly for minerals with higher flow rates in the concentrate than the tailing
•• the values are bounded between zero per cent and 100 per cent
•• a consistent approach is provided.
There are several practical advantages in obtaining the recovery of minerals, aside from the
previously stated observation that, because minerals are being separated, logic dictates that
mineral data are the most relevant.
When metallurgical data are analysed on the basis of elements, some or all of the solid in
the non-sulfide gangue category may be ignored. Further, knowledge of an ore’s behaviour
via minerals allows ready calculation of the effects from changing ore head grade, assuming
all other properties remain fixed. Because this approach requires a knowledge of the
stoichiometry of each mineral and because the concentrations of minor and trace elements
in the lattice of each mineral should also be found in detailed analyses, more complete
information on the real limits to a separation can be obtained, based on the properties of the
minerals. For example, most sphalerite contains some iron in solid solution and the mineral
marmatite reflects naturally occurring zinc sulfide with very high levels of iron.
2b. calculate the recovery value for each mineral in each size fraction, noting that the
recovery values with respect to the actual feed and recalculated feed will be the same
because the data were adjusted statistically to provide internally consistent assays in 2a.
When there is size reduction inside the section considered, it is only possible to employ
methods 1a and 1b for the recovery calculation for each mineral in each size fraction, there
being no basis for data smoothing. It is also relatively common that no sized data exist for
the feed stream. This can arise because a sampling point did not exist or the available sample
was not selected for sizing. In this case, the recoveries can only be calculated by use of the
recalculated feed.
The next step, which allows interpretation of the recovery-size information more readily,
is to summarise the information in a graph of mineral recovery (y-axis) and the mid-point of
the size fraction (x-axis) using a logarithmic scale on the x-axis. All the minerals should be
plotted on the one graph for a given processing stage and the values for the mid-points on the
x-axis should reflect the specific gravity of the minerals when the sizing device operates on
the basis of hydraulic equivalents such as for a cyclosizer, its predecessor (infrasizer) or for
sizings by beaker decantation. When the sizing is obtained by sieving, the mid-points for all
minerals will be the same.
To fully appreciate recovery-size data for minerals, the flow rates of each mineral in each
size fraction in the recalculated/actual feed also need to be reviewed. The values for mineral
recovery are key indicators of metallurgical performance but the practical and economic
significance of the recovery values depends on the quantity of each mineral in each size
fraction on which the recovery value acts. While it is less common to include this information
in the recovery-size curves, this can be done by way of a histogram along the base of the
recovery-size curve. Alternatively, tables or other means may be devised to allow convenient
recognition of this important information.
with the superimposed effects from incomplete liberation, which affect all size fractions,
but which typically affect the less liberated coarse size fractions to the greatest extent for
a given ore. It must be noted that by obtaining liberation data as a next step the effects of
this complication can be understood for a given ore. This avoids having full reliance on
deductions from recovery-size graphs, particularly for an ore whose liberation characteristics
are not well known.
It is also assumed that a particle that reports to a given size fraction in the recovery-size
curves existed in that size fraction in the flotation process. The existence of a fully dispersed
pulp is therefore assumed.
It is useful to summarise the general form of some basic relationships (Pyke, Fornasiero and
Ralston, 2003) between the efficiency of each subprocess in the pulp zone and particle diameter
(Figure 3). Figure 3 arises from ongoing investigations at the Ian Wark Research Institute. The
authors used the following equation to describe the collection or capture efficiency (Ecoll) for a
particle and bubble in terms of the efficiencies of the subprocesses in the pulp zone:
Ecoll = EcEaEs
where:
Ec = collision efficiency
Ea = adhesion efficiency
Es = stability efficiency
For each of the efficiencies in the subprocesses (Ec , Ea and Es), Pyke, Fornasiero and Ralston
(2003) provided equations relating them to properties of the flotation system.
For the particle/bubble collision step, there is a direct relationship between collision
efficiency and size (see Figure 3). This arises because small particles approaching a bubble
tend to be swept along the stream-lines around bubbles while larger particles with higher
FIG 3 – Calculated relationships between collision efficiency Ec, adhesion or attachment efficiency Ea and stability efficiency
Es and particle diameter dp, and the resulting calculated relationship between the first-order rate constant k and particle
diameter dp for the pulp zone, where cases a, b and c are for advancing contact angles of 50°, 65° and 80° respectively.*
* Reprinted from J Colloid and Interfacial Science, Vol 265, B Pyke, D Fornasiero and J Ralston, Bubble particle
heterocoagulation under turbulent conditions, pp 141–151, Copyright (2003), with permission from Elsevier.
momentum have an increased ability to cross the stream-lines and complete a collision. This
inability to cross stream-lines means that the probability is low for small particles to approach
a bubble with sufficient proximity for a collision to have occurred, ie with sufficient proximity
for the adhesion process to commence. For a particle of a given diameter, its momentum and
its ability to cross the stream-lines is increased if it has a higher density.
The second subprocess (known as adhesion or attachment) commences after a particle and
bubble come into close contact. Pyke, Fornasiero and Ralston (2003) described the approach
of the particle and the bubble:
Should they approach quite closely, within the range of attractive surface forces, the intervening
liquid film between the bubble and particle will drain, leading to a critical thickness at which
rupture occurs. Movement of the three-phase contact line (the boundary between the solid
particle surface, receding liquid phase, and advancing gas phase) then occurs, until a stable
wetting perimeter is established.
There is an inverse relationship between the adhesion efficiency and particle diameter
(Figure 3), noting that this subprocess is referred to as attachment in the reference (Pyke,
Fornasiero and Ralston, 2003). Small particles slide more slowly over the surface of a bubble
as they are ‘protected’ due to their low diameter by existing in the more slowly moving
boundary region of the water phase near the surface of the bubble. The lower sliding velocity
for smaller particles allows a greater time for the adhesion subprocess to be successful, ie for
the actual contact time between the particle and the bubble to exceed the needed contact time
for adhesion, known as the induction time.
Similarly, for the transportation step in the pulp zone, there is also an inverse relationship
between the stability efficiency and particle diameter (Figure 3). Small particles are subjected
to lower forces of detachment and have a higher probability of successful transportation as a
particle/bubble aggregate to the base of the froth phase.
In Figure 3, Pyke, Fornasiero and Ralston (2003) provided the summation of the relationships
for the three described subprocesses by calculating the relationship between the first-order
rate constant (k) and particle diameter for the hydrophobic mineral for a set of conditions.
This calculation involved the use of typical parameters for the induction time.
The first-order rate constant (k) for a mineral is closely related to its recovery (see section
entitled Process Analysis with Down-the-Bank Flotation Data) and summarises the propensity
for the flotation process for a mineral to proceed in the pulp region for a set of conditions. The
rate constant for the pulp region is linked to the efficiencies in the three subprocesses by the
following equation (Pyke, Fornasiero and Ralston, 2003):
k = z Nb Ec Ea Es
where:
z = frequency of particle bubble collision
Nb = number of bubbles per unit volume
It can be seen that the relationship between the rate constant and particle diameter in the
example has a maximum as a result of the interaction of the direct relationship between
efficiency and particle diameter for the collision subprocess and the two indirect relationships
discussed for the adhesion and transportation subprocesses. In some cases, a small plateau
region may result from the interactions of the three relationships. The importance of a high
particle hydrophobicity in the adhesion subprocess and in avoiding detachment in the
The entrainment efficiency value for size i has been called the classification function
(Johnson, 1972) or the classification vector (Lynch et al, 1981). It is commonly described by
the term ENTi in the present literature and the topic was reviewed by Johnson (2005). Typical
values for ENTi for a range of size fractions of siliceous non-sulfide gangue are provided in
Review
To understand the behaviour of a particular ore, there is often a benefit from observing its
flotation behaviour in the absence of collector but with a normal frother addition (and possibly
depressant addition). The benefit can be increased if the recovery-size behaviour is obtained
for the recovered minerals. Minerals that display normal flotation through the particle/bubble
collision sequence in the absence of collector can be readily recognised (recovery values in
some size fractions considerably larger than water recovery) and the size fractions in which
TABLE 2
Values for entrainment efficiency factors (ENTi) for siliceous non-sulfide gangue (Lynch et al, 1981).
Size fraction (µm)
-11 -16 +11 -23 +16 -33 +23 -44 +33 -75 +44 +75
Entrainment efficiency 0.83 0.44 0.24 0.11 0.04 0.03 0.0
this occurs can be readily identified (Johnson and Jowett, 1982). Minerals that display this
behaviour can be valuable sulfides, non-valuable sulfides and non-sulfide gangue (eg talc).
Hydrocarbons present during ore formation and which remain associated with a portion of
one mineral (eg pyrite with a rimming of hydrocarbons) can cause this behaviour (Croxford,
Draper and Harraway, 1961).
As a result of the various mechanisms for flotation and the steps involved in each mechanism,
recovery-size curves for the valuable mineral being recovered by flotation adopt a general form
with many variations, of which a few are demonstrated in Figure 4. It is given that the curves
were observed after high residence times, ie their position and shape were changing only very
slowly with additional residence time. For the valuable mineral in all graphs, two scenarios
(Johnson, 2006) are shown for the fine fractions. In scenario 1, the diminished recoveries are
typical of those arising from deficiencies in the collision subprocess. In scenario 2, the more
greatly diminished recoveries result from an additional effect beyond the collision deficiency
likely to result from an imbalance in the ratio of adsorbed hydrophobic/hydrophilic species
on the valuable mineral.
For the minerals for which recovery is not being sought, an extremely wide range of
positions and shapes exist (Figure 4). For the hypothetical fully liberated feed, the following
are demonstrated:
•• entrained liberated non-sulfide gangue (cases A, C and E)
•• entrained and hydrophobic, liberated sulfide gangue due to collector adsorption (case C)
•• entrained and naturally hydrophobic liberated sulfide gangue (case E).
For the realistic cases (B, D and F) with an imperfectly liberated feed, cases are shown with
increased recoveries of sulfide and non-sulfide gangue due to their recovery in composites,
along with lower recovery of coarse valuable mineral due to the lower hydrophobicity of
composite particles containing some valuable mineral. In case B, two examples are given for
the relationship between non-sulfide gangue recovery and size (log scale); no examples for
sulfide gangue are given in case B or case A.
Case A Case B
100
100 100
100
9090 9090
8080 8080
7070 7070
Recovery (%)
Recovery (%)
6060 6060
Recovery (%)
Recovery (%)
5050 5050
4040 4040
3030 3030
2020 2020
Water recovery (10%) Water recovery (10%)
1010 1010
10 10
11 1010
100
100
11 10
10
100
100
Case C Case D
100
100 100
100
9090 9090
8080 8080
7070 7070
Recovery (%)
Recovery (%)
6060 6060
Recovery (%)
Recovery (%)
5050 5050
4040 4040
3030 3030
2020 2020
Water recovery (10%) Water recovery (10%)
1010 1010
11
0 1 0
1 1010
100
100 11 1010
100
100
Case E Case F
100
100
100
100
90
90 9090
80
80 8080
70 7070
Recovery (%)
Recovery (%)
70
60 6060
Recovery (%)
Recovery (%)
60
50 50 5050
40
40 4040
30 30 3030
20
20
2020
Water recovery (10%)
Water recovery (10%)
10
10 1010
10 10
11 10
10
100
100 11 10
10
100
100
Valuable mineral (Scenario 2) – additional deleterious effect beyond ‘collision effect’ for fine valuables
Non-sulfide gangue (Case 2) – lowered contribution from recovery in composites with valuable mineral
compared to case 1
Sulfide gangue
FIG 4 – Examples of some mineral recovery – particle diameter (log scale) curves for a feed with
perfect liberation (cases A, C, E) and for a feed with acceptable liberation (cases B, D, F) for which
recovery of unwanted minerals in composites becomes a possible mechanism.
the unwanted minerals can be recognised, providing strong clues to their mechanism of
recovery. Equally, the liberation state of the valuable mineral is determined. If significant
losses of valuable minerals in composites are observed, corrective steps involving grinding or
regrinding may be assessed.
Liberation data allow the flow rate of each mineral in each size fraction of each product
(obtained from the recovery-size (log scale) level of analysis) to be distributed between a
number of categories (typically from five to ten). The performance of the various categories
becomes the basis of the next level of analysis.
Liberation data for a mineral are supplied in one of two basic forms. The traditional point
counting method with use of an optical microscope provides liberation data in the general
format as shown in Table 3 for a size fraction. Examples of graphs derived from this type of
data are given in Figures 5, 7, 8 and 9.
Automated scanning electron microscopes can also provide the data in the point counting
format illustrated in Table 3. The data from these devices are also commonly supplied in
another format based on the percentage of the mineral of interest in each category. An example
is shown in Table 4. An example of a graph derived from this type of data is given in Figure 6.
In the type of data indicated in Table 4, some grouping of the original data has been
performed. Typically, the data are provided with increments of ten per cent from category to
category; in practice, some grouping of the higher quality classes is often used while narrow
classes are employed for the very important lower quality composites, as shown in Figure 6.
The reader is reminded that several calculation steps, as described in preceding sections
of this chapter, are the necessary precursors to incorporation of liberation data. These steps
are listed:
1. Calculation of the mass or solid balance for the circuit. The water balance should also be
calculated.
2. Use of size distributions and assays for the size fractions of the concentrate and tailing
(as the minimum relevant streams), along with the solid flows from item 1, to calculate
the flow rates of minerals in each size fraction of the concentrate and tailing. The mineral
assays for each size fraction may be from conversion of elemental assays to minerals
TABLE 3
Example of some point count data for sphalerite in one size fraction.
Liberation data (categories)
% Liberated % in binary composite with listed minerals % in ternary
Galena Iron sulfide Non-sulfide gangue compositesa
% in each category 60 16 4 11 9
a. Includes other multiphase composites, eg quaternaries.
TABLE 4
Example of second form of liberation data for sphalerite in a size fraction.
Liberation data (% of mineral of interest defined in each category)
100 90–100 70–90 50–70 30–50 20–30 10–20 0.1–10
Percentage in category 61.2 30.4 4.1 2.1 1.1 0.4 0.4 0.3
100
90
80
70
60
Recovery (%)
Total Sphalerite
50 Liberated Sphalerite
Binary with Galena
40 Binary with Iron Sulfide
Binary with NSG
Ternaries
30
20
10
0
1 10 100
Particle Diameter (µm - log scale)
FIG 5 – Relationship between total sphalerite recovery and particle diameter (µm – log scale) for data
from the recovery-size level of analysis and from the recovery-size-liberation level of analysis illustrating
the recovery of the following sphalerite liberation categories in a sphalerite rougher – liberated, binaries
with galena, binaries with iron sulfide, binaries with non-sulfide gangue (NSG) and ternaries.
100
90
80
Copper Sulfide Recovery (%)
70 100
100‐80
60
80‐60
50
60‐40
40
40‐25
30 25‐15
20 15‐5
10 5‐0
0
1 10 100 1000
Particle Diameter (µm ‐ log scale)
FIG 6 – Relationship between copper sulfide rougher recovery and particle diameter (µm – log scale) for data at the
recovery-size-liberation level of analysis, showing the performance of a range of volume per cent classes for copper sulfide
(data from one rougher bank of the Phu Kham operation, provided with the kind permission of Pan Aust Limited).
assays, from modal analyses of the size fractions by an automated electron microscope or
by other means.
3. Calculation of the recovery of each mineral in each size fraction with respect to the
recalculated feed.
Steps 2 and 3 represent the calculations required for graphing the mineral recovery – particle
size (log scale) curves as discussed in earlier sections. The next level of analysis (liberation
level) is now discussed. An example in the literature with the same methodology can also be
reviewed (Johnson, 1987).
To provide an example of the type of calculations required for the liberation level of analysis,
some sphalerite liberation data for a single selected size fraction are presented for a system
with such data for the concentrate and tailing only. In Table 5, the calculations for one size
fraction of one mineral are illustrated.
The flow rate of sphalerite in the concentrate and tailing were 6188.2 and 1155.9 kg/h
respectively. These flow rates (column A) were multiplied by each liberation value (each
divided by 100) in the five columns labelled as B to distribute the mineral flow rates amongst
the various liberation categories (columns labelled as C).
The recalculated feed is obtained in the table by summation of the flow rates in each
liberation category. In a further step, the distribution of sphalerite between the concentrate
and tailing in each liberation category is calculated (see values in brackets). In other words,
the recovery of mineral in each liberation category has been calculated with respect to the
recalculated feed.
Although the calculations are straightforward, the presentation of the calculated quantities
is more difficult for the listed reasons:
•• there is interest in information on the feed, concentrate and tailing streams
•• there are many size fractions and minerals
•• there are many liberation categories of relevance.
For compactness in some examples of presentation methods, data are used for a bank
where only four size fractions encompass all the solid in the concentrate and tailing. The first
presentation uses tables only (see Table 6). This table provides the flow rates and recoveries
for all the minerals in the various liberation categories and size fractions in the concentrate.
This table is particularly valuable for recognition of the major sources of gangue dilution in
the concentrate.
Therefore, the dominant flow rates for the gangue minerals (>200 kg/h) in the concentrate
have been highlighted. In reading the flow rates for minerals in a typical binary in Table 6, the
TABLE 5
Example of processing of point count form of some sphalerite liberation data for one size fraction.
Product Mineral Liberation data (B) Mineral flow in categories (kg/h) (C)
flow Liberated % in binaries with Tern Liberated Binaries with Tern
(kg/h) category category
Galena Iron NSG Galena Iron NSG
(A)
sulfide sulfide
Concentrate 6188.2 75.6 5.1 12.2 2.4 4.6 4678.3 315.6 755.0 148.5 284.7
(97.0) (90.4) (73.0) (54.8) (33.2)
Tailing 1155.9 12.7 2.9 24.2 10.6 49.6 146.8 33.5 279.7 122.5 573.3
(3.0) (9.6) (27.0) (45.2) (66.8)
Recalculated 7344.1 65.7 4.8 14.1 3.7 11.7 4825.1 349.1 1034.7 271.0 858.0
feed (100.0) (100.0) (100.0) (100.0) (100.0)
Notes: Tern = category containing ternary and quaternary composites. NSG = non-sulfide gangue.
Recalculated feed = recalculated feed from summation of flows in concentrate and tailing as shown in columns labelled as (A) and (C).
Liberation values for each category in the row ‘Recalculated feed’ and in columns labelled as (B) calculated using the flow rates for the recalculated
feed in columns labelled as (C) and the total flow rate of sphalerite (7344.1 kg/h) in column (A).
TABLE 6
Listing of flow rates (kg/h) and recoveries for sphalerite, galena, iron sulfide and non-sulfide gangue
in the various liberation categories and size fractions for a zinc rougher. All gangue mineral flow rates
>200 kg/h have been highlighted to indicate the dominant gangue mineral flow rates.
Mineral Liberated Galena Sphalerite Iron sulfide NSG Ternary Size fraction*
(Flow in binary with)
Flow rate of mineral in all liberation categories in concentrate
Galena 0.6 - 61.1 0.0 108.3 117.8
Sphalerite 1199.3 858.5 - 281.8 180.2 756.9
+38 µm
Iron sulfide 43.4 0.0 117.8 - 0.0 256.4
NSG 84.5 65.2 297.6 71.3 - 362.8
Galena 3.9 - 61.8 10.8 25.5 127.8
Sphalerite 4678.3 315.6 - 755.0 148.5 284.7
-38 µm +11 µm
Iron sulfide 266.5 13.1 339.1 - 13.9 139.1
NSG 214.3 0.0 267.2 80.7 - 641.6
Galena 7.0 - 11.0 1.3 4.4 4.7
Sphalerite 1336.7 29.4 - 29.4 63.3 13.3
-11 µm +8 µm
Iron sulfide 116.9 3.5 46.4 - 18.9 0.0
NSG 143.9 12.4 87.5 43.7 - 0.0
Galena 56.5 - 89.3 10.3 35.6 37.9
Sphalerite 5978.3 131.7 - 131.7 283.1 59.3
-8 µm
Iron sulfide 575.3 17.4 228.3 - 93.2 0.0
NSG 1206.4 103.8 733.5 366.7 - 0.0
Recovery of mineral in all liberation categories in concentrate
Galena # - # 0.0 12.9 6.9
Sphalerite 80.5 92.3 - 39.1 9.9 19.9
+38 µm
Iron sulfide 1.6 0.0 13.5 - 0.0 5.0
NSG 0.2 2.8 12.7 0.3 - 2.6
Galena # - 79.3 20.0 12.9 29.4
Sphalerite 97.0 90.4 - 73.0 54.8 33.2
-38 µm +11 µm
Iron sulfide 3.3 10.5 31.0 - 1.0 11.1
NSG 0.9 0.0 33.1 1.9 - 29.6
Galena # - # # 9.1 24.2
Sphalerite 92.4 # - 66.5 63.3 44.1
-11 µm +8 µm
Iron sulfide 7.2 # 37.3 - 12.3 ##
NSG 2.5 28.1 58.0 8.2 - ##
Galena # - # # 5.3 15.1
Sphalerite 74.6 # - 32.5 29.4 16.0
-8 µm
Iron sulfide 5.0 # 29.0 - 8.8 ##
NSG 2.0 24.3 53.2 6.9 - ##
Notes: NSG = non-sulfide gangue; = gangue flow >200 kg/h; # = insufficient observations of this category in tailing for reliable value;
## = insufficient observations of this category in concentrate for reliable value. * Values are for sphalerite.
reader is reminded of the following example for the sphalerite/iron sulfide binary (+38 µm
fraction): 281.8 kg/h of sphalerite and 117.8 kg/h of iron sulfide. The table also provides the
recovery values for all the categories in the various size fractions.
The recovery values are particularly valuable for examination of metallurgical behaviour of
the various liberation categories of the valuable mineral in the concentrate. The recovery values
represent the metallurgical performance for the valuable mineral in the category. However, it
must be noted that the significance of a high or low recovery value for the valuable mineral in
a category depends also on its flow rate in the feed. For example, a category with a high flow
rate in the feed and a metallurgical recovery of 95 per cent may result in a much higher flow
rate of valuable mineral to the tailing than another category with a low flow rate in the feed
and a low metallurgical recovery of 20 per cent.
Some recovery values have not been entered in Table 6. In collecting data from a standard
number of particles, the number of observations in some liberation categories may be too
small to provide reliable recovery values. For example, galena was present in the tailing in
small amounts only and no observations existed for some liberation categories in the tailing,
implicating a recovery of 100 per cent. A second phase of data collection recording only
information on a selected mineral (in this case for galena) in some liberation categories is
required in such circumstances.
The graph of recovery-particle diameter (log scale) for the valuable mineral can be updated
with the additional liberation categories for the valuable mineral as shown in Figure 5, where
the curve for sphalerite contained six data points. With the grouping of size fractions, the
curves for the various liberation categories contained four size fractions in this example. It can
be noted that the curve for liberated sphalerite displayed higher recoveries than for the overall
sphalerite curve. Further, the points available for sphalerite-galena binaries were in a similar
region to the liberated sphalerite because both the galena and sphalerite were hydrophobic in
these binary particles. For the other liberation categories containing sphalerite in composites
with less hydrophobic unwanted minerals, lower recoveries were observed in general. Similar
patterns of behaviour are often observed in data sets of this type.
For examination of rougher data for the valuable mineral, particularly where the rougher
tailing is directed to the final tailing, an informative type of graph utilising volume per cent
classes for the valuable mineral is illustrated in Figure 6. This type of graph is based on data of
the type discussed earlier and shown in Table 4. Figure 6 illustrates clearly the deterioration
in the recovery of the copper sulfide-bearing particles for larger and lower quality composite
particles with respect to their volume per cent of copper sulfides.
Figures 7 and 8 demonstrate the use of simple effective bar charts for illustrating process
weaknesses for sized liberation data. Figure 7 illustrates the flows of a gangue mineral in
various occurrences (liberated and several types of composites) in a concentrate. Figure 8
illustrates the flows of a valuable mineral in various occurrences (liberated and several
types of composites) in a tailing. The use of simple effective methods in reporting and in
presentations is an important aspect of communicating the findings from data at the mineral
recovery-size-liberation level of analysis.
Graphical presentations can be extended to three-dimensional graphs (with axes of size
fraction / liberation category / measure of quantity) of the types listed:
1. distribution of a selected mineral in the feed, concentrate or tailing
2. flow rate of a selected mineral in the feed, concentrate or tailing
3000
Ternary
2500 Binary with Iron Sulfide Gangue
Non‐Sulfide Gangue Flow (kg/hr) Binary with Sphalerite
Binary with Galena
2000
Liberated
1500
1000
500
0
F7 F6 Size Fraction ‐38um +F5 +38um
FIG 7 – Simple and clear method for illustrating the flows of one gangue mineral (non-sulfide gangue) in various
occurrences in various size fractions of a concentrate (F7 = -10 µm, F6 = -14 +10 µm and -38 µm +F5 = -38 +14 µm).
6000
Ternary
5000 Binary with Non‐Sulfide Gangue
Binary with Iron Sulfide
Binary with Galena
4000
Sphalerite Flow (kg/hr)
Liberated
3000
2000
1000
0
F7 F6 Size Fraction ‐38um +F5 +38um
FIG 8 – Simple and clear method for illustrating the flows of the valuable mineral (sphalerite) in various occurrences
in various size fractions of a tailing (F7 = -8 µm, F6 = -11 +8 µm and -38 µm +F5 = -38 +11 µm).
3. recovery values for a selected mineral (often the valuable mineral) in the concentrate or
tailing.
The types of graphs in item 1 are demonstrated in Figure 9 because their structure is difficult
to explain without an actual demonstration. However, the graph types in items 2 and 3
are not demonstrated because their structure is straightforward. The graph type in item 2
is suitable for demonstrating how each gangue mineral is diluting a concentrate and also
for demonstrating the flow rate of the valuable mineral in each category and size fraction
A
100
Recalculated Zinc Rougher Feed 90
80
Distribution of Sphalerite in
70
60
50
40
30
20 Total
Ternaries
10 Sphalerite in Binaries with NSG
Sphalerite in Binaries with Iron Sulfide
0 Sphalerite in Binaries with Galena
Liberated Sphalerite
Size Fraction
B
100
90
Distribution of Sphalerite in Zinc
80
Rougher Concentrate (with
70
60
Rougher Feed)
50
40
30
20 Total
Ternaries
10 Sphalerite in Binaries with NSG
Sphalerite in Binaries with Iron Sulfide
0 Sphalerite in Binaries with Galena
Liberated Sphalerite
Size Fraction
C
100
90
Recalculated Zinc Rougher Feed)
Distribution of Sphalerite in Zinc
Rougher Tailing (with respect to
80
70
60
50
40
30
20 Total
Ternaries
10 Sphalerite in Binaries with NSG
Sphalerite in Binaries with Iron Sulfide
0 Sphalerite in Binaries with Galena
Liberated Sphalerite
Size Fraction
FIG 9 – Examples of usage of three-dimensional graphs with axes of size
fraction/liberation: (A) recalculated feed, (B) concentrate, (C) tailing.
in the tailing. It can be noted that items 2 and 3 effectively represent a means of graphical
presentation of the type of data presented in tabular form in Table 6.
For non-sulfide and sulfide gangue minerals in each size fraction, liberation data can be
used to determine the contribution from both liberated particles and composites (often with
the valuable minerals) to their recovery. For fine-grained ores, recovery of non-sulfide and
gangue minerals in composites can continue to be significant in the fine fractions. For non-
sulfide gangue minerals, the data for the liberated form in each size fraction is the relevant
form for analysis of the contribution from entrainment. For sulfide gangue minerals and
certain non-sulfide gangue minerals such as talc, the recovery values for the liberated form in
each size fraction will be from entrainment, which can be estimated from the water recovery
and ENTi values, and any additional contribution from its recovery due to its hydrophobicity.
In some unusual systems, the values for ENTi for the liberated non-sulfide gangue can be
greater than 1.0 for the finest fraction and continue to exceed greatly the expected values for
ENTi in the intermediate and coarse size fractions. Entrapment is the term for this additional
mechanism ‘beyond entrainment’, which can arise from steric hindrance in the regions in
the froth phase containing solid particles and water, causing restrictions on the drainage of
liberated gangue. One example from Vianna (2004) is discussed in Johnson (2005).
Liberation data can also be used to identify the extent of liberation achieved at each grinding
or regrinding stage at the commencement or within a processing circuit. This information
can be obtained by including samples of the combined feed and product for a grinding or
regrinding circuit; such use of liberation data is extremely informative but is rarely used in
the industry. Measurement of the liberation values for the minerals in each size fraction of
the feed and product allows the overall liberation value for minerals in each stream to be
calculated. For each mineral, the increase in liberation across the grinding circuit is calculated
by difference. The method for calculating the liberation value within a given stream is
illustrated in Table 7.
The taking of a few extra samples in a plant survey may allow more reliable quantification
of the changes in liberation at size reduction steps. For example, to improve the quality of
the data, it is preferable that a single combined feed sample and a single product sample
be taken for an overall regrinding system, which often will contain a mill for size reduction
and a classification device. It is advisable that a survey plan is reviewed for the directness by
which liberation data may ultimately be obtained. In other words, the technical viability of a
‘liberation survey’ within the larger plant or circuit survey needs to be addressed separately
in the planning steps. Equivalent samples can be taken in pilot plant work and in laboratory
batch or cycle tests involving regrinding.
Historically, metallurgists have often sought to determine the benefits of regrinding by
examination of the separation results (often using the flotation process) with various levels
of regrinding. Particularly for regrinding to fine sizes, it is much safer to establish initially
if the regrinding step is causing a significant increase in the liberation level of the valuable
mineral and the gangue minerals. If a significant increase in liberation is measured, the necessary
conviction is provided to seek conversion of the increase in liberation into an increase in performance of
the following separation. Such increases in performance may not emerge after initial cursory separation
tests, even if a significant improvement in liberation is resulting from the regrinding. Optimisation of
the chemical conditions during regrinding and the following flotation separation may require many
months of tests for some systems.
The data from two-dimensional mounts may overestimate the levels of liberation of each
mineral because a two-dimensional intersection of a particle may, by chance, occur in one
mineral only, even if there is more than one mineral in the particle (Jones, 1987). This issue
is well known from the area of stereology. From the examination of liberated and composite
particles from real ores in the current literature on this stereological issue, there is no need
TABLE 7
Calculation method for the total liberation of a mineral in a given stream.
Column 1 Column 2 Column 3
Size Mineral flow Liberation value Flow of liberated mineral
-105 µm +53 µm 5 20 1
-53 µm +CS2 8 25 2
-CS2 +CS3 10 40 4
-CS3 +CS4 10 60 6
-CS4 +CS5 10 75 7.5
-CS5 +CS6 20 80 16
-CS6 +CS7 37 80* 29.6
Total 100 → A 66.1 → B
Input information:
Column 1 – Flow of mineral in each size fraction (from size distribution and assays)
Column 2 – Liberation value for mineral in size fraction
The liberation value is calculated as:
B # 100
A
ie 66.1 # 100 = 66.1%
100
This is simply the weighted average of the liberation values in Column 2.
Note: in this example, observations could not be made on the CS7 fraction. The liberation value for this size
fraction (denoted by *) was assumed to equal the value for the C6 fraction because the liberation values had
essentially reached a plateau region.
for a stereological correction of 2D liberation data, or only a minor need. This finding applies
to real ores, which produce complex composite particles with a wide range of compositions
from very low quality to high quality with respect to the valuable mineral, and which produce
liberated particles of the valuable mineral along with barren particles containing no valuable
mineral. Hence, uncorrected two-dimensional data may be used with caution and with other
checks to improve understanding of the process for the listed reasons.
Firstly, in calculation of the recovery values for liberation-based species in Table 6, there may
be correction factors (a and b) needed to the flow rates in the numerator and denominator of
the following equation:
In percentage terms, the magnitude of the correction factor (slightly less than or equal to 1)
for the tailing flow rate (b) may differ from the correction factor for the concentrate flow rate
(a), ie b does not equal a. However, because the direction of the correction must be the same for
both the numerator and denominator, there is at least partial cancellation of the stereological
effect. When the flow rate for the concentrate in the equation is very large compared to the
flow rate for the tailing, the cancellation of the stereological effect is more complete.
Secondly, if a metallurgist is suspicious of the flow rates in Table 6 provided by uncorrected
liberation data, it is possible with care and patience to section a given set of particles in a
briquette at various levels to obtain, for practical purposes, three-dimensional liberation data.
Note that this step is a possibility for particularly unusual circumstances or when very high
confidence is required in the liberation data.
Thirdly, the liberation level (two-dimensional basis) for the valuable mineral(s) in a process
feed is used to judge if the sizing of the feed is at a value where an acceptable separation
could be expected. On a two-dimensional basis, the guidelines shown in Table 8 can be used
in process engineering where all the liberation is occurring in a grinding circuit at the start of
the circuit, ie no regrinding exists inside the flotation circuit. With liberation data converted
from 2D to 3D values, a new set of slightly different boundaries may exist.
Of course, obtaining a high level of liberation does not guarantee an efficient separation as
the appropriate settings for the process variables must also be obtained. For an efficient froth
flotation separation, both a high level of liberation and suitable settings for the physical and
chemical variables in the process are needed.
In porphyry copper circuits, liberation levels for the copper minerals may sometimes be less
than 50 per cent in the rougher feed. However, the regrinding steps in the circuit must raise
the liberation level of the copper minerals such that the guidelines in Table 8 are achieved for
the recalculated feed (summation of the final concentrate(s) and tailing(s)). This applies to
other circuits with major regrinding and liberation steps in the flotation circuit.
For an ore with multiple valuable minerals reporting to a single concentrate, it is necessary
to consider the liberation of the combined valuable minerals in the feed or recalculated feed in
conjunction with the guidelines in Table 8. For example, for an ore containing several copper
sulfide minerals, the liberation of the combined copper sulfides should be calculated because
its value will exceed the values for the individual copper sulfides; a particle containing only
chalcopyrite and bornite is a liberated copper sulfide particle.
The use of liberation data in conjunction with recovery-size data can be employed in
existing plants for surveys, monthly or yearly composite samples on feed and products, or
for samples from supporting laboratory or pilot plant development work. Plotting of the level
of liberation of the valuable mineral(s) in the recalculated feed of a concentrator versus time
(monthly or quarterly) has been demonstrated to be very useful in understanding declines
and monitoring improvements in plant performance (Young et al, 1997). Equally, for a new
deposit, the technique can be applied to laboratory test work and pilot plant surveys as a part
of flow sheet development and optimisation.
THE ROLE OF SURFACE AND SOLUTION ANALYSES FOR ORGANIC AND INORGANIC SPECIES
The use of recovery-size data in conjunction with liberation data will sometimes reveal
the recovery of a liberated gangue mineral in quantities that indicate it was recovered due
TABLE 8
The maximum potential quality of separation possible at various liberation levels (two-dimensional data)
for the flotation feed (circuit without regrinding) or the recalculated feed (circuit with regrinding).
Liberation level (%) (2D data) Potential quality of separation
>90% Very high, extremely efficient separation possible
80% to 90% High, very efficient separation possible
70 to 80% Sound, moderate separation possible
<70% Poor, inadequate separation only possible
The froth recovery Rf of a mineral in a size fraction is defined as the ‘fraction of particles
attached to bubbles entering the froth phase that are collected in the concentrate launder’
(Vianna, 2004) and this term is included in the modelling framework for the first-order rate
constant (k) for the overall process (combined pulp and froth regions) as used at the Julius
Kruttschnitt Mineral Research Centre:
k = P Sb Rf
where:
P = mineral floatability in pulp region
Sb = bubble surface area flux
Rf = froth recovery
The key additional step requires a device for capturing and measuring the upward flow
rate of minerals in particles attached to bubbles at the top of the pulp zone, just before the
bubbles enter the froth region; other related data gathering steps are required to obtain the
froth recovery for a given cell in a plant. A plant metallurgist can now request froth recovery
measurements on all or selected cells in a plant as part of a plant survey.
The obtaining of the froth recovery across the froth region is relevant for both conventional
flotation machines (including tank cells) and cells with water addition to the froth region.
Novel measuring devices (Seaman, 2006) for determining the transfer or flow rate of
attached particles at the top of the pulp region at steady state during the survey have the
advantage that there is no disruption to the plant. Some other methods require alteration of
the froth height after the survey and collection of data for the cell in question at each froth
height. Methods that require such disruption to the plant are also time consuming and are
much less practical.
Froth stability and therefore froth recovery (Rf) are believed to depend on factors such as
frother type and concentration and particle properties such as size, number, hydrophobicity
and shape. The role of particle hydrophobicity is of particular interest because two recent
studies have shown there is a maximum in the relationship between froth recovery and
hydrophobicity (Ata, Ahmed and Jameson, 2003; Schwarz and Grano, 2005). These findings
corroborate broadly with earlier more fundamental findings reported by Dippenaar
(1982). While increases in hydrophobicity assist the subprocesses of adhesion and stable
transportation in the pulp zone, such increases may only assist in the froth region up to an
intermediate hydrophobicity, indicated by a contact angle in the region 60 to 70 degrees.
Cumulative Grade
of Valuables (%)
of Sulfide Gangue (%)
Cumulative Recovery
of Non-sulfide Gangue (%)
Cumulative Recovery
FIG 10 – Analysis of down-the-bank flotation data for a single rougher or cleaner bank using
the grade recovery curve for the valuables and selectivity curves for each gangue mineral.
1 2 3 4 5 6 7 8
Rougher First Cleaner First Cleaner A
Concentrate Tailing
A
B
8 7 6 5 4 3 2 1 C
Second Cleaner
D
1 2 3 4 5 6 7 8
Major Water Addition Third Cleaner
(for dilution cleaning)
45
1 B
40 1-3
Cumulative Grade of Valuable Element (%)
1-2 1-5
35 1-4 1-7 Combined Third Cleaner Concentrate
1 1-6 1-8
30 1-3
1-2 1-5
25 1 1-4 1-6
1-7 Combined Second Cleaner Concentrate
20 1-2
1-3 1-4 1-5 1-8
15 1-6 Combined First Cleaner Concentrate
1-7
10 1-8
5
Rougher Concentrate
0
0.00 20.00 40.00 60.00 80.00 100.00 120.00
Cumulative Recovery of Valuable Element (%)
45
1
1-3
C
40
1-5
1-2
Cumulative Grade of Valuable Element (%)
1-4 1-6
35
1-7 Combined Third Cleaner Concentrate
1
1-8
30 1-3 1-4
1-2 1-5 Combined Second Cleaner Concentrate
25 1-7
1-6
1
Second Cleaner Feed (C+D)
20 1-8
1-2 1-3 1-6
1-4 1-5
15
1-7 Combined First Cleaner Concentrate (C)
1-8
10
5
First Cleaner Feed (A+B)
FIG 11 – Examination of down-the-bank flotation data for three stages of closed-circuit dilution cleaning using grade recovery curves
for the valuable element using a straightforward method and a more advanced method. (A) Three-stage closed circuit cleaning system
referred to in (B) and (C). (B) Grade recovery curve of valuable element in cleaning system. (C) More complete grade recovery curves
for valuable element in cleaning system (providing additional information on magnitude and grade of circulating loads of valuables).
(i) Batch flotation process (ii) Series of perfectly mixed plant cells
Time (min)
Case 1 Nominal Residence Time (min)
0 5 10 15 20 25 30 35 40 45 0 5 10 15 20 25 30 35 40 45
1 1
Case 1 Case 1
Fraction remaining
Fraction remaining
0.1 0.1
Case 1 Case 1
0.01 0.01
Time (min)
Cases 2 and 3 Nominal Residence Time (min)
0 5 10 15 20 25 30 35 40 45 0 5 10 15 20 25 30 35 40 45
1 1
Case 2 Case 2
Case 3 Case 3
Fraction remaining
Fraction remaining
Case 2 Case 2
0.1 0.1
Case 3 Case 3
0.01 0.01
FIG 12 – Demonstration of various relationships between fraction remaining for the valuable
mineral (log scale) and (i) time (min) for a batch process and (ii) nominal residence time (min)
for a series of perfectly mixed plant cells and with various rate constant components.
where:
kf, ks, km = first order rate constant for high, low and medium rate constant components
(min-1)
ϕs, ϕm = fraction of component with low or medium rate constant
t = time for batch operation
λ = nominal residence time for each of the N continuous cells in series for continuous
operation
The behaviour of each mineral may be expressed in kinetic terms where the standard graph
(Figure 12) is the fraction or percentage remaining in the bank (log scale) on the y-axis and
nominal residence time (average residence time per cell multiplied by the number of cells to
that point in the bank) on the x-axis for plant data. For batch laboratory tests, the cumulative
flotation time is employed on the x-axis. Such relationships for the batch and plant cases
allow recognition of the kinetic components, which describe the behaviour of each mineral
in the bank. The number of kinetic components, the first-order rate constant of each and the
proportion of the mineral with that rate constant value can be determined. Such graphs can
be on an overall, unsized basis or they can be used for minerals or elements in size fractions.
It must be emphasised that the rate constant components can only be determined from
data with progressive values for fraction remaining and residence time. If the relationship
between mineral recovery and particle diameter (log scale) has been determined along with
the residence time at the end of a bank or a batch flotation stage, without progressive data on
reaching the overall performance for the stage, a single rate constant value can be calculated
for that mineral. The rate constant calculated by this method will represent a weighted average
of more than one rate constant component in many cases. While the single rate constant value
can be useful for some purposes, its limitations must also be recognised.
Some examples are provided in Figure 12. In the simplest case in Figure 12, the graph of
fraction remaining (log scale) versus time for a single rate constant is depicted, providing a
straight-line relationship. This case is followed by an example where the kinetic behaviour
with two, quite different values (high and low) for the rate constant is depicted. In this case,
two components existed for the mineral. It can be noted that a closely related case with two
components exists when the low rate constant is a non-recoverable component with a rate
constant of zero min-1 (this case is not illustrated in Figure 12). A third case is also depicted in
Figure 12 with three components, ie three quite different values (high, low and medium) for
the rate constant.
By comparing the batch and continuous cell examples (left and right-hand side of Figure 12,
respectively), it can be seen that a higher value for the fraction remaining (ie lower recovery)
exists at a given residence time for a given set of rate constants for the plant cells. By comparing
cases 2 and 3 and noting the apparent similarity of the fraction remaining (log scale) and
residence time relationships, it can be surmised that there may be difficulty in accepting the
existence of the third component, ie a fit deemed acceptable may still be obtained using the
equation in the form with two rate constant components.
Valuable liberated mineral with a medium rate constant will be recovered in a rougher
section with adequate installed residence time. The diminished rate constant component of
the valuables (medium or even lower rate constant) may reflect solely the deficiencies of
fine particles in the collision subprocess, or it may reflect an additional issue arising from an
imbalance in hydrophilic and hydrophobic species on the surface of the component. Such
liberated valuable mineral would be expected to report to the final concentrate on the basis
of its liberation state. However, the medium rate constant valuable mineral may have more
difficulty in being sufficiently competitive during conventional cleaning for the majority of it
to report to a high-grade final concentrate.
The design of laboratory batch rougher or cleaner tests with collection of a series of
concentrates also allows collection of traditional cumulative concentrate grade/cumulative
concentrate recovery curves and kinetic data as just described for a bank in a plant. The
usefulness of such data depends on the times selected for collection of each concentrate in a
laboratory batch test.
For both laboratory and plant processes, it is advisable to have many data points at the
commencement of a stage to allow checking for anomalous behaviour because the process was
galena 50
grade % galena grade /recovery
40
4
30
10
10
mnulative flotation times ��
20
10
galena recovery %
□---
sphalerifl! 80
recovery¾
60
sp"11erite selectivity
/
/;if
40
0/
20 /
galena recovery %
allowed to commence before the feed was fully prepared. It is also often useful to survey the
last cell individually to obtain directly the concentrate grade for the final cell. For a laboratory
batch test, a separate concentrate for a short time at the end of the test is the equivalent.
Experience with surveying improves a metallurgist’s judgement on the appropriate level of
detail in down-the-bank data collection.
Down-the-bank data can be collected for a series of bank surveys or laboratory tests with
changed conditions for each test. An example is given in Figure 13 (Grano, Ralston and
Johnson, 1988), which applies the graphing method in Figure 10. The following points can be
made in establishing the changes that occurred in the performance of each mineral for each
survey or laboratory test:
•• A graph of cumulative grade/cumulative recovery for the valuable element or mineral
provides a summary of the performance of the bank and indicates the trade-off between
grade and recovery in each portion of the bank. The graph does not allow analysis of
changes in the behaviour of the unwanted minerals when a series of tests is being compared.
•• A convenient method for analysis of changes in behaviour of the unwanted minerals
is the use of selectivity curves. There is more than one method for plotting selectivity
curves. A common feature of all methods is that the recovery of each unwanted mineral
(or an equivalent measure) is plotted against the recovery of the valuable mineral in a
series of graphs.
•• From the use of selectivity curves of the type illustrated in Figures 10 and 13, changes in
behaviour of the other minerals relative to the valuable mineral can be easily recognised.
These changes in conjunction with changes in the flotation rate constant of the valuable
mineral itself can be used to understand the origins of changes in position of a series of
grade-recovery curves for the valuable mineral.
•• The selectivity curves shown in Figure 13 are legitimate for convenient comparison of the
recovery of each gangue mineral at a given recovery for the valuable mineral, the primary
purpose of the described analysis. (However, for considering a single test in isolation, this
type of selectivity curve is not legitimate for evaluating changes in selectivity between the
valuable mineral and gangue mineral under examination from the commencement to the
end of the test. For this purpose on a single test, the logarithm of the recovery value for
both the valuable mineral and the gangue mineral in question must be plotted. A straight-
line relationship indicates no change in selectivity along the bank or during the batch test.)
The set of curves in Figure 13 shows that the selectivity of sphalerite relative to galena was
improved clearly with the use of zinc sulfate. The reagent also caused a slight decrease in
selectivity of iron sulfide and non-sulfide gangue relative to the galena. From Figure 13 alone,
no discussion is possible on changes in selectivity between the galena and each unwanted
mineral during each test.
by summing the liberation information from all size fractions in the feed. Some laboratories
use the measured two-dimensional liberation data to construct the theoretical grade recovery
curve; other laboratories use three-dimensional liberation data from conversion of the two-
dimensional liberation data. Comparisons of the observed and theoretical performance for
each size fraction can also be obtained.
The physical meaning of key points on the theoretical grade recovery curve is provided in
Box 5.
For the feed information, liberation data for a new orebody or deposit, or the feed to a plant
or a bank in a plant could be used. In some cases, the liberation data for a recalculated feed
Consider the listed particles as separator feed (note: darkened mineral is the valuable mineral – V):
Liberated (group 1–40% of V) High quality composites (group 2–30% of V)
Medium quality composites (group 3–20% of V) Low quality composites (group 4–10% of V)
B C B C
CUMULATIVE GRADE of valuable
90
CONCENTRATE 90
CONCENTRATE
D GRADE (100%) D GRADE (100%)
mineral in concentrate (%)
80 80
70 70
THEORETICAL GRADE-RECOVERY
60
E 60
CURVE FOR VALUABLE MINERAL E
50
THEORETICAL GRADE-RECOVERY 50
30 30
20 20
THEORETICAL GRADE-RECOVERY
A FEED GRADE OF CURVE FOR VALUABLE ELEMENT A FEED GRADE OF
10 VALUABLE 10 VALUABLE
MINERAL (10%) 0
MINERAL (10%)
0
0 10 20 30 40 50 60 70 80 90 100 0 10 20 30 40 50 60 70 80 90 100
CUMULATIVE RECOVERY of valuable mineral in concentrate (%) CUMULATIVE RECOVERY of valuable mineral/element in
concentrate (%)
group 1 group 1
group 1+2 group 1+2
group 1+2+3 group 1+2+3
group 1+2+3+4 group 1+2+3+4
BOX 5 – Illustration of the concepts in construction of a theoretical grade recovery curve for separation of a valuable mineral
(dark phase) based on the group of particles in a size fraction of the feed to the separation (step 1). The theoretical grade recovery
curve for the valuable element in the valuable mineral is included (step 2) based on the stoichiometry of the valuable mineral.
must be used to calculate the relevant theoretical performance curve. This situation would
arise for a number of reasons:
•• for a bank, liberation data on the concentrate(s) and tailing only were obtained for a bank,
possibly to reduce costs
•• for a total plant or section of a plant, grinding or regrinding existed within the section
and hence a ‘feed’ sample would not represent the total achieved liberation, unlike the
summed concentrate(s) and tailing (ie the recalculated feed), which would represent the
total achieved liberation
•• for a laboratory batch separation, only the concentrate(s) and tailing exist for that particular
sample tested.
As a further step, bands can be indicated between the actual and theoretical performance
curves to indicate the various sources of inefficiency from unwanted minerals. Examples are
the contribution of each liberated sulfide gangue mineral, each liberated non-sulfide gangue
mineral and composite particles containing no valuable mineral (all mentioned categories
contain no valuable mineral and were not included in construction of the theoretical grade
recovery curve) to creation of the difference in position for both the actual and theoretical
performance curves. The non-recovery of liberated valuable mineral can be another
contributor, by ‘pushing’ the observed grade recovery curve to the left.
ACKNOWLEDGEMENT
The important contribution of Suzanne Munro (Mineralurgy Pty Ltd/Mineralis) in
preparation of diagrams and of other aspects for the content in this chapter is acknowledged
with gratitude.
REFERENCES
Ata, S, Ahmed, N and Jameson, G, 2003. A study of bubble coalescence in flotation froths, Int J Miner
Process, 72:255–266.
Croxford, N J W, Draper, N and Harraway, D H, 1961. Some aspects of the carbonaceous fraction of
Mount Isa lead concentrate, The AusIMM Proceedings, 197:149–161.
Dippenaar, A, 1982. The destabilization of froth by solids, 1. The mechanism of film rupture, Int J Miner
Process, 9:1–14.
Grano, S R, Ralston, J and Johnson, N W, 1988. Characterization and treatment of heavy medium
slimes in the Mount Isa Mines lead-zinc concentrator, Minerals Engineering, 1(2):137–150.
Johnson, N W, 1972. The flotation behaviour of some chalcopyrite ores, PhD thesis, The University of
Queensland.
Johnson, N W, 1987. Application of kinetics and liberation data to analysis of an industrial flotation
process, CIM Bulletin, 80(899):113–117.
Johnson, N W, 2005. A review of the entrainment mechanism and its modelling in industrial flotation
processes, in Proceedings Centenary of Flotation Symposium, pp 487–496 (The Australasian Institute of
Mining and Metallurgy: Melbourne).
Johnson, N W, 2006. Liberated 0 - 10 µm particles from sulphide ores, their production and separation
– recent developments and future needs, Minerals Engineering, 19:666–674.
Johnson, N W and Jowett, A, 1982. Comparison of lead primary rougher behaviour of several Mount
Isa lead zinc ores, in Proceedings Mill Operators’ Conference, pp 271–285 (The Australasian Institute of
Mining and Metallurgy: Melbourne).
Jones, M P, 1987. Applied Mineralogy – A Quantitative Approach (Graham and Trotman: London).
Lynch, A J, Johnson, N W, Manlapig, E V and Thorne, C G, 1981. Mineral and Coal Flotation Circuits –
Their Simulation and Control (Elsevier: Amsterdam).
Pyke, B, Fornasiero, D and Ralston, J, 2003. Bubble particle heterocoagulation under turbulent
conditions, J Colloid and Interfacial Science, 265:141–151.
Restarick, C J, 1976. Pulp sampling techniques for steady state assessment of mineral concentrators, in
Proceedings Sampling Symposium, pp 161–168 (The Australasian Institute of Mining and Metallurgy:
Melbourne).
Rumball, J A and Richmond, G D, 1996. Measurement of oxidation in a base metal flotation circuit by
selective leaching with EDTA, Int J Miner Process, 48:1–20.
Schwarz, S and Grano, S, 2005. Effect of particle hydrophobicity on particle and water transport across
a flotation froth, Colloids and Surfaces, A: Physiochem Eng Aspects, 256:157–164.
Seaman, D, 2006. Selective transport of attached particles across the froth phase, PhD thesis, The
University of Queensland.
Vianna, S M, 2004. The effect of particle size, collector coverage and liberation on the floatability of
galena particles in an ore, PhD thesis, The University of Queensland.
Young, M F, Pease, J D, Johnson, N W and Munro, P D, 1997. Developments in milling practice at the
lead/zinc concentrator of Mount Isa Mines Limited from 1990, in Proceedings Sixth Mill Operators’
Conference, pp 3–12 (The Australasian Institute of Mining and Metallurgy: Melbourne).
TABLE A1
Guidance for selection of size range factor (g).
Description of size range Value for d/d' Resulting value for g
Broad >4 0.25
Medium breadth 2 to 4 0.50
Narrow <2 0.75
Uniform 1 1.00
The calculation can be repeated for coarser streams with the same particle properties. For
streams with top sizes (d) of 0.2, 2.0 and 20 cm associated with grinding and crushing circuits,
the calculated values for ms increase to 23 g, 2269 g and 226 889 g respectively (assuming d1
is 0.02 cm and letting parameter b equal 1.0). For the coarser solid in grinding circuits and
particularly in crushing circuits, care is therefore required such that sufficient sample mass
is obtained to ensure a given fundamental error is not exceeded. For each stream in crushing
and grinding circuits, the appropriate calculations for the sample size must be performed
using the real properties of the particles in the population being sampled.
To comply with sample correctness as discussed earlier, one aspect is that the width of the
parallel opening of the sampler must be commensurate with the top size of the particles in
the stream. Large particles must not have a lower probability than smaller particles entering
the sample because of interference with their motion. Standards exist for the key dimensions
of the samplers in such cases. Further, the samplers must be emptied in a way that ensures all
particles irrespective of their size or density are removed from the sampling device or system
and continue to be part of the sample.
The equation for ms can be expressed with the fundamental error as the unknown. Holmes
(1992) noted that:
… a fundamental characteristic of s2FSE is that it diminishes very quickly when d is reduced
and not so quickly when ms is increased. However, it can never be eliminated no matter what
crushing and homogenisation procedures are used, although for fine flotation concentrates it is
negligible when the sample mass exceeds about 100 g.
In collection of samples from a flotation circuit with several rounds of sampling and using
conventional types of samplers, the resulting sample mass is usually unavoidably in the
region of 500 g (or larger). Hence, as dictated by the use of conventional sampling devices in a
flotation circuit, the sample mass tends to be in a region where the value for the fundamental
error is at acceptable low values.
To obtain an assay for a sample or a size fraction, a very small mass (typically 0.25 g to 1 g)
has to be sampled from a much larger mass of dry solid in the laboratory. The fundamental
error for this step can also be calculated using the equation from Gy (1982) as described earlier.
If the fundamental error is unacceptably high, the sample or size fraction can be pulverised to
lower the top size, and greatly lower the mass of sample providing an acceptable fundamental
error. Of course, if a property that is affected by size reduction (eg liberation state) is to be
measured, this approach cannot be taken.
For a particular operation, it is clearly more effective to obtain a plant feed sample from the
flotation circuit feed, rather than the coarser grinding circuit feed or the even coarser crushing
circuit feed. For a given sampling error, the minimum mass of the required sample is much
smaller for the flotation circuit feed and cost savings result from handling of the smaller
sample, aside from the convenience. In collecting a sample, a large number of frequent
correctly executed sampling events with collection of the same small sample weight in each
sampling event is superior to a small number of infrequent correctly executed sampling
events, which produces the same final sample weight.
Some additional sources of information on sampling principles can be accessed (Weiss,
1985; Birnbaum, 1992).
LIP SAMPLING
The remainder of the appendix describes the lip sampling procedure, which provides various
types of data:
1. a reliable sizing and assay of the solid being recovered at the lip
2. a reliable value for the per cent solid of the pulp being recovered at the lip
3. an estimate of the solid flow rate passing over the lip.
To satisfy requirements 1 and 2, the sampler should be moved at right angles to the lip
across the full width of the cell lip at constant speed to intercept the discharging froth. The
speed should be selected to ensure that overflowing of the sampler is not possible and to
ensure that uninhibited entry of the discharging froth into the sampler is possible during
the sampling. Such lip samplers have typically the following dimensions: width 9 cm, height
30 cm and length 20 cm. The method was described by Restarick (1976).
To satisfy requirement 3, the time for which the sampler was intercepting the stream has to
be recorded. Therefore, the lip sampling procedure is performed using a crew of two people
with one recording the sampling times. The following points can be noted:
•• The requirements for the starting and finishing points in the movement of the lip sampler
are illustrated in Figure A1. These starting and finishing points arise because of the
requirement for determination of the solid flow rate per unit width of the sampler as
expressed in Equation A1:
Total weight (g) 3600 Lip width (mm)
TPH Solid = # # (A1)
Total time (sec) 106 Cutter width (mm)
That is, the full width of the sampler must be collecting sample during all the timing
period. (This requirement differs subtly from the correct sampling motion of starting and
finishing the sampler’s motion outside the flowing concentrate.)
•• The time taken for the sampler to intercept the concentrate stream has to be approximately
the same for each traverse to obtain reliable estimates of the solid flow rate.
•• For a group of cells being sampled together, the time taken for the sampler to intercept the
concentrate stream for each cell also has to be approximately the same for each traverse
Lip Width
Cell
Starting Finishing
Position Position
Cutter Width
FIG A1 – Illustration of the lip sampling technique as described in the appendix for one cell. The lip
sampler is positioned under the lip from which concentrate is discharging in this plan view.
of each cell. Initially, all the cells in the grouping have to be considered in selection of the
required time for the group of cells.
For obtaining the solid flow rate, the details of the correct lip sampling procedure for a
single cell are illustrated in Figure A1, and Equation A1 for conversion of the measurements
of the total mass of collected solid and the total sampling time (for all traverses on the various
rounds of sampling) was provided earlier. The total mass of collected solid divided by
the total sampling time is effectively the observed flow rate of solid per unit width of the
sampler. Movement of the sampler across the cell width at fixed velocity allows this value
to be averaged across the entire concentrate flow. In Equation A1, the ratio lip width/cutter
width then allows scaling up of the observed flow rate of solid per unit width of the sampler
to the full width of the cell lip.
To this point, all the discussion of lip sampling to obtain the solid flow rate has been based
on movement of the lip sampler across the cell lip at right-angles at fixed speed as indicated
in Figure A1. This movement of the sampler addresses any variations in concentrate assay
and flow rate across the full width of the lip. For a cell in good mechanical condition, such
variations are usually minimal.
However, given the large size and shape of some cells, the existence of floor grating over
the top of the cells, the design and maintenance of the cell lips or the existence of water pipes/
sprays or other obstacles in the launder, it is not possible to move the lip sampler across the
entire lip width or a portion of the lip. In this situation, an estimate of the observed flow rate
of solid per unit width of the sampler can be obtained by placing the sampler in one or more
fixed positions where a portion of the cell lip is accessible. The observed value can then be
scaled up to the full length of the cell lip.
Large tank cells can be one example of the described situation for several reasons:
•• the existence of relatively inaccessible external and internal launders and sometimes cross
launders
•• the existence of floor grating over the top of cells with possibly some small trapdoors for
access.
For sampling a single cell, the discharge of concentrate into the various types of launders
may have to be treated as separate but parallel sampling steps for that cell. It is also worth
noting that, depending on the overall sampling scheme, the objective from sampling a single
unit cell may be a sample of the combined concentrate for analysis without the need for an
estimate of the solid flow rate. While such an objective appears less onerous, timing of the
concentrate collection with the sampler in various fixed positions may still be required.
For example, taking a cell with one internal launder and one external launder, if the
concentrate discharging to the internal launder happened to be at a higher flow rate per unit
length of the lip than for the external launder, its assay may differ from the concentrate at an
external launder for the same cell. To obtain the assay of the combined concentrate for the
entire cell, two approaches could be taken:
1. The same lip sampling time for the concentrate discharging into the internal and external
launders could be used as the basis for a concentrate sample representing the entire
concentrate.
2. The assays for the solid at the internal and external launders along with the flow rate
estimate of each could be used to calculate the assay for the combined concentrate.
(Clearly, it is preferable that the plant designer provides a sampling point to enable
sampling of the entire concentrate from all launders at a single point, to obtain the
overall assay of the stream.)
It can be noted that timing and recording of the lip sampling times would assist in correct
execution of approach 1 even though obtaining a value for the lip tonnage is not the objective.
Further, timing and recording of the lip sampling times at the internal and external launders
is an integral part of execution of approach 2.
REFERENCES
Birnbaum, P M, 1992. An evaluation of sampling errors in a mineral concentrator, in Proceedings Sampling
Practices in the Mineral Industry, pp 49–58 (The Australasian Institute of Mining and Metallurgy:
Melbourne).
François-Bongarçon, D and Gy, P, 2002a. Critical aspects of sampling in mills and plants: a guide to
understanding sampling audits, J S Afr Inst Min Metal, Nov/Dec, pp 481–484.
François-Bongarçon, D and Gy, P, 2002b. The most common error in applying ‘Gy’s Formula’ in the
theory of mineral sampling, and the history of the liberation factor, J S Afr Inst Min Metal, Nov/Dec,
pp 475–479.
Gy, P M, 1982. Sampling of Particulate Materials – Theory and Practice, second edition (Elsevier: Amsterdam).
Holmes, R J, 1992. Sampling of mineral process streams, in Proceedings Sampling Practices in the Mineral
Industry, pp 33-37 (The Australasian Institute of Mining and Metallurgy: Melbourne).
Restarick, C J, 1976. Pulp sampling techniques for steady state assessment of mineral concentrators, in
Proceedings Sampling Symposium, pp 161–168 (The Australasian Institute of Mining and Metallurgy:
Melbourne).
Weiss, N L, 1985. SME Mineral Processing Handbook (Society of Mining Engineers, American Institute of
Mining, Metallurgical, and Petroleum Engineers: New York).
CHAPTER 3
Mass balancing flotation data
Rob Morrison
ABSTRACT
A flotation circuit survey will typically produce a substantial volume of data – usually as
metal assays and flow rates.
In general, these measured data will not be consistent around any separation unit or junction
within the circuit. A fair assessment of circuit performance requires a set of numerically
consistent data. Otherwise key performance indicators (KPIs), such as recovery in each
section, will depend on which measured values are chosen as the basis of the calculation.
Generating a set of self-consistent data provides a basis for a single set of KPIs. The set of
self-consistent data should in some way be a ‘best estimate’ of the state of the circuit during
the survey.
This chapter discusses strategies for selection of flow measurement and sampling points.
Various ways of estimating measurement accuracy are considered. Lastly, some calculation
methods are described. These are supported by some simple examples.
INTRODUCTION
The basis of mass balancing is that all measurements are subject to statistical variation. If we
could make each measurement (sample, assay or flow rate) many times, the results would
have a spread of values. In practice, we can typically only afford to take a few replicate
samples at each measurement point. If we have some knowledge of the degree of variation
expected in each measurement, we can analyse all the data together to try to find a best
estimate of the mass balance. Data that is self-consistent with this mass balance can be used
to assess and compare performance and as the basis of mathematical models of the process.
The mass balancing process can also incorporate redundant data to produce not only mass
split factors but a further estimate of how well that mass split is defined. The ‘traditional’
approach, using a two-product formula based on a single assay, will also produce an estimate
of the mass split. However, it provides no estimate of the accuracy of that mass split and
therefore provides at best a poor basis for decision-making.
Note that these techniques are not at all new. The first general purpose system is due to
Wiegel (1972) who developed the original MATBAL code. This chapter focuses on how to
attack the simpler cases using spreadsheet technology.
The other key factor is that the flow sheet being balanced must be operating in reasonable
‘balance’ during the measurement period; that is, operating at as close as possible to steady
state for both flow rates and separation processes. Otherwise, the fundamental assumption
that ‘what goes in is equal to what goes out’ is not justified.
Objectives
We wish to produce sets of self-consistent data to suit a range of measurement strategies to
characterise the performance of a flotation circuit or a part of it and to generate estimates of
the self-consistency of the data and of the accuracy of the flow rate measurements.
For the general case we can express these objectives in mathematical terms as three types
of data:
1. xi is a measurement of some sort – assay, flow rate, size fraction, liberated mineral fraction,
etc. These values are generally not self-consistent.
2. x*i is the true (unknown) value of the measured quantity. We can estimate this value at
various levels of accuracy but can never know it exactly. These values are self-consistent.
V i is an adjusted value of xi that satisfies all constraints, that is, is self-consistent and is in
3. x
some way a best estimate of the true value.
Hence, we can define an adjustment Δi of each measurement as:
Δi = (xi - xU i)
The simplest way to find a ‘best’ set of adjusted data is to find a strategy that minimises the
sum of squares (SSQ) of the adjustments. However, this strategy does not take any account
of how well (or badly) a measurement may be defined. The estimated or measured standard
deviation of each measurement σi provides us with an estimate of how well each piece of
data is defined. Hence, a better criterion for minimisation is a sum of squares where each
adjustment is divided (or weighted) by our estimate of the standard deviation of its associated
measurement.
SSQ = ∑ (Δi)
The weighted sum of squares (WSSQ) is the sum of squares of the adjustments where
‘weighted’ means that the adjustment is scaled in terms of the expected variation of that data.
Di 2
WSSQ = / f p
i
vi
It is also necessary that the adjusted values satisfy all flow sheet and summation constraints.
That is, what goes in is equal to what comes out and the various kinds of assays and subassays
add up to one or to the assay in the next level of measurements.
ACCURACY CONSIDERATIONS
We can estimate the standard deviation (‘sd’ or error for short) σi from experience or from
repeated sampling and assaying or repeated measurements against a known standard.
In practice this means taking five to ten replicate samples at several key points within a circuit
and subjecting all of them to the same sample preparation and assaying process that you
intend to use in the actual test work. The formula for sampling variance or standard error
then provides an estimate of standard deviation at each measurement point:
v 2x = / _x - x i / (n - 1)
i
2
Strictly speaking, the standard deviation refers to the complete distribution (which we can
never measure completely) while the standard error is the same property (the square root of
the variance) of the sampling distribution (which we can measure). However, the two terms
are often used interchangeably.
There are some suggestions in the examples for sd estimation based on experience, but it is
usually worth doing some repeatability testing to obtain actual estimates.
In general, this does not mean doing a full sampling tree (AS 2884.4-1997 Heavy mineral
sand concentrates – sampling moving streams) to split the errors into their components.
If you cannot reduce process and assaying variability to less than a few per cent (relative),
there is very little point in carrying out detailed test work as the results will mean very little.
If in general the measurement variations are small, then the required adjustments will also
be small and be drawn from the same population of differences. Therefore, we expect the
average value of (xi - xU i) / σi to be about equal to one and, as a consequence, the WSSQ to be
roughly similar to the number of measurements. However, each time we apply a flow sheet
or a summation constraint, we lose one degree of freedom. This reduces the expected value of
the WSSQ by one. Hence, we now have a quite general way of looking at data quality. If the
WSSQ is of the same order as the number of measurements, our data is likely to be suitable
for further analysis.
A note of caution: if you have taken multiple data sets, a few of these may balance well
by pure chance and still be nonsense. Balances that are self-consistent by chance are rarely
sensible in terms of other criteria such as the size by recovery response.
The WSSQ can also be used to provide an estimate of the standard deviation of each calculated
result xU i. If a small change in xi (or of a flow rate estimate) causes a large change in the sum of
squares, then xU i is well defined. If it causes no change, then it is not defined at all. This property
is particularly useful for checking whether calculated flow rates are well defined.
The formal name for this rule is the ‘propagation of variance’. An alternative approach is
to apply some artificial (random) variation to each measured value and solve the problem
many times. This is called a Monte Carlo method. The variation in results provides an
estimate of the variation that might be expected from a single experiment and analysis if
repeated many times.
This overall approach is usually called the ‘minimisation of weighted squared errors’ or,
simply, ‘least squares’. If for some reason the measured values are drawn from a highly
asymmetric probability distribution, outlying values may cause biases in the calculations.
An alternative approach to finding xU i is to use what are called ‘maximum likelihood methods’.
For measurement and analysis of flotation data, the probability distributions are sufficiently
symmetrical for least squares methods to be quite adequate.
B* = b * A* and C * = (1 - b *) A*
This equation for β* offers the useful insight that, even with perfect data, this method is not
going to work for splitters where we can expect: ai = bi = ci to within measurement variation.
For this case β* is undefined as zero/zero, which is not zero but undefined. Beware of flow
split estimates that are only based on experimental noise. If the flow split is defined to any
reasonable degree, then the splitter is functioning as some kind of separator, which is not
usually desirable. This general approach also works for measured data and provides a useful
way to do an initial evaluation of a set of data.
A B
C
FIG 1 – The simplest case is a single node that represents any process with one input and two outputs.
We can set up the preliminary balance around our separator in MS Excel following the
general structure shown in Table 1. A spreadsheet containing some real data from a section of
a flotation circuit is shown in Table 2.
From the measured flow rate of copper cleaner concentrate, the mass split to concentrate
(beta) should be about one per cent. Note that 890 kg/h is not too difficult to measure as the
flow is only about 15 kg/min.
The estimates of mass split due to the change in each assay vary from 30 per cent to
-66.7 per cent, but perfect data should generate identical values that would be the same as the
perfect mass split.
The calculated mass split based on the copper assays is almost exactly one per cent. The
split based on the silver assay is 1.5 per cent. In each of these cases, the process strongly
concentrates the mineral containing that element.
Our estimate of beta is the ratio of two differences. If these differences are small compared
with the accuracy of measurement, then the ratio will be poorly defined – as it is here.
TABLE 1
Example of a preliminary balance set up around the separator in MS Excel.
Feed Product Reject (ai - ci) (bi - ci) Beta
assay assay assay (ai - ci)/(bi -ci)
A B C
Flows
Assayi
Assay 1 ' ' '
Assay 2 ' ' '
Assay 3 ' ' '
…
Assay n
TABLE 2
Spreadsheet containing some real data from a section of a flotation circuit.
Stream Float feed Cu clnr conc Cu rghr tail
(A) (B) (C)
Measured t/h 87.90 0.89
Relative flow 1 Beta (1-Beta) Beta for
each assay
ai bi ci (ai - ci) (bi - ci) (ai - ci)/(bi - ci)
Measured Pb (%) 4.02 4.07 4.04 -0.02 0.03 -0.667
Measured Zn (%) 15.80 7.80 15.50 0.30 -7.70 -0.039
Measured Cu (%) 0.43 22.40 0.20 0.23 22.20 0.010
Measured Fe (%) 14.30 16.10 13.50 0.80 2.60 0.308
Measured Au (g/t) 1.05 47.50 0.94 0.11 46.56 0.002
Measured Ag (g/t) 142 3023 97 45.00 2926.00 0.015
If the differences are large compared with the accuracy of sampling and assay, then the ratio
should be well defined. This is almost certainly not the case for the lead assays, where both
concentrate and tailing assays exceed the feed value.
It is also clear that zinc is being rejected from the concentrate and that the zinc grade of the
tailings is apparently reduced as a result. A more reasonable possibility is that the zinc assay
of the tails is higher than that in the feed, but our measurement accuracy is insufficient to
demonstrate this effect.
The absolute value of an assay is also important. One gram per tonne is 0.0001 per cent or
one per cent is 10 000 ppm. Hence, the silver assays are really 0.0142 per cent, 0.3032 per cent
and 0.0097 per cent.
The gold assays are two orders of magnitude smaller again and the mass split of 0.2 per cent
is a reflection of good sampling and assaying at those concentrations.
Error models
There are some useful generic ways to ‘estimate’ likely measurement errors based on assay
magnitude. The most popular method is to assume a constant relative (or per cent) error. This
approach implies that the best defined values would be the tailings assays for gold and silver
and is plainly nonsense.
The next commonly used approach is to assume that the expected error is constant or
(usually) one. This simplifies the arithmetic and is a more sensible estimate. The larger assays
are now assumed to have the smallest relative error.
The JKMRC approach is to measure the error distribution through repeated sampling or to
use a well-proven heuristic; that is, a rule that should behave in a sensible manner. This rule
is often attributed to Bill Whiten, who disclaims ownership.
We assume that for assays (or size fractions) in the per cent range that the minimum sd
is 0.1 per cent (absolute) and that for assays greater than nine per cent the absolute sd is
one per cent. In between those two limits, the sd is 0.1 plus 0.1 times the assay value. Hence,
the relative error at a measured value of 0.1 per cent is 200 per cent. The relative error at an
assay value of ten per cent is also ten per cent.
At a measured value of 0.5 per cent, the absolute error is 0.1 plus 0.05 per cent or 30 per cent
relative. Some people refer to this as a ‘two-term’ error model, where 0.1 is the fixed error and
0.1 is the fractional error between zero and nine per cent. Another term is a ‘one over x’ model
because the relative error curve has much the same shape as that function.
If the analysis method used is very different, we can expect the error models to change.
For example, the fire assays typically used for precious metals should be intrinsically more
accurate at ppm values than standard techniques such as X-ray fluorescence (XRF) and
atomic absorption spectroscopy (AAS) for base metals in the per cent ranges would be at ppm
values. The reason for this is that the assaying technique itself uses some preconcentration
before measurement.
Note that only one β value is required for all the n assays and that Δi should be zero for
perfect data. Hence, a process that makes a large difference to the stream assays will have a
better-defined flow split than one that only makes a small difference, such as a final cleaner
or scavenger bank. Similarly, if a splitter is working well, it is essential to measure or estimate
its flow split as mass balancing will not be helpful.
Transposing and simplifying, we get for each component:
D i = (a i - c i) - b (b i - c i)
If we square both sides and sum up all the components:
We might reasonably expect the value of β that minimises SSQ to be a reasonable estimate
of β*, which we can call bU .
As for the first case, this can use a quite generic spreadsheet format as shown in Table 3.
Set up this spreadsheet as shown in Table 4. Then work through it with estimates of β of 0,
0.005, 0.01, 0.02, 0.1, 1.0, as we know from the measured flow rate that the mass split is about
one per cent.
The gold assays have been divided by 100 and the silver assays by 1000 to bring them closer
to per cent values. (As an exercise, check out the minima with the unadjusted values.) If your
set-up is correct, the SSQ should go from 0.783 at 0.0 to 550 at 1.0 with a minimum value of
0.742 at 0.01.
It is also a useful exercise to tabulate β and plot it against SSQ as shown in Figure 2. This
graph shows a well-defined minima at β is approximately 0.01. It is well defined because
making a small change (±0.005) in β makes a ±10 per cent change in the SSQ. However, it is
necessary to plot the log of the SSQ to be able to see the minima.
Go through the β sequence again and watch how the SSQ or (Δi)2 for each assay varies. They
do not all have a minimum at the same value of β. How should we decide which value of β
to use? Obviously, the value that has the most accurately measured data should dominate.
TABLE 3
Generic structure of a spreadsheet balance around a single node using the method of mass flow errors.
Estimated beta Initial estimate
Feed Product Reject (ai - ci) Β*(bi - ci) Δi Δi2
assay assay assay
Measured A B C
Flows
Assayi
Assay 1 ' ' '
Assay 2 ' ' '
Assay 3 ' ' '
Assay 4
…
Assay n
Total SSQ =
TABLE 4
An example of the method of mass flow errors.
Estimated beta = 0.0036 A B C
Stream Float feed Cu clnr con Cu rghr tail
Measured t/h 87.90 0.89
Relative flow 1.000 0.004 0.996 Delta Delta squared
ai bi ci ai - β* bi - (1-β) *ci
Measured Pb (%) 4.02 4.07 4.04 -0.020 0.000
Measured Zn (%) 15.80 7.80 15.50 0.328 0.108
Measured Cu (%) 0.43 22.40 0.20 0.149 0.022
Measured Fe (%) 14.30 16.10 13.50 0.790 0.625
Measured Au (g/t) 0.0105 0.475 0.0094 -0.012 0.000
Measured Ag (g/t) 0.142 3.023 0.097 0.020 0.000
SSQ = 0.759
100
10
Log of SSQ
SSQ
1
0 0.05 0.1 0.15 0.2 0.25 0.3
0.1
Estimated Beta
FIG 2 – Log of the sum of squared errors at a range of estimates of the mass split – beta.
Let us first consider the case where each of the minima is quite similar. Clearly the weighting
will have very little effect on the value of the flow split. In this case, there is little room for
debate about the value of the flow split.
In general, it will be more useful to derive an estimate of the standard deviation of the flow
split error based on our measured standard deviations. This estimate can be used to ‘weight’
the contribution of each flow split error to the sum of squares that is to be minimised.
Di 7a i - bb i - (1 - b) c iA
vD i = vD i
The rule for ‘propagation of variance’ is that the variance of a function is the sum of the
product of the variance of each input parameter and the square of its partial derivative – see
Deming (1938) or almost any statistics reference.
Dropping the i for the moment:
2D 2 2D 2 2D 2
v 2D = : 2a D (v 2a) + : 2b D (v b2) + : 2c D (v c2)
= v a2 + b 2 v b2 + (1 - b) 2 v c2
If we add some additional columns to our spreadsheet, we can derive a weighted sum of
squares WSSQ of the mass flow errors (Table 5).
As before, run through in increments of 0.001 for β from 0.00 to 0.02 and watch how the
minima change. For the first case, assume the sd of each assay is one. For the second case
assume they are each five per cent of the assay value. For the third case, assume ten per cent.
If your spreadsheet is right, the unit weighting case will look quite like the original but be
driven by the larger assay values. For the second and third cases, the smallest assays will
dominate. Neither seems very sensible and we will consider a better balanced approach in
the section on error models.
AN ANALYTICAL SOLUTION
In the prehistoric days before PCs and Solver, a simple analytical solution was very useful.
Recalling some high school calculus, if we wish to find the minimum (or maximum) of a
function, we take the first derivative with respect to the variable of interest and set the result
to zero.
SSQ = 7(ai - ci) - b (bi - ci)A
/ 2
dSSQ
db
=2 / 7(a - c ) - b (b - c )A(b - c )
i
i i i i i i
0= / 7(a - c ) (b - c ) - b (b - c ) A
i i i i i i
2
b=
/ (a - c ) (b - c )
i i i i
/ (b - c ) i i
2
This is an interesting result. When ai = bi = ci, it is also undefined. The assays with the largest
difference will dominate each sum. Hence, the target of the separation process (for example,
copper in a copper circuit) provides the best defined balance.
As for the earlier cases, this approach is easy to set up in a spreadsheet. We already have
columns of (ai - bi) and (bi - ci) in the first example. Hence, we only need to add a column
each for their product and the second term squared. Divide the sums of these columns for an
analytical estimate of beta as shown in Table 6.
TABLE 5
Additional columns required for Table 3 to include a weighted sum of squares (WSSQ).
σa σb σc β2 (1 - β)2 D 2i
WSSQ =
TABLE 6
Spreadsheet balance including the analytical solution for the ‘best fit’ mass split.
Data A B C
Flotation feed Cu cleaner conc Cu rougher tail
t/h 87.90 0.89
1 0.010 0.990 Analytical Estimate
ai bi ci (ai - ci) (bi - ci) (ai - ci)*(bi - ci) (bi - ci)2
Pb (%) 4.02 4.07 4.04 -0.02 0.03 -0.001 0.001
Zn (%) 15.80 7.80 15.50 0.30 -7.70 -2.310 59.290
Cu (%) 0.43 22.40 0.20 0.23 22.20 5.106 492.840
Fe (%) 14.30 16.10 13.50 0.80 2.60 2.080 6.760
Au (g/t)/100 0.001 0.475 0.009 -0.01 0.47 -0.004 0.217
Ag (g/t)/1000 0.142 3.023 0.097 0.05 2.93 0.132 8.561
Sum 5.003 567.669
Beta 0.008813553
The accuracy with which the flow split is defined depends on how rapidly the SSQ increases
as we move away (in either direction) from the best fit value of beta. Once again from high
school calculus, we recall that the curvature of a function is its second derivative:
2SSQ
2b
= / ((a - c ) (b - c )) - / (b - c )
i i i i i i
2
i i
2 2 SSQ
2b 2
=- / (b - c ) i i
2
i
Hence:
v D2 = v 2b / (b - c ) i i
2
In the Excel case, Solver does not provide this value for parameters but we can easily change
beta by ±1 per cent and note the change on the sum of squares to test for a well-defined value.
The Excel function RAND() draws a value from a calculated random distribution between
zero and one. We can scale the range of values between ‘min’ and ‘max’ by multiplying by
((max-min) + min). This generates a uniform distribution of random values between min and
max. However, what we actually need is a random set of variations drawn from a probability
distribution similar to the distribution we were sampling by making measurements or taking
samples. This is a little trickier than it sounds at first.
If we consider the cumulative (or integral) form of any probability distribution, by definition,
it will start at zero and end at one. Hence, we need the value of our cumulative distribution that
corresponds to each random value between zero and one. This is the inverse of the probability
distribution. This will work well most of the time, but it is worth remembering that the normal
distribution extends from negative to positive infinity. Most chemists will not accept this range
of variation and will simply not report numbers that are statistically possible but not plausible.
If we reduce the range of possible random values to say 0.025 to 0.975, then the inverse values
will lie between about plus and minus two standard deviations of the mean value.
To test out this process, set up four blocks of similar structure as shown in Table 7. The first
one is the measured data block. The second one is the block of estimated standard deviations.
The third one is our scaled random number. The last one is the block of perturbed data,
which we can use to rerun our balance (and other calculations) to generate an estimated
range of variation.
TABLE 7
Spreadsheet example of a Monte Carlo analysis.
Measured Data A B C
Stream Flotation feed Cu cleaner conc Cu rougher tail
Measured t/h 87.90 0.89
Relative flow 1 0.010 0.990
ai bi ci
Measured Pb (%) 4.02 4.07 4.04
Measured Zn (%) 15.80 7.80 15.50
Measured Cu (%) 0.43 22.40 0.20
Measured Fe (%) 14.30 16.10 13.50
Measured Au (g/t)/100 0.001 0.475 0.009
Measured Ag (g/t)/1000 0.142 3.023 0.097
Standard deviation Model A B C
Stream Flotation feed Cu cleaner conc Cu rougher tail
Standard deviations t/h 4.40 0.04 5%
ai bi ci
Standard deviations Pb (%) 0.50 0.51 0.50 0.1 +10% Or 1 if > 9
Standard deviations Zn (%) 1.00 0.88 1.00 0.1 +10% Or 1 if > 9
Standard deviations Cu (%) 0.14 1.00 0.12 0.1 +10% Or 1 if > 9
Standard deviations Fe (%) 1.00 1.00 1.00 0.1 +10% Or 1 if > 9
Standard deviations Au (g/t)/100 0.050 0.050 0.050 0.05
Random values RAND – min 0.025 RAND – max 0.975
Synthetic Data A B C
Stream Flotation Cu cleaner Cu rougher
feed conc tail
Measured t/h 80.67 0.90 Analytical Estimate
ai bi ci (ai - ci) (bi - ci) (ai - ci)* (bi - ci)2
(bi - ci)
Synthetic Pb (%) 3.51 3.54 3.61 -0.10 -0.07 0.008 0.006
Synthetic Zn (%) 15.13 9.12 16.04 -0.91 -6.92 6.277 47.819
Synthetic Cu (%) 0.37 22.19 0.17 0.20 22.02 4.403 484.766
Synthetic Fe (%) 13.13 16.24 12.57 0.56 3.67 2.060 13.460
Synthetic Au (g/t)/100 0.02 0.52 0.00 0.02 0.52 0.010 0.267
Synthetic Ag (g/t)/1000 0.18 2.98 0.11 0.07 2.87 0.207 8.243
Sum 12.964 554.560
Beta 0.023377
Each time the sheet is executed (press F9 to force a recalculation), this process will
generate an equally likely ‘plausible’ set of synthetic data, which our formula will then
solve for beta. After each run, jot down the resulting beta. If you edit the spreadsheet, it will
generate another value unless you switch to manual recalculation (Tools/Options/Manual
Recalculation).
In addition to the scaled RAND() function, you will need to use the NORMAL.INV (inverse
x, mean of x and sd) function to generate a deviation to add to the measured value. An
example is shown in Table 7.
Do as many runs as you like – but no less than ten. Now use the mean and sd functions
(AVERAGE(Range) and STDEV.S(Range)) with your row of β values as the argument. If the
mean is very different from your original W
b , there is a problem in the spreadsheet. The sd of
the synthetic betas provides an estimate of the sd of the calculated beta. Some mass balance
programs only provide this way of estimating how well the balance is defined for the balanced
flow rates.
Reconstitution
If the flow split is well defined, adding the reject and concentrate together in the flow split
ratio has much to recommend it. If the weighted sum estimate of the feed assays is within
±1 standard deviations of the feed measurement, this realisation will be quite adequate for
further analysis. In practice, we can insert a single column to the right of the feed column
and add β times the concentrate plus (1 - β) times the tailing assay. Note that the summation
constraints are satisfied (Table 8). This strategy has the advantage that two thirds of our
numbers are exactly the ones that we measured.
Data adjustment
Persons of a statistical bent will favour least squares adjustment of all of the assays. This does
lead us back towards the general case, so it is useful as an exercise. We can start with the unit
weighted case. Recall that:
Di = ai - bbi - (1 - b ) ci
If we want an exact solution, we need some strategy to apportion Δi across the measured
assays. Strictly speaking, we should use the best fit value of beta, W
b . However, as we are
W
going to do a combined minimisation a little later using b it is probably confusing. Some
possible simple solutions are to divide delta by three or to proportion delta according to
the flow in each stream. The mathematically preferred approach is to minimise the required
adjustment in some way – and the sum of squares is a good general approach. Let Δa, Δb and
Δc be the minimum adjustments and omit the i – as we can consider each assay as a separate
case. Our adjusted data must balance if:
0 = (i - Da) - b (b - D b) - (1 - b) (c - Dc)
TABLE 8
A self-consistent data set generated by reconstituting products at the best fit mass split.
Data A B C
Flotation feed Reconstituted feed Cu cleaner conc Cu rougher tail
t/h 87.90 0.89
Beta 1 0.0088 0.991
ai bi ci
Pb (%) 4.02 4.040 4.07 4.04
Zn (%) 15.80 15.432 7.80 15.50
Cu (%) 0.43 0.396 22.40 0.20
Fe (%) 14.30 13.523 16.10 13.50
Au (g/t)/100 0.001 0.014 0.475 0.009
Ag (g/t)/1000 0.142 0.123 3.023 0.097
D = Da - bD b - (1 - b) Dc
We can use a Lagrange multiplier λ to impose the constraint by setting the constraint
equation to zero and adding it to the SSQ. This is called a ‘modified’ sum of squares. We then
minimise each adjustment and λ as well.
SSQ = (Da) 2 + (D b) 2 + (Dc) 2 - 2m (- D + Da - bDb - (1 - b) Dc)
Then:
2SSQ
= 2Da - 2m = 0
2Da
2SSQ
= 2bDb + 2mb = 0
2Db
2SSQ
= 2Dc + 2m (1 - b) = 0
2Dc
2SSQ
= 2 (- D + Da - bDb - (1 - b) Dc) = 0
2m
As we already know the value of β and Δ for each assay, we can easily calculate a set of
adjustments. For this case, the adjustment is proportional to the flow of that component (assay)
in each stream. Table 9 shows the adjusted data and a check that it actually does add up.
Examining the adjusted values for lead and zinc shows that in this case it would be very
difficult to distinguish the likely change in the tailings assays as a result of the concentrate
assays; that is, the difference imposed by the process. By way of contrast, copper, iron and
silver are all changed significantly.
If we wish to scale each adjustment by its own standard deviation, the solution is quite
similar (subject to the same constraints as before):
WSSQ = (Da/ va) 2 + (Db/ vb) 2 + (Dc/ vc) 2
2WSSQ
= 2Da/ va - 2m = 0 or Da = mva
2Da
2WSSQ 2Db
= + 2mb = 0 or Db = - mbvb
2Db vb
TABLE 9
A self-consistent set of adjusted data generated by minimising the sum of squares of the required adjustments.
Beta 0.008813553 A B C 1.982528
Stream Flotation Cu cleaner Cu rougher
feed conc tail
Measured t/h 87.90 0.89
Relative flow 1 0.010 0.990 Check Beta
aI bi ci Delta Lambda Delta Check
Adjusted Pb (%) 4.03 4.07 4.03 -0.0203 -0.01 0 0.008814
Adjusted Zn (%) 15.61 7.80 15.68 0.3679 0.19 0 0.008814
Adjusted Cu (%) 0.41 22.40 0.22 0.0343 0.02 0 0.008814
Adjusted Fe (%) 13.91 16.10 13.89 0.7771 0.39 0 0.008814
Adjusted Au (g/t)/100 0.01 0.47 0.003 -0.0125 -0.01 0 0.008814
Adjusted Ag (g/t)/1000 0.13 3.02 0.11 0.0192 0.01 0 0.008814
2WSSQ 2Dc
= + 2 (1 - b) m = 0
2Dc vc
2WSSQ
= 2 (- D + Da - bDb - (1 - b) Dc)
2m
Da = mva
Db = - bmvb
Dc = (1 - b) mvc
D = mv a + b 2 mv b + (1 - b) 2 mv c
m = D/ _v a + b 2 v b + (1 - b) 2 v ci
TABLE 10
A self-consistent set of adjusted data generated by minimising the sum of squares
of the adjustments weighted by the sd estimates from the Monte Carlo example.
Beta 0.010778355 A B C
Stream Flotation Cu cleaner Cu rougher
feed conc tail
Measured t/h 87.90 0.89
Relative flow 1 0.010 0.990 Check Beta
ai bi ci Delta Lambda Delta Check
Weighted adj Pb (%) 4.03 4.07 4.03 -0.0203 -0.02 0 0.010778
Weighted adj Zn (%) 15.61 7.80 15.69 0.3830 0.19 0 0.010778
Weighted adj Cu (%) 0.44 22.40 0.20 -0.0093 -0.04 0 0.010778
Weighted adj Fe (%) 13.91 16.10 13.89 0.7720 0.39 0 0.010778
Weighted adj Au (g/t)/100 0.01 0.47 0.003 -0.0134 -0.14 0 0.010778
Weighted adj Ag (g/t)/1000 0.14 3.02 0.10 0.0135 0.14 0 0.010778
adj – adjustment.
TABLE 11
A mass balance that combines the search for the best fit mass split with the minimised weighted assay adjustments.
Adjusted Beta 0.0107783 A B C
Weighted Stream Flotation Cu cleaner Cu rougher
feed conc tail
Data Measured t/h 87.90 0.89
ai bi ci Delta Lambda Delta
Pb (%) Wtd adj Pb (%) 4.03 4.07 4.03 -0.0203 -0.02 0
Zn (%) Wtd adj Zn (%) 15.61 7.80 15.69 0.3830 0.19 0
Cu (%) Wtd adj Cu (%) 0.44 22.40 0.20 -0.0093 -0.04 0
Fe (%) Wtd adj Fe (%) 13.91 16.10 13.89 0.7720 0.39 0
Au (g/t) Wtd adj Au (g/t)/100 0.01 0.47 0.003 -0.0134 -0.14 0
Ag (g/t) Wtd adj Ag (g/t)/1000 0.14 3.02 0.10 0.0135 0.14 0
Adjustments A B C
Stream Flotation Cu cleaner Cu rougher
feed conc tail
Measured t/h 87.90 0.89 Weighted
ai bi ci SSQ
Pb (%) Wtd adj Pb (%) -0.02 0.000 0.02 0.0008
Zn (%) Wtd adj Zn (%) 0.19 -0.002 -0.19 0.0741
Cu (%) Wtd adj Cu (%) -0.04 0.000 0.04 0.0025
Fe (%) Wtd adj Fe (%) 0.39 -0.004 -0.39 0.3012
Au (g/t) Wtd adj Au (g/t)/100 -0.14 0.001 0.13 0.0361
Ag (g/t) Wtd adj Ag (g/t)/1000 0.14 -0.001 -0.13 0.0366
WTDSSQ 0.4514
become essential – each elemental assay will generate a flow split estimate, but they will not
define mineral behaviour.
will be essential to refine your sampling and assaying techniques before attempting more
detailed test work.
SSQ = /D i
2
As stream E is not measured for this case, we can use Solver to find ‘best fit’ values of β
and γ. Similarly, we can use Solver to find the minimum squared data adjustments subject to
the condition that the adjusted data add up to zero at the specified values of β and γ. As before,
the adjustments can be weighted by their estimated standard deviations. However, tackling
the complete problem in Solver will rapidly exceed the maximum number of constraints
available with the add-in version. Various expanded versions are available for purchase.
Note:
•• You will need at least two independent assays to solve this balance at all as there are two
unknowns.
•• If stream C is a middlings stream with similar composition to the feed stream, the flow rate
in stream C is not defined. Almost any flow measurement estimate (including a ‘calibrated
eyeball’) will be more useful than the result of the balance.
For the second configuration (in which E is measured) there are two separations or nodes.
Hence, we need two error equations:
D1i = ai - bbi - (1 - b) ei
D2i = (1 - b) ei - cci - (1 - b - c) di
SSQ = / 7(D1 ) i
2
+ (D2i) 2A
i
This case is also quite straightforward to set up for Solver – with or without weighting the
errors. As before, if your data is accurate and self-consistent, there is much to recommend
simple reconstitution to generate a self-consistent set of data for further analysis. However,
A E D
B C
FIG 3 – A two-node, rougher scavenger circuit.
it is also fairly straightforward to do a least squares adjustment one assay at a time. We will
work through the second case as it has two constraints.
For each value of i, we wish to minimise:
SSQ = Da 2 + Db 2 + Dc 2 + Dd 2 + De 2
subject to:
D1 = Da - bDb - (1 - b) De
and:
D2 = (1 - b) De - cDc - (1 - b - c) Dd
+ 2m 2 (D2 - (1 - b) De - (1 - b - c) Dd - cDc)
/ 2Dc = 2Dc + m 2 2c = 0
/ 2De = 2De - 2m 2 (1 - b) = 0
m 2 = (De - m1 (1 - b)) / (1 - b)
m 2 = De / (1 - b) - m1
and calculate both Lagrange multipliers. This gives a direct solution for each adjustment.
As before, we can extend this to a direct solution of the minimum sum of squared adjustments
or of weighted squared adjustments.
The following set of data in Table 12 can be used as an exercise.
TABLE 12
Measured assay data for the circuit shown in Figure 3.
Stream ID Pb (%) Zn (%) Cu (%) Fe (%)
Measured assay Measured assay Measured assay Measured assay
Pb rougher feed A 11.89 24.02 0.25 11.14
Pb rougher conc B 20.20 30.10 0.31 9.10
Pb rougher tail E 1.25 16.60 0.16 13.50
Pb scavenger conc C 8.20 39.80 0.37 11.70
Pb scavenger tail D 0.51 14.60 0.14 13.70
This problem provides a different kind of constraint. We have an estimate of the ratios of the
flow rate from each cell but we do not know the total flow rate. We may or may not know the
total assay of the concentrate. An analytical solution for this multiconcentrate plus each cell
tail was developed by the author and published in Chapter 7 of Lynch (1977). It is outlined in
Figure 4. The data from that example are shown in Table 13.
The flow ratios to A2, A3 can be designated as alpha 2, alpha 3 and so on. The solution is a
somewhat complicated matrix inversion:
Rm11 m12 0 0 VW SRa2VW RSx1VW
S
Sm21 m22 m23 0 W Sa3W S 0 W
S 0 m32 m33 m34W $ S W = S 0 W
S W Sa4W S W
S0 0 m43 m44W Sa5W Sx4W
T X T X T X
Each of the ‘m’ and ‘x’ terms is a sum of differences of assays. When a circuit reaches this
level of complexity it is recommended that the user switch to a more general approach or to
a commercial package.
The flow ratio case is better suited to a numerical approach. If we have measured flow rates,
fi, f2 →fm we can assume that they will be related to the total functional flow rate by some ratio
t. Hence, our node equation becomes:
D1 = a - bb - (1 - b) c
tb /f j
conc j l for βb.
j
TABLE 13
Measured assay data for the circuit shown in Figure 4.*
A1 A2 A3 A4 A5 B1 B2 B3 B4 B5
Copper 3.03 1.0 1.15 0.75 0.52 21.2 22.5 17.1 9.7 15.5
Iron 8.2 7.1 6.4 6.0 5.7 23.8 25.9 23.4 20.2 21.4
Sulfur 6.5 5.3 4.4 3.8 3.3 26.5 29.2 25.9 20.8 23.0
Insoluble 46.6 49.0 51.0 51.8 52.8 14.4 11.4 18.4 27.6 21.6
* Reprinted from Mineral Crushing and Grinding Circuits, A J Lynch, 350 p, Copyright (1977), with permission from Elsevier.
In either case, we are solving the original problem for the flow weighted average of the
concentrate assays and this will usually be a good first step to see if the lip sample data
are reasonably self-consistent. If they are, weighted addition and reconciliation should be
adequate for further analysis. Alternatively, a similar constrained least squares adjustment
can be used – either weighted or unweighted.
TABLE 14
Connection matrix for the circuit shown in Figure 4. Note that ‘B’ streams
are numbered 6 to 10 and zeros are omitted for clarity.
Nodes Streams
1 2 3 4 5 6 7 8 9 10
1 1 -1 -1
2 1 -1 -1
3 1 -1 -1
4 1 -1 -1
5 1 1 1 1 -1
A1 A2 A3 A4 A5
B1 B2 B3 B4
B5
FIG 4 – A typical ‘down-the-bank’ equivalent circuit for detailed sampling.*
* Reprinted from Mineral Crushing and Grinding Circuits, A J Lynch, 350 p, Copyright (1977), with permission from Elsevier.
Similarly, each assay times its stream flow rate times the relevant stream/node entry in
the connection matrix will generate a vector of mass flow errors. Hence, we can call [Ci] a
connection matrix where each row is multiplied by assay type i, then:
[Ci] . f = Δi for each assay i.
As MS Excel has a built-in function for matrix multiplication, it is straightforward to set up
a spreadsheet to define constraints and the SSQ of mass flow errors for each assay type. The
product of the Connection Matrix (Table 14) and the copper assay matrix is shown in Table 15.
The best fit solution for all flows (based on f1 = 100) is shown with the copper SSQ (Table 16).
The best fit flow rates can be based on any one or any combination of assays by including
its SSQ in the target for Solver to minimise. Weighting and sensitivity estimates can be added
in much the same way as for the single node case. Once the best fit flow rates have been
established, we can use a similar strategy to find each set of data adjustments. In this case, the
adjustments become the parameters for Solver, the SSQ of the adjustments is the target for
TABLE 15
Product of the Connection Matrix (Table 14) and the copper assay matrix.
Nodes Streams
1 2 3 4 5 6 7 8 9 10
1 3.03 -1 0 0 0 -21.2 0 0 0 0
2 0 1 -1.15 0 0 0 -22.5 0 0 0
3 0 0 1.15 -0.75 0 0 0 -17.1 0 0
4 0 0 0 0.75 -0.52 0 0 0 -9.7 0
5 0 0 0 0 0 21.2 22.5 17.1 9.7 -15.5
minimisation and our connection matrix multiplied by the flow rate of the selected component
must be constrained to zero. The results for copper are shown in Table 17, followed by the
constraints in Table 18, which are near enough to zero.
TABLE 16
Best fit flow rate solution for all flows (based on f1 = 100) is shown with the copper SSQ.
Flows Estimates Matrix product SQD
1 100 64.4250163 4150.58
2 93.13985 -78.0880874 6097.74
3 90.13671 3.37058247 11.3608
4 88.13768 -11.8850601 141.254
5 84.63477 43.0064029 1849.5
6 6.860148
7 3.003143 SSQ
8 1.999028 12250.5
9 3.502911
10 15.36523
TABLE 17
Comparison of adjusted and measured copper assay data.
Stream 1 2 3 4 5 6 7 8 9 10
Adjustment -0.091 0.664 -0.166 -0.125 -0.250 -0.949 -0.437 -0.287 -0.498 2.139
Adjustment squared 0.008 0.440 0.028 0.016 0.062 0.900 0.191 0.082 0.248 4.576
Adjusted Cu 2.939 1.664 0.984 0.625 0.270 20.251 22.063 16.813 9.202 17.639
Measured 3.030 1.000 1.150 0.750 0.520 21.200 22.500 17.100 9.700 15.500
Total SSQ 6.552
TABLE 18
Best fit flow rate and constraint residuals around each
node of the balance. The residuals are effectively zero.
Flows Estimates Matrix product
1 100 -3.7E-14
2 93.13985 6.35E-14
3 90.13671 -1.5E-14
4 88.13768 -2.1E-14
5 84.63477 1.78E-15
6 6.860148
7 3.003143
8 1.999028
9 3.502911
10 15.36523
100.00
Pb
Zn
10.00
Cu
Fe
1.00
0.10 1.00 10.00 100.00
0.10
FIG 5 – Parity graph comparing measured total assays for each metal in each stream
with those based on reconstitution of the assays for each size fraction of each stream.
TABLE 19
Measured sizes, total assays and assays reconstituted from size fraction assays based on samples taken from zinc circuit.
Measured data A B C
Sizing Wt % Wt % Wt %
ai - ci bi - ci Beta (ai - ci) * (bi - ci) (bi - ci)2 Beta bar
+75 20.85 11.87 18.48 2.36 -6.61 -0.3571 -15.623 43.746
+53 12.46 15.73 8.58 3.88 7.16 0.5424 27.770 51.197
+38 11.53 14.22 7.07 4.46 7.15 0.6242 31.910 51.124
-38 55.17 58.18 65.87 -10.71 -7.69 1.3920 82.343 59.155
Totals 126.400 205.222 0.616
Pb 0.59 1.19 0.39 0.20 0.80 0.2438 0.156 0.640
Zn 15.34 59.41 0.86 14.48 58.55 0.2472 847.511 3428.103
Cu 0.15 0.29 0.10 0.05 0.19 0.2368 0.009 0.036
Fe 13.70 5.38 16.44 -2.74 -11.06 0.2477 30.304 122.324
Rem 70.24 33.73 82.21 -11.98 -48.48 0.2470 580.548 2350.310
Recon totals 1458.528 5901.413 0.247
Pb 0.71 1.49 0.41 0.30 1.08 0.2801 0.326 1.163
Zn 15.43 57.50 1.10 14.32 56.40 0.2540 807.833 3180.405
Cu 0.14 0.34 0.12 0.02 0.22 0.0861 0.004 0.047
Fe 13.97 5.14 15.81 -1.84 -10.67 0.1729 19.668 113.755
Remainder 69.75 35.53 82.55 -12.80 -47.02 0.2722 601.971 2211.343
1429.802 5506.712 0.260
Overall 3014.731 11613.34 0.260
There are many more sophisticated approaches but they will not help very much with poor
data – other than to spread the problems across the data set and make them more difficult to
identify.
Note that for diagnostic purposes, you can do a balance for each within each size fraction to
see if the mass split within each size fraction is self-consistent. The mass split calculated this
way should be close to the mass split in the size fraction, which is the concentrate size fraction
multiplied by the overall mass split divided by the feed size fraction.
There are a range of mathematically complex ways to approach the general case (Hodouin
and Everel, 1980; Gay, 1999).
be useful to investigate whether all of the size fractions are likely in error or a particular
one is at fault. We can investigate by applying the 25 per cent split to the concentrate size
distribution and the 75 per cent split to the tails.
This generates mass flows of 2.97 and 13.86 respectively. The mass split within each size
fraction can be estimated from the assays within that size fraction. Table 20 estimates the
+75 micron mass split as 0.144, which gives flow rates of 3.00 and 17.85 respectively.
This simple analysis suggests that the feed and concentrate are consistent with the assay
split but the tailings are not. Hence, investigating the collection and sample preparation for
the tailings stream would be a good idea before doing more test work.
TABLE 20
Assay mass balance within the +75 micron size fraction.
Mass splits in size fractions
Measured data A B C
Sizing Wt % Wt % Wt% Component
+75 ai - ci bi - ci Beta (ai - ci) * (bi - ci) (bi - ci)2
Pb 0.60 1.30 0.40 0.20 0.90 0.222 0.180 0.810
Zn 8.90 55.20 1.50 7.40 53.70 0.138 397.380 2883.690
Cu 0.17 0.30 0.18 -0.02 0.12 -0.125 -0.002 0.014
Fe 7.90 5.00 7.60 0.30 -2.60 -0.115 -0.780 6.760
Rem 82.44 38.20 90.32 -7.88 -52.12 0.151 410.966 2716.494
Fraction total 100 100 100 807.7444 5607.768
Beta bar 0.144040
TABLE 21
Assay mass balance within the +53 micron size fraction.
Measured data A B C
Sizing Wt % Wt % Wt% Component
+53 ai - ci bi - ci Beta (ai - ci) * (bi - ci) (bi - ci)2
Pb 0.65 1.00 0.50 0.15 0.50 0.300 0.075 0.250
Zn 18.3 59.00 1.30 17.00 57.70 0.295 980.900 3329.290
Cu 0.23 0.35 0.20 0.03 0.15 0.200 0.005 0.023
Fe 17.8 4.80 25.40 -7.60 -20.60 0.369 156.560 424.360
Rem 63.02 34.85 72.60 -9.58 -37.75 0.254 361.645 1425.063
Fraction total 100 100 100 1499.184 5178.985
Beta bar 0.289474
TABLE 22
Adjusted assays for the +75 micron size fraction.
Adjusted A B C
Sizing Wt % Wt % Wt% Exp Check
+75 Delta Lambda Sum
Pb 0.57 1.31 0.43 0.06 0.03 0.000
Zn 9.58 55.09 0.93 -1.17 -0.68 0.000
Cu 0.18 0.30 0.16 -0.03 -0.02 0.000
Fe 7.49 5.07 7.95 0.72 0.41 0.000
Remainder 82.18 38.24 90.53 0.44 0.25 0.000
Fraction total 100.00 100.00 100.00
TABLE 23
Assay adjustments for the +75 micron size fraction.
SSQ A B C
Sizing Wt % Wt % Wt % Delta
+75 SSQ
Pb 0.03 -0.01 -0.03 0.00
Zn -0.68 0.11 0.57 0.79
Cu -0.02 0.00 0.02 0.00
Fe 0.41 -0.07 -0.35 0.30
Remainder 0.25 -0.04 -0.21 0.11
Fraction total SSQ 1.20
We can now generate a complete set of data adjustments within each size fraction (weighted
or not). To link the size fractions together and tie back to the total assays it is important to
realise that the mass split within each size fraction has to be identical for the size-fraction-based
calculation and the assay-based calculations. Hence, we can multiply each feed size fraction
by that mass to generate a set of concentrate flows and by one minus that split to generate
a set of tailings size fraction flows and then convert the flow back to size distributions. This
strategy makes another set of constraints implicit.
The last constraint can be imposed by calculating one of the feed size fractions (usually the
fines) by difference. We can use the other feed size fractions and each of the size fraction split
factors as a set of parameters to minimise the total SSQ for all sets of data. One way to set
this up is shown in Table 24. The parameters to minimise are identified in bold font as are the
adjusted total assays.
This approach is somewhat tedious to set up but does provide a very detailed view of
the data. For example, the problem with the tailings size distribution identified earlier is
now very clear. That set of sizing data contributes more than half the SSQ. We can omit that
contribution and reminimise to see how much the results change. As might be expected, the
changes are mostly in the tailings size distribution but they are not very large. Hence, either
set of data would be similarly useful. However, it does suggest that taking samples optimised
for equi-probable selection is essential for this kind of work. That means cutting through the
stream to be sampled – and avoiding any kind of dip sampling.
TABLE 24
A simple spreadsheet strategy for a complete size by assay balance.
Calc size A B C B C
Flows Wt % Flow Flow Wt % Wt %
Adj A Adj B Adj C
+75 fit 18.89 3.02 15.87 11.81 21.31
+53 fit 13.05 3.74 9.31 14.63 12.50
+38 fit 12.08 3.76 8.31 14.74 11.16
-38 55.99 15.01 40.98 58.82 55.02
Totals 100.00 25.53 74.47 100.00 100.00
Size Splits
Beta 75 0.15963
Beta 53 0.28631
Beta 38 0.31148
Beta-38 0.26814
Adjusted Assays
Recon Flows A B C
Pb 69.76 38.14 31.62 0.698 1.494 0.425
Zn 1553.57 1468.72 84.85 15.536 57.540 1.139
Cu 16.84 8.54 8.30 0.168 0.334 0.111
Fe 1379.62 133.11 1246.51 13.796 5.215 16.737
Rem 6980.21 904.03 6076.18 69.802 35.417 81.587
Totals 10000.0 2552.54 7447.46 100.00 100.00 100.00
RECAP
The suggested analysis sequence follows the order of the chapter. Run a check around each
node using the method of mass flow errors detailed in earlier sections ‘The simplest case’ and
‘The method of mass flow errors’.
The formulae derived in the sections ‘An analytical solution’ and ‘Estimating the accuracy of
the flow split’ are useful for a quick check on the flow split and how well it is defined.
Where the data are very different in magnitude (ppm and per cent assays), it is usually
worth using weighted least squares for both the flow split and the data adjustments. If the
required data adjustments are large (more than ten per cent of the measurement) for more
than a few of the measurements, repeat the experiment. The total weighted sum of squares
should always be of the same order as the number of measurements, node by node or across
a complete circuit. Hence, it is worth examining each total by node and for each set of assays.
If the adjustments are small, there is a lot to be said for reconstituting concentrate and tailings
to generate a set of self-consistent data – at least two thirds of your numbers will be the ones
you measured.
Once you have checked each node for self-consistency, you can continue on to the multinode
balancing strategies outlined in the section on ‘Practical application – single level balancing’;
however, as the circuit becomes more complex, it is well worth considering a commercial
product.
The single node checks will often pinpoint measurement problems that may be concealed
by doing the complete balance first. Some commercial systems (such as JKSimMet) will let
you select each node – or a broader subset of the flow sheet – for a check balance.
Plotting measured data against adjusted data is a way of compressing a lot of data into a
single graph. It is also an excellent way of detecting bias. If all of the small assays are adjusted
upwards, there is likely a problem with low assay analyses.
For more complex data, such as size by assay or size by mineral, the same guidelines apply.
Consider the balance around each node with each size fraction before progressing on to total
assay balances across the circuit and then total size by assay balances across the circuit.
It is usually a poor strategy to attack the complete problem from the start as least squares
balancing can spread a single wrongly measured (or transcribed) piece of data across the
complete set. Such errors are much easier to identify and weed out in single node slices.
Keeping backup splits of samples allows for reassaying of dubious values when the rest of
the set are self-consistent.
For mineral-based analysis (MLA or Q*S) be aware of the number of particles or sections
that contribute to your measurement. Low number errors follow the Poisson distribution and
even 100 sections will have a relative sd of ten per cent.
Overall, try to use a systematic approach, starting at the node level and working across the
circuit as you test each data point for its actual contribution to the balance. While we may
never be able to know the ‘true’ value, careful measurement and analysis will produce data
that can provide a sound basis for flotation modelling and for decision-making.
CONCLUSIONS
The techniques outlined in this chapter provide a reasonably simple yet quite powerful
approach to analysis of flotation data. Using a spreadsheet for the process avoids capital
outlay but like all spreadsheet activity requires extreme care in checking and including checks
on constraints. While setting up these spreadsheets is a good way to learn about the analysis
process, they will not generally transfer between users.
For shared use – or anything at all complex – investing in one of the commercial, flow-
sheet-based products has much to recommend it. Some relevant product websites are listed
at the end of this chapter. Compared with even a very optimistic estimate of the true cost of
professional labour, the cost of commercially available mass balancing systems is really very
reasonable. A few days of time saved would pay for most of them.
Unless you are quite an expert user, building your own spreadsheets for more than the single
node cases is a high-risk endeavour. Using someone else’s spreadsheet is an excellent way to
generate a misleading analysis as the subsequent users do not know what modifications have
been made by others. If you are obliged to use a spreadsheet, the author strongly recommends
that you also use it to generate checks for self-consistency of your calculated or adjusted data
and graphs of all measured and adjusted data.
The well documented and secured spreadsheet offered by Luttrell (2004) may be a workable
compromise in some cases as it should not be subject to unauthorised and undocumented
‘improvements’.
ACKNOWLEDGEMENTS
Except as otherwise noted the data used for the examples in this chapter were provided
by Dr Chris Greet. The author gratefully acknowledges permission granted by the Elsevier
Scientific Publishing Company to reproduce Figure 7.8 and Table 7-IV from Lynch (1977).
REFERENCES
Boas, M L, 1966. Mathematical Methods in the Physical Sciences (Wiley).
Deming, W E, 1938. Statistical Adjustment of Data (Dover).
Gay, S L, 1999. Avoiding the zero-flow solution in mass-balance equations, Transactions of the Institutions
of Mining and Metallurgy, Mineral Processing and Extractive Metallurgy, 108:C121–126.
Hodouin, D and Everel, M D, 1980. A hierarchical procedure for adjustment and material balancing of
mineral processing data, International Journal of Mineral Processing, 7:91–116.
Luttrell, G H, 2004. Reconciliation of excess circuit data using spreadsheet tools, Coal Preparation,
24:35–52.
Lynch, A J, 1977. Mineral Crushing and Grinding Circuits, 340 p (Elsevier: Amsterdam).
Standards Australia, 1997. AS 2884.1-1997 Heavy mineral sand concentrates – sampling moving streams.
Wiegel, R L, 1972. Advances in mineral processing material balances, Canadian Metallurgical Quarterly,
11:413–419.
FURTHER READING
Montgomery, D C, Runger, G C and Hubele, N F, 2000. Engineering Statistics, second edition (Wiley).
Morrison, R D, 1991. Material balance techniques, in Evaluation and Optimization of Metallurgical
Performance, pp 209–218 (American Institute of Mining, Metallurgical, and Petroleum Engineers:
Denver).
Morrison, R D (ed), 2008. An Introduction to Metal Balancing and Reconciliation (Julius Kruttschnitt
Mineral Research Centre: Brisbane).
Morrison, R D and Richardson, J M, 1991. JKMBal – the mass balancing system, in Proceedings Second
Canadian Conference on Computer Applications for the Mineral Industry (CAMI ‘91), volume 1, pp 275–
286, Vancouver, September.
Napier-Munn, T J, 2014. Statistical Methods for Mineral Engineers: How to Design Experiments and Analyse
Data (Julius Kruttschnitt Mineral Research Centre: Brisbane).
Wills, B A and Napier-Munn, T J, 2006. Mineral Processing Technology, chapter 3 and associated
spreadsheet examples (Elsevier).
USEFUL WEBSITES
BilMat: <http://www.tpt.com/bilmat>.
JKSimMet, JKMBal, JKSimFloat: <http://www.jktech.com.au>.
CHAPTER 4
A practical guide to some aspects of
mineralogy that affect flotation
Alan R Butcher
ABSTRACT
This chapter is designed to introduce some of the basic aspects of mineralogy that can
influence flotation behaviour, with particular emphasis on sulfide flotation. Since the
invention of flotation in the early 1900s, the minerals industry has continued to refine and
continuously improve the many and varied chemical reagents and flotation cell technologies
during this period, yet it is only in the last decade or so that the role of mineralogy has truly
been realised as another important component in the better understanding of the flotation
separation process.
Most of the information required by mineral processing engineers involved in flotation
can now be provided by geologists and mineralogists using well-established optical, X-ray,
laser and electron-microbeam technologies. Metallurgically-relevant observations for
flotation processing applications can be made at any time during the life of a mine, from
first exploration drill core, which can flag-up problematic minerals and textures, through the
start-up phases of a mine, to full production, which can allow fine tuning of a concentrator
to produce acceptable grades and recovery. Finally, mineralogy can be used to optimise
recycling and/or disposal of tailings during the closure phase of the mining life cycle.
An understanding of the geological context, mineral assemblage and textures of an ore is
absolutely key to understanding its potential amenability to the flotation process. Equally
important is the measurement of sizes, composition, locking and liberation, and flotation
behaviour of particles as they pass through a circuit from blasting, crushing, grinding,
flotation to final concentration, using the appropriate mineralogical techniques, at the right
scale of observation.
This chapter will review the many and varied mineralogical techniques currently available,
and will guide the reader through when and where to use them. Examples will be drawn
from case studies that illustrate how mineralogy has allowed for a better understanding and
improvement in the flotation process on a range of ore types.
INTRODUCTION
An understanding of the geological context, mineral assemblage and texture of an ore is
absolutely key to efficient and effective unlocking and concentration of ore minerals during
blasting, crushing, grinding and flotation. Whilst this opening statement might appear
somewhat obvious to an experienced process mineralogist, it is seldom appreciated by
geologists, let alone by mining and minerals engineers. Let us explore why this is so.
The reasons for this are not always clear, or even logical, but experience has shown that it is
due in part to one or more of the following factors:
•• Historical – in the past, only mining of high-grade ores or easy-to-process ores was practised
and, generally speaking, these did not require strong geological input. In fact, in some of
the early mines, a geologist, let alone a mineralogist, may not even have been on staff.
•• Managerial – mineralogy is typically only considered and commissioned during the
exploration, feasibility and start-up stages, as part of due diligence exercises. It is not that
common for a mine manager to agree to regular mineralogical audits once an operation is
up and running, except of course when grade and recovery drop below acceptable limits.
Also, mining geologists rarely make it to mine manager level where they can influence
behaviour patterns in such matters.
•• Financial – there is real perception (which is also sometimes based on truth) that mineralogy
is costly, requires expensive equipment and involves highly qualified personnel who
write not particularly timely and relevant reports, which are all too often full of jargon,
and serve only to alienate the very audience for which they were intended.
Of course, the modern mining industry is quite different. Gone are the days where one could
choose to mine only the near-surface, high-grade, easy-to-process ores. With volatile metal
prices, heightened expectations from shareholders and unprecedented industrial growth
from China and India, the race is on to extract as much as possible from the earth to keep
up with demand. This must, of course, take place in a world where the awareness of human
safety and the environmental footprint of mining are coming under ever-increasing scrutiny.
All of these factors are challenging the viability of companies and mining operations every
day as we move to a world where mining lower-grade, more-difficult-to-process resources
will be the norm.
Given that our mantra as professionals involved in the earth resources industry could
be to ‘Take apart what Mother Nature has put together. Safely … with due regard for the
environment … and economically’, we really need to make effective use of the armoury
available to us to succeed in this combat exercise.
The science of mineralogy, and in particular that area of mineralogy dedicated to the
application of mineralogical information to mineral processing and smelting in order to
improve understanding, solve problems and improve efficiency of extraction, commonly
referred to as Process Mineralogy, is one such effective weapon in our mining armoury on
which we can draw. There is no doubt that process mineralogy, when used judiciously,
can make a major difference to a mining operation. In the author’s opinion, this was first
brought to the world’s attention in a seminal paper by Henley (1983), although many
others followed shortly thereafter (Hiemstra, 1984; Petruk, 1984; Baum et al, 1989). A text
book by Petruk attempts to pull together his experiences as a process mineralogist and is
a good source of case studies (Petruk, 2000). More recently, the Julius Kruttschnitt Mineral
Research Centre (JKMRC) has published a monograph on process mineralogy that contains
a number of case studies for a range of commodities, with chapters written by domain
experts (Becker, Wightman and Evans, 2016).
Before we move on, let us consider a hypothetical, but typical situation, which hopefully
will keep the sceptics interested and the novices alert.
The situation involves an ore type from the Eureka Concentrator. It needs investigating
because it is new. Routine assaying by the geologists in the mine suggests this ore type
contains a three weight per cent copper head grade, similar to all existing production ores.
The temptation is to go ahead, mine and crush the ore, and hope for the best. You might be
lucky. But a quick inspection of the ore by a mineralogist can immediately establish if the
copper is deporting to the usual phases (say sulfides), at the same grain size (which affects
grind size) and in the same host gangue (which affects flotation conditions) as the production
ore. If conditions are the same, fine. If they have changed, you could have saved yourself a lot
of expensive experimentation, not to mention embarrassment.
Mineral processing plants process minerals, and for mostly practical reasons, it is
accepted that conditions are often based on chemical assays rather than on mineralogical
information. Hopefully, your opinion on the use of mineralogy will change after reading
through this chapter.
accepted hypothesis as to how the mineral deposit formed. So we typically end up with a
physically descriptive classification that includes a descriptor of how the deposit formed.
Once one understands that certain ore types produce distinctive mineral assemblages and
textures (and geologists are trained to know this), then it is easier to focus on issues that are of
specific interest. Figures 1 and 2 summarise some of the basic ore deposit types and structures
likely to be encountered by the professional minerals engineer. For a more thorough review,
the interested reader is referred to Guilbert and Park, 1986; Evans, 1993; Roberts and Sheahan,
1988; Eckstrand, 1984; Peters, 1978 and McKinstry, 1948.
FIG 1 – Summary diagram illustrating the main types of ore deposit, classified
according to mode of origin, host-rock and commodity, with real examples.
Mineralisation
Stratiform Stratabound
Vein-hosted Fault-hosted
• Exploration methods
• Mining method
• Sampling Strategies
Disseminated Replacement
FIG 2 – Cartoons to illustrate how the geological structure of an orebody controls how you explore for it, how you mine it and how
you sample it. Here, six typical types are presented. Black is mineral or ore of interest. No scale (after Butcher and Trudu, 1999).
Take, for example, the deposit on which Eureka Mine is positioned. We know at the most
basic level that it is a polymetallic sulfide ore, and is mined underground. Armed with just this
information, an experienced geologist and process mineralogist can mentally map this to
be most likely a sediment-hosted, volcanogenic massive sulfide deposit, with an economic
mineral assemblage of sphalerite (Zn), galena (Pb), chalcopyrite (Cu).
With more information (such as age, geographic locality, geological model, tectonic history,
historical company reports), a mental checklist can be immediately created, and forms the basis
on which the mineralogist starts his or her campaign. It may include the following: look for
oxidation effects of sulfides; watch for grain size reduction if metamorphosed; look for talc
if sheared; check for more than one type of sphalerite; etc. More on this later in the chapter.
Ore mineral
FIG 3 – Cartoons to illustrate the basics of ore textures, with special reference to flotation. At the simplest
level, ores can be considered either equigranular (all grains same size) or inequigranular (not all same size).
The mineral of interest is shown in black. Grain boundaries are straight or gently curved. No scale.
phases – as this will control breakage mechanisms during blasting, crushing and grinding
(intragranular versus intergranular), and will make it either relatively easy, or difficult (or
impossible) to liberate and float effectively. It is not uncommon for an ore to have multiple
textural deportments for the element of particular interest (see Figure 5).
Finally, and for completeness, a further concept is now introduced. It was mentioned above
that orebodies are generally contained within old rocks. Practically, this means that they have
been around long enough to experience post-formation modification, which can take one of many
forms (metamorphism, deformation, oxidation and weathering, to name a few). Sometimes
these so-called secondary processes can overprint or obliterate those that developed at the
time the deposit formed (so-called primary features). Metamorphism, for example, is generally
bad news for the metallurgist as it often involves growth of new (and unwanted) minerals
Rims Inclusions
Interstitial Disseminated
Exsolved Laminated
FIG 4 – Further cartoons illustrating complex ore textures, with special reference to those which affect
liberation and flotation performance. Mineral of economic interest shown in black. No scale.
FIG 5 – One major outcome of a textural analysis of an ore is to ascertain the deportment of elements of economic
interest. In this case, the element of interest, shown in black (say copper) occurs as: chalcopyrite intergrown with
pyrite (upper left); chalcopyrite blebs included within sphalerite (upper right); discrete native copper (centre);
chalcocite rimming pyrite (bottom left); bornite inclusions within silicate (bottom right); and native copper within
a cross-cutting vein (bottom left to top right). Scale: width 10–100 microns (after Butcher and Trudu, 1999).
(such as mica, talc, graphite, serpentine, etc), which have an annoying habit of floating along
with the ore minerals to contaminate the concentrate – referred to here as menace minerals.
Deformation can lead to fracturing of the ore, recrystallisation and grain-size reduction, which
can change the comminution behaviour of an ore in a negative way (production of slimes).
Oxidation and weathering, however, can be both good and bad for a minerals engineer – it
can either upgrade (laterites, gossans) or downgrade (silcretes, calcretes and regoliths) the ore
depending on which process is dominant.
Figure 6 attempts to simplify what is obviously a very complex concept. The main point
being made here is that it is the job of the process mineralogist to: identify the presence of
these processes; then unravel them; and finally, understand how each one may impact on the
physical and chemical processing of the ore under investigation, and report these results in a
clear, easy to understand and practical way to the metallurgist.
Ore minerals
Pristine Ground Water Heat & geofluids Heat, pressure & fluids
FIG 6 – Cartoon illustrating some of the processes which are known to affect ores after initial formation. The
fate of three pristine ore textures is considered (one equigranular and two inequigranular), which undergo
progressively more aggressive modification from left to right (oxidation, hydrothermal alteration to metamorphism).
The job of the process mineralogist is to identify the presence of these processes, unravel them and understand
how each one may impact on the physical and chemical processing of the ore under investigation.
Hand-specimen analysis
Every self-respecting geologist and mineralogist should be able to gain information from
examination of materials in hand-specimen, using the naked eye or aided only by a hand
FIG 7 – Cartoon illustrating the effect of particle size reduction and increasing liberation from a complex, equigranular, ore texture,
where the mineral of economic interest is shown in black. Note that even the finest particle sizes still contain locked particles.
lens, a scratch plate and a bottle of acid. However, once materials become reduced in size
to the point where they are barely visible to the naked eye, other techniques clearly need
to be used.
Optical techniques
This probably still remains the most widely used technique worldwide, and not surprisingly,
it has been around almost as long as the science of geology. It turns out that cleverly designed
combinations of reflected and transmitted light sources, in both polarised and unpolarised states,
allow minerals to be identified by their characteristic optical behaviour, especially when
viewed under magnification on a petrographic polarising microscope. Ore minerals (such
as sulfides, precious metals and oxides) are best viewed in reflected light as they tend to
be opaque, whereas gangue phases (such as silicates, carbonates and phosphates) are best
examined in transmitted light as they are often translucent. Sample presentation normally
takes the form of polished sections or petrographic thin-sections, although loose grains can also
be usefully observed with a basic stereo microscope. Apart from the identification of minerals,
the modal mineralogy of a sample can be determined by point counting. Textures can be
observed and recorded photographically (photomicrographs) and typically are used to
augment reports and illustrate features of the sample. A useful text on this topic is by Gribble
and Hall (1992).
In each case, optical techniques obviously require a competent person to operate the
microscope and make meaningful observations, typically a graduate in geological sciences
with a few years of relevant experience. Great microscopists, however, are a greying
workforce and continue to be in serious decline and good microscopists are getting harder
to find. It is becoming more and more difficult to encourage young graduates to move into
optical mineralogy. Automation of phase identification, modal point counting and texture
analysis goes some way to redress this situation (see Clemex Technologies Inc, 2009; Carl
Zeiss Microscopy, 2018).
A clear advantage of this method is that, in the right hands, a quick (and often relatively
cheap) prognosis can be obtained for both unbroken ore as well as particulates. And while that
might be good enough, like all techniques, it must be used judiciously, and in combination
with other methodologies.
In certain situations, it can be a relatively quick and inexpensive method to solve a problem.
It is best used, however, in combination with other mineral techniques.
XRD, of course, is not to be confused with X-ray fluorescence spectrometry (XRFS), which
is another established X-ray beam technique, but unlike XRD, provides whole-rock chemical
data. It is most commonly used for grade determination (say, Fe-, Pb- and Zn-content of ores,
Ti-contents of mineral sands, P-content of iron ores, etc). Latest scanning XRFs now have the
capability of creating both elemental and mineral maps, on drill cores and particle mounts
(Bruker, 2018).
An emerging area of imaging and analysis is X-ray computed tomography – known in
short form as either X-CT, micro-X-CT, or nano-CT – depending on the size of the sample
to be analysed and the voxel resolution required. Although commonly used in the oil
industry, typically on drill cores, it is finding more application within mining, especially
with regards to understanding ore textures (Sayab et al, 2017) and 3D liberation (Garcia, Lin
and Miller, 2009).
Other techniques
Other technologies used by mineralogists include those that utilise proton beams, ion beams,
laser beams and infra-red and near infrared beams. A proton microprobe is ideal for mapping
ultra-trace elements within minerals. An example application at Eureka Mine would be to
examine the galena by PIXE to ascertain if silver was present in solid solution (Goodall and
Scales, 2006; Goodall, Scales and Butcher, 2005). A secondary ion mass spectrometer (SIMS)
can be used for accurately measuring invisible gold in sulfides – gold that is not possible to
see with SEM or EPMA microbeams as it is in solid solution (Vaughan and Corrans, 1977).
Fine grinding is not an option for this texture, but leaching could be.
Logging of cores, chips and powders for mineralogical composition can now be undertaken
using scanners that utilise near infrared radiation, which is phase specific. This is ideal for
on-site situations where rapid (near real-time) bulk mineralogical information is required to
troubleshoot processing problems or predict required conditions of operation as ores and
products pass along conveyors (Huntington, 2006; Shreeve et al, 2017).
FIG 8 – Optical photomicrographs, taken using a digital camera attached to a reflected polarising
petrographic microscope, of processed particles from an Australian lead–zinc operation (McArthur River),
which contain textures that historically would have been too fine for conventional flotation, and now require
combinations of staged and fine-grinding methods to separate the galena and sphalerite.
Methodology
First up, we need decent primary samples to examine. These must be representative of the
problem we are trying to fix. A good starting point is to always organise for the collection of
a feed, concentrate and tail for each of the main saleable products. Experience has shown that
you might need to involve all relevant personnel in all key areas of the mines’ activity in the
sampling, and to make them aware of the common pitfalls and problems in sampling mineral
processing plants (many of these issues are covered in this book in other chapters, and by
other publications, for example Annels (1991)). Sampling a dynamic system such as a mineral
processing plant, where crushing throughputs can be in the region of 100 kt/d, is certainly a
challenging exercise. Results of any kind always need to be viewed with this in mind.
Next, we need to create secondary samples from the primary samples, using appropriate
methods, which normally take the form of one or more of the following, all of which are
familiar to the metallurgist: cone and quartering, splitting, riffling, sieving, cyclosising and
de-sliming. From these secondary samples, we are then able to produce genuine replicate
aliquots, which can be used for all mineralogy and allied testing and analysis (assaying,
etc). Typically, these aliquots then get made into a number of products depending on what
happens next. Polished thin-sections and polished blocks are best as these can be used for
optical work, SEM and EPMA work, as well as automated SEM techniques.
Then, we need to generate quality petrographic information from the prepared products.
A typical checklist that most mineralogists might run through when examining samples
could look something like the following:
•• complete inventory of all known mineral phases present (Figure 9)
•• detailed modal analysis for major minerals (>5 per cent), minor phases (>1 per cent –
<5 per cent) and trace phases (<1 per cent)
•• textural information on both gangue and ore minerals, which includes:
•• grain size estimates (used to determine optimum liberation grain sizes and predict or
prevent problems such as poor liberation, over-grinding of the valuables, production
of slimes)
•• mineral associations (important for optimising the separation process, say galena from
sphalerite, chalcopyrite from sphalerite)
•• grain boundary relationships (curved, straight or irregular boundaries will affect behaviour
during crushing and grinding as this imparts preferential breakage mechanisms – for
example along or across boundaries)
•• elemental deportment information (important for tracking how metals behave during
processing; covered by Figure 5)
•• photomicrographs of typical textures.
All of the above can be generated manually from careful traditional optical-SEM-EPMA
interactive analysis. Automated techniques have enabled this type of information to be
collected and reported in a fast, robust, operator independent, time-efficient and scalable
manner.
Waste
Typical Process Overview
chalcopyrite – fast floating; pyrrhotite – slow floating depressants for carbonates and talc
Mineralogical Data
Roughers Liberation
Particle size
Particle composition
Cleaners Particle surface area
Problematic particles
Penalty elements
Re-grind Mill Mineral associations
Metallurgical Implications
• Reagent regime
Scavengers • Surface preparation
• Residence time
• Mechanical entrainment
• Flash flotation options
• Staged grinding
• Grinding media
FIG 10 – Idealised view of what mineralogical data can be extracted from samples taken from a mineral
processing circuit, and the implications these data may have on metallurgical behaviour and practices.
Lead-zinc Galena
deposit
-
+
2
6
0
0
/
0
+
0
/
/
+
+
2
5
0
4
5
5
6
0
G a l e n a B i - S i z e D i s t r i b u t i o n
FIG 11 – An image grid of galena-containing particles (red) from crushed drill core, organised by galena grain-size estimate,
as measured by a QEMSCAN® and plotted using iExplorer™. The distribution shows a distinct bi-modal pattern.
Lead-zinc Sphalerite
deposit
-
+
2
6
0
0
/+
/+
0
+
0
/+
/+
/
+
+
2
5
0
4
5
5
6
S p h a le r it e S i z e D i s t r ib u t i o n M T G A
FIG 12 – An image grid of sphalerite-containing particles (blue) from crushed drill core, organised by sphalerite grain‑size
estimate, as measured by a QEMSCAN® and plotted using iExplorer™. The graph shows a normal distribution.
Answer
Given that a two-stage grinding circuit was not possible at Eureka, an optimal grind size of
212/+106 microns was trialled during bench-scale grinding tests, and was found to achieve
the required grade and recovery.
Answer
Historic samples taken on a good day were compared with those containing high iron contents.
The results showed that Fe deports to both pyrite and Fe-rich sphalerite. Both phases were
more abundant in the final concentrate that had higher Fe-content (Figure 13).
Fe-Sphalerite
Pyrite
High Fe-Concentrate
CS2 CS3+4 CS5 CS2 CS3+4 CS5
100
90
80
70
60
50
40
30
20
10
FIG 13 – Comparison of samples from the final zinc concentrate, good and bad, in terms of Fe-content.
It illustrates that when iron increases, it is due to increases in both Fe-sphalerite and pyrite.
Background
Fe Sulfides
Sphalerite
Galena
Other Sulfides
NSG
FIG 14 – Image grid of particles from a feed sample, sorted according to the size fraction from which they were
measured, and according to the texture they exhibit. The results show that the locking characteristics of galena are quite
complex and involve: locking with sphalerite (high- and low-grade ternary), Fe-sulfide, non-sulfide gangue (NSG).
Both barren and fully liberated particles are also present. Measured on QEMSCAN® and plotted using iExplorer™.
Answer
Finer grinding of the feed ore is perhaps suggested. This will liberate the composite particles.
Installation of secondary grinding equipment might be the solution, such as stirred mills,
which will help reduce particle size to the required liberation size range. Bench-scale testing
could validate this theory.
CONCLUDING REMARKS
It is the hope of the author that, after having gone through this chapter, the reader will take
something away of interest, and most importantly, something of use, which can be directly
applied to their everyday work.
No one expects or wants minerals engineers to become mineralogists or geologists – far
from it. What the author encourages, however, is that metallurgists, like geologists, mining
engineers and environmentalists, should take some of the responsibility for developing a
culture of cross-pollination of professional skill areas within the modern minerals industry.
Mineralogy is clearly useful when applied in the correct way. Rocks contain secrets that
traditionally only geologists and mineralogists have been able to unlock. With the advent of
REFERENCES
Annels, A E, 1991. Mineral Deposit Evaluation, first edition, 436 p (Chapman and Hall).
Barbery, G, 1991. Mineral liberation: measurement, simulation and practical use in mineral processing,
Quebec Editions GB, p 341.
Baum, W, Lotter, N O and Whittaker, P J, 2004. Process mineralogy – a new generation of ore
characterisation and plant optimization, SME Annual Meeting, Denver, 23–25 February 2004, pp 1–5.
Baum, W, Sanhueza, J, Smith, E and Tufar, W, 1989. The use of process mineralogy for plant optimization
at the El Indio gold-silver-copper operation (Chile), Erzmetall, 42(9):373–378.
Becker, M, Wightman, E M and Evans, C L, 2016. Process Mineralogy, SMI-JKMRC Monograph No 6
(Julius Kruttschnitt Mineral Research Centre: Indooroopilly), p 470.
Bruker, 2018. Bruker web site. Available from: <https://www.bruker.com/products/x-ray-diffraction-
and-elemental-analysis/micro-xrf-and-txrf/m4-tornado/overview.html>.
Butcher, A R and Trudu, A G, 1999. Ore characterisation and its relevance to the crushing and grinding
of ores, presented to AJM Crushing and Grinding Conference, Kalgoorlie, 27–28 April 1999.
Carl Zeiss Microscopy, 2018. Zeiss web site. Available from: <https://www.zeiss.com/microscopy/int/
solutions/raw-materials.html>.
Clemex Technologies Inc, 2009. Clemex web site. Available from: <http://www.clemex.com>.
Craig, J R and Vaughan, D J, 1994. Ore Microscopy and Ore Petrography, second edition, 434 p (John Wiley
and Sons, Inc).
Dou, M, 1998. The segmentation of BSE images of mineral particles using an object based approach, in
Proceedings Seventh JKMRC Conference, Brisbane.
Eckstrand, O R, 1984. Canadian Mineral Deposit Types: A Geological Synopsis, Geological Survey of Canada,
Economic Geology Report 36, 86 p (Canadian Government Publishing Centre).
Evans, A M, 1993. Ore Geology and Industrial Minerals – An Introduction, third edition, 389 p (Blackwell
Scientific Publications Inc).
Garcia, D, Lin, C L and Miller, J D, 2009. Quantitative analysis of grain boundary fracture in the
breakage of single multiphase particles using X-ray microtomography procedures, Minerals
Engineering, 22(3):236–243.
Goodall, W R and Scales, P J, 2006. Proceedings MEI Automated Mineralogy ’06, Brisbane.
Goodall, W R, Scales, P J and Butcher, A R, 2005. The use of QEMSCAN and diagnostic leaching in the
characterisation of visible gold in complex ores, Minerals Engineering, 18:877–866.
Gottlieb, P, Wilkie, G, Sutherland, D, Ho-Tun, E, Suthers, S, Perera, K, Jenkins, B, Spencer, S,
Butcher, A and Rayner, J, 2000. Using quantitative electron microscopy for process mineralogy
applications, JOM, April, 52:24–25.
Grant, G, Hall, J S, Reid, A F and Zuiderwyk, M, 1977. Characterisation of particulate and composite
mineral grains by on-line computer processing of SEM Images, in Proceedings 15th APCOM Symposium,
Brisbane, pp 159–170.
Gribble, C D and Hall, A J, 1992. Optical Mineralogy, Principle and Practice, p 295 (UCL Press).
Gu, Y, 2003. Automated scanning electron microscope based mineral liberation analysis – an introduction
to JKMRC/FEI mineral liberation analyser, Journal of Minerals and Material Characterisation and
Engineering, 2(1):33–41.
Guilbert, J M and Park Jr, C F, 1986. The Geology of Ore Deposits, 985 p (W H Freeman and Company).
Henley, K J, 1983. Ore Dressing Mineralogy – A Review of Techniques, Applications, and Recent Developments,
special publication no 7, pp 175–200 (Geological Society of South Africa: Marshalltown).
Hiemstra, S A, 1984. The role of mineralogy in ore beneficiation as illustrated by case histories, in
Proceedings Second International Congress on Applied Mineralogy, Los Angeles, 22–25 February, pp 9-20.
Huntington, J, 2006. The hi-logger scanner for automated mineralogy, presented to MEI Automated
Mineralogy ’06, Brisbane.
Ixer, R A and Duller, P R, 1998. Virtual atlas of opaque and ore minerals in their associations. Available
from: <http://www.smenet.org/opaque-ore>.
Jones, M P, 1984. Recent developments in the rapid collection of quantitative mineral data, in Proceedings
Second International Congress on Applied Mineralogy, Los Angeles, 22–25 February, pp 141–156.
Jones, M P, 1987. Applied Mineralogy: A Quantitative Approach, p 253 (Graham and Trotman, Springer).
Keith, J, 1998. Mineral identification using X-rays, in Proceedings Seventh JKMRC Conference, Brisbane.
King, R P and Schneider, C L, 1993. An effective SEM-based image analysis system for quantitative
mineralogy, KONA, 11:165.
McKinstry, H E, 1948. Mining Geology, 680 p (Prentice Hall: Englewood Cliffs).
Miller, P R, Reid, A F and Zuiderwyk, M A, 1982. QEM*SEM image analysis in the determination of
modal assays, mineral associations and mineral liberation, presented to XIV International Mineral
Processing Congress, Toronto.
Oxford Instruments, 2018. Oxford Instruments web site. Available at: <https://www.oxford-instruments.
com/products/microanalysis/energy-dispersive-x-ray-systems-eds-edx/eds-for-sem/eds-software-
aztec>.
Peters, W C, 1978. Exploration and Mining Geology, 696 p (John Wiley and Sons Inc).
Petruk, W, 1984. Image analysis measurements and data presentation for mineral dressing applications,
in Proceedings Second International Congress on Applied Mineralogy, Los Angeles, 22–25 February,
pp 127–140.
Petruk, W, 2000. Applied Mineralogy in the Mining Industry, p 286 (Elsevier Science).
Pirrie, D, Butcher, A R, Power, M P, Gottlieb, P and Miller, G L, 2004. Rapid quantitative mineral and
phase analysis using automated scanning electron microscopy (QEMSCAN®); potential applications
in forensic geoscience, Journal of the Geological Society of London, 232:123–136.
Roberts, R G and Sheahan, P A, 1988. Ore Deposit Models, 194 p Geoscience Canada, Reprint Series 3
(Geological Association of Canada).
Sandmann, D, 2015. Method development in automated mineralogy, Doctor Rerum Naturalium
Dissertation, Technischen Universität Bergakademie Freiberg, p 152.
Sayab, M, Miettinen, A, Aerden, D and Karell, F, 2017. Orthogonal switching of AMS axes during type‑2
fold interface: insights from integrated X-ray computed tomography, AMS and 3D petrography,
Journal of Structural Geology, 103:1–16.
Schneider, C L and Neumann, R, 2004. Measurement of exposed area by image analysis particle cross-
section (eds: Pecchio et al), in Proceedings 2004 ICAM, Sao Paulo, pp 1035–1038.
Shreeve, J W, Shreeve, B, Pitel, J, Palix, E and Souf, A, 2017. The use of non-destructive core logging
and X-ray imaging techniques to resolve a complex geological stratigraphy for a planned offshore
windfarm, in Proceedings Eighth International Conference of Offshore Site Investigation and Geotechnics,
12–14 September 2017 (Royal Geographical Society: London).
Spencer, S and Sutherland, D N, 2000. Stereological correction of mineral liberation grade distributions
estimated by single sectioning of particle images, Image Anal Sterel, 19:175–182.
Sutherland, D, Gottlieb, P, Jackson, R, Wilkie, G and Stewart, P, 1988. Measurement in section of
particles of known composition, Minerals Engineering, 1(4):317–326.
Sutherland, D N, Wilkie, G and Johnson, C R, 1989. Simple prediction of processing characteristics of
ores using QEM*SEM, in Proceedings Mineralogy and Petrology Symposium, pp 81–84 (The Australasian
Institute of Mining and Metallurgy: Melbourne).
Tescan, 2018. Tescan web site. Available at: <https://www.tescan.com/en-us/technology/special-
solutions/tima-x>.
ThermoFisherScientific 2018. ThermoFisherScientific web site. Available at: <https://www.fei.com/oil-
gas/products-services/>.
Vaughan, J and Corrans, I J, 1997. Estimation of submicroscopic gold in ores by SIMS and ultrafine
milling: a comparative investigation, (eds: G Gillen, R Lareau, J Bennett and F Stevie), in SIMS XI
Annual Workshop, Orlando, 7–12 September (Wiley and Sons).
Xstrata Technology, 2005. Quote from Xstrata Technology company profile published in Proceedings
Centenary of Flotation Symposium (The Australasian Institute of Mining and Metallurgy: Melbourne).
CHAPTER 5
Some troublesome sulfide
and non-sulfide minerals
Bill Johnson and Peter Munro
ABSTRACT
The abundance of each mineral in an ore sample allows early recognition of likely problems in
development of a processing circuit for a new orebody or in improvement of the performance
of an existing processing circuit.
This chapter discusses various aspects of the properties of selected valuable sulfide minerals
which affect the ease of preparation of a concentrate of marketable grade suitable for safe
transportation.
This chapter also reviews the properties of two common sulfide gangue minerals (pyrite and
pyrrhotite) that are known to cause processing problems. Troublesome non-sulfide gangue
minerals are also reviewed, eg those displaying hydrophobicity, those prone to causing
penalty payments for the concentrate and those with unusual or troublesome properties.
INTRODUCTION
Quantitative data on the percentage by weight of each valuable and gangue mineral present
in an ore is one type of data that is mandatory before commencement of the development of
a processing circuit. Such mineral abundance data allow early recognition of likely issues in
development of a processing circuit:
•• problems in both the preparation of a concentrate of marketable grade and its transportation
and sale, due to solid solution series, the presence in the valuable minerals of elements
incurring a penalty or the presence of elements affecting its fitness for safe transport
•• recognition of sulfide gangue minerals with some properties causing processing problems
and other properties that may provide a basis for process solutions
•• recognition of non-sulfide gangue minerals known to be hydrophobic in other ores, non-
sulfide gangue minerals containing elements incurring a penalty and non-sulfide gangue
minerals with unusual properties that cause processing problems.
20
15
10
FIG 1 – Calculated relationship between the per cent by weight of gangue minerals that can be present in the
zinc concentrate while still producing a concentrate with 52 per cent by weight of zinc (y axis) and the per cent by
weight of Fe and other elements in solid solution in the sphalerite/marmatite solid solution series (x axis).
This is a realistic target because typical commercial zinc concentrates contain in the region
of 90 per cent valuable mineral by weight.
of 26 per cent by weight (Riley, 1974, 1976; Knights and Patterson, 1988; Knights and Johnson,
2016). If freibergite contains a significant portion of the silver in an ore, the silver recovery
will be dictated by the liberation state of the freibergite and, for the non-liberated portion,
the behaviour of the mineral with which it occurs in composites. The composites with galena
(the lead mineral) will be recovered in the desired concentrate for silver, the lead concentrate,
while the composites with sphalerite (the zinc mineral) are likely to be recovered in the zinc
concentrate, which is less desirable commercially. The composites with iron sulfide gangue
minerals will be largely lost to the tailing.
Chalcocite (Cu2S)
Why can chalcocite be troublesome? Chalcocite oxidises rapidly and releases copper ions
which can have deleterious effects on some other minerals in the system, particularly pyrite,
pyrrhotite and sphalerite.
Impact on feasibility study/existing flotation plant: It is necessary to investigate the influence of
conditioning pH (from pH 5 to 11.5) and the use of NaHS in restoring the surface of oxidised
chalcocite to its original form just before the addition of collector for the flotation process.
In the flotation testing, it is necessary to obtain chalcocite recovery – size (log scale) curves
to compare the recovery obtained in the fine and intermediate size fractions ensuring that
the recovery of the chalcocite in the fine size fractions is maximised relative to that in the
intermediate fraction.
Solution: Use the findings from the described approach to testing to establish the necessary
chemical conditions for flotation of the chalcocite or its closely related minerals in the ore.
Chalcocite is one of the most important source minerals for copper. A well-known occurrence
of chalcocite is in the supergene enriched zone of sulfide deposits. Another is in iron oxide-
hosted copper-gold deposits where it occurs as a primary mineral in a well-defined zone, not
from supergene enrichment. An unusual property of chalcocite is that, with suitable values
for Eh and pH, it will oxidise rapidly, in minutes, to other copper sulfide phases and release
copper ions into the system which can have deleterious effects on the other minerals present
in the system.
The oxidation of chalcocite occurs through a series of non-stoichiometric copper sulfides as
shown in Table 1 (Woods, personal communication, 2015), based on copper sulfide phases
identified in the literature (Goble, 1985).
From Table 1, in conversion of chalcocite to djurleite, 1.5 per cent of the copper initially
present is released; in conversion of chalcocite to djurleite and then to digenite, a total of
10 per cent of the copper initially present is released. Eventually, 50 per cent of the copper is
released after moving from chalcocite (Cu2S) through the series of non-stoichiometric copper
sulphides to covellite (CuS), as dictated by the formulae of chalcocite and covellite. The values
TABLE 1
The oxidation of chalcocite (Cu2S) through a series of non-stoichiometric copper sulfides.
Copper sulfide Chalcocite Djurleite Digenite Anilite Geerite Spionkopite Yarrowite Covellite
phase
Formula Cu2S Cu1.97S Cu1.8S Cu1.75S Cu1.6S Cu1.4S Cu1.12S CuS
Copper released by
oxidation of Cu2S 0 1.5 10 12.5 20 30 40 50
(% of total in Cu2S)
for the released copper are based on conversion of all the chalcocite in the starting material,
ie both the surface and bulk chalcocite.
During the oxidation sequence in Table 1 through the non-stoichiometric copper sulfide
phases, the copper ions are very mobile in the bulk and surface regions of the phases.
In contrast, for the end point of the sequence, covellite, the copper ions are not mobile, ie it
has a ‘frozen’ lattice.
For the oxidation of chalcocite through the non-stoichiometric copper sulfide phases
to covellite, the thermodynamics of the direct conversion from chalcocite to covellite is
misleading and should not be considered (Woods, personal communication, 2015).
The copper which is released from the copper sulfide phases (ie not part of the non-
stoichiometric copper sulfide phases) may contribute to an acid soluble copper assay for a
sample containing nominally chalcocite.
For the Ok Tedi deposit, industrial experiences and supporting research work with chalcocite
(principally occurring as digenite in one orebody) were carefully documented (Orwe, Grano
and Lauder, 1997; Orwe, Grano and Lauder, 1998); for the Eh and pH conditions in the Ok
Tedi plant, the chalcocite was likely to be unstable, ie undergoing surface oxidation.
The authors studied the influence of conditioning pH and the use of NaHS in restoring
the surface of the oxidised chalcocite to its original form (Orwe, Grano and Lauder, 1998)
before introducing collector into the system for adsorption. However, surface re-oxidation
before the flotation process should be minimised. After exposure to oxidising conditions, the
authors found that fine particles of chalcocite (-10 µm) suffered greater adverse effects on
recovery when compared with the intermediate size fractions of the same mineral. The use of
NaHS minimised the collector required for flotation as it precipitated the hydrophilic metal
hydroxides, present at both the chalcocite surface and in solution, as a very insoluble metal
sulfide (Orwe, Grano and Lauder, 1997), because there are less hydrophilic species on the
chalcocite surface to lower the hydrophobicity conferred by the adsorbed collector.
2. process changes to introduce washed froth cleaning technology or increase the proportion
of the concentrate produced by washed froth cleaning technology
3. process changes incorporating points 1 and 2
4. process changes to regrind to a finer size and increase the liberation of the non-sulfide
gangue
5. process changes incorporating point 4 and points 1 and/or 2.
In points 1, 2, 3 and 5, the process improvement lessens the contribution from recovery
by the entrainment mechanism of non-sulfide gangue containing sources of radioactivity.
If sufficient lowering of the radioactivity level is not achieved by these options, other methods
directed at the valuable sulfide mineral itself may be needed; such options are topics for
research projects at present and cannot be addressed here.
with the framboidal texture represents a metallurgical problem in the copper and zinc-lead
ores at and close to Mount Isa (Croxford, Draper and Harraway, 1961) and for the zinc-
lead ores of other deposits in the Carpentaria Inlier (McArthur River, Century and Dugald
River). The spheroidal pyrite in the framboidal texture contains hydrocarbons derived from
bacteria at the time of ore formation. In these orebodies, the other minerals are affected only
slightly by the hydrocarbons in comparison with the spheroidal pyrite. The spheroidal pyrite
contains fine grains, typically 5–15 microns in diameter, and has the hydrocarbons existing on
its surfaces after breakage, causing this portion of the pyrite to display an intermediate level
of hydrophobicity (in the presence of frother and in the absence of collector) (Johnson and
Jowett, 1982). Dilution of the flotation concentrate occurs with this type of hydrophobic pyrite
in the feed as it is not affected by the variables used to control the euhedral pyrite in flotation.
This spheroidal pyrite is sometimes given the name ‘carbonaceous pyrite’; it is porous and
highly reactive.
Porous pyrite with high reactivity can also exist in some ores in a colloform texture which
results from hot metal bearing solutions contacting cold ocean water during ore formation,
conditions allowing little time for crystals to form. The pyrite formed has been described
(McArthur, personal communication, 2018) as the mineral melnikovite, a primitive pyrite.
Later recrystallisation results in voids, ie the porosity noted earlier. This component can be
a carrier of elevated levels of precious metals such as gold and it is important to understand
this occurrence.
Pyrite has a polymorph, marcasite, which can be present with pyrite in an ore, usually in
small amounts, marcasite being much less common than pyrite in general. Pyrite (specific
gravity 5.02) belongs in the cubic crystal system while marcasite (specific gravity 4.89) is
in the orthorhombic crystal system. Marcasite is less stable than pyrite, decomposing more
readily than pyrite into melanterite (FeSO4 7H2O).
For metallurgical characterisation of ores with a gangue mineral displaying moderate
hydrophobicity such as those with spheroidal pyrite in the Carpentaria Inlier, the direct
method is to use a standardised pre-flotation rougher test; this requires multiple staged
concentrates to be collected with frother as the necessary reagent in the test. It is necessary to
use the assay results on a mineral basis, not on an element basis; this requires an element assay
to mineral assay conversion system for the ore in question. For the ores in the Carpentaria
Inlier, the important pieces of information are listed:
•• pyrite head grade
•• percentage of the pyrite in the feed which is recovered in the pre-flotation test, ie that has
natural hydrophobicity (typically in the range ten per cent to 35 per cent).
The worst case is potentially an ore with a high pyrite head grade with a high percentage
of the pyrite displaying hydrophobicity as measured in the pre-flotation test. It is believed
that the distribution of the hydrocarbon is important in maximising the quantity of solid
displaying unwanted hydrophobicity, and may be more important than the quantity of
hydrocarbon in the feed.
Forbes, Smith and Vepsalainen (2018) compared the galvanic interaction of chalcopyrite with
two distinctly different types of pyrite recognised using several textural and electrochemical
characterisation methods. In flotation tests with the chalcopyrite and pyrite 1 and with the
chalcopyrite and pyrite 2 as the feed, the flotation response of the chalcopyrite was the same
but, for pyrites 1 and 2, the flotation response differed. This finding indicates that additional
characterisation methods for pyrite may need to be used in the future.
TABLE 2
Data for naturally occurring pyrrhotites (from Multani and Waters, 2017).
Ideal Vacancy (%) Charge neutral Ideal atomic composition (%) Magnetic
composition formula Fe S properties
Fe11S12 8.3 Fe2 Fe9 S12
3+ 2+ 2-
47.83 52.17 Non-magnetic
Fe10S11 9.1 Fe23+Fe82+S112- 47.62 52.38 Non-magnetic
Fe9S10 10.0 Fe2 Fe7 S10
3+ 2+ 2-
47.37 52.63 Non-magnetic
Fe7S8 12.5 Fe2 Fe5 S8
3+ 2+ 2-
46.67 53.33 Magnetic
magnetic properties and the importance of understanding the types present in an orebody
and variation in the relative amounts of these types over the deposit can be recognised.
In the past, magnetic separators were common in processing ores in which pyrrhotite was
a gangue sulfide mineral and the varieties of pyrrhotite were directed into separate magnetic
and non-magnetic circuits (Table 2), based on their magnetic properties, and subjected to the
flotation process in independent circuits. More recently, operations have tended to eliminate
magnetic separation in the processing plants with the varieties of pyrrhotite being processed
in a single flotation circuit where it has been recognised that there is a wide range in flotation
properties for the varieties of pyrrhotite in real systems (Multani and Waters, 2017).
In the area of flotation properties, the focus has been on Fe7S8 pyrrhotite and Fe9S10 pyrrhotite,
being the most common and occurring together in intergrown mixtures in the ore texture.
It has been necessary to examine the recovery – size (log scale) graphs of each to understand
the differences in their behaviour in real systems. For Fe9S10 pyrrhotite, a recovery >60 per cent
is observed for the intermediate size fractions, while this is not the case for Fe7S8 pyrrhotite
which produces a poor response in the fine, intermediate and coarse size fractions (Multani
and Waters, 2017).
‘Clay Minerals in General’, brief reference is made to a recent publication (Grafe et al, 2017)
on clay minerals.
Sericite
Why can sericite be troublesome? Sericite can contribute the penalty element, F, to the concentrate
by the mechanisms of entrainment and in composites with the valuable mineral. Sericite,
a fine-grained member of the mica group, may be capable of interacting with the valuable
mineral, reducing the flotation rate of the valuable mineral.
Impact on feasibility study/existing flotation plant: Investigate the main minerals contributing
to the total F in the concentrate. If sericite is one of the main minerals, staged options exist
in Solution. Investigate if the sericite can adhere to the valuable mineral or interact with it in
other ways if the flotation rate of the liberated valuable mineral is observed to be diminished.
Solution: Investigate the benefits of reduced entrainment of liberated sericite (more dilute
pulps in conventional flotation cells and/or use of cells designed for washed froth cleaning)
and, if necessary, the benefits from further regrinding to increase sericite liberation.
Sericite has the composition of muscovite and it is the occurrence of muscovite in some
schistose rocks as fibrous aggregates of minute scales with a silky lustre that has resulted in
the use of the term, sericite (Klein, 2002). For muscovite and sericite, fluorine can substitute
in the locations normally occupied by the hydroxyl group. Various other elements can also
substitute to a limited extent in the locations occupied normally by potassium and aluminium.
For a particular ore, sericite can contribute to the total fluorine content of the final concentrate.
If the dominant occurrence of the sericite is in composite particles with the valuable mineral in
the concentrate, additional regrinding will be required to improve its liberation and increase
its rejection. If the dominant occurrence of sericite in the concentrate is in fine liberated
particles, steps will need to be taken to lower the contribution of the entrainment mechanism
in the recovery of sericite. As part of this procedure, process data should be analysed to
confirm that the sericite is reporting to the final concentrate in each size fraction as expected
from the entrainment mechanism.
Silvester et al (2011) noted that sericite has been viewed historically as a troublesome gangue
mineral. The authors also noted that it was possible to control its recovery by ‘judicial use of
dilution cleaning and/or froth washing’; this finding suggested that the sericite was being
recovered by the entrainment mechanism. For ore with elevated sericite content, it is therefore
important to establish if the liberated sericite follows the guidelines for entrainment in each
size fraction, noting that the sericite particles may possess an unusual shape.
such as carboxy methyl cellulose (CMC), derivatives of guar gums, locust bean gums and
others where these reagents are added in cleaning and/or roughing. Additional approaches
to remove these minerals are a pre-flotation step preceding the rougher section and reverse
flotation of the final concentrate.
At the commencement of ore characterisation in laboratory testing of an ore, a standardised
pre-flotation rougher test should be used to evaluate the extent to which minerals in the feed,
including talc or pyrophyllite, are being recovered in quantities in excess of those expected
from entrainment. The pre-flotation test requires the use of frother and the absence of
collector; it is possible that a depressant for a gangue sulfide mineral may be used depending
on the system. A standardised pre-flotation rougher test as an ore characterisation method
was discussed earlier in the section on the mineral pyrite.
Both talc and pyrophyllite are industrial minerals which are often discussed together
because of their similarity (Tomaino, 2018).
Chrysotile (Mg3Si2O5(OH)4)
Why can chrysotile be troublesome? Chrysotile displays a positive surface charge at alkaline
pH values, causing it to adhere to the negatively charged valuable mineral in this pH region,
lowering the hydrophobicity and flotation rate of the valuable mineral and also lowering
concentrate grade. Additionally, it is fibrous, forming cylindrical tubes, and has an extremely
detrimental effect on pulp rheology. Its most common occurrence is in serpentinitic nickel
sulfide ores.
Impact on feasibility study/existing flotation plant: Investigate the use of depressants which
negate the positive surface charge on the chrysotile and investigate circuitry and conditions
which reduce its detrimental effects such as separate flotation of size fractions and the use of
very dilute pulps.
Solution: Implement the most cost effective methods for reducing the deleterious effects.
Most base metal sulfide ores are processed by froth flotation in the alkaline region.
Fortunately, most sulfide minerals (both valuable and gangue) and most non-sulfide gangue
minerals possess a negative surface potential in the alkaline region. This contributes to the
dispersion of the particles due to similarity of the sign of their surface charge.
An exception to the described situation is the mineral chrysotile (or closely related minerals)
which possesses a positive surface potential in the alkaline region and results in it adhering
to the other negatively charged particles, including the particles containing the valuable
mineral. This hydrophilic mineral which is adhering to the valuable mineral increases the
difficulty in obtaining a sufficiently hydrophobic surface on the valuable particles by collector
addition, thus reducing their flotation response.
Chrysotile is classified in the serpentine group of minerals (Klein, 2002), which contains
three common polymorphs: antigorite, lizardite and chrysotile. The occurrences of antigorite
and lizardite are commonly massive and fine grained while the occurrence of chrysotile is
fibrous, forming cylindrical tubes. Another deleterious effect from chrysotile’s structure is
that it has the most extreme effect on pulp rheology of all the clay minerals (Forbes, Ma and
Bruckard, 2017).
The recognition of chrysotile or closely related minerals in an ore from initial determination
of the ore’s mineral abundance data must be used to provide a warning of potential
metallurgical difficulties in process development. The most common occurrence of this
problem is with sulfide nickel ores with serpentine gangue minerals and for which the effect
of chrysotile is documented (Edwards, Kipkie and Agar, 1980).
ACKNOWLEDGEMENT
The important contribution of Suzanne Munro (Mineralurgy Pty Ltd/Mineralis) in
preparation of diagrams and of other aspects for the content in this chapter is acknowledged
with gratitude.
REFERENCES
Becker, M, de Villiers, J and Bradshaw, D, 2010a. The mineralogy and crystallography of pyrrhotite
from selected nickel and PGE ore deposits, in Economic Geology, 105:1025–1037.
Becker, M, de Villiers, J and Bradshaw, D, 2010b. The flotation of magnetic and non-magnetic pyrrhotite
from selected nickel ore deposits, Mineral Engineering, 23:1045–1052.
Croxford, N W J, Draper, N and Harraway, D H, 1961. Some aspects of the carbonaceous fraction of
Mount Isa lead concentrate, The AusIMM Proceedings, 197:149–161.
Edwards, C, Kipkie, W and Agar, G, 1980. The effect of slime coatings of the serpentine minerals,
chrysotile and lizardite, on pentlandite flotation, International Journal of Mineral Processing, 7:33–42.
Ekmekci, Z, Becker, M, Bagci Tekes, E and Bradshaw, D, 2010. The relationship between the
electrochemical, mineralogical and flotation characteristics of pyrrhotite samples from different Ni
ores, Journal of Electroanalytical Chemistry, 647:133–143.
Forbes, E, Ma, M and Bruckard, W, 2017. Clay minerals in flotation and comminution operations, in
Clays in the Minerals Processing Value Chain (eds: M Grafe, C Klauber, A J McFarlane and D J Robinson),
pp 302–326 (Cambridge University Press).
Forbes, E, Smith L and Vepsalainen, M, 2018. Effect of pyrite type on the electrochemistry of
chalcopyrite/pyrite interactions, Physicochem Probl Miner Process, 54(4):1116–1129.
Goble, R J, 1985. The relationship between crystal structure, bonding and cell dimensions in the copper
sulfides, in Canadian Mineralogist, 23(1):61–76.
Grafe, M, Klauber, C, McFarlane, A J and Robinson, D J (eds), 2017. Clays in the Minerals Processing
Value Chain (Cambridge University Press).
Johnson, N W and Jowett, A, 1982. Comparison of lead primary rougher behaviour of several Mount
Isa lead zinc ores, in Proceedings Mill Operators’ Conference, pp 271–285 (The Australasian Institute of
Mining and Metallurgy: Melbourne).
Kerr, A, 2002. Overview of recent developments in flotation technology and plant practice for nickel
ores, in Proc of Minerals Processing Plant Design, Practice and Control Vol 1 (eds: A Mular, D Halbe and
D Barratt), pp 1142–1158 (Society for Mining, Metallurgy, and Exploration: Littleton).
Klein, C, 2002. Mineral Science (The 22nd Edition of the Manual of Mineral Science, after James D. Dana) (John
Wiley and Sons, Inc).
Knights, J G and Johnson, N W, 2016. Silver mineralogy and metallurgy of lead-zinc-silver deposits at
Mount Isa, in Process Mineralogy (Chapter 24) (eds: M Becker, E M Wightman and C L Evans), JKMRC
Monograph Series in Mining and Mineral Processing: No 6, pp 344–364 (SMI-JKMRC, The University
of Queensland).
Knights, J G and Patterson, D J, 1988. Process mineralogy of silver in lead-zinc-silver ores at Mount Isa,
Australia, in Proceedings International Conference ‘Silver-exploration, Mining and Treatment’, pp 101–110
(Institution of Mining and Metallurgy).
Multani, R S and Waters, K E, 2017. A review of the physicochemical properties and flotation of
pyrrhotite superstructures (4C – Fe7S8/ 5C – Fe9S10) in Ni-Cu sulphide mineral processing, in Canadian
Journal of Chemical Engineering, 96(5):1185–1206.
Orwe, D, Grano, S R and Lauder, D W, 1997. Chalcocite oxidation and its influence on fine copper
recovery at the Ok Tedi Concentrator, Papua New Guinea, in Proceedings Sixth Mill Operators’
Conference, pp 85–95 (The Australasian Institute of Mining and Metallurgy: Melbourne).
Orwe, D, Grano, S R and Lauder, D W, 1998. Increasing fine copper recovery at the Ok Tedi
Concentrator, Papua New Guinea, Minerals Engineering, 11(2):171–187.
Riley, J F, 1974. The tetrahedrite-freibergite series, with reference to the Mount Isa Pb-Zn-Ag Orebody,
Mineralium Deposita, 9:117–124.
Riley, J F, 1976. Silver mineralogy and its metallurgical significance at Mount Isa, Queensland, Australia,
in Transactions of the Institution of Mining and Metallurgy, 85:C102–C104.
Silvester, E J, Heyes, G W, Bruckard, W J and Woodcock, J T, 2011. The recovery of sericite in flotation
concentrates, Transactions of the Institutions of Mining and Metallurgy, 120:C10–C14.
Tomaino, G P, 2018. Talc and pyrophyllite, Mining Engineering, July, pp 86–89.
Torrisi, C, 2001. Leaching of fluorine bearing minerals from lead and zinc concentrates, Minerals
Engineering, 14(12):1637–1648.
CHAPTER 6
Characterisation measurements
in industrial flotation cells
Greg Harbort, Sarah Schwarz and Thu Nguyen
ABSTRACT
In recent years, a significant volume of research has focused on flotation machine parameters
and the impact they have on metallurgical performance. Technology has been developed as
part of the Australian Mineral Industries Research Association (AMIRA) P9 project, to conduct
characterisation measurements (such as gas dispersion and froth recovery) in industrial
flotation cells.
Gas dispersion measurements include gas hold-up, superficial gas velocity and bubble
size, and are a good indication of the hydrodynamics within the pulp phase. Each of these
measurements can be made in industrial-sized cells using specialised equipment developed
within the AMIRA P9 project. These measurements have been made in a wide range of type
and size of flotation cells throughout the world.
Various techniques have also been developed to indicate the efficiency of the froth phase,
known as froth recovery. These techniques range from physically changing the froth height,
through to performing mass balances across the froth phase, and collecting samples of the
material attached to the bubbles within the pulp phase.
This chapter discusses methods for obtaining these characterisation measurements, as well
as other necessary measurements within an industrial flotation plant environment. These
measurements can be used by plant metallurgists to analyse and improve the performance of
their flotation circuits.
INTRODUCTION
The range of flotation cell types and sizes in flotation plants has varied over the years,
from older circuits containing many small volume cells (<3 m3), through to newer circuits
containing smaller numbers of larger cells (>100 m3). The typical and/or optimum operating
conditions of flotation cells must be taken in context of the cell type and size. The following
lists the common types of cells currently in use in the flotation industry:
•• mechanical air-induced cells, eg Outotec, Metso, Bateman
Bubble size
The bubble size within the pulp phase of a flotation cell significantly influences the efficiency
of particle-bubble contact and attachment. In general, smaller bubbles are more efficient for
smaller particles and larger bubbles are more suited for larger particles (Biswal, Reddy and
Bhaumik, 1994). Also, it has been acknowledged that smaller bubbles within the pulp phase
improve the flotation kinetics (Diaz-Penafiel and Dobby, 1994). While these points have
been in general knowledge for a while, the problem has been actually measuring the bubble
size in the pulp phase of industrial flotation cells. Many bubble measurement techniques
are available, consisting of either invasive or non-invasive techniques. The most common
technique developed to measure bubble size in industrial flotation cells currently is the
McGill bubble size analyser, due to its simplicity and robustness. This is discussed in detail in
the following sections, with other measurement devices discussed briefly, highlighting their
advantages and disadvantages.
Window
Camera
Diffuser
Light source
Sample chamber
Sampling tube
Froth
Flotation cell
FIG 1 – Schematic of the McGill bubble size analyser and typical set-up.
zone. The set-up is filled with water dosed with enough frother to prevent bubble coalescence
in the tube. Bubbles from the flotation cell rise up through the sampling tube and slide up
the inclined window of the viewing chamber. The sloped window ensures a single layer
of bubbles passes through the image area, captured by a digital video camera. The light
source illuminates the image area, resulting in each bubble casting a shadow, which can be
detected by the software associated with the bubble viewer. Full details of the apparatus and
methodology can be found elsewhere (Chen, Gomez and Finch, 2001; Hernandez-Aguilar,
Gomez and Finch, 2002; Hernandez-Aguilar et al, 2004). Further refinements have since been
made to the apparatus, with LED lights operating from a battery and a digital still camera
used instead of a video camera. The probe is known as the Anglo Platinum Bubble Sizer, or
APBS and has been described in Taute and McClelland (2006).
pierced is registered. They can be used in three-phase high air hold-up conditions (Magaud,
Souhar and Boisson, 2001). The most commonly used in industrial environments is the UCT
bubble probe (Figure 2) (Tucker et al, 1994). This analyser consists of a sampling system and
an optical/electronic detector system. The sampler enables incoming bubbles to be cleaned of
any attached particles before they enter the capillary, where they are deformed from spheres
to cylinders. The length of the bubbles is detected by a pair of optical sensors.
Conductivity probes
Conductivity probes register the variation in conductivity between bubbles and the
surrounding pulp. Immersion conductivity probes are also used and can measure bubble
size without extracting bubbles from the flotation machine (Sanaullah, Zaidi and Hills, 2001;
Wu et al, 2001).
The atmosphere
Three-way valve
Flush water
Burette
Water reservoir
Photo-detector MPU timer Laptop
Capillary amplifier detector computer
Peristaltic
Plant water
Pump 2
The atmosphere
Pinch valve
1.6
1.2
1.0
0.8
0.6
0.4
0.2
0.0
0 1 2 3 4 5 6 7 8
Bubble diameter (mm)
FIG 3 – Normalised distribution of bubble sizes obtained from four cells within a flotation circuit.
The average bubble diameter is calculated from Equation 1 and the Sauter mean bubble
diameter is determined from Equation 2:
N
/d i
db = i=1
(1)
N
/d 3
i
d32 = i=1
(2)
/d 2
i
i=1
where:
di is the measured bubble diameter
N is the number of bubbles measured
Generally, the Sauter mean bubble diameter is quoted from these measurements, as this has
been shown to be a more accurate representation of the bubble size distribution. Currently
there is some debate about whether the ratio of the Sauter mean to the arithmetic mean should
also be quoted, to indicate the width of the bubble size distribution.
30
Frequency (%)
20
15
10
0
0.5 0.75
1 1.25
1.5 1.75
2 2.25
2.5 2.75
3 3.25
Bubble size (mm)
FIG 4 – Range of Sauter mean bubble size obtained in mechanical, column,
Jameson and Flash flotation cells (from Schwarz and Alexander, 2006).
shown to give a good indication of how well the air is dispersed in the cell (Gorain, Franzidis
and Manlapig, 1996). Superficial gas velocity can be measured by various means, ranging
from the air flowmeter readings through to sophisticated instruments, and these will be
discussed in this section.
Water
Atmosphere
Compressed Air
Feed 3-way Valve
Mark 1
Mark 2
Perspex
Tube
Pinch
Valve
Nozzle
Tailings
the length of the probe and the distance the probe is immersed in the pulp phase. Full details
of the use and set-up of this probe can be found in Gorain, Franzidis and Manlapig (1996) and
Power, Franzidis and Manlapig (2000).
Worked example
Repeat measurements have been made using a Jg probe in a single location in a flotation cell.
The following is a list of the times taken for the water level to drop 20 cm: 14.0 s, 19.1 s, 17.7 s,
15.7 s, 14.8 s. The length of the probe from the bottom valve to the bottom line was 2.9 m
and the distance from the bottom valve to the top of the pulp phase was 1.2 m. Calculate the
average superficial gas velocity of this cell.
Calculations
The experimental Jg value can be obtained using the following equation:
Jg = L
t
where:
L is the distance between marks
t is the time taken for the water level to drop between the marks
This value is calculated for each measurement, for example:
3 3
-JVHQVRU 5HIHUHQFH
/
)/27$7,21
0$&+,1(
80
Changing slope P1 P1
PRESSURE (cm H2O)
60
Difference
40
Correct slope
P2 – level changes
20 P2
0
0 50 100 150
TIME (s)
where:
Po is the atmospheric pressure
PL is the pressure at the bottom of the tube, which changes as the level changes
tb is the density of the aerated pulp
Gas hold-up
Gas hold-up measures the volume fraction of air contained within the pulp phase of the
flotation cell. Generally, increasing the gas content within the cell results in more bubbles per
unit volume and hence more available surface area for particle-bubble interactions to occur,
producing improved kinetics in the pulp phase (Ahmed and Jameson, 1989). However, there
is a maximum value of gas hold-up before the cell capacity is limited, resulting in reduced
residence time. Again, there are various measurement techniques available to measure both
overall and local gas hold-up values, and these are discussed in the following section.
25
Conventional cells
15
10
0
0.25 0.5
0.75 1 1.25 1.5
1.75 2 2.25 2.5
2.75 3
Superficial gas velocity (cm/s)
FIG 8 – Range of superficial gas velocity values obtained from conventional,
column, Jameson and Flash flotation cells (from Schwarz and Alexander, 2006).
f g = DH (5)
H
where:
ΔH is the change in pulp level after switching off the gas
H is the pulp level with gas
This method has the advantage that no specialised equipment is necessary and it also gives
the overall gas content within the whole flotation cell. However, it can often be quite difficult
to obtain the initial measurement of the pulp level with gas present in the cell.
This technique is simple and rapid, giving almost instantaneous results. However, the
results are for a localised region of the cell, and measurements must be repeated across the
cross-section of the flotation cell to give the average gas hold-up across the whole cell. It is
generally recommended that at least three measurements of gas hold-up are performed in
each location within the cell to ensure representative results.
Three-way valve
Air line
Compressed air
Worked example
Repeat measurements have been made using a JKMRC gas hold-up probe. The volume of the
probe is 1070 ml. The following is a list of slurry volumes obtained from the same location in
the flotation cell: 1000 ml, 1010 ml and 1000 ml. Calculate the average gas hold-up from this
position within the flotation cell.
Calculation
Using Equation 6, the gas hold-up from the first measurement can be calculated as:
fg = s
k
(7)
1 + 0.5 k0
s
where:
ko is the specific conductivity of the slurry with bubbles present
ks is the specific conductivity of the slurry with no bubbles
Again, this technique produces gas hold-up values for a given location within the flotation
cell, and can be repeated across the cell to give the average overall gas hold-up. This technique
can also be used for continuous monitoring of the gas content within the cell.
Syphon cell
Open cell (slurry, no gas)
FIG 10 – Schematic and arrangement of the McGill online gas hold-up probe.
Ultrasonic techniques
Doppler ultrasonic probes have been used for many years for measuring liquid velocity
in pipes. Their use is based on the propagation of sound waves, generated within a piezo-
ceramic disc. Techniques can either rely on ultrasound transmittance through an aerated pulp
(Waniewski, Hunter and Brennen, 2001), or reflectance from bubble surfaces (Kulmyrzaev,
Cancelliere and McClements, 2000).
Acoustic emission
Acoustic emission techniques operate with a similar principle to the use of microphone
recordings to determine a grinding mill load. Pulp, with a specific air hold-up will generate a
unique acoustic signal that can be recorded through a hydrophone (Boyd and Varley, 2004).
Radiation techniques
Radiation sources have for many years been in use for purposes of in-stream analysis. Gases,
liquids and solids have different adsorption coefficients of X-ray or γ-ray radiation or neutron
beam radiation. Radiation released by a radioactive source is passed through the phase being
evaluated. The attenuation of radiation can be related to the air hold-up (George et al, 2001).
45
40
Column, Jameson, Flash
Frequency (%)
35
30 Conventional cells
25
20
15
10
5
0
6 12 18
24 30 36
42 48 54
60 66 72
Air hold-up (%)
There is a strong linear relationship between bubble surface area flux and recovery across
the pulp phase (Gorain, Franzidis and Manlapig, 1997).
The bubble surface area flux can also be estimated using impeller dimensions and speed as
well as particular properties of the pulp phase in an empirical relationship (Gorain, Franzidis
and Manlapig, 1999).
Also useful to note is that when treating the same type of ore, this linear relationship is
independent of impeller type, as shown in Figure 12. The slope of the relationship is related
to the floatability of the ore – with higher slopes indicating faster-floating material.
Typical bubble surface area flux values range from 30 s-1 to 70 s-1 for both mechanical and
column flotation cells, as shown in Figure 13.
0.8
Flotation Rate Constant (1/min)
0.7 Chile-X
Pipsa
0.6 Outokumpu
0.5
Dorr-Oliver
0.4
0.3
0.2
0.1
0
0 50 100 150 200 250
Bubble Surface Area Flux (m²/m² sec)
30
Frequency (%)
Conventional cells
20
15
10
0
10 20 30
40 50 60
70 80 90
100 110
120
-1
Bubble surface area flux (s )
FIG 13 – Range of bubble surface area flux values obtained from conventional,
column, Jameson and Flash flotation cells (from Schwarz and Alexander, 2006).
Rf = k (9)
kc
Various authors including Vera, Franzidis and Manlapig (1999) indicated the value for kc
by extrapolation of the overall rate constant (k) versus froth residence time (τf) graph to the
y-axis, where τf = 0 (the collection zone) (Feteris, Frew and Jowett, 1987; Gorain et al, 1998).
Froth residence time can be defined in several ways. The definition referred to in this chapter
is based on the concentrate flow rate. As such, the froth residence time is defined as (Mathe,
Harris and O’Connor, 2000):
0.2
0.15
0.1
0.05
0
0 20 40 60 80 100
FIG 14 – Typical variation of flotation rate constant with froth residence time.
(1 - f g) Vfroth
FRT = (10)
Qconc
where:
εg is the gas hold-up in the froth phase
Vfroth is the volume of the froth phase
Qconc is the volumetric flow rate of the concentrate
By varying the froth residence time, through either varying froth height or airflow and
measuring recovery, a set of flotation rate constants at various froth residence times can be
calculated and plotted. A typical example of this is shown in Figure 14. From this graph
an estimated flotation rate constant (kc) of 0.13 min-1 is determined. At 20 seconds residence
time the flotation rate constant is approximately 0.05 min-1. Therefore, at 20 seconds froth
residence time the froth recovery will be 38.5 per cent.
In operations where the flotation circuit operates with flotation cells of various sizes an
adjustment based on the horizontal distance the froth travels to the concentrate launder may
be required. This is done by dividing the froth retention time by a typical cell length as shown
below; where τfs may be thought of as ‘specific’ froth residence time, which is independent
of cell size:
x fs = FRT (11)
L
where:
τfs = specific froth residence time (s/m)
L = perpendicular distance from impeller to launder (m)
The technique is non-disruptive and involves taking samples of feed, concentrate, tail, froth
and air hold-up. These samples are assayed and a mass balance of attached and entrained
particles at various size fractions is conducted, as discussed below.
The air hold-up sample taken directly below the pulp/froth interface with the superficial
gas velocity measurement is also used to estimate the mass flow rate of attached particles
entering the froth phase, based on the following equation:
M ATT
Pulp
= 3.6 $ QAir $ BPulp (12)
where:
M ATT
Pulp
= mass flow rate of attached particles in the pulp zone
Qair = volumetric flow rate of air (m3/s)
BPulp = bubble load in the pulp phase (grams of solids/litre of air)
The volumetric flow rate of air can be measured directly in each cell or estimated using the
superficial gas velocity measurements across the cell cross-sectional area:
QAir = A $ Jg (13)
where:
Jg = superficial gas velocity (cm/s)
A = open area of cell (m2)
The bubble load is defined as the mass of attached solids per volume of air and involves
obtaining a sample using the air hold-up probe and using the assumptions outlined above.
Air bubbles (with particles attached) and pulp (with particles entrained) pass through the
air hold-up sampler in the open position. When the air hold-up probe is closed, the attached
particles become detached from the air bubbles and are included with the pulp containing
entrained particles and water. The volume of air and the mass and grade of the pulp sample
are noted. The bubble load can then be calculated using:
matt $ 1000
BPulp = (14)
V
where:
matt = mass of attached particles in the air hold-up probe (g)
V = volume of air measured in the air hold-up probe (ml)
The mass of attached particles in the air hold-up probe is estimated using the following
mass balance equation:
mtotal. $ (Gtotal - Gent)
matt = (15)
(Gatt - Gent)
where:
mtotal = total mass of particles in the air hold-up probe (g)
Gent = grade of entrained particles in the air hold-up probe
Gtotal = grade of all particles in the air hold-up probe
Gatt = grade of attached particles in the air hold-up probe
In order to calculate the mass of attached particles (matt), the grade of attached particles (Gatt)
is taken as that of the grade of the top of froth sample, while the grade of entrained particles
(Gent) is taken to be equivalent to the grade of the tailings. This is in accordance with the
assumptions made at the beginning of this section.
The estimate of the mass of attached particles in the air hold-up sample obtained from
Equation 15 can then be used to calculate the mass flow rate of attached particles into the
froth phase (Equation 12).
The mass flow rate of attached particles in the concentrate can be calculated using a similar
mass balance equation to Equation 15 applied to the concentrate sample:
mtotal
Conc
$ (Gtotal - Gent)
m att
conc
= (16)
(Gatt - Gent)
where:
m Att
Conc
= mass flow rate of attached particles in the concentrate stream (t/h)
mtotal
Conc
= mass flow rate of all particles in the concentrate stream (t/h)
Gent = grade of entrained particles in the concentrate
Gtotal = concentrate grade
Gatt = grade of attached particles in the concentrate
The froth recovery of attached particles can then be calculated using the following equation:
m Att
Conc
Rf = (17)
m Att
Pulp
where:
Rf = froth recovery of attached particles
m Conc
Att
= mass flow rate of attached particles in the concentrate stream (t/h)
m Pulp
Att
= mass flow rate of attached particles in the pulp (t/h)
Worked example
A flotation machine operates with an air rate of 396 m3/h treating a feed of 6.5 t/h solids. The
average air hold-up volume is 83.3 ml, giving an air hold-up of 7.79 per cent and the mass of
solids in the air hold-up probe is 28.3 g. The following zinc assays have been recorded: feed
– 11.05 per cent Zn; concentrate – 33.72 per cent Zn; tailing – 8.80 per cent Zn; air hold-up –
9.28 per cent Zn; froth – 36.1 per cent Zn. Calculate the froth recovery.
Calculations
The overall concentrate produced can be calculated either by concentrate measurements, or
from the two-product formula:
6.5 # (11.05 - 8.8)
mtotal
Conc
= = 0.59 t/h
(33.72 - 8.8)
The total mass of attached particles in the air hold-up probe is:
(9.28 - 8.8)
m att = 28.3 # = 0.50 g
(36.1 - 8.8)
The bubble load is:
B Pulp = 0.50 # 1000 = 6.00 kg/m 3 air
83.3
M ATT
Pulp
= 3.6 # 396 # 6.00 = 8555 kg/h = 8.56 t/h
FIG 15 – Arrangement of the bubble load device (from Seaman, Franzidis and Manlapig, 2004).
The bubble load technique is also non-invasive and generates only two samples per cell
(concentrate and bubble load sample). This technique is performed in conjunction with a
general survey to give the flow rates of the individual cell concentrate streams. There have
been some issues with the size of the sample collected, with long sampling periods sometimes
required to obtain enough sample for sizing analysis to be performed.
The bubble load device has been shown to give similar froth recovery measurements to
the mass balance technique, but significantly different results to the froth height or froth
residence time methods. This is mainly due to the different basis of calculations – the froth
residence time method is concerned with processes occurring within the froth phase only,
while the mass balance and bubble load techniques consider processes occurring at the pulp/
froth interface as well. The majority of bubble-particle detachment has been found to occur
at the pulp/froth interface, due to the sudden change in momentum of the bubble-particle
aggregates. Since the froth residence time method does not take this section of the flotation
system into account, the froth recovery measurements are routinely higher than for the same
cell and operating conditions using the other techniques. A comprehensive evaluation of
different froth recovery measurement techniques was given by Runge et al (2010).
Standard units of the operating variables are used in the model viz m/s for Ns, cm/s for Q/A,
which in theory is equal to superficial gas velocity, dimensionless for As and microns for P80,
which is typically used to represent particle size in flotation operations.
A database of cell characterisation values for over 2500 flotation machines has been
developed at the Julius Kruttschnitt Mineral Research Centre (JKMRC) for flotation cells
of different sizes and designs operated using a variety of impeller types, and for industrial
scale cells on operating plants. Other substantial cell characterisation activities have also been
conducted by Metso Process Technology & Innovation, the Outotec Flotation Division and
the Mining and Metals Consulting division of Amec Foster Wheeler. Initial investigations
indicate that the same general form can be used for the Jameson cell, using jet velocity and
the downcomer diameter to orifice plate diameter ratio.
Decrease in
Particle Size Increase in
Bubble Loss to
Tailing
Decrease in
Solids SG Decrease in
Increase in Air
Hold-up Residence Time
Decrease in
Increase in Recovery
Viscosity Increase in
Decrease in
Bubble Size
Kinetic Rate
Increase in % Increase in
Solids Concentrate
Solids Loading
Increase in
Residence
Decrease in
Time Increase in
Volumentric
Flow Recovery
Decrease in
Bubble Loss to
Tailing
Decrease in Air
Hold-up
Decrease in Solids
Short Circuit to
Tailing
FIG 16 – Interactions of various factors on gas characteristics and their affect on recovery.
The viscosity of a non-Newtonian fluid (as in the case of mineral slurry) at a specified
point is referred to as ‘apparent viscosity’. The apparent viscosity of slurry depends on the
microscopic structures of suspended particles in the slurry and is a function of the volume
fraction of the solid, particle shape, particle interaction and the aggregation of the suspended
solid (Yoshida et al, 2013).
Rheology or viscosity has an important influence on the flotation process. However,
measuring the rheological behavior is difficult as particles tend to settle during measurement.
It has been shown that the measurement of froth rheology in situ can be done using the vane
rheometer but only at a limited vane head speed (Chao et al, 2015).
Two key determinants of pulp viscosity are the per cent solids by weight and the solids
specific gravity, as shown in Figure 17. If gangue changes from sulfide- or iron-dominant
gangue to siliceous gangue, the solids specific gravity decreases, resulting in an increase in
viscosity, which will alter gas characteristics.
2.0
1.9
1.8
Viscosity (mPa.s)
1.7
1.6
1.5
1.4
1.3
1.2
1.1
1.0
0 10 20 30 40 50 60
% Solids (w/w)
50
Viscosity (mPa.s) 40
30
20
10
10 20 30 40
Volume concentration (%)
Spheres Grains Discs Rods
FIG 18 – Effect of particle shape on viscosity (from Barnes, Hutton and Walters, 1989).
increase in viscosity for the same phase volume. Figure 18 (Barnes, Hutton and Walters, 1989)
details the effect of particle shape on viscosity for the following cases:
1. spheres – grinding via attrition mills such as the Vertimill® or IsaMill® generates essentially
spherical particles
2. grains – grinding via impact mills such as SAG mills, ball mills or rod mills generates
granular particles
3. discs – the presence of discs generally comes from material that breaks along plate lines
such as mica
4. rods represent the most severe particle shape to affect viscosity and can be fibrous, such
as asbestos.
Particle
Diameter (um)
101 0.05
102
101
100
Particle Coagulation Region
10‐1
0.1 0.2 0.3 0.4 0.5 0.6
Solids volume fraction
FIG 19 – Formation of agglomerates (from Barnes, Hutton and Walters, 1989).
1.8
1.6
1.4
Sauter mean bubble size (mm)
1.2
1.0
0.8
0.6
0.4
0.2
0.0
1 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8
Viscosity (mPa.s))
1.8
1.6
1.4
Sauter mean bubble size (mm)
1.2
1.0
0.8
0.6
0.4
0.2
0.0
0 10 20 30 40 50
% Solids (w/w)
FIG 21 – Effect of per cent solids on bubble size.
30
25
Air hold‐up (%)
20
15
10
0
0 10 20 30 40 50 60
Per cent solids (w/w)
FIG 22 – Effect of per cent solids on air hold-up.
35
30
25
Air hold‐up (%) 20
15
10
0
0 50 100 150 200
Sb (s‐1)
JKTech Database Matiolo et al
FIG 23 – Effect of Sb on air hold-up.
Increases in volumetric pulp flow result in an increase in pulp superficial velocity, Jp, which
in turn changes the operating point where the transition to the foam flow regime occurs.
Figure 24 details the relationship at Jp values of 0.0 cm/s and 0.66 cm/s (Matiolo et al, 2011) and
1.5 cm/s (Amec Foster Wheeler database). Typically, once familiar with the issue, operators
will reduce air to prevent flooding. The loss in recovery from reduced kinetics is typically far
less than the loss due to cell flooding.
CONCLUSIONS
Measurement of gas dispersion characteristics in industrial flotation plant cells has aided
troubleshooting and identified problem areas such as worn impellers, non-uniform gas
dispersion, etc. Being able to measure these parameters quickly and relatively easily has
enabled plant metallurgists to make adjustments to their cell operation and improve the plant
performance significantly.
Measurement of the froth performance has indicated that most flotation cells operate at
reasonably low froth recovery, with a significant proportion of material that entered the froth
30
25
FOAM TRANSITION POINT
20
Air hold‐up (%)
15
10
0
0 20 40 60 80 100 120
Sb (s‐1)
Jp = 0.0 cm/s Jp = 0.66 cm/s Jp = 1.5 cm/s
phase dropping back down to the pulp phase. These measurements have also identified that
the processes within the froth phase are still relatively unknown, but quite significant in the
overall flotation cell performance.
While a great deal of flotation plant operation has developed based on experience and
instinct, it is envisaged that with the various techniques now available to measure critical
parameters within the flotation cells, plant metallurgists will have a greater understanding
of the flotation process. Once the process is better understood, there is a greater potential for
improving and operating the flotation cells at their optimum levels for the best metallurgical
performance possible.
REFERENCES
Ahmed, N and Jameson, G J, 1989. Flotation kinetics, Mineral Processing and Extractive Metallurgy
Review, 5:77–99.
Alexander, D J, Franzidis, J-P and Manlapig, E V, 2003. Froth recovery measurement in plant scale
flotation cells, Minerals Engineering, 16(11):1197–1203.
Aslan, M M, Crofcheck, C, Tao, D and Mengüç, M P, 2006. Evaluation of micro-bubble size and gas
hold-up in two-phase gas–liquid columns via scattered light measurements, Journal of Quantitative
Spectroscopy and Radiative Transfer, 101(3):527-539.
Barnes, H A, Hutton, J F and Walters, K, 1989. An Introduction to Rheology, Chapter 7 – Rheology of
suspensions, pp 115–137 (Elsevier: Amsterdam).
Biswal, S K, Reddy, P S R and Bhaumik, S K, 1994. Bubble size distribution in a flotation column, The
Canadian Journal of Chemical Engineering, 72(1):148–152.
Boyd, J W R and Varley, J, 2004. Acoustic emission measurement of low velocity plunging jets to
monitor bubble size, Chemical Engineering Journal, 97:11–25.
Boyer, C, Duquenne, A-M and Wild, G, 2002. Measuring techniques in gas-liquid and gas-liquid-solid
reactors, Chemical Engineering Science, 57:3185–3215.
Chao, L, Farrokhpay, S, Shi, F and Runge, K, 2015. A novel approach to measure froth rheology in
flotation, Minerals Engineering, 71:89–96.
Chen, F, Gomez, C O and Finch, J A, 2001. Technical note: bubble size measurement in flotation
machines, Minerals Engineering, 14(4):427–432.
Diaz-Penafiel, P and Dobby, G S, 1994. Kinetic studies in flotation columns: bubble size effect, Minerals
Engineering, 7(4):465–478.
Engelbrecht, J and Woodburn, E T, 1975. The effects of froth height, aeration rate, and gas precipitation
on flotation, Journal of the Southern African Institute of Mining and Metallurgy, October, pp 125–132.
Feteris, S M, Frew, J A and Jowett, A, 1987. Modelling the effect of froth depth in flotation, International
Journal of Mineral Processing, 20:121–135.
George, D L, Shollenberger, K A, Torczynski, J R, O’Hern, T J and Ceccio, S L, 2001. Three-phase
material distribution measurements in a vertical flow using gamma-densitometry tomography and
electrical-impedance tomography, International Journal of Multiphase Flow, 27:1903–1930.
Gorain, B K, Franzidis, J-P and Manlapig, E V, 1995a. Studies on impeller type, impeller speed and
air flow rate in an industrial scale flotation cell – Part 1: Effect on bubble size distribution, Minerals
Engineering, 8(6):615–635.
Gorain, B K, Franzidis, J-P and Manlapig, E V, 1995b. Studies on impeller type, impeller speed and air
flow rate in an industrial flotation cell – Part 2: Effect on gas hold-up, Minerals Engineering, 8:1557–
1570.
Gorain, B K, Franzidis, J-P and Manlapig, E V, 1996. Studies on impeller type, impeller speed and air
flow rate in an industrial flotation cell – Part 3: Effect on superficial gas velocity, Minerals Engineering,
9:639–654.
Gorain, B K, Franzidis, J-P and Manlapig, E V, 1997. Studies on impeller type, impeller speed and
air flow rate in an industrial flotation cell – Part 4: Effect of bubble surface area flux on flotation
performance, Minerals Engineering, 10:367–379.
Gorain, B K, Franzidis, J-P and Manlapig, E V, 1999. The empirical prediction of bubble surface area
flux in mechanical flotation cells from cell design and operating data, Minerals Engineering, 12(3):309–
322.
Gorain, B K, Napier-Munn, T J, Franzidis, J-P and Manlapig, E V, 1998. Studies on impeller type,
impeller speed and air flow rate in an industrial scale flotation cell. Part 5: Validation of k-Sb
relationship and effect of froth depth, Minerals Engineering, 11(7):615–626.
Gourram-Badri, F, Conil, P and Morizot, G, 1997. Measurements of selectivity due to coalescence
between two mineralized bubbles and characterisation of MIBC action on froth flotation, International
Journal of Mineral Processing, 51:197–208.
Hernandez-Aguilar, J R, Coleman, R G, Gomez, C O and Finch, J A, 2004. A comparison between
capillary and imaging techniques for sizing bubbles in flotation systems, Minerals Engineering,
17(1):53–61.
Hernandez-Aguilar, J R, Gomez, C O and Finch, J A, 2002. A technique for the direct measurement of
bubble size distributions in industrial flotation cells, in Proceedings 34th Annual Meeting of the Canadian
Mineral Processors of CIM, pp 389–402 (Canadian Institute of Mining, Metallurgy and Petroleum:
Montreal).
Jameson, G J and Allum, P, 1984. A survey of bubble sizes in industrial flotation cells, report to AMIRA
Ltd (Australian Mineral Industries Research Association Limited).
Kilickaplan, A, 2007. The effect of pulp rheology on flotation, Master of Applied Science thesis,
Hacettepe University, 2007.
Koh, P T L and Schwarz, M P, 2008. Computational fluid dynamics modelling of slimes flotation at
Mt Keith Operations, in Proceedings Metallurgical Plant Design and Operating Strategies (MetPlant 2008),
pp 325–337 (The Australasian Institute of Mining and Metallurgy: Melbourne).
Kulmyrzaev, A, Cancelliere, C and McClements, D J, 2000. Characterisation of aerated foods using
ultrasonic reflectance spectroscopy, Journal of Food Engineering, pp 235–241.
Laplante, A R, Toguri, J M and Smith, H W, 1983. The effect of air flow rate on the kinetics of flotation.
Part 2: The transfer of material from the froth over the cell lip, International Journal of Mineral Processing,
11:221–234.
Magaud, F, Souhar, M and Boisson, N, 2001. Experimental study of bubble column hydrodynamics,
Chemical Engineering Science, 56:4597–4607.
Mathe, Z T, Harris, M C and O’Connor, C T, 2000. A review of methods to model the froth phase in
non-steady state flotation systems, Minerals Engineering, 13(2):127–140.
Matiolo, E, Testa, F, Yianatos, J and Rubio, J, 2011. On the gas dispersion measurements in the collection
zone of flotation columns, International Journal of Mineral Processing, 99:78–83.
Nesset, J E, Zhang, W and Finch, J A, 2012. A benchmarking tool for assessing flotation cell performance,
in Proceedings 44th Annual Canadian Mineral Processors Operators Conference, Ottawa, 17–19 January,
pp 183–209.
O’Connor, C T, Randall, E W and Goodall, C M, 1990. Measurement of the effects of physical and
chemical variables on bubble size, Int J Min Proc, 28:139–149.
Power, A, Franzidis, J-P and Manlapig, E V, 2000. The characterisation of hydrodynamic conditions in
industrial flotation cells, in Proceedings Seventh Mill Operators’ Conference, pp 243–255 (The Australasian
Institute of Mining and Metallurgy: Melbourne).
Rodrigues, R T and Rubio, J, 2003. New basis for measuring the size distribution of bubbles, Minerals
Engineering, 16:757–765.
Runge, K, Crosbie, R, Rivett, T and McMaster, J, 2010. An evaluation of froth recovery measurement
techniques, in Proceedings XXV International Mineral Processing Congress (IMPC), pp 2313–2324 (The
Australasian Institute of Mining and Metallurgy: Melbourne).
Sanaullah, K, Zaidi, S H and Hills, J H, 2001. A study of bubbly flow using resistivity probes in a novel
configuration, Chemical Engineering Journal, 83:43–53.
Savassi, O N, Alexander, D J, Johnson, N W, Franzidis, J-P and Manlapig, E V, 1997. Measurement
of froth recovery of attached particles in industrial flotation cells, in Proceedings Sixth Mill Operators’
Conference, pp 149–155 (The Australasian Institute of Mining and Metallurgy: Melbourne).
Schubert, H and Bischofberger, C, 1978. On the hydrodynamics of flotation machines, International
Journal of Mineral Processing, 5:131–142.
Schwarz, S and Alexander, D A, 2006. Gas dispersion measurements in industrial flotation cells,
Minerals Engineering, 19:554–560.
Seaman, D, Franzidis, J-P and Manlapig, E V, 2004. Bubble load measurement in the pulp zone of
industrial flotation machines – a new device for determining the froth recovery of attached particles,
Mineral Processing and Extractive Metallurgy Review, 5:123–145.
Taute, J J and McClelland, A J, 2006. Introduction to the Anglo Platinum Bubble Sizer, presented to the
South African Institute of Mineral Processing Conference, August.
Tucker, J P, Deglon, D A, Franzidis, J-P, Harris, M G and O’Connor, C T, 1994. An evolution of a
direct method of bubble size distribution measurement in a laboratory batch flotation cell, Minerals
Engineering, 7(5/6):667–680.
Vera, M A, Franzidis, J-P and Manlapig, E V, 1999. Simultaneous determination of collection zone
rate constant and froth zone recovery in a mechanical flotation environment, Minerals Engineering,
12(10):1163–1176.
Waniewski, T A, Hunter, C and Brennen, C E, 2001. Bubble measurements downstream of hydraulic
jumps, International Journal of Multiphase Flow, 27:1271–1284.
Wu, Q, Welter, K, McCreary, D and Reyes, J N, 2001. Theoretical studies on the design criteria of
double-sensor probe for the measurement of bubble velocity, Flow Measurement and Instrumentation,
12:43–51.
Yoshida, Y, Katsumoto, T, Taniguchi, S, Shimosaka, A, Shirakawa, Y and Hidaka, J, 2013. Prediction of
viscosity of slurry suspended fine particles using coupled DEM-DNS simulation, Chemical Engineering
Transactions, 32:2089–2094.
CHAPTER 7
Chemical measurements during plant
surveys and their interpretation
Stephen Grano
ABSTRACT
This chapter addresses simple chemical measurements and their interpretation in the
optimisation of base metal sulfide flotation plants. The measurements include pH, Eh,
dissolved oxygen, dissolved oxygen demand, extractable metals, and solution species. For
each measurement, the basic theory and practical examples of application are discussed.
Simple hydrolysis and Eh/pH diagrams are described which aid interpretation. The key
sampling points and methodology for chemical measurements during a plant survey are
outlined. The critical importance and method of correlating the laboratory conditions and
performance to the plant is discussed. Two practical case studies highlight the utility of these
chemical measurements for process troubleshooting are discussed.
INTRODUCTION
This chapter is divided into the following three sections:
1. chemical measurements during a plant survey
2. analysis and interpretation of these chemical measurements
3. two practical case studies that highlight the utility of these chemical measurements for
process troubleshooting.
For each measurement type, a brief theoretical background will firstly be discussed. This
will then be followed by a brief discussion on the importance of the measurement, how the
measurement may be made, and some practical guidelines.
The combined measurements of Eh and pH will, to a significant extent, define the
thermodynamic stability of bulk species (Garrels and Christ, 1965). This is because most equilibria
are principally controlled by pH (eg hydrolysis) and Eh (eg oxidation/reduction). The other
important variable when considering the predominant equilibrium species is concentration
(eg the total metal ion or collector concentration in the system), though this parameter may
be very difficult to define and measure. However, the strict application of thermodynamic
equilibria is often not appropriate for reactions that have not come to equilibrium, or at a mineral
surface where the interfacial pH and concentrations may not be at their bulk solution values
(Woods, 1984). Putting aside these considerations, it is still very valuable to measure both pH
and Eh during plant surveys and to use these measurements to ‘predict’ species predominance
(Natarajan and Iwasaki, 1973a). Several examples are presented which show the utility of such
an approach. It is always important to measure both pH and Eh at key sample points, such as
the ball mill discharge, flotation feed and process water in a plant survey.
pH
This is the most common and easiest to measure of the chemical measurements. pH is
relatively easy to measure as it does not rely on particle contact with the electrode to ensure
an accurate measurement, unlike Eh, which does. The definition of pH is given below (Garrels
and Christ, 1965):
pH = -log [H+]
This definition is a formal one as it involves the activity of a single ion, a quantity that cannot
be measured directly. As a result, the measured pH is defined in terms of an operational
procedure. For all practical purposes, the measurement of pH is accomplished using an
electrochemical cell that consists of a glass electrode dipping into the solution, together with
a reference electrode to complete the circuit. The reference electrode is usually a mercury-
mercurous chloride (calomel) electrode in saturated KCl solution, connected to the solution to
be measured by a salt bridge of saturated KCl solution. The glass electrode consists of a glass
bulb containing an acid solution, and an inner electrode of fixed potential, usually Ag-AgCl,
to conduct electrons, into and out of solution. When the bulb is immersed in a solution, an
electrical potential is developed between the inner and outer solutions which is proportional
to log [H+] in the external solution. The reference electrode has a constant potential, whilst
that of the glass electrode will vary with [H+]. The overall cell may be represented as: glass
electrode, sol. x| KCl(sat.), AgCl (s); Ag (l).
The half-cell (electrode) potential of the glass electrode is given as:
The overall cell potential is given by the difference of the two-half cell potentials:
E x = C H - 2.303 RT pH x - E ref
F
2+ 2+
Cu2(OH)2 Cu 4-
Cu(CO3 ) 3
what species are in solution at specific points in the circuit. For example, a thickener overflow,
which contributes to the process water, containing 5 ppm of copper at pH 7, is sufficient to
inadvertently activate some sphalerite during galena flotation and increase contamination of
the lead concentrate by zinc, reducing also zinc recovery to the zinc concentrate.
Increasing the pH further will dissolve Cu(OH)2 (s) as Cu(CO3)32-. It is important to note that
Cu(OH)2 (s) will still be present in the system to pH values greater than ten, though equilibrium
considerations suggest it will be at a lower concentration. Because the Cu(OH)2 (s) is in
equilibrium with negatively charged species in solution, the zeta potential of the Cu(OH)2 (s)
will be negative at pH values greater than approximately 8–8.5 (pHiep). To dissolve copper
hydroxide it is necessary to decrease the pH to less than ~6, though this will depend on the total
concentration of copper in solution. Copper in solution may be derived from the oxidation
of copper minerals, which is dependent upon the pulp Eh, while hydrolysis of the copper is
pH dependent. The adsorption of copper and iron onto copper sulfide minerals controls their
flotation properties to a very large extent. An excellent case study of the effect of pH on iron
hydrolysis and the zeta potential of pyrite is given by Fornasiero, Eijt and Ralston (1992),
while the ensuing effect of pH on collector adsorption is discussed by Montalti, Fornasiero
and Ralston (1991). Kristall et al (1994) has described the influence of pH on the selective
dissolution of zinc hydroxide on sphalerite while leaving iron (III) hydroxide on pyrite intact
in a copper/zinc flotation plant.
Eh
The oxidation/reduction potential, Eh, is defined as the potential of a half-cell, referred to
the standard hydrogen halfcell, the potential of the standard halfcell being taken as zero at
all temperatures, by definition (Garrels and Christ, 1965). Oxidation/reduction potential is
measured with an electrode pair consisting of an inert electrode and a reference electrode.
The same reference electrode is used for most Eh measurements as for pH measurements –
the saturated calomel. The inert electrode used most commonly is bright platinum, although
gold electrodes are also used. The role of the calomel electrode is to supply a known potential
and to make electrical connection with the system to be measured. The inert electrode acts
as an electron acceptor or donor to the ions in the measured solution. When connected to the
calomel electrode, the platinum electrode can accept electrons from dissolved ionic species,
or it can give them up, depending upon whether the potential of the half-cell involving the
dissolved species is greater or less than that of the calomel reference electrode.
The half-cell reaction for every oxidation-reduction system can be written as:
An example is:
Fe3+ + e- = Fe2+
As an example, it is possible to calculate, using the Nernst equation, the theoretical potential
of the cell:
Pt (inert) | Fe3+ (aq) 0.01 M || Hg2Cl2 | Hg
The value of Ehref is constant at a given temperature and pressure, and for the saturated
calomel cell (Hg2Cl2) at 25°C is 0.2444 v. The half-cell potential for the Fe2+, Fe3+ couple is
given by:
[Fe 3+] 2
Eh Fe2 + ,Fe3+ = EoFe2 + ,Fe3+ + 0.0592 ln
2 [Fe 2+] 2
Therefore:
EMF cell = 0.830 - 0.244 = 0.586 v
The measured potential of an inert electrode will be close to the ‘reversible’ potential
calculated by the Nernst equation, provided:
•• All the species involved in the oxidation/reduction are in equilibrium.
•• The indicator electrode is truly an inert electrode and does not itself become oxidised or
reduced.
•• There is only one couple, which dominates the potential response. Often there are two or
more redox couples, such as the reduction of water, which may influence the potential.
In this case, the developed potential is referred to as a ‘mixed’ potential (Rand and
Woods, 1984; Natarajan and Iwasaki, 1974).
In Eh measurements most of the difficulties result from contamination of the platinum
electrode (Natarajan and Iwasaki, 1973b). It is useful to have two indicator or inert electrodes
set up with the same reference electrode to check for differences due to contamination.
Solutions containing sulfide ion often ‘poison’ Eh electrodes. Eh measurements under ideal
conditions are typically only reproducible to ±5 mV, while for measurements in slurries
it is only reproducible to ±10 mV. In slurries, the Eh will also depend upon contact of the
inert electrode with conducting particles of sulfide minerals (Rand and Woods, 1984). It is
important to maintain a reproducible technique to measure the Eh, which usually involves
rapid stirring of the pulp to ensure adequate particle contacting with the electrode.
The degree to which the oxidised species predominates over the reduced species depends
on the difference between the potential of the system (Eh) and the standard potential (Eo)
(Figure 2). Eo is the potential at which all species are at unit concentration. For single electron
transfer reactions (n = 1), if the potential is more than 60 mV greater than the Eo, then
90 per cent of the species considered will most likely be in their oxidised state (Figure 2),
assuming equilibrium has been achieved. If the potential is more than 60 mV less than the
Eo, then 90 per cent of the species considered will likely be in their reduced state (Figure 2),
again assuming equilibrium has been achieved. In the above case of the Fe2+, Fe3+ couple,
over 90 per cent of the iron is in the Fe3+ oxidation state at a potential 50 mV greater than the
standard half-cell potential.
For systems which are open to the atmosphere, the reduction of oxygen at the inert electrode
is very important. For example, slurry exiting a ball mill may be in a highly reducing state
100
90
80
70
n=1
% Oxidised
60
n=2
50
n=3
40
30
20
10 Eo
0
-120 -80 -40 0 40 80 120
Potential (mV) from Eo
in situ of the pulp, but with sampling and stirring the Eh of the sample can increase due to the
introduction of air into the pulp. The Eh of the plant pulp can also increase through pumping,
cycloning and its introduction to flotation. The reduction of oxygen can be written as:
PO 2 [H +] 4
Eh = E o + 0.059 Log
4 [H 2 O] 2
Substituting -pH for [H+], the activity of pure water being unity, and with Eo = 1.23 v yields:
1.2
PO
1 2 =1
PO PH
0.8 2 =10 -2 2 =10 -41
0 .6
PH
Eh versus SHE (V)
0.6 PO 2 =10 -3
2= 10 -40 1.6
0.4 PO PH
2= 2 =10 -4
10 -40 1.6
0.2 PH
PO 2 =10 -1
2 =10 -8 1.6
3.1
0
PH
2 =1
-0.2
-0.4
-0.6
-0.8
0 2 4 6 8 10 12 14
pH
Finally, it is possible to summarise the various redox and hydrolysis reactions for a system
in the form of an Eh/pH diagram (Figure 4 in the case of iron). In the case of vertical lines ‘A’
and ‘E’, these represent the equilibrium between Fe3+/Fe (OH)3 and Fe2+/Fe (OH)2, respectively.
These equilibriums are hydrolysis reactions only and so do not depend upon Eh, only pH.
In the case of horizontal lines ‘B’ and ‘C’, these represent the equilibrium between Fe3+/Fe2+
and Feo/Fe2+, respectively. These equilibria are redox reactions only and so do not depend
upon pH, only Eh. In the case of lines ‘D’, ‘F’ and ‘G’, these represent the equilibria between
Fe2+/Fe(OH)3, Fe(OH)2/Fe(OH)3 and Feo/Fe(OH)2, respectively. These equilibria involve both
redox and hydrolysis reactions and therefore have slopes which are neither simply horizontal
nor vertical. For the case of a ball mill discharge (Figure 4) in which there is oxidation of iron,
redox reactions involving this species will tend to dominate. This is reflected in the low Eh
(-200 mV, pH 7) of this stream. Fe is stable at low Eh values and will tend to decrease the Eh.
However, with pumping and cycloning, the introduction of air into the pulp raises the Eh at
the flotation feed sample point. This is because of:
•• the removal of the slurry away from reducing environment in the ball mill
•• the introduction of another electrochemical couple, O2/H2O.
It is possible to show that the Eh increase is not due to a change in pH over these two
sample points, as the Eh increases without any appreciable change in pH. The change
in Eh shows there must also be a change in the dissolved oxygen content of the pulp
(Figure 3). The initially depressed Eh in flotation, and changing Eh with aeration, have
consequences in flotation. Johnson, Jowett and Heyes (1982) have shown that the initial
flotation rate of the valuable mineral can be retarded due to the low Eh in the initial stages
of flotation. With aeration, the Eh increases above a critical value for collector adsorption.
This is a consequence of the Eh dependency of sulfide mineral flotation with thiol collectors
(Richardson and Walker, 1985).
Dissolved oxygen
Dissolved oxygen is the next most important variable to measure. However, its value in
a sampled pulp is often not at steady state. This is because of the continuous removal of
oxygen from the pulp due to mineral and media oxidation. The introduction of oxygen
1.5
‘A’
1 Fe 3+
‘B’ O2
‘D’ Fe (OH)3
0.5
Eh versus SHE (V)
Flotation
Fe 2+ Feed
0
H2 Ball Mill
‘F’
Discharge ‘E’
-0.5 Fe (OH)2
‘C’
Fe ‘G’
-1
0 2 4 6 8 10 12
pH
FIG 4 – Eh/pH diagram for the Fe system. [Fe]tot = 0.01 M. Also shown are values of Eh
and pH taken from an Australian gold mine. The stability limits of water are also shown.
in flotation systems raises the Eh of the pulps. Thus, when a sample is removed from a
pulp stream, the pulp is being taken away from the dynamic equilibrium which had been
established between the removal of oxygen in solution and the introduction of oxygen. In
some cases, it may be worthwhile to attempt to measure the dissolved oxygen (and Eh)
in situ of the pulp.
The dissolved oxygen electrode consists of a gold cathode and Ag/AgCl anode, placed
in an electrolyte solution. The solution is contained behind a plastic membrane, usually
polyethylene. A polarising potential of about 800 mV is applied between the two electrodes.
The gold electrode is placed close to the membrane and because of the polarising potential,
oxygen diffusing through the membrane will be reduced at the gold electrode:
O2 + 2H+ + 2e- = H2O2
The reduction of oxygen will produce a current through the electrode. A load thermistor
situated in the electrode itself converts this current into a voltage proportional to the oxygen
partial pressure. The thermistor has a temperature coefficient, which can compensate for
changes in ambient temperature, though this will depend upon the thickness and type of
plastic membrane. It is therefore always important to use the correct plastic membrane that is
matched to the particular electrode. In some ore systems containing galena or sphalerite and
large quantities of pyrrhotite, the critical variable appears to be dissolved oxygen, rather than
Eh, due to the oxidising influence of Fe3+ on Eh. In this case, collectors may be used which are
able to adsorb onto galena in a low oxygen environment (eg alkyl or aryl phosphinates), or
alternatively oxygen may be supplied to the pulp to oxidise pyrrhotite and increase the Eh,
allowing conventional xanthate collectors to be used.
O2 Concentration (ppm)
7
0
0 30 60 90 120 150 180
Aeration Time (minutes) in the Supply Chamber
FIG 5 – Dissolved oxygen concentration as a function of time, after oxygen purging the slurry for different periods of time. For
each determination of oxygen demand, the initial oxygen concentration was set to approximately 6–8 mg/L with air purging,
then the oxygen discontinued and the oxygen concentration measured with time. Slurry sample is grinding cyclone overflow.
d [O2] t
= - k1. [O2] t - k2. ([O2] t - [O2] eqm)
dt
d [O2] t
= - (k1 + k2) . [O2] t + k2. [O2] eqm
dt
Integration yields:
k2. [O2] eqm k2. [O2] eqm
[O2] t = + e[O2] o - o exp - (k1 + k2) t
(k1 + k2) (k1 + k2)
where:
[O2]o is the initial concentration of oxygen at t = 0 minutes when the oxygen purging gas is
turned off
[O2]eqm is the concentration of oxygen in the liquid when it is in equilibrium with the
atmosphere presented to it, having an approximate value of 8 mg/L when it is in
equilibrium with air
When k2 = 0 (ie the demand chamber is closed to the atmosphere) the oxygen concentration
is simply given by:
[O2]t = [O2]o exp − (k1)t
This equation has been used to fit the experimental data points shown in Figure 6, and the
value of rate constant k1 determined for the various amounts (ie purging time) of oxygen
delivered. It is possible to quantify the oxygen delivered with air delivery time in the
supply chamber by measuring the rate of oxygen dissolution in the liquid phase only at the
experimental temperature and pressure. The value of this rate constant is plotted against
time of air purging in the supply chamber (Figure 7). As the time increases, the rate of oxygen
depletion decreases. These experiments also show the effect of adding iron to the system.
Adding iron increases the oxygen demand of the slurry. The rate of decrease of oxygen
demand with aeration time is similar for the ore without iron and with iron, suggesting that
the rate of removal of oxygen is controlled by media oxidation.
4.0
PCO
3.0 3kg/t Fe
6kg/t Fe
2.5
15kg/t Fe
2.0
1.5
1.0
0.5
0.0
0 20 40 60 80 100 120 140
Aeration time (minutes)
FIG 6 – Rate of oxygen removal from solution (k1) as a function of aeration time. The various experiments
are for a plant cyclone overflow (PCO), a sample of ore taken during the measurement of the PCO and
ground in a laboratory mill (no Fe), and with various addition of iron to the laboratory mill.
Conditioned
Reagents Rougher Feed
Process Process
Water Primary X X X
Water Cyclone O/F
Rougher
X Block
Primary
SAG Mill X X
Ball Mill X
SAG X Includes regrind
pump box and cyclone
Feed
Reagents
Regrind
Reagents Ball Mill
Reagents
X
Reagents
Process Water
X
Cleaner
Block
X
X = Sampling Point
FIG 7 – Typical points (X) for measuring pulp chemistry values in a flotation
plant, showing typical reagent and process water addition points.
Alternatively, in the absence of the oxygen demand equipment described, it is possible also
to measure the oxygen demand in a beaker. In this case the procedure is to carefully place a
sample into a plastic beaker from a process stream and stirring for two minutes during which
time the dissolved oxygen is recorded with time. The dissolved oxygen can be measured over
this period and the rate of depletion measured.
K, for the complexes. This means that there is a strong driving force for the formation of metal
complexes with EDTA. Note that complexes of EDTA have an intermediate stability between
the sulfide and hydroxide. This means that the metal hydroxides adsorbed or precipitated
on the mineral surface will be dissolved with EDTA, but that EDTA will not dissolve the
unoxidised metal sulfide. This statement is true, provided that EDTA does not catalyse the
oxidation of the sulfide minerals. In this method, the slurry sample is first purged of oxygen
to reduce oxidation during the EDTA extraction stage. A prepurged solution of EDTA is then
added to the slurry to make an overall concentration of ~10-2 M. It is important that the EDTA
concentration be high enough to ensure the soluble species which can be complexed by EDTA
(eg calcium) do not consume EDTA, such that there is sufficient available to complex with the
products of surface oxidation. The amount of EDTA metal ions is determined by filtration and
measurement of the solution phase, with mass balancing. It is very important to quote the
EDTA extractable metal ions as both:
•• the concentration of EDTA-metal ions taking into account both sample dilution and metal
ions already in solution prior to extraction
•• the percentage EDTA-metal of the total metal in the head sample.
2. Pulp chemical measurements and sampling, including pulp samples for bulk chemical
and surface analysis and process water samples. This could also include specialised
measurements for reagents in solution and at the mineral surface. It should always include
the measurement of all reagent flows during the survey.
3. Cell characterisation measurements, including superficial gas velocity, bubble size and
residence time distribution by tracer tests (eg an estimation of energy dissipation in the
cell should also be made (usually in the range of 1–10 w/kg of pulp).
4. Batch flotation tests on process streams for the purpose of correlating laboratory conditions
to the plant and/or to build a circuit model based on component floatability. The batch
flotation tests will determine the fractions of different floatability in the feed.
5. Sample of SAG feed and circuit water for the purpose of laboratory testing. This is not
often carried out in conventional plant surveys, but it is very useful for optimisation at
laboratory scale tests. Obtaining these samples allows the plant conditions to be mimicked
at laboratory scale in grinding and flotation experiments.
The following section will focus on aspects of items 2, 4 and 5 above.
importance of water quality is the use of fresh water in the zinc cleaning stages of a lead/zinc
flotation plant to reduce lime consumption to achieve a set point pH value and, consequently,
improve selectivity between sphalerite and pyrite. The normal process water at this plant
is supersaturated in calcium sulfate (Grano et al, 1997), suggesting that calcium sulfate has
probably precipitated in the pulps. This has a pH buffering effect, meaning that higher lime
additions are required to achieve a set pH (~11), and that more calcium sulfate will precipitate
due to lime addition, a source of calcium. By using fresh water in the cleaning stages, lower
lime additions are required, and better selectivity is achieved. In the future, use of treated
water sources for special applications (eg cleaning, fine grinding) will likely become more
common. Another example is when thickeners are operated with dirty overflows and there is
residual polymeric flocculant adsorbed on the particles in the thickener overflow. This may
interfere with flotation due to the adsorption of these hydrophilic slime particles onto the
value mineral.
For block surveys, samples and measurements contrasting the products of the major
separation should also be undertaken. This includes those across the rougher block (ie rougher
concentrate, rougher tailing reporting to final tailing, and rougher feed) and cleaner block
(ie final concentrate, cleaner block tailing reporting to final tailing, and cleaner feed). EDTA
extractions and surface analysis samples (not discussed here) would contrast the oxidation
state of these streams. EDTA extraction and other simple measurements can be taken more
routinely and correlated to plant flotation performance. In general, the valuable metal in the
tailing streams has a greater proportion in an oxidised state than the valuable metal in the
concentrate streams, both extractions measured on a mass of metal basis when using the
EDTA technique. Sulfide flotation, under conventional conditions (eg without sulfidisation)
with thiol collectors, is always selective against oxidised forms. Changing oxidation levels in
the conditioned feed gives rise to changing recoveries in the plant.
It is very important to measure the reagent flows during the pulp chemical survey. This
should include all reagents and all addition points (eg collector stage added points, usually in
the scavengers, frothers – often overlooked in surveys, activators and depressants, including
any test reagents). The solution strength of the reagents should be noted, including taking a
sample for later analysis, and the reagents additions quoted as mls/min, g/t of total plant feed
and g/t of the specific process stream to which the reagents are added. This is very important
for any laboratory tests to be conducted on the sample of SAG feed collected during the
survey. A mass balance on reagents should be attempted, which may identify problems of,
for example, frother adsorption onto particles and high residual frother in the process tailings
and the continuing presence of frother in the process water (Tsatouhas, Vera and Grano, 2005).
The type of plant survey, which if comprehensive, should be the beginning point of any
optimisation program. It is important to undertake multiple, simpler block surveys at key
sample points, both before and after the comprehensive survey to establish plant stability and
the normal operating range of the plant. These simple surveys may be called quality control
surveys and are aligned to the comprehensive survey in their reporting. The quality control
surveys increase confidence that the comprehensive survey was conducted under conditions
which reflect the plant operating range at that time.
lime in conditioning and not to the mill in the laboratory, while lime is added to the mill in the
plant). To be more confident about successfully transferring the results of laboratory studies
to plant scale it is, at least, necessary to mimic the plant pulp chemical conditions and particle
size distribution at laboratory scale. This has the advantage that it will focus the initial stages
of laboratory investigations onto the effect of simple parameters, which may be relatively
easy to manipulate at plant scale, and which may acutely affect plant performance. This could
include the effect of water quality or water source, grind pH, aeration and standard plant
reagent additions, etc. The procedure outlined below adds more confidence to the testing of
future ores, which may have not yet been processed through the plant.
During the comprehensive survey, a SAG mill feed sample is taken (Figure 8) and treated
in the normal way to produce samples suitable for laboratory flotation testing. Storage of the
crushed and riffled sample should aim to minimise oxidation, though subjecting the sample
to the same standard laboratory flotation conditions with storage time may examine this,
along with routine EDTA extractions. During the comprehensive survey, flotation tests on the
conditioned rougher feed should ideally be undertaken in triplicate under known conditions
of the physical variables, such as cell type and mechanism, gas flow rate (Jg), impellor speed,
scrape rate and bubble size. These same physical conditions would then be used in latter testing
of the SAG feed sample to ensure that differences from the plant to the laboratory are only due
to differences in the feed (properties) presented to the flotation cell. During the comprehensive
survey, a large sample of circuit water should also be collected and used in laboratory flotation
testing (Figure 8). However, circuit water samples may not be stable due to the continuing
oxidation of thio-salts and microbiological activity. In this case, a simulated circuit water
sample should be prepared based on the inorganic composition of the circuit water, matching
all the major constituents of the circuit water during the survey including the total dissolved
solids (TDS). Unfortunately, it is very difficult to match the organic components that may
include, for example, collector decomposition products, bacteria and residual polymer from
dewatering. Comparing these water types may be very informative as well. Also, removing
Ore
(SAG Feed)
Reagents Reagents
Matching Conditions
FIG 8 – Procedure for assimilating the laboratory to the plant conditions of pulp chemistry.
some, or all, of the organic components from the circuit water (eg by adsorption onto activated
carbon) may determine the effect of these components on flotation performance.
Once these samples have been collected and treated appropriately, the SAG feed,
water samples and reagents may be used at laboratory scale. Special mills are available
(eg Magotteaux Mill®, Greet et al, 2004) which allow the pH to be continuously controlled
during grinding. This may be important in matching the laboratory to the plant as the pH
of the feed to most plant primary ball mills (ie the cyclone underflow) is the same pH as the
plant ball mill discharge. Most batch mills have wide variations in pH from the beginning
of grinding to the end, with the mill discharge pH usually being set as the target pH value.
This special mill (Magotteaux Mill®) allows continuous purging with different gases during
grinding to ensure that the laboratory mill discharge Eh is at the same value as the plant ball
mill discharge. Gas purging during grinding and post-grinding can be very informative in
terms of optimisation. Once grinding is complete the slurry is conditioned with the same
reagents and at the same addition level as during the plant survey. Laboratory flotation is
then conducted under the same physical conditions as that in the laboratory flotation tests
conducted during the plant survey. With care it is possible to duplicate the same performance
in the flotation tests during the plant survey. Differences will highlight variables which subtly
affect plant performance. These may include:
•• grind pH
•• grind Eh
•• flotation pH
•• flotation Eh
•• water quality
•• pulp temperature
•• grinding media
•• reagent quality, etc.
The parameters which require matching between the laboratory and plant trials are shown
in Figure 9. As well as Eh and pH, the particle size distribution of the ore and individual
minerals need to be reasonably matched. If there are significant deviations in the mineral size
distribution, then it may be necessary to analyse the laboratory flotation behaviour for both
tests on the plant conditioned feed and the SAG mill feed on a size-by-size basis. This can
be advantageous in providing flotation performance by particle size. The dissolved oxygen
demand and EDTA extractable metal ions can also be matched. However, these parameters
are dependent upon Eh and pH of both the grinding and conditioning stages. Matching the
values of Eh and pH will go a significant way towards matching dissolved oxygen demand
and EDTA extractable metal ions, provided the SAG feed sample has not oxidised during
storage (Pietrobon, Grano and Greet, 2004). It is also important to match the Eh and pH of
the mill discharge separately from the flotation feed Eh and pH values. This is because the Eh
during grinding is usually much lower than the Eh during flotation. However, if the Eh during
grinding is not mimicked then the conditions during laboratory grinding may not be realistic.
The grind and flotation Eh and pH values should be considered separately in the study.
GRINDING FLOTATION
FIG 9 – Parameters that need to be matched from the plant to the laboratory.
plant which normally treats porphyry copper ore (ie copper in chalcocite and chalcopyrite
in igneous host rock), but for which the proportion of the run of mine (ROM) ore from
sulfide skarns is generally increasing throughout the remainder of the mine life. Sulfide
skarns emanate from the intrusion of the igneous into the country rock and generally
contain higher copper and gold concentrations, but also much higher pyrite and magnetite
contents (Orwe, Grano and Lauder, 1997). The flotation plant encountered problems when
treating feeds with high sulfide skarn content such as low copper recovery (<75 per cent) and
low final concentrate grade (<20 per cent Cu). Ore blending was only partially effective and
cannot be used in the latter stages of the mine life due to the preponderance of sulfide skarn.
The low concentrate grade was a consequence of both high pyrite recovery and low copper
recovery. The normal conditions in the plant (lime addition to SAG feed to pH 11–11.5 and
using a dithiophosphate collector in the primary ball mill) were inadequate in the flotation
separation of sulfide skarn ores. The copper-bearing minerals were adequately liberated
from gangue being ~80 per cent fully liberated, across all size ranges, at the normal grind
size of 150 μm in flotation feed. The level of copper mineral liberation was the same or
greater than for porphyry sulfide ores. The reason for the low copper concentrate grade and
recovery were related to pulp chemistry differences between the porphyry and sulfide skarn
ores, rather than differences in mineral liberation.
In the preceding section, the stability diagram of iron was discussed. The stability diagram
shows the thermodynamically most stable species (ie the species at the highest concentration)
at various Eh and pH values while assuming a certain total concentration of the species.
Attention will now turn to the Eh-pH diagram for the Cu-S-H2O system (Figure 10). The Eh/
pH measurements at the cyclone overflow from two plant surveys on a porphyry ore are also
shown (Orwe, Grano and Lauder, 1998). The principal copper-bearing sulfide minerals are
chalcocite and chalcopyrite. The normal pH for this plant is 11.5. The Eh/pH diagram shows
that chalcocite is only stable to approximately Eh 0 mV at pH 11, while above this Eh value
it may begin to oxidise. However, the Eh of the pulp at the cyclone overflow is +150 mV
(Survey 1), suggesting that chalcocite oxidation may occur. Oxidation was confirmed by
EDTA extraction and surface analysis (Orwe, Grano and Lauder, 1998). In this system, further
aeration of the pulp in flotation does not appreciably increase the Eh, suggesting that the pulp
800
O2
H2 O o
CuO + S
CuS + Cu S Plant
400 2-x Cyclone O/F
Survey 1
Eh / mV 0 Cu2 S Cu S + CuO
2-x
H2 O
Cu2 S Plant
H2
-400 Cyclone O/F
Survey 2
o o 2-
Cu + H 2S Cu + HS- Cu + S
-800
0 2 4 6 8 10 12 14
pH
FIG 10 – Plant cyclone overflow pulp potential (Eh) as a function of pH in plant surveys on porphyry ore blends (head grade
2.05 per cent Cu, 5.6 per cent S, Cu/S 0.37, six per cent pyrite) for Survey 1 without sulfidisation and Survey 2 with sulfidisation.
is fully aerated under these high pH conditions (ie DO is high and the DO demand of the
minerals and ore does not outweigh oxygen delivery to the pulp). In Survey 2, NaHS (~100 g/t)
was added to the plant cyclone overflow to achieve an Eh of -50 mV. This decreased the Eh of
the cyclone overflow as shown into the stability region of Cu2S (Jones and Woodcock, 1978).
Oxidation would be prevented while oxidation products on chalcocite (copper hydroxide)
would be reduced to a ‘Cu-S’ phase. With aeration in the plant flotation cells the Eh increases
to +150 mV, which is an acceptable value for collector adsorption onto chalcocite (assuming the
sulfidised chalcocite does not re-oxidise) (Trahar, 1984). Copper recovery increased with this
treatment though the extent of this increase depended upon copper mineralogy (chalcocite
oxidises more than chalcopyrite and so the effect is less marked for chalcopyrite dominated
feeds), the feed particle size (fine chalcocite is more oxidised than coarser chalcocite), the
sulfidisation potential and other factors.
Another example of the application of presenting pulp chemical data in an Eh/pH
representation is shown in Figure 11. Presenting data in this way gives order to the data as it
allows changes in Eh, which may have been caused by changes in pH, to be detected. In this
example, both plant and laboratory data are shown (Putubu, Grano and Morey, 2001). The
Eh and pH experimental values from plant surveys conducted in October (on predominantly
porphyry ore) and March (on predominantly skarn ore) 2000, are compared to laboratory
data for moderate (shown as 20/11/98 SAG feed) and high pyrite skarn feed types (shown
as PSM). The plant Eh and pH values for skarn ore (high pyrite) surveys are lower than the
porphyry ore survey, but also lower than the Eh values after laboratory grinding of skarn
samples. It should be noted that for high pyrite skarn ores, higher (~4 kg/t) than normal
(~1 kg/t) lime additions are required to achieve the target pH (11.5). In spite of the higher
lime additions, the pH value is often lower than the target pH for high pyrite skarns. This is
because of pyrite oxidation, which generates acid in solution.
In the latter cases, laboratory grinding at high pH causes adequate chalcopyrite recovery
and selectivity from a high pyrite feed (PSM). Surface oxidation of pyrite in the laboratory
experiments assists depression of copper activated pyrite and the promotion of chalcopyrite
recovery. Chalcopyrite recovery is increased due to increased collector adsorption on its
surface due to reduced collector adsorption onto pyrite. It is also useful to note that without
pH control, the pH levels of the pulps generally decrease with aeration conditioning
450
60
Aeration
30 O2
400 10
Time
H2O
60
30
350
Eh / mV
3 4 5 6 7 8 9 10 11 12 pH
20/11/98 SAG feed sample Pit Skarn Magnetite (PSM) sample
Head Grade 2.05% Cu, 5.6% S, Head Grade 1.59% Cu, 10.7% S
Cu/S 0.37, 6% Pyrite Cu/S 0.15, 17% Pyrite
FIG 11 – The pulp potential (Eh) as a function of pH in plant surveys on porphyry ore blends (17 October 2000) () and sulfide
skarn blends (March 2000) (), as well as for laboratory tests on 20 November 1998 SAG feed sample (head grade 2.05 per cent
Cu, 5.6 per cent S, Cu/S 0.37, six per cent pyrite) (open symbols) and on pit skarn magnetite (PSM) sample (head grade
1.59 per cent Cu, 10.7 per cent S, Cu/S 0.15, 17 per cent pyrite) (closed symbols) for aeration tests at different initial grind pH values
(~11.5, ~10 and natural pH). The numerical values are aeration times in minutes after grinding the ores in the laboratory.
(Figure 11). In the case of grinding at pH 11.5 and 10 this is accompanied by increases in
Eh, which may be accounted for by changes in pH through the oxygen/water equilibrium.
This is because with aeration, the pH/Eh relationship parallels the oxygen/water equilibrium.
However, in the case of grinding at natural pH, the Eh increases markedly after grinding
suggesting an increase in the dissolved oxygen content of the pulp with aeration. The pH/
Eh relationship does not parallel the oxygen/water equilibrium for test at natural pH. In this
case, the oxygen content of the pulp increases with aeration. In these laboratory experiments,
pyrite very quickly passivates at alkaline pH values. It is noted that the optimum potential
for chalcopyrite flotation occurs in the Eh range of 200 to 250 mV (SHE) (Trahar, 1984). When
the pyrite is not adequately passivated (at alkaline pH), the pyrite consumes oxygen (and
collector), which retards chalcopyrite flotation in this system.
100
Dissolved Oxygen / %
Cell 15 Tail
80 Cell 9 Tail
Cell 3 Tail
60
Cell 15 Tail
Rougher Feed Cell 9 Tail
40
Cell 3 Tail
20 Rougher Feed
A B
0
0 0 30 60 90 120 0 30 60 90 120 Time / sec
Porphyry ore blend Sulfide skarn blends
17/10/2000; average for 15/3 to 21/3/2000;
Head Grade Head Grade
0.6% Cu, 1.2% S, Cu/S 0.5 ~1.2% Cu, ~4.0% S, ~Cu/S 0.3
FIG 12 – Dissolved oxygen concentration as a function of time (dissolved oxygen demand) after the discontinuation of
air purging for plant rougher streams in surveys on (A) porphyry ore blends (17 October 2000; head grade 0.6 per cent
Cu, 1.2 per cent S, Cu/S 0.5) and (B) sulfide skarn blends (average for 15–21 March 2000; head grade ~1.2 per cent Cu,
~4.0 per cent S, ~Cu/S 0.3). The acid soluble copper and Cu/S ratio is a very important parameter in porphyry sulfide ores.
minerals, most probably pyrite, at alkaline pH values and under laboratory grinding and
flotation conditions is demonstrated in Figure 13. The high pyrite skarn feed (PSM) requires
a longer aeration time to achieve complete oxidation and that grinding under alkaline
conditions is important for pyrite oxidation.
Following these plant surveys, flotation tests were performed on samples of SAG feed.
Attention turned to the effect of grind pH on copper recovery and selectivity against pyrite for
laboratory tests of SAG feed (Figure 14). For tests with lime addition after the mill (Figures 14a
and 14b), pyrite readily responds to collector addition in spite of lime addition to pH 11.5.
Under these conditions, copper shows lower recovery and rate. This flotation behaviour is
characteristic of problems encountered under normal plant conditions during treatment of
high pyrite skarn ores. In the plant it is not possible to increase collector addition to counter
low copper recovery due to the increased pyrite recovery when treating sulfide skarns under
normal plant conditions. Radically different results are obtained with lime addition to the
laboratory mill for pH values above ten (Figures 14c and 14d). In this case, pyrite recovery
is markedly decreased, and copper maximum recovery and rate greatly increased. It seems
clear that without appropriate control of the grinding environment, in particular grind pH
and grind Eh, pyrite activation ensues, and copper recovery is very low.
In terms of plant practice, it was recommended that lime addition be made to the primary
ball mill feed, as well as SAG feed, and that collector addition be made after the grinding
circuit. The objective here was to passivate, or oxidise, pyrite prior to collector addition so that
most of the collector would then be directed towards chalcopyrite, and not pyrite. It was also
recommended that further work should investigate stage-adding collector, not only for coarse
composite recovery, but also for pyrite control. The effect of using the first cell in flotation as an
aerator to passivate pyrite, prior to collector addition, also needed to be investigated for high
pyrite skarn feeds. Once conditions are established which ensure effective pyrite depression,
collector additions may then be generally increased, in accordance with the higher copper
grades of the sulfide skarn ores. It was also recommended that dissolved oxygen demand
and EDTA extraction should be used to monitor pyrite oxidation and copper activation, and
100
90 20/11/98 SAG Feed
80 Pit Skarn Magnetite Ore
Dissolved Oxygen / %
70
60
50 A
20/11/98 SAG Feed
90
Pit Skarn Magnetite Ore
80
70
Aeration
60
Time
50 B
30 min
90 10 min 20/11/98 SAG Feed
80 3 min
30 min
70
60 10 min Pit Skarn Magnetite Ore
50 C 3 min
40
0 20 40 60 80 100 120 Time / seconds
FIG 13 – Dissolved oxygen concentration as a function of time (dissolved oxygen demand) after the
discontinuation of air purging for laboratory tests on 20 November 1998 SAG feed sample (closed symbols)
and on pit skarn magnetite (PSM) sample (open symbols) for aeration tests at different initial grind pH values,
(A) 11.5, (B) 10 and (C) natural pH. Aeration times of three (Δ), ten (), and 30 () minutes shown.
6 30
pH 6.8
Pyrite Recovery / % Copper Grade / %
25
4 pH 7.5
20
pH 8 pH 9.3
15
pH 10 pH 10.6
2 10
pH 11 pH 11.5
A pH 11.5 5
C
0 0
80 40
60 30
40 20
Lime after mill Lime in mill
20 10
B D
0 0
0 20 40 60 80 100 0 20 40 60 80 100
Copper Recovery / % Copper Recovery / %
FIG 14 – Copper grade/recovery relationship (A, C) and selectivity of copper against pyrite recovery (B, D) for
laboratory flotation tests on pit skarn magnetite (PSM) sample head grade 1.59 per cent Cu, 10.7 per cent S,
Cu/S 0.15, 17 per cent pyrite with d80 144 microns 30 g/t 7249 collector added after grinding and pH adjustment,
for tests with lime added after milling at natural pH (A, B) and for lime added to the mill (C, D).
this data correlated to plant performance. It was recommended that laboratory tests should
continue to focus on correlating laboratory conditions to the plant, while the effect of Eh,
controlled by gases in the mill and conditioning stages, should also be investigated. The
possibility of increasing the Eh from +170 mV to +230 mV during the treatment of high pyrite
sulfide skarn blends should also be investigated. Based on the plant Eh/pH relationship
(Figure 11), it is unlikely that simply increasing the pH to above 11.5 would be adequate to
achieve this desired increase in Eh.
Zinc Concentrate
Pyrite Cleaner
Tailing (from
Pyrite Cleaner)
SO2 Conditioning
Heat Conditioning
Upgraded
Air Roughing Zinc Concentrate
Pyrite Rougher
Concentrate
(to Pyrite Cleaner)
FIG 15 – Simplified circuit diagram for the plant zinc reverse cleaning process.
400
300 22oC
SO2 Conditioning
Pulp Eh (mV)
-100
3 4 5 6 7 8 9 10 11 12
Pulp pH
FIG 16 – Pulp Eh-pH diagram for the plant zinc reverse cleaning (BMS).
heat conditioning stage, there was a considerable change towards more reducing conditions
in comparison to those after addition of sulfur dioxide (Figure 16). This change is easily
seen in the Eh-pH diagram because there was no complicating simultaneous large change
in pH. The main feature for the flotation stage with air was an increase in the Eh value.
Unlike the two previous process steps where the Eh values had moved further away from
the experimental water/oxygen line, the Eh values moved back towards the water/oxygen
line (Figure 18).
Attention will now turn to the pilot plant studies at Mt Isa lead/zinc concentrator (Johnson
and Munroe, 1988). At the commencement of the pilot plant testing of zinc reverse, only a
single pyrite rougher stage was used (ie no cleaning of the reverse concentrate), and the
feed to the reverse process typically assayed 23–24 per cent Zn. In this case, the zinc reverse
process was applied to a plant stream which contained high levels of slow floating sphalerite.
This type of sphalerite is difficult to upgrade in conventional cleaning because of its small
flotation differential against gangue sulfide minerals. In reverse flotation, the rate constant
for the sphalerite in conventional cleaning is irrelevant, and therefore the reverse flotation
approach provides one possible solution to the problem of upgrading this type of sphalerite.
In the initial stage of the pilot plant study, the pyrite rougher stage only occasionally
performed to expectation, and it was apparent that it was insufficiently developed to
routinely maintain the expected daily result. The zinc grade was expected to increase by
0.8 units for each unit of zinc loss, but the observed increase was only approximately 0.38
(Table 1, large cells). However, surveys of Eh/pH values in the pyrite rougher bank and
conditioning stages provided very important clues to the cause of the inconsistent depression
of sphalerite (Figure 17). The data revealed a very large increase in Eh (~+200 mV) between
the last conditioner and the first cell of the pilot plant pyrite rougher bank (Figure 17,
unsuccessful process). This behaviour differed from the previously described behaviour for
plant processes where an Eh increase of only ~50 mV was observed for the entire pyrite
rougher flotation stage.
The large increase in Eh meant that even the first cell of the pyrite rougher was destroying
the deliberately created chemical environment in the conditioners. The pulp flowed a
horizontal distance of four meters by gravity from the last conditioner and the first cell, which
was completely enclosed from the atmosphere. It was reasoned that there was little scope
for accidentally raising the Eh during transportation through the pipe and that the oxygen
in the flotation cell was destroying the reducing environment created in the conditioners.
Therefore, a solution to the problem (at that time) was the use of smaller flotation cells with
weirs at intervals to lower exposure of the pulp to oxygen in the initial cells (Johnson and
Munroe, 1988). The use of weirs was necessary to eliminate any possibility of back mixing
TABLE 1
Summary of some daily metallurgical results for the pyrite rougher before and after changing the size of the cells in the pilot plant.
Cells Assays/zinc (%) Zinc recovery (%) Zinc grade Zinc grade increase
Feed Concentrate Tailing to concentrate increase (%) per unit zinc loss
Large 23.7 43.1 16.6 48.7 19.4 0.38
Small 21.9 49.7 10.1 67.6 27.8 0.86
400
-100
3 4 5 6 7 8 9 10 11 12
Pulp pH
FIG 17 – Pulp Eh-pH diagram for two configurations of the zinc reverse flotation process.
of comparatively aerated pulps from the latter cells (Johnson and Munroe, 1988). Another
possibility was to use nitrogen instead of air as the carrier gas in flotation.
A robust (ie a consistent performance irrespective of feed condition) process was obtained
at the expected level after the change to smaller cells with weirs (Table 1). Surveys of the
Eh/pH values of the new pyrite rougher bank showed that a much smaller increase in Eh
values (~100 mV) existed between the last conditioner and the first cell (Figure 17, successful
process). Further, the increase in Eh for the remainder of the pyrite rougher was quite
small, ~20 mV. The zinc concentrate (45–52 per cent Zn) from the pyrite rougher concentrate
typically assayed 6.5 - 8.5 per cent iron after successful removal of the pyrite. Later inclusion
of cleaning to the process improved its performance further. There is an indication that the Eh
in flotation should not exceed +150 mV to ensure depression of sphalerite in reverse flotation.
This work demonstrated the importance of measuring both Eh and pH to solve flotation
separation problems. It also suggested that on-line measurement of these variables may
provide a more stable and robust process. It was also speculated, though not examined,
that use of an inert gas (eg nitrogen) in flotation may benefit flotation by either improving
selectivity or reducing the need for pulp heating, which may improve the process economics.
the percentage of solids may have beneficial effects on molybdenite recovery due to a
number of mechanisms related to the factors outlined above. At least one is to reduce the
concentration of fine gangue particles which may potentially adsorb onto molybdenite.
The effect of percentage solids has been investigated in three different rougher flotation
circuits and as a function of particle size (Figure 18). Significant increases in molybdenite
recovery are noted, and are greatest in the coarse particle size fraction where molybdenite
losses are greatest, occurring in predominantly fully liberated form to 250 microns. This
effect is related to both reduced particle interactions, an effect related to ore type, and also
to increased particle-bubble collision efficiency and frequency at low percentage solids.
At lower percentage solids, the rheology of the medium is reduced increasing the effective
shear rate. Molybdenite recovery increases even though the cell residence time decreases
due to higher water additions.
CONCLUSIONS
This chapter has dealt with simple pulp chemical measurements which should be taken with
most plant surveys. The basic theory, interpretation and utility of these measurements are
demonstrated.
100
Reduced %solids
81% Mo Recovery
80 Across all Size Fractions
60
Normal %solids
71% Mo Recovery
40 Across all Size Fractions
Molybdenite Recovery / %
20
A
0
Reduced %solids
80 85% Mo Recovery
Across all Size Fractions
60
Normal %solids
40 73% Mo Recovery
Across all Size Fractions
20
B
0
Reduced %solids
85% Mo Recovery
80 Across all Size Fractions
60
Normal %solids
75% Mo Recovery
40 Across all Size Fractions
20 Particle Size
C / Microns
0
10 100 1000
FIG 18 – Molybdenite recovery as a function of particle size across three plant rougher:
(A) Plant A, (B) Plant B, (C) Plant C, for different per cent solids in the rougher flotation feed.
REFERENCES
Ametov, I, Grano, S R, Zanin, M, Gredelj, S, Magnuson, R, Bolles, T and Triffet, B, 2008. Copper and
molybdenite recovery in plant and batch laboratory cells in porphyry copper rougher flotation, in
Proceedings 24th International Mineral Processing Congress (IMPC), Beijing, 24–28 September, pp 1129–
1137.
Broman, P G, Hultqvist, J and Marklund, U, 1985. Experience from the use of SO2 to increase selectivity
in complex sulphide ore flotation, in Developments in Mineral Processing, Flotation of Sulphide Minerals
(ed: K S Forssberg), pp 277–291 (Elsevier: Stockholm).
Fornasiero, D, Eijt, V and Ralston, J, 1992. An electrochemical study of pyrite oxidation, Colloids
Surfaces, 62:63–73.
Garrels, R M and Christ, C L, 1965. Solutions, Minerals and Equilibria (Harper and Row: New York).
Grano, S R, Johnson, N W and Ralston, J, 1997. Control of the solution interaction of metabisulphite
and ethyl xanthate in the flotation of the Hilton ore of Mount Isa Mines Limited, Australia, Minerals
Engineering, 10(1):17–39.
Grano, S R, Lauder, D W, Johnson, N W and Ralston, J, 1997. An investigation of galena recovery
problems in the Hilton Concentrator of Mount Isa Mines Limited, Australia, Minerals Engineering,
10(10):1139–1163.
Greet, C J, Small, G L, Steinier, P and Grano, S R, 2004. The Magotteaux Mill®: investigating the effect of
grinding media on pulp chemistry and flotation performance, Minerals Engineering, 17(7–8):891–896.
Johnson, N W, 1988. Application of electrochemical concepts to four sulphide flotation separations,
in Proceedings International Symposium Electrochemistry in Mineral and Metal Processing II
(eds: P E Richardson and R Woods), vol 88–21, pp 36–48 (The Electrochemical Society).
Johnson, N W, Jowett, A and Heyes, G W, 1982. Oxidation-reduction effects in galena flotation:
observations on Pb-Zn-Fe sulphides separation, Transactions of the Institution of Mining and Metallurgy,
Mineral Processing and Extractive Metallurgy, 91:C32–C37.
Johnson, N W and Munroe, P D, 1988. Eh-pH measurements for problem solving in a zinc reverse
flotation process, The AusIMM Proceedings, 293(3):53–58.
Jones, M H and Woodcock, J T, 1978. Evaluation of an ion-selective electrode for control of sodium
sulphide additions during laboratory flotation of oxidised ores, Transactions of the Institution of
Mining and Metallurgy, Mineral Processing and Extractive Metallurgy, 87:C99–C106.
Jones, M H and Woodcock, J T, 1984. Application of pulp chemistry to regulation of chemical
environment in sulphide mineral flotation, in Proceedings Principles of Mineral Flotation (The Wark
Symposium) (eds: M H Jones and J T Woodcock), pp 147–183 (The Australasian Institute of Mining and
Metallurgy: Melbourne).
Kant, C, Rao, S R and Finch, J A, 1994. Distribution of surface metal ions among the products of
chalcopyrite flotation, Minerals Engineering, 7(7):905–916.
Kristall, Z, Grano, S R, Reynolds, K, Smart, R St C and Ralston, J, 1994. An investigation of sphalerite
flotation in the Murchison Zinc concentrator, in Proceedings Fifth Mill Operators’ Conference, pp 171–
179 (The Australasian Institute of Mining and Metallurgy: Melbourne).
McTavish, S, 1980. Flotation practices at Brunswick Mining, CIM Bulletin, February, pp 115–120.
Montalti, M, Fornasiero, D and Ralston, J, 1991. UV-visible spectroscopic study of the kinetics of
adsorption of ethyl xanthate on pyrite, Journal of Colloid and Interface Science, 143:440–450.
Natarajan, K A and Iwasaki, J, 1973a. Practical implication of Eh measurements in sulfide flotation
circuits, Transactions of the American Institute of Mining, Metallurgical, and Petroleum Engineers (AIME),
254:323–328.
Natarajan, K A and Iwasaki, I, 1973b. Effect of poisoning of platinum electrodes on Eh measurements,
Transactions of the American Institute of Mining, Metallurgical, and Petroleum Engineers (AIME), 254:22–28.
CHAPTER 8
Electrochemical aspects of
sulfide mineral flotation
Ronald Woods OAM
ABSTRACT
Flotation is considered from an electrochemical perspective. Electrode potential is a
determining parameter in recovery and can be monitored and controlled in the laboratory
and the plant. The underlying electrochemical science is reviewed and its application to plant
practice explored.
oxygen’. That is, the interaction of thiol collectors with mineral surfaces comprises a mixed-
potential mechanism in which an anodic oxidation involving the collector transfers electrons
to the mineral and the simultaneous cathodic reduction of oxygen returns this charge to the
solution phase. Most sulfide minerals are semiconductors with high electronic conductivity
and hence can readily sustain such processes. Nixon (1957) suggested that ‘prominent
theories of flotation could be reconciled’ by the electrochemical approach. This is illustrated
in Figure 1, in which the requirement of a neutral surface species espoused by Cook (Cook
and Nixon, 1950) can be met in different ways. The anodic reaction can be adsorption as
proposed by Wark (Figure 1a). Also, it can be the formation of a metal thiol compound as
proposed by Taggart (Taggart, Taylor and Knoll, 1930). The latter process can occur as a single
step as in 1b, or through separate surface oxidation and ion exchange processes as shown in
1c and 1d. In addition, the anodic process can be the formation of the dithiolate as suggested
by Nixon (1957) and illustrated in 1e.
Each of these mechanism types has been identified experimentally and shown to occur in
flotation systems. They all impart hydrophobicity to mineral surfaces. Chemisorption is the
thermodynamically favoured reaction for most mineral/collector systems, which means it
occurs first. As will be shown later, it is sufficient, in itself, to induce efficient flotation.
Chemisorption involves bonding between collector molecules and metal atoms in the
mineral surface. The attachment of ethyl xanthate to a galena surface is illustrated in Figure 2.
The lower spheres are sulfur (white) and lead (black) atoms in the galena surface. The sulfur
atoms of the xanthate are smaller than those in the mineral because they have no ionic character.
The xanthate sulfurs are equally bonded to lead atoms in the mineral surface. This bonding
is similar to that occurring in lead xanthate molecules, but in chemisorption, the metal atoms
remain part of the mineral lattice. Various spectroscopic investigations have confirmed this
mode of attachment of xanthate to mineral surfaces (Buckley, Hope and Woods, 2003). With
collectors such as 2mercaptobenzothiazole and butylethoxycarbonylthiourea, bonding with
metal atoms in the mineral surface was found to occur through nitrogen atoms as well as
sulfur atoms (Goh et al, 2006; Hope et al, 2006).
The electrochemical mechanism has two important implications. First, the potential across
the mineral/solution interface will be an important parameter in determining flotation
recovery. Second, the reaction imparting floatability, the anodic process involving the
collector, is amenable to investigation using electrochemical techniques. From the viewpoint
of the flotation engineer, it is the former of these implications that is more important and this
attribute will be the focus of this chapter. Since the potential can be measured in flotation cells
in the laboratory and in flotation plants, it can provide the engineer with a diagnostic tool
for developing flotation strategies and for alleviating flotation problems. Nowadays, most
laboratories apply potential measurement to derive optimum flotation strategies for operation
in the plant. As pointed out by Hintikka and Leppinen (1995) ‘monitoring of potential with
noble metal electrodes is now almost routinely carried out in laboratory testwork and full-
scale processes’. The reader is referred to Chapter 7 in the present volume for information on
the practical aspects of measuring potentials in laboratory cells and in flotation plants. Early
examples of the application of potential measurements in developing flotation strategies are
those of Johnson (1988) at Mt Isa Mines in Queensland. More recent examples are those of
Leppinen, Hintikka and Kalapudas (1998), in which the influence of potential on flotation
separation in copper and zinc from complex sulfide ores was considered, and of Smith and
Bruckard (2007) and Smith, Davey and Bruckard (2012), in which flotation strategies for the
M X X- S MX2 X
-
S M
M X
S S MX2
M X M SO42-
e- S e- S MX2
M M
S SM
M
S 02 MS O2
M SM
MS
0H- OH
Anodic Anodic
X- Xods + e- 2X- + MS + 4H2O MX2 + SO42- + 8H+ + 8e-
Cathodic Cathodic
O2 + 2H2O + 4e- 4OH- O + 2H O + 4e-
2 2
4OH-
S -
MSO4 S MX2 X
M H+ M
S MSO4 S MX2
M M
e- S MSO4 S MX2 SO42-
M M
S M S M
M S O4 M S
S M S M
OH-
Anodic
MS + 4H2O MSO4 + 8H+ + 8e-
Cathodic
O2 + 2H2O + 4e- 40H-
M X X-
S I
M X
S X
e- M I
S X
M
O2
S
M
OH-
Anodic
2X- X2 + 2e-
Cathodic
O2 + 2H2O + 4e- 2OH-
FIG 1 – Schematic representation of the mixed potential mechanism for the interaction of thiol collectors (X-) with sulfide minerals
in which the anodic process is (A) chemisorption; (B) reaction to form a metal collector compound; (C and D) the reaction in (B)
occurring in two stages, (C) oxidation of the mineral and (D) ion exchange with the collector; and (E) formation of the dithiolate.
Pb
S(Pbs)
S(xanthate)
O
H
FIG 2 – Schematic of the attachment of chemisorbed xanthate to a galena surface.
separation of copper-arsenic minerals such as tennantite and enargite from chalcopyrite and
bornite were developed.
ELECTROCHEMICAL POTENTIAL
The electrochemical potential is the driving force for processes in which electrons are
transferred between a conducting solid and species in solution. The web-based Electro
chemical Dictionary (https://www.corrosion-doctors.org/Dictionary/Dictionary-E.htm) says
the electrical potential:
… is somewhat analogous to the difference in height in a waterfall that causes the water to fall,
or the difference in pressure in a pipeline that causes the gas to flow.
Electrode potentials are measured against a reference electrode and are often converted to
the standard hydrogen electrode (SHE) scale. For example, potentials measured against the
silver/silver chloride 3.5 mol dm3 potassium chloride reference electrode at 25°C require the
addition of 0.205 V to convert them to the SHE scale. Similarly, for the calomel electrode at
25°C, 0.244 V has to be added to make them the potential vs SHE. Potentials on the SHE scale
are often referred to as the Eh.
The equilibrium potential of an electrochemical process is the potential at which no overall
reaction can take place. As the potential is taken more positive than the equilibrium value,
an anodic oxidation reaction occurs, in which electrons are transferred from solution species
to the electrode. Conversely, as it is taken more negative, the reverse, a cathodic reduction
reaction takes place in which electrons are transferred from the electrode to solution species.
The potential is a mixed potential when it is determined by two or more different
electrochemical processes occurring on the same electrode surface. Mixed potential systems
are non-equilibrium systems and hence are determined by the kinetics of potential-
determining reactions as well as the thermodynamics. The situation in which two redox
couples are involved is illustrated in Figure 3.
E1 and E2 in Figure 3 are the reversible potentials of the two couples and correspond to
equilibrium between forward and back reactions for each of the couples in the absence of the
other. When both are present, the potential will take up a value at which the two processes
give equal and opposite currents so that the net charge transfer across the interface is zero.
This potential is the mixed potential. It can be seen that the system involves an anodic
oxidation and a cathodic reduction reaction proceeding at finite rates and hence an overall
chemical reaction is taking place.
Anodic
E2
E1
CURRENT
Mixed
potential
Cathodic Potential
It should be pointed out that both processes that make up the mixed potential system can
proceed on a homogeneous surface. The surface is then acting both as a source and a sink for
electrons. On the other hand, a heterogeneous surface may have areas of different activity for
the individual processes and in this case the anodic and cathodic reactions occur on different
surface sites. The potential is the key parameter in determining whether the important
reaction can take place. It can be seen from Figure 3 that, for the oxidation of the species
characterised by E1 to take place, the potential must be greater than E1.
Mixed potential mechanisms are important in a range of reactions of metallurgical
interest, for example corrosion of metals, cementation (metal displacement) reactions,
hydrometallurgical processing, supergene alteration of sulfide orebodies, spontaneous
oxidation of sulfide minerals in transport and storage, as well as in sulfide mineral flotation.
The important potential determining flotation is that occurring across the mineral/solution
interface. It is generally measured by inserting a noble metal electrode into the pulp, in which
case it is often referred to as the pulp potential. It is presumed that sufficient mineral particles
contact the noble metal surface for the measured potential to be that of the mineral of interest.
The solution potential is often determined in metallurgical systems and is measured by a noble
metal electrode inserted into filtered solution. This is different from the potential across the
mineral/solution interface and may not relate to flotation recovery. Only when a very strong
redox couple is present in solution will the two potentials be the same and this is due to
the redox couple determining the potential of the indicator electrode as well as that of the
mineral/solution interface (Rand and Woods, 1984).
GALVANIC INTERACTIONS
It was pointed out above that a heterogeneous surface may have areas of different activity for
the individual processes in a mixed potential system and in this case the anodic and cathodic
reactions occur on different surface sites. When two dissimilar conducting solids are in contact,
it is possible for the anodic process to occur on one solid and the cathodic on the other. This
situation is utilised in corrosion prevention through cathodic protection of a metal structure
such as a ship or a bridge with a sacrificial anode composed of a more electrochemically
active metal such as magnesium. The manner in which this procedure operates is illustrated
in the schematic shown in Figure 4.
Anodic
CURRENT
A
Potential
Cathodic
The solid anodic curve represents corrosion of a metal to form an ionic species or an oxide.
The solid cathodic line represents the reduction of oxygen derived from the atmosphere. The
mixed (corrosion) potential will be at A and it can be seen that corrosion will be proceeding
at the current flowing at point A. When a sacrificial anode is attached to the corroding metal,
oxidation of this metal occurs to give the dashed curve and the mixed potential now shifts to
B. It can be seen from Figure 4 that, at this potential, there is a negligible contribution to the
anodic current at the mixed potential from the metal to be protected. Thus, corrosion of the
structure of interest does not take place.
Similar galvanic interactions occur in mineral flotation systems. Grinding of sulfide ores is a
prime example. The grinding media galvanically protects the sulfide minerals in an ore from
oxidation that would impair interaction with a collector in the subsequent float cells. On the
other hand, oxidation of the grinding media results in iron oxy-hydroxy species that inhibit
flotation through coating the minerals of interest with a hydrophilic layer. Galvanic interaction
between grinding media and sulfide ores has been the subject of many investigations reported
in the literature, from those of Iwasaki and co-workers in the 1980s (Adam, Natarajan and
Iwasaki, 1984) to those today (for example Grano and Huang, 2006; Peng and Grano, 2010).
Similarly, sulfide minerals can undergo galvanic interaction with each other and this
can enhance oxidation of a mineral and affect selective flotation. For example, pyrite has a
relatively high rest potential and also strongly supports oxygen reduction (Biegler, Rand and
Woods, 1977). If a sulfide such as galena is in contact with pyrite in a flotation pulp, then
its oxidation is not only driven by oxygen reduction on itself, but also that on pyrite and,
consequently, is increased significantly. This can result in loss in flotation recovery. Again,
this phenomenon has been the subject of many investigations reported in the literature, from
the 1960s (Majima, 1969) to today (Mu, Peng and Lauten, 2018). Galvanic interaction between
sulfides can also promote spontaneous heating in mining, storage, shipping and disposal of
sulfide minerals (Payant et al, 2012).
20 μA/cm2
Anodic
CURRENT
-0.4 0 0.4
Cathodic
POTENTIAL Eh
FIG 5 – Voltammograms for a galena electrode at pH 9.2; scan rate 10 mV s-1; ethyl
xanthate 9.5 × 10-3 mol dm-3 (solid line), 0 (dashed line) from Woods (1971).
via charge transfer to form a metal-sulfur bond and, in situations for which the reversible
potential of the formation of the bulk phase is known, at underpotentials, that is, at
potentials below the reversible value.
The electrochemical quartz crystal microbalance (EQCM) provides a means of studying
mass changes at the nanogram level and hence can detect chemisorption. Woods and Jeffrey
(2007) applied this technique to elucidate the interaction of the collector, ethyl xanthate,
with silver surfaces. The induced oscillations of the quartz crystal on which the electrode
material under investigation is deposited follow changes in the mass of the electrode and
hence measure adsorption. The oscillations follow changes in mass of the electrode and
hence measure adsorption. Slippage at the interface occurs when the surface is hydrophobic
but not when it is hydrophilic and hence the EQCM allows changes in wettability of the
surface to be determined.
The atomic force microscope (AFM) provides a means of probing surfaces at the atomic
level and hence can monitor adsorption. In a recent review of the application of AFM to
flotation, Xing et al (2018) pointed out that, in this technique, a sharp probe mounted near to
the end of a cantilever scans over the sample surface and provides a high-resolution three-
dimensional topographic image. In addition, the AFM tip can be used as a force sensor to
detect local properties such as adhesion, stiffness, charge etc.
Electrochemical impedance spectroscopy has also been applied to the study of flotation
collector adsorption on mineral surfaces (Ertekin, Pekmez and Ekmekci, 2016). Those authors
report that microflotation, induction time measurement and electrochemical methods were
used to measure adsorption of flotation reagents and to derive a proxy to correlate with
flotation behaviours of sulfide minerals.
FIG 6 – Flotation recovery of chalcocite with ethyl xanthate as a function of potential: –– data from Richardson
et al (1984); –– curves from Heyes and Trahar (1979); dashed curve from Basilio, Pritzker and Yoon (1985).
Dotted curve is a cyclic voltammogram at 5 mV s-1 for a chalcocite bed electrode (O’Dell et al, 1984).
It can be seen from Figure 6 that flotation does not take place below ~ -0.2 V and that
efficient flotation occurs at ~0 V. It can also be seen that there is good agreement between the
three groups as to the potential of onset of flotation even though the experimental conditions
employed differed. The dotted curve is a voltammogram obtained with a chalcocite particle
bed electrode in the flotation cell used by Richardson et al (1984) to record their flotation
recovery data. The characteristics of the voltammogram are the same as that found with
electrodes made from single chalcocite specimens. The anodic peak at ~ -0.15 V arises from
the chemisorption of a monolayer of ethyl xanthate on the mineral surface, while the anodic
current above 0 V is due to interaction with the mineral to form copper(I) xanthate. It can
be seen that the onset of flotation occurs in the potential region of chemisorption, and that
efficient flotation still occurs when copper(I) xanthate is also formed.
As illustrated for galena in Figure 5 and chalcocite in Figure 6, chemisorbed xanthate occurs
before metal xanthate formation; that is, the chemisorbed collector is the thermodynamically
more favourable species. Furthermore, it is sufficient to induce flotation.
Figure 7 shows the flotation recovery/potential curves for chalcocite, bornite, chalcopyrite
and pyrite in the presence of ethyl xanthate reported by Richardson and Walker (1985). It can
be seen that, in each case, the mineral changed from a non-float to a float condition at a
characteristic potential. The potential of onset of flotation of pyrite is similar to that found by
Gardner and Woods (1973) with their particulate bed system.
Guy and Trahar (1984) applied their modified Denver flotation cell to the galena/ethyl
xanthate system. Flotation of galena from mixtures with quartz was found to be a function of
potential, but the potential dependence varied with the mode of pretreatment of the mineral
(Figure 8). The left-hand diagram in Figure 8 shows the potential dependence after grinding
the mineral in a ceramic or a stainless-steel mill and flotation is observed from the most
negative potential induced. When a mild-steel mill was used (middle diagram), negligible
flotation was observed below about 0 V. The difference between these two relationships
could be explained in terms of the redox conditions in the mill. In the ceramic and stainless-
100
40
20
0
-0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4
POTENTIAL (V vs SHE)
FIG 7 – Dependence of flotation recovery on conditioning potential; conditioning time
ten minutes; flotation time two minutes; pH 9.2; xanthate 1.44 × 10-5 mol dm-3 for Cu S and
(2 ± 0.1) × 10-5 mol dm-3 for the other minerals (from Richardson and Walker, 1985).
100 100 100
GALENA GALENA
RECOVERY PER CENT IN FIRST MINUTE
90 90 90
pH 8 pH 8
80 80 KEX 80 PPG 400
STEEL MILL KEX
70 70 ALCOHOL FROTHERS 70
STAINLESS STEEL MILL
GALENA
NO 2 S IN MILL
60 60 60
pH 8
PPG 400
50 KEX 50 50
40 40 40
CERAMIC MILL
STAINLESS STEEL MILL
30 30 30
SAMPLE 1
20 SAMPLE 2 20 20
10 10 10
0 0 0
-500 -400 -300 -200 -100 0 100 200 300 400 500 -500 -400 -300 -200 -100 0 100 200 300 400 500 -500 -400 -300 -200 -100 0 100 200 300 400 500
steel mills, the mineral is free to oxidise by reaction with atmospheric oxygen, but the mild-
steel mill will provide cathodic protection and keep the mineral at a low potential. A similar
situation can be achieved by grinding in the presence of sodium sulfide. The right-hand
graph in Figure 8 shows the results obtained with a stainless-steel mill in the presence of
10-2 mol dm-3 sodium sulfide solution. It can be seen that the recovery/potential relationship
observed is similar to that for the mild-steel mill.
Figure 9 shows the correlation made by Guy and Trahar between their flotation recovery
data for galena in the presence of ethyl xanthate and the voltammetry of Woods (1971, 1972).
It can be seen that, for mineral ground under reducing conditions, the onset of flotation occurs
at the same potential as does the anodic peak due to chemisorbed xanthate (cf Figure 5).
When galena is pre-oxidised, some sulfur is lost from the surface as thiosulfate or sulfate.
Thus, when the potential is subsequently taken to low values, the oxidised surface species
100 100
90 RECOVERY 90
60 20
CURRENT
50 0
PREPARED IN
40 REDUCING -20
CONDITIONS
RECOVERY
30 PREPARED IN -40
REDUCING
20 CONDITIONS -60
10 -80
0 -100
-500 -300 -100 0 100 300 500
EPt (mV)
FIG 9 – Anodic current and flotation response of galena in the presence of ethyl xanthate
following preparation under reducing or oxidising conditions from Guy and Trahar (1985).
are reduced to lead metal on the surface. On the subsequent potential scan, an anodic peak is
observed that occurs at potentials expected for the reaction of lead with xanthate to form lead
xanthate. In the work of Guy and Trahar for mineral prepared under oxidising conditions, the
flotation correlates with the current peak and, in this case, lead xanthate would appear to be
the species inducing flotation. The flotation recovery increases significantly, however, when
the potential is in the region of chemisorption of xanthate on galena.
The potential dependence of the flotation of galena has also been determined in a pilot plant
operating with Mt Isa Mines lead/zinc ore (Johnson, Jowett and Heyes, 1982) using hydrazine
to adjust the potential and the results are shown in Figure 10.
Comparison of Figures 8–10 shows that the potential of onset of flotation in the pilot plant
is the same as that in the laboratory investigations in which the mineral was ground under
reducing conditions. Thus, such a pretreatment would appear to be the one most relevant to
flotation practice. It is interesting to note that flotation of galena was found to be initiated at
0.1 V in the particle bed electrode cell used by Gardner and Woods (1973).
Grano, Ralston and Smart (1990) also carried out studies on the potential dependence of the
flotation of galena from Mt Isa lead-zinc ore (Figure 11). It can be seen from the figure that
the recovery/potential relationship they obtained is similar to that in Figures 8–10. It can be
seen in Figures 8 and 9 that there is an upper as well as a lower potential limit of flotation
of galena and that it is independent of the pretreatment conditions. A similar upper limit of
flotation was observed for the flotation of chalcocite with ethyl xanthate (Heyes and Trahar,
1979; Basilio, Pritzker and Yoon, 1985). The results of Heyes and Trahar (1979) are shown in
Figure 12; note that this figure covers a wider potential range than that covered in Figure 6.
It can be seen from Figure 12 that the potential at which the upper flotation edge occurs varies
with change in pH. For both galena and chalcocite, the upper limit is close to the potential at
which the metal xanthate is known to oxidise to the metal oxide and diethyl dixanthogen, and
displays the expected pH dependence. Since dixanthogen is more hydrophobic than metal
xanthates, oxidation to this compound and a metal oxide would not, alone, be expected to
50
30
20
10
0
-500 -400 -300 -200 -100 0 100 200 300 400 500
EPt (mV)
100
EX-0.0 kg/t
80 EX-0.10 kg/t
EX-0.20 kg/t
RECOVERY %
60
40
20
Eh (mV)
FIG 11 – Potential dependence of galena flotation from Mt Isa lead/zinc
ore with ethyl xanthate at pH 8.1 from Grano, Ralston and Smart (1990).
inhibit flotation. Oxidation of the mineral itself also occurs in the relevant potential range and
it is likely that, once the mineral surface is no longer passivated by a metal xanthate, oxidation
occurs to form significant quantities of metal oxide and this renders the surface hydrophilic.
Roos, Celis and Sudarsono (1990a) determined the potential dependence of the xanthate-
induced flotation of chalcopyrite and (1990b) of chalcocite and covellite. Onset of the flotation
100
80
70
60
CHALCOCITE
50
KEX 7.6 mg/L
PPG 400
40
pH 11
pH 8
30
20
10
0
-600 -400 -300 -200 -100 0 100 200 300 400 500 600
FIG 12 – Flotation recovery of chalcocite with ethyl xanthate as a function of potential at pH 8 and 11 from Heyes and Trahar (1979).
of chalcopyrite and of chalcocite was found to occur at the potentials previously reported
by Richardson and Walker (1985). The threshold flotation potential for covellite was about
50 mV more positive than that for chalcocite under the same conditions. An anodic shift is to
be expected since the activity of copper in covellite is less than that in chalcocite.
Researchers at the Central South University, Changsha, China determined the potential at
which optimum flotation recovery of sulfide minerals occurs (Hu, Sun and Wang, 2009). They
then established ore grinding conditions and chemical compositions in the flotation pulp that
give rise to the derived optimum potential. This approach, which is referred to as ‘Original
Potential Control Flotation (OPCF)’, has been applied in a number of flotation plants and it
is reported (Hu, Sun and Wang, 2009) that improvements in recovery of target minerals is
significantly greater than with conventional procedures.
A B 100 1.0
100 1.0
Flotational recovery / (%)
60 0.6 60 0.6
40 0.4 40 0.4
20 0.2 20 0.2
0 0 0.0
-0.3 -0.2 -0.1 0 0.1 0.2 -0.40 -0.30 -0.20 -0.10 0.00
Potential / V(vs SHE) Potential / V(vs SHE)
C
Flotational recovery / (%)
100 1.0
FIG 13 – Comparison of potential dependence of flotation recovery and collector coverage (A) galena/
ethyl xanthate, (B) chalcocite/ethyl xanthate, and (C) chalcocite diethyldithiophosphate.
75 100
40
25
20
0 0
and Trahar, 1989). It was pointed out by Woods (1991) that the potentials reported by Healy
and Trahar for 50 per cent flotation recovery of these minerals in one minute correlated with
the equilibrium potentials for the oxidation of each mineral to form sulfur. This suggests that
the important species in the self-induced flotation of sulfides is a sulfur species similar to that
formed on chalcopyrite.
Attempts were made by Control International to exploit self-induced flotation commercially
in Arizona copper mines in the late 1980s (Arbiter and Gebhardt, 1992). Flotation recoveries
in the complete absence of collector failed, however, to match those achieved in conventional
practice and process development was terminated. A more optimistic application of
collectorless flotation of chalcopyrite was recently reported by Heyes (2007). A strategy to
separate chalcopyrite from chalcocite and bornite was developed based on work at CSIRO in
the late 1970s (Heyes and Trahar, 1977, 1979). Flotation in the absence of collectors produced
a concentrate that was predominantly chalcopyrite and the other copper sulfide minerals
were then floated following collector addition. This procedure was devised to maintain the
provision of a copper concentrate with a copper-sulfur ratio suitable for direct to blister
smelting technology as the chalcopyrite content of the ore is increased as a result of the mine
being converted from underground to open cut operation.
It should be pointed out that self-induced flotation will occur whenever the potential of the
mineral of interest is in the region of surface oxidation. Thus, self-induced flotation acts as a
complication in collector-induced flotation.
Flotation of sulfide minerals in the absence of conventional collectors also occurs in the
presence of sulfur(-II) species in solution (Hu, Sun and Wang, 2009).
(b2)
(c)
Anodic
CURRENT
C
A
Potential
Cathodic
(b1)
FIG 15 – Schematic diagram showing the effect of various factors on the mixed potential of collector/oxygen systems.
The solid lines represent collector oxidation (anodic) and oxygen reduction (cathodic)
occurring at a mixed potential A. This system can be modified in the following ways:
•• The mixed potential can be shifted to B if the oxygen current can be decreased to give
curve b1, for example by adding a reagent such as sulfite that diminishes the oxygen
content. In this case the collector oxidation current becomes zero, no oxidised collector
species are formed, and flotation is depressed.
•• If the collector oxidation current is increased, that is, the curve becomes b2, the new mixed
potential B is characterised by greater currents and flotation will be enhanced. This can
be achieved by increasing the collector concentration or substituting a homologue of the
collector with a longer alkyl chain length.
•• If collector oxidation is inhibited such that the anodic wave becomes curve c, flotation
will be depressed. In this case, the mixed potential will be moved to point C, at which
value negligible collector is oxidised. The effect of cyanide in depressing pyrite, in which a
surface iron cyanide species is formed, is an example of this mechanism.
•• Flotation can be depressed by the addition of a species that oxidises at a lower potential
than that for the collector. Thus, if the additional reaction is represented by curve b2, the
mixed potential B, at which the rate of reaction is equal and opposite to oxygen reduction,
corresponds to zero collector oxidation and hence no flotation. The use of NaSH to depress
chalcopyrite in copper/moly plants follows this mechanism.
•• Flotation can also be depressed if the solution conditions are changed such that the mineral
oxidises in a potential region below that for the collector. That is, mineral oxidation
becomes represented by curve b2 and the mixed potential, again, involves negligible
collector oxidation. Change in pH can bring about this situation. Another example is the
action of cyanide on copper minerals in which situation oxidation of the mineral occurs
at lower potentials as a result of the formation of copper cyanide complexes in solution.
Each of the mechanisms discussed above have been confirmed by electrochemical
investigations. For example Castro and Larrondo (1981) validated the mechanism of the
depressant action of cyanide for chalcocite in the presence of propyl xanthate. These authors
correlated measurements of flotation recovery with the potential of a chalcocite electrode
in the solution. They found that depression of the mineral was accompanied by a decrease
in potential and the appearance of copper in solution. The dissolution of chalcocite to form
a copper cyanide complex ion shifted the potential below the region in which xanthate
reactions occur.
100
90
80
70 Enargite
Chalcopyrite
Recov ery, %
60 Pyrite
Chalcocite
50
40
30
20
10
0
280 300 320 340 360 380 400 420
Potential, m V (SHE)
FIG 16 – The effect of oxidation pre-treatment with hydrogen peroxide at pH 9.8 on the flotation recovery of energite,
chalcopyrite and pyrite with isopropyl xanthate at pH 6.8 (from Castro, Balttierra and Hernández, 2003).
100
90
80
70
Recovery (%)
60
50
40
30
20 Chalcopyrite
10 Enargite
0
-0.6 -0.4 -0.2 0 0.2 0.4 0.6
Potential (V vs SCE)
FIG 17 – Effect of potential on the flotation recovery of enargite and chalcopyrite at pH 10 in the
presence of 7 × 10-5 mol dm-3 potassium amyl xanthate (from Guo and Yen, 2005).
be floated efficiently at a potential higher than +0.2 V versus SCE (0.444 versus SHE), while
chalcopyrite was completely depressed at potentials higher than this value. Selective flotation
revealed that enargite can be successfully removed from chalcopyrite through controlling
the pulp potential in the range +0.2 to +0.55 V versus SCE, with the most effective separation
being achieved at 0.5 V (0.744 versus SHE). These authors noted that the flotation response of
enargite was greater in mineral mixtures with chalcopyrite than when floated alone. The work
of Guo and Yen illustrates that, although single mineral studies may not give precise data for
mixtures, they can be used to identify conditions for selective flotation in polymetallic ores.
Kaye (2006) reported that Leanne Smith (née Shannon) and co-workers of CSIRO, building
on the many years of experience in investigating the potential dependence of flotation at their
Reductant
Concentrate Tailing
(chalcolite +
cuprite) Air
bleed
Tailing Concentrate
(chalcopyrite) (enargite +
chalcocite +
cuprite)
PRODUCT 1
Air bleed
Oxidant
Tailing Concentrate
(chalcocite + (enargite)
cuprite
PRODUCT 3 PRODUCT 2
copper minerals at particular pulp potential windows. They showed that, for each ore
investigated, the identified procedure can produce a saleable low-arsenic copper concentrate,
which can be directly smelted without penalty, and a high-arsenic copper concentrate, which
will require further processing.
Smith, Davey and Bruckard (2012) reported that the pulp potential at which separation of
tennantite from chalcopyrite and bornite was possible with one of the ores tested was lower
than that expected from single mineral data, assuming that bornite behaves in a similar manner
to chalcopyrite The threshold potential for the flotation had been shifted about 100 mV SHE
in a more reducing direction. These authors considered a number of reasons for this shift in
the pulp potential, including that there was a difference in the potential dependence of the
flotation of bornite and chalcopyrite. Indeed, it can be seen from the findings of Richardson
and Walker (1985), shown in Figure 7, that the threshold potential for bornite is, indeed,
100 mV below that for chalcopyrite.
In a study of the treatment of copper/nickel ore from Sudbury, He et al (2008) found that
pentlandite recovery, and its selectivity against pyrrhotite, is considerably improved by
controlling pulp potential. In later work, Manouchehri, Lawson and Taylor (2010) applied the
Outotec Chena analyser (vide infra) to monitor pulp potential in the flotation of nickel ores
and were able to identify optimum conditions.
Lotter and Bradshaw (2010) reviewed the usage of mixtures of collectors in sulfide flotation
– both in the application and in research – in order to gain insights and prepare guidelines
for the development of a methodology for use in a predictive capacity. McFadzean, Mhlanga
and O'Connor (2013) reported that most collector mixtures achieved greater recoveries than
would be expected based on their additive performance alone. This synergistic behaviour
was most notable for mixtures of strong and weak collectors. Antagonistic behaviour was
noted for mixtures of two strong collectors.
Buckley et al (2017) utilised XPS to identify the species adsorbed on Cu sulfide ore minerals
from alkaline or near neutral solutions of different blends of the flotation collectors dibutyl
dithiophosphate (BDTP) and 2-mercaptobenzothiazole (MBT). Depending on the collector
concentration and pH, direct (albeit ex situ) surface chemical characterisation established
the adsorption of different species from a BDTP-MBT blend compared with sequential
conditioning in the individual collectors, and hence potentially different hydrophobicities.
The identity and coverage of each collector species depended on the relative concentration of
the two collectors in the blends investigated.
Thus, pH needs to be recorded as well as the potential of the redox electrode for proper
control of NaSH additions. In practice, redox potentials in flotation pulps are recorded with an
oxidation/reduction cycle (ORC) electrode, which contains a platinum redox electrode and an
internal reference electrode, which is usually silver/silver chloride. Commercial electrodes often
combine Eh and pH measurement in the same probe structure. If the redox potential is required
on the standard hydrogen scale, the measured value needs adjusting by adding 205 mV at 25°C
(vide supra). The correction at other temperatures is given by 205 - 0.77 (T - 25°C) mV.
Aravena (1987) reported that NaSH consumption was reduced by 48 per cent with the
introduction of nitrogen at the Chuquicamata mine in Chile and a further saving of 21 per cent
was achieved when depressant additions were controlled by electrode measurements.
Meadows et al (2013) claimed that several of the larger copper projects recently built in the
Americas have included a molybdenum plant due to favourable market conditions, and that
these plants apply ORC electrodes to control depressant additions.
in operation at the Hitura nickel mine since 1984, at the Vammala nickel and Vihanti Cu/
Zn/Pb mines since 1988 and was installed at the Pyhasalmi Cu/Pb/Zn/FeS2 mine in the early
1990s. At Hitura, the replacement of pH control by potential control was reported to result in
an improvement in nickel recovery by up to two per cent depending on ore type (Ruonala,
Heimala and Jounela, 1997). At Vihanti, it was claimed that lime and collector consumption
has been decreased to one third of the values experienced prior to the installation of the OK-
PCF system. It is claimed that profitability has increased by ten to 20 per cent since potentials
have been monitored.
Outokumpu proposed the use of electrodes composed of synthetic or natural sulfide
minerals, usually of the composition being floated. For each valuable mineral, it is claimed
(Ruonala, Heimala and Jounela, 1997) that there may be one or more combinations of
variables for the most economic processing, and this requires integration of on-stream X-ray
fluorescence analysis, electrochemical potential measurements and collector concentrations.
Outokumpu insisted (Ruonala, Heimala and Jounela, 1997) that, because different sulfide
minerals can express different potentials in the same slurry, electrodes prepared from the
sulfide minerals present in the flotation pulp must be used to maximise separation and
recovery. The company marketed a plant audit procedure (Sotka et al, 2003; Oravainen
et al, 2005) that includes the measurement of potentials in the flotation plant with mineral
electrodes. The application of Outokumpu’s electrochemical knowledge would now appear
to be focused on its audit procedure rather than online control.
Outotec Oyj supplemented the Outokumpu flotation monitoring knowledge in 2007
through the acquisition of the Chena® (Chemistry Navigator) trademark from Liqum Oy
and then began to market the Chena® Electropotential Measurement System for monitoring
flotation circuits. This system was promoted by the company for use online to provide
flotation plant operators with potentials measured with a number of electrodes as well as
the pH and, together with assays from other on-stream analysers, suggested it could be used
for expert control of the circuit. Manouchehri, Lawson and Taylor (2010) used the Chena
system incorporating a range of metal and sulfide mineral electrodes to monitor changes in
pulp chemistry in a laboratory study of copper/nickel ore from Clarabelle, Sudbury, Canada.
These authors concluded that the Chena analyser ‘offers a possibility to control the flotation
characteristics of pentlandite and pyrrhotite through the understanding and optimising of the
pulp chemistry’. Manouchehri (2014) reported that, using the Outotec Chena analyser in the
treatment of pentlandie/pyrrhotite ores, the rest potential displayed by pentlandite electrodes
in low oxidising pulps is still high enough to allow the formation of dixanthogen and hence
promote nickel flotation. Controlling pulp potential at slightly lower values, pentlandite
recovery from pyrrhotite was still found to be satisfactory.
Etelapaa et al (2012) and Mashevsky et al (2012) proposed the application of a molybdenum
electrode to monitor Eh using the Outotec Chena system as they perceived problems with
the use of platinum. The role of an Eh electrode in monitoring flotation pulps is to act as
an inert probe that will adopt the mixed potential being expressed by the sulfide mineral
particles through multiple collisions. As pointed out above, Rand and Woods (1984)
demonstrated that platinum can operate in this way. In this regard, the problems Etelapaa
et al (2012) and Mashevsky et al (2012) perceived with the use of platinum for measuring
the relevant Eh value are not pertinent. Certainly, the replacement probe they adopted,
molybdenum, will not be inert; it will adopt the potential of the molybdenum/molybdenum
oxide couple at the pH of the pulp. Indeed, molybdenum electrodes have been shown to act
as pH probes (Nomura and Ujihira, 1987). The actual values, however, could be influenced
by other components of the pulp solution, but not the mineral content. Any correlations
found between flotation recovery and the potential displayed by a molybdenum electrode,
therefore, relate to pH rather than Eh.
CONCLUDING REMARKS
In this chapter, it has been shown that the electrochemical nature of the interaction of flotation
collectors with sulfide minerals is well established. Also, that combining electrochemical
methods with a range of spectroscopic techniques provides valuable detail on collector-
mineral interaction. Furthermore, monitoring of electrochemical parameters in flotation pulps
can identify optimum conditions that yield the highest concentrate grades and recoveries.
As pointed out above, the potential-dependence of the flotation of a sulfide mineral present
in a sulfide assemblage has been found to differ from that of the mineral on its own (see
above). It would be advantageous to have a better understanding of the mineral-mineral
interactions responsible for such differences.
REFERENCES
Adam, K, Natarajan, K A and Iwasaki, I, 1984. Grinding media wear and its effect on the flotation of
sulfide minerals, Int J Miner Process, 12:39–54.
Aravena, J J, 1987. Column flotation applications at Chuquicamata’s molybdenite flotation plant, in
Proceedings Copper ’87 (eds: A Mular, G Gonzalez and C Barahona), Universidad de Chile, Santiago,
vol 2, pp 155–169.
Arbiter, N and Gebhardt, J E, 1992. Requirements for industrial collectorless flotation of sulfide
minerals, in Proceedings International Symposium Electrochemistry in Mineral and Metal Processing III
(eds: R Woods and P E Richardson), pp 1–13 (The Electrochemical Society: Pennington).
Basilio, C, Pritzker, M D and Yoon, R-H, 1985. Thermodynamics, electrochemistry and flotation of
the chalcocite – potassium ethyl xanthate system, SME-AIME Annual Meeting, New York, preprint
no 85–86.
Biegler, T, Rand, D A J and Woods, R, 1977. Oxygen reduction on sulphide minerals, in Trends in
Electrochemistry (eds: J O’M Bockris, D A J Rand and B J Welch), pp 291–302 (Plenum: New York).
Buckley, A N, Hamilton, I C and Woods, R, 1985. Investigation of the surface oxidation of sulphide
minerals by linear potential sweep voltammetry and X-ray photoelectron spectroscopy, in Developments
in Mineral Processing (ed: K S E Forssberg), vol 6, pp 41–60 (Elsevier: Amsterdam).
Buckley, A N, Hope, G A, Parker, G K, Steyn, J and Woods, R, 2017. Mechanism of mixed dithiophosphate
and mercaptobenzothiazole collectors for Cu sulfide ore minerals, Minerals Engineering, 109:80–97.
Buckley, A N, Hope, G A and Woods, R, 2003. Metals from sulfide minerals: the role of adsorption
of organic reagents in processing technologies, in Solid-Liquid Interfaces Macroscopic Phenomena –
Microscopic Understanding (eds: K Wandelt and S Thurgate), Topics in Applied Physics 85, pp 61–96
(Springer-Verlag: Berlin).
Buckley, A N and Woods, R, 1984. An X-ray photoelectron spectroscopic study of the oxidation of
chalcopyrite, Aust J Chem, 37:2403–2413.
Buckley, A N and Woods, R, 1991. Adsorption of ethyl xanthate on freshly exposed galena surfaces,
Colloids Surf, 53:33–45.
Buckley, A N and Woods, R, 1993. Underpotential deposition of dithiophosphate on chalcocite,
J Electroanal Chem, 357:387–405.
Buswell, A M, Bradshaw, D J, Harris, P J and Ekmekci, Z, 2002. The use of electrochemical measurements
in the flotation of a platinum group minerals (PGM) bearing ore, Minerals Engineering, 15:395–402.
Castro, S H, Balttierra, L and Hernández, C, 2003. Redox conditions in the selective flotation of enargite,
in Proceedings International Symposium on Mineral and Metal Processing VI (eds: F M Doyle, G H Kelsall
and R Woods), p 27 (The Electrochemical Society: Pennington).
Castro, S H and Larrondo, J, 1981. An electrochemical study of depression of flotation of chalcocite by
cyanide and iron-cyanide complexes, J Electroanal Chem, 118:317–326.
Chadwick, J, 2011. Great Mines – Zambia, Kanshansi, International Mining, August, pp 8–10, 12, 14, 16,
18–20.
Chander, S and Fuerstenau, D W, 1983. Electrochemical flotation separation of chalcocite from
molybdenite, Int J Miner Process, 10:89–94.
Cook, M A and Nixon, J C, 1950. Theory of water-repellent films on solids formed by adsorption from
aqueous solutions of heteropolar compounds, J Phys Colloid Chem, 54:445–712.
Corin, K C, Kalichini, M, O’Connor, C T and Simukanga, S, 2017. The recovery of oxide copper
minerals, Minerals Engineering, 102:15–17.
Dunne, R, Levier, M, Acar, S and Kappes, R, 2009. Keynote address: Newmont’s contribution to gold
technology, Symposium Series S60, World Gold Conference 2009, pp 221–240 (The Southern African
Institute of Mining and Metallurgy: Johannesburg).
Ertekin, Z, Pekmez, K and Ekmekci, Z, 2016. Evaluation of collector adsorption by electrochemical
impedance spectroscopy, International Journal of Mineral Processing, 154:16–23.
Etelapaa, M, Petrov, A V, Romanenko, S A, Sufianov, F S and Balanova, A Z, 2012. New approach
to regulating the flotation process of selecting sulphide minerals from pyrite in a lime environment,
in Proceedings XXVI International Mineral Processing Congress, New Delhi, 24–28 September, Paper
No 1083, 8 p.
Fornasiero, D, Fullston, D, Li, C and Ralston, J, 2001. Separation of enargite and tennantite from non-
arsenic copper sulfide minerals by selective oxidation and dissolution, Int J Miner Process, 61:109–119.
Gardner, J R and Woods, R, 1973. The use of a particulate bed electrode for the electrochemical
investigation of metal and sulphide flotation, Aust J Chem, 26:1635–1644.
Gardner, J R and Woods, R, 1977. An electrochemical investigation of contact angle and of flotation in
the presence of alkylxanthates. II. Galena and pyrite surfaces, Aust J Chem, 30:981–991.
Gardner, J R and Woods, R, 1979. An electrochemical investigation of the natural floatability of
chalcopyrite, Int J Miner Process, 6:1–16.
Gaudin, A, 1939. Principles of Mineral Dressing, 554 p (McGraw-Hill: New York).
Gebhardt, J E, Dewsnap, N F and Richardson, P E, 1985. Electrochemical conditioning of a mineral
particle bed electrode for flotation, Report Invest – US Bureau Mines, RI 8951, 14 p.
Gebhardt, J E and Richardson, P E, 1987. Differential flotation of a chalcocite-pyrite particle bed by
electrochemical control, Miner Metall Process, 4:140–145.
Goh, S W, Buckley, A N, Lamb, R N and Woods, R, 2006. The ability of static secondary ion mass
spectrometry to discriminate submonolayer from multilayer adsorption of thiol collectors, Minerals
Engineering, 19:571–581.
Grano, S, Ralston, J and Smart, R St C, 1990. Influence of electrochemical environment on the flotation
behaviour of Mt Isa copper and lead-zinc ore, Int J Miner Process, 30:69–97.
Grano, S R and Huang, G, 2006. Improving the flotation behaviour of a sulphide ore by controlling
electrochemical interactions during grinding, in ECS Transactions, 2(3):9-20, Electrochemistry in
Mineral and Metal Processing VII (eds: F M Doyle, G H Kelsall and R Woods) (The Electrochemical
Society: Pennington).
Guo, H and Yen, W-T, 2005. Selective flotation of enargite from chalcopyrite by electrochemical control,
Minerals Engineering, 18:605–612.
Guy, P J and Trahar, W J, 1984. The influence of grinding and flotation environments on the laboratory
batch flotation of galena, Int J Miner Process, 12:15–38.
Guy, P J and Trahar, W J, 1985. The effects of oxidation and mineral interaction on sulphide flotation, in
Flotation of Sulphide Minerals (ed: K S E Forssberg), pp 91-110 (Elsevier: The Netherlands).
He, S, Grano, S, Manouchehri, H R, Lawson, V and Taylor, A, 2008. Critical influence of pulp oxygen
content on the separation of pentlandite from pyrrhotite in two process streams of the Clarabelle Mill
in Vale Inco, Sudbury, Canada, in Proceedings XXIV International Mineral Processing Congress, Beijing
(eds: W D Zhou, S C Yao, W F Liang, Z L Cheng and H Long), pp 1028–1038.
Healy, T W and Trahar, W J, 1989. Challenges in sulfide flotation fundamental, in Challenges in Mineral
Processing (eds: K V S Sastry and M C Fuerstenau), pp 3–14 (Society for Mining, Metallurgy and
Exploration / The American Institute of Mining, Metallurgical, and Petroleum Engineers: Littleton).
Hecker, C, Ramirez, J, Beas, E and Cartes, F, 1999. Development of an on-line Eh-pH electrochemical
sensor for the flotation process control, in Proceedings International Conference Copper ’99 – Cobre ’99
(eds: B Hancock and M Pon), 2:83–94 (The Minerals, Metals and Materials Society: Warrendale).
Heimala, S and Jounela, S, 2000. Effective flotation of oxidised Cu, Zn, Ni and Co-minerals, paper
presented at Obogashchenije – 2000, St Petersburg, June.
Herrera-Urbina, R, Sotillo, F J and Fuerstenau, D W, 1999. Effect of sodium sulfide additions on the
pulp potential and amyl xanthate flotation of cerussite and galena, Int J Miner Process, 55:157–170.
Heyes, G W, 2007. The differential separation of copper sulphides by flotation, presented to Flotation
07, Cape Town, November.
Heyes, G W and Trahar, W J, 1977. The natural flotatability of chalcopyrite, Int J Miner Process, 4:317–
344.
Heyes, G W and Trahar, W J, 1979. Oxidation-reduction effects in the flotation of chalcocite and cuprite,
Int J Miner Process, 6:229–252.
Hintikka, V V and Leppinen, J O, 1995. Potential control in the flotation of sulphide minerals and
precious metals, Minerals Engineering, 8:1151–1158.
Hope, G A, Woods, R, Parker, G K, Watling, K M and Buckley, F M, 2006. Spectroelectrochemical
investigations of flotation reagent–surface interaction, Minerals Engineering, 19:561–570.
Hu, Y, Sun, W and Wang, D, 2009. Electrochemistry of flotation of sulphide minerals, Tsinghua
University Press and Springer-Verlag, Berlin Heidelberg, 304 p.
Johnson, N W, 1988. Application of electrochemical concepts to four sulfide flotation separations, in
Proceedings International Symposium Electrochemistry in Mineral and Metal Processing II (eds: R Woods
and P E Richardson), pp 131-149 (The Electrochemical Society: Pennington).
Johnson, N W, Jowett, A and Heyes, G W, 1982. Oxidation-reduction effects in galena flotation:
observations on Pb-Zn-Fe sulphides separation, Transactions of the Institution of Mining and Metallurgy,
Mineral Processing and Extractive Metallurgy, 91:C32–C37.
Jones, M H, 1991. Some recent developments in the measurement and control of xanthate, perxanthate,
sulphide, and redox potential in flotation, Int J Miner Process, 33:193–205.
Jones, M H and Woodcock, J T, 1978. Evaluation of ion-selective electrode for control of sodium sulfide
additions during laboratory flotation of oxidized ores, Transactions of the Institution of Mining and
Metallurgy, Mineral Processing and Extractive Metallurgy, 87:C99–C105.
Kaye, T, 2006. Separation process boosts copper production, Process, CSIRO research in mineral
processing and metal production, February, p 8.
Leppinen, J O, Hintikka, V V and Kalapudas, R P, 1998. Effect of electrochemical control on selective
flotation of copper and zinc from complex ores, Minerals Engineering, 11:39–51.
Long, G, Peng, Y and Bradshaw, D, 2012. A review of copper-arsenic mineral removal from copper
concentrates, Minerals Engineering, 36–38:179–186.
Lotter, N O and Bradshaw, D J, 2010. The formulation and use of mixed collectors in sulphide flotation,
Minerals Engineering, 203:945–951.
Luttrell, G H and Yoon, R-H, 1984. Surface studies of the collectorless flotation of chalcopyrite, Colloids
Surf, A, 12:239–254.
Majima, H, 1969. How oxidation affects selective flotation of complex sulphide ores, Can Metall Quart,
8:269–273.
Manouchehri, H, Lawson, V and Taylor, A, 2010. Electrochemistry measurements using Chena
Analyser, to investigate selective separation of pentlandite and pyrrhotite, in Proceedings XXV
International Mineral Processing Congress (IMPC), pp 2153–2164 (The Australasian Institute of Mining
and Metallurgy: Melbourne).
Manouchehri, H R, 2014. Pyrrhotite flotation and its selectivity against pentlandite in the beneficiation of
nickeliferous ores: an electrochemistry perspective, Minerals and Metallurgical Processing, 31(2):115–125.
Mashevsky, G N, Petrov, A V, Romanenko, S A, Sufyanov, F S and Balmanova, A Zh, 2012. New
approach to regulating the flotation process of selecting sulphide minerals from pyrite in a lime
environment, Oboggaschenie rud, 1:12–16.
McFadzean, B, Mhlanga, S S and O’Connor, C T, 2013. The effect of thiol collector mixtures on the
flotation of pyrite and galena, Minerals Engineering, 50–51:121–129.
Meadows, D, Jensen, D, Traczyk, F, Yu, S and Riffo, L, 2013. Molybdenum flotation practice – cell type
selection and design considerations, SME Annual Meeting, 24–27 February, Denver, Preprint 13–157, 8 p.
Mu, Y, Peng, Y and Lauten, R A, 2018. The galvanic interaction between chalcopyrite and pyrite in
the presence of lignosulfonate-based biopolymers and its effects on flotation performance, Minerals
Engineering, 122:91–98.
Nixon, J C, 1957. Discussion, in Proceedings Second International Congress of Surface Activity, vol 3, p 369
(Butterworth: London).
Nomura, K and Ujihira, Y, 1987. Magnetite coated iron wire and molybdenum wire as pH sensor,
Analytical Sciences, 3(2):125–129.
O’Dell, C S, Dooley, R K, Walker, G W and Richardson, P E, 1984. Chemical and electrochemical
reactions in the chalcocite–xanthate system, in Proceedings International Symposium on Mineral and
Metal Processing (eds: P E Richardson, S Srinivasan and R Woods), pp 81–95 (The Electrochemical
Society: Pennington).
Oravainen, H, Heimala, S, Lamberg, P, Liipo, J, Lyyra, M and Sotka, P, 2005. The plant process audit,
mineral electrodes and metals recovery, in Proceedings Centenary of Flotation Symposium, pp 523–588
(The Australasian Institute of Mining and Metallurgy: Melbourne).
Panayatov, V, Kovachev, K, Önal, G, Dovan, J, Diner, H, Bulut, G, Panayotov, M and Ninova, V,
2000. Technology for copper-zinc flotation by electrochemical treatment, in Developments in Mineral
Processing 13, Proceedings XXI International Mineral Processing Congress (ed: P Massacci), vol B,
pp B8b103–108 (Elsevier: Amsterdam).
Payant, R, Rosenblum, F, Nesset, J E and Finch, J A, 2012. The self-heating of sulfides: galvanic effects,
Minerals Engineering, 26:57–63.
Peng, Y and Grano S, 2010. Effect of grinding media on the activation of pyrite flotation, Minerals
Engineering, 23:600–605.
Rand, D A J and Woods, R, 1984. Eh measurements in sulphide mineral slurries, Int J Miner Process,
13:29–42.
Richardson, P E, Stout, J V III, Proctor, C L and Walker, G W, 1984. Electrochemical flotation of sulfides:
chalcocite–ethyl xanthate interactions, Int J Miner Process, 12:73–93.
Richardson, P E and Walker, G W, 1985. The flotation of chalcocite, bornite, chalcopyrite, and pyrite in
an electrochemical-flotation cell, in Proceedings XV International Mineral Processing Congress, Cannes,
vol II, pp 198–210.
Roos, J R, Celis, J P and Sudarsono, 1990a. Electrochemical control of metallic copper and chalcopyrite-
xanthate flotation, Int J Miner Process, 28:231–245.
Roos, J R, Celis, J P and Sudarsono, 1990b. Electrochemical control of chalcocite and covellite-xanthate
flotation, Int J Miner Process, 29:17–30.
Ruonala, M, Heimala, S and Jounela, S, 1997. Different aspects of using electrochemical potential
measurements in mineral processing, Int J Miner Process, 51:97–110.
Senior, G D, Guy, P J and Bruckard, W J, 2006. The selective flotation of enargite from other copper
minerals – a single mineral study in relation to beneficiation of the Tampakan deposit in the Philppines,
Int J Miner Process, 81:15–26.
Shchukarev, A V, Kravets, I M, Buckley, A N and Woods, R, 1994. Submonolayer adsorption of alkyl
xanthates on galena, Int J Miner Process, 41:99–114.
Smith, L K and Bruckard, W J, 2007. The separation of arsenic from copper in the Northparkes copper-
gold ore using controlled-potential flotation, Int J Miner Process, 84:15–24.
Smith, L K, Davey, K J and Bruckard, W J, 2012. The use of pulp potential to separate copper and arsenic
– an overview based on selected case studies, in Proceedings XXVI International Mineral Processing
Congress, New Delhi, 24–28 September, Paper No 321, 12 p.
Sotka, P, Heimala, S, Lyyra, M and Oravainon, H, 2003. Process optimisation with Outokumpu’s plant
audit procedure, extended abstract, Flotation ’03 Conference, Helsinki, March.
Symmons, G L, Orlich, J N, Lenz, L C and Cole, J A, 1999. Implementation – start up of N2Tec flotation at
the Lone Tree mine, SME/AIME Annual Meeting, Denver, 1–4 March (Society for Mining, Metallurgy
and Exploration / The American Institute of Mining, Metallurgical, and Petroleum Engineers: Littleton).
Taggart, A F, Taylor, T C and Knoll, A F, 1930. Chemical reactions in flotation, Trans Amer Inst Metall
Engrs, 87:217–260.
Wark, I W and Cox, A B, 1934. Principles of flotation, I – an experimental study of the effect of xanthates
on contact angles at mineral surfaces, Trans Amer Inst Metall Engrs, 112:189–244.
Woodcock, J T and Hamilton, J K (eds), 1993. Australasian Mining and Metallurgy, The Sir Maurice Mawby
Memorial Volume, vol 1 (The Australasian Institute of Mining and Metallurgy: Melbourne).
Woods, R, 1971. The oxidation of ethyl xanthate on platinum, gold, copper, and galena electrodes:
relation to the mechanism of mineral flotation, J Phys Chem, 75:354–362.
Woods, R, 1972. The anodic oxidation of ethylxanthate on galena electrodes, Aust J Chem, 25:2329–2335.
Woods, R, 1991. Electrochemistry of sulfide minerals – Stokes medal address, Chemistry in Australia,
58:392–395.
Woods, R and Jeffrey, M I, 2007. Frequency changes observed with the EQCM resulting from
hydrophilic/hydrophobic transitions, Colloids and Surfaces A, 302:488–493.
Woods, R, Kim, D S and Yoon, R-H, 1993. The potential dependence of flotation of chalcocite with
diethyl dithiophosphate, Int J Miner Process, 39:101–106.
Woods, R, Young, C A and Yoon, R-H, 1990. Ethyl xanthate chemisorption isotherms and Eh-pH
diagrams for the copper/water/xanthate and chalcocite/water/xanthate systems, Int J Miner Process,
30:17–33.
Xing, Y, Xu, M, Gui, X, Cao, Y, Babel, B, Rudolph, M, Weber, S, Kappl, M and Butt, H-J, 2018. The
application of atomic force microscopy in mineral flotation, Advances in Interface and Colloid Science,
256:373–392.
INTRODUCTION
Oxidation of the sulfide minerals proceeds throughout a flotation separation circuit, with a
measurable increase in the extent of oxidation from the commencement of roughing to the
end of flotation. This appendix provides an overview of the products from oxidation of the
sulfide minerals, both valuable and gangue, and shows how these products influence the
flotation process.
Equation A1 depicts the transfer of a metal cation from the surface of the sulfide mineral
and the retention of the cation in the aqueous phase. Normally, the conditions in a sulfide
flotation circuit are slightly or strongly alkaline. Under alkaline conditions, many metal
cations are no longer soluble because the solubility product for the hydroxide of the
metal cation (M(OH)2) is exceeded due to the increased hydroxide ion concentration in
the aqueous phase. Table A1 shows the pH value at which various metal hydroxides are
precipitated (Vogel, 1962). It can be seen that various metal ions, but not all, are precipitated
below or at pH 7 while others are precipitated in the slightly alkaline region.
TABLE A1
pH values for precipitation of various hydroxides (Vogel, 1962).
pH value Metal ion precipitated
3 Fe3+, Sn2+, Zr4+
4 Th4+
5 Al3+
6 Cu2+, Zn2+, Cr3+
7 Fe2+
8 Ni , Co2+, Cd2+
2+
Hence, under alkaline conditions, the initial oxidation of a sulfide mineral (MS) in a flotation
circuit can be described by Equation A2 (Witika and Dobias, 1995):
Equation A2 introduces into Equation A1 the formation of the metal hydroxide, which is
retained in the system. In fact, the metal hydroxide precipitates can adhere on the surfaces
of the particles in the flotation pulp with process implications. Both Equations A1 and A2
show elemental sulfur (So), which is hydrophobic, as a product of oxidation. Its presence is
one explanation of the observed limited hydrophobicity of sulfides during initial oxidation,
another being the existence of a metal-deficient sulfide on the surface.
From flotation experiments, the following ranking (Trahar, 1984) was produced for the
floatability of the sulfide minerals from initial oxidation in the absence of collector (most
floatable minerals in the group on the left):
Chalcopyrite, Galena, Pyrrhotite, Pentlandite > Bornite > Chalcocite > Pyrite > Sphalerite
Ongoing oxidation in the alkaline region increases the quantity of hydrophilic hydroxide
precipitates on the surfaces, as shown in Equation A2, overriding the limited hydrophobicity
from initial oxidation.
Further oxidation under alkaline conditions of the sulfide mineral (MS), depleted in metal
cations on its surface, may lead to the production of oxy-sulfur species in the aqueous phase
(Witika and Dobias, 1995) according to Equation A3:
For the oxidation reactions involving the minerals in Equations A1, A2 and A3 to continue,
the electrons produced in Equations A1, A2 and A3 must be consumed; reactions of the
type in Equation A4 provide this. Note that such reactions can occur on pyrite and a high
pyrite feed grade causes greater oxidation of the other sulfide minerals. Sulfide oxidation is
electrochemical in nature from the combination of anodic (Equations A1 to A3) and cathodic
(Equation A4) reactions.
An example is given of one anodic reaction for pyrite, the most widespread and abundant of
all sulfide minerals. This reaction (Equation A5) applies above pH 6.7 where Fe2+ is no longer
stable (Hiskey and Schlitt, 1982):
In Equation A5, the products from the iron and sulfur in the pyrite are insoluble iron
hydroxide and soluble sulfate, respectively, and these two products are of the type indicated
in the general Equations A2 and A3 for the metal iron and sulfur respectively. Importantly,
the generation of hydrogen ions in Equation A5 shows that the pH can be expected to move
in the acid direction as a result of the oxidation of pyrite and sulfide minerals in general.
Table A2 lists the rest potentials for six of the sulfide minerals for which a ranking was given
earlier for their floatability after initial oxidation, in the absence of collector. In general, the
minerals with low rest potentials, that is minerals more prone to oxidation showed high self-
induced floatability from greater formation of sulfur (So), as shown in Equations A1 and A2
TABLE A2
Rest potential values of sulfides and related behaviour for self-induced floatability (also known as collectorless or natural
floatability) and sodium-sulfide-induced floatability (modified from Hu, Sun and Wang, 2009 and Wills and Finch, 2016).
Mineral Rest potential Self-induced floatability Sodium-sulfide-induced
(volts, SHE) (no collector) floatability (no collector)
Pyrite (noblest) 0.66 Low High
Chalcopyrite 0.56
Sphalerite 0.46
Pentlandite 0.35
Pyrrhotite 0.31
Galena 0.28 High Low
SHE – standard hydrogen electrode as the reference electrode.
(Hu, Sun and Wang, 2009). In experiments, chalcopyrite displayed much higher self-induced
floatability than expected from its rest potential in Table A2.
Pyrite, the noblest mineral based on the rest potential values in Table A2, displayed low
self-induced floatability and, commensurately, it is observed to display high sodium sulfide
induced floatability (Hu, Sun and Wang, 2009), as indicated in Table A2, because it requires
assistance from the sodium sulfide by providing a potential source of sulfur (So).
From Table A2, sulfide mineral pairs with a large difference in rest potential are expected to
demonstrate a greater driving force for galvanic oxidation in flotation pulps, in comparison
with a pair with a small difference. The same approach is also taken in consideration of the
propensity for oxidation and dangerous self-heating to occur in sulfide mineral concentrates
under storage or during transport (Payant et al, 2012).
SURFACE POTENTIAL
0 pH
Mn (OH)2
Cu (OH)2
Zn (OH)2 (s)
Sulfide
–
Gangue
FIG A1 – Schematic representation (Healy, 1984) of the variation of surface potential of a typical sulfide mineral and a typical
non-sulfide gangue mineral and of colloidal hydroxide precipitates of Zn(OH)2, Cu(OH)2 and Mn(OH)2. The typical relationship for
a sulfide and a non-sulfide mineral is a positive surface potential in the acid region with the potential declining and becoming
negative, while still in the acid region, with the value becoming more negative with neutral and alkaline pH values.
+
SURFACE POTENTIAL
Sulfide + Zinc
0 pH
Gangue + Zinc
FIG A2 – Schematic representation (Healy, 1984) of the effect of adsorption of a colloidal hydroxide
precipitate of Zn(OH)2 on the surfaces of a typical sulfide mineral and a typical non-sulfide gangue mineral.
Both minerals now show the behaviour of Zn(OH)2 in the region in which Zn(OH)2 forms. The solid lines in
Figure A1 are replaced by broken lines in Figure A2 for ease of comparison of the two related graphs.
•• the surface properties of the two example minerals adopt the properties of the precipitated
colloidal Zn(OH)2 and therefore the surface properties of the two example minerals
become similar
•• depending on the pH value, the example minerals have regions of both positive surface
potential and negative surface potential.
Oxidation of an iron or steel grinding media in grinding or regrinding in a flotation circuit
creates the same difficulty; the oxidation products from the grinding media provide a
potentially large source of hydrophilic species that adhere to the surfaces of the valuable
minerals and greatly increase the difficulty in preparing a hydrophobic surface on the
valuable mineral. It is useful to consider the relative magnitudes of the rest potentials for the
sulfide minerals (Table A2) with mild steel grinding media (Wills and Finch, 2016), noting the
large difference between the values for mild steel and galena:
(Noble end) Pyrite > Chalcopyrite > Sphalerite > Pentlandite >
Pyrrhotite > Galena >> Mild Steel (Active end)
The large difference in the rest potentials for mild steel and for the various sulfide minerals
indicates that there is large driving force for galvanic oxidation of the mild steel in grinding
systems, producing large quantities of iron species as detrimental hydrophilic hydroxides/
oxy-hydroxides adsorbed by particles in the system, as discussed earlier. Grinding balls
with elevated levels of chromium have a higher rest potential than mild steel and reduce the
driving force for galvanic oxidation in grinding systems, in comparison with mild steel balls.
Increasing the chromium content in the grinding balls heightens this effect by driving its rest
potential further from the value for mild steel.
Another outcome of the oxidation step for the ore minerals is that the released metal ions
are capable of interacting with, and altering the surfaces of the other sulfide and gangue
minerals in the flotation system. Some examples are:
•• Metal ions from oxidation of a valuable mineral in a copper circuit or a lead circuit will
transfer to other sulfide minerals and change their surfaces, for example uptake of copper
ions and lead ions on sphalerite in a copper circuit or a lead circuit, noting that alteration of
the sphalerite surface is only required in a zinc circuit that follows copper or lead circuits.
Premature unintentional activation of the sphalerite will cause losses of sphalerite because
it reports to the wrong concentrate.
•• Metal ions from oxidation of a valuable mineral in a copper circuit will activate gangue
sulfide minerals, for example copper ions will activate pyrrhotite and if present as Cu+
ions under reducing conditions, will activate pyrite. Unintentional activation of the pyrite
will cause dilution of the concentrate containing the valuable mineral.
As described earlier, the oxidation of a sulfide ore creates a large amount of hydrophilic
hydroxide/oxy-hydroxide species on the surfaces of all minerals, including the valuable
mineral. Collector must be selectively adsorbed on the valuable sulfide to produce surfaces
that are sufficiently hydrophobic for recovery by the flotation process. Hydrophobic
surfaces on the valuable sulfide must be obtained despite the presence of the described
hydrophilic species from various sources also on the surfaces. This uptake of collector is also
electrochemical in nature where the oxidation of the collector is the anodic reaction. For the
flotation metallurgist, the consequence is that the pulp potential (see earlier in this appendix)
will be an important parameter and must exceed a critical value to allow collector uptake; this
pulp potential is dependent on the mineral and the type of collector.
For some ores with a high head grade of iron sulfide minerals, that is pyrite and pyrrhotite,
it may be necessary to aerate the pulp extensively before rougher flotation to raise the pulp
potential to the critical value, allowing collector uptake. Even for ores with a low head grade
of iron sulfides, the head grade of iron sulfides in some internal streams being reground in the
flotation circuit may be quite high and pulp aeration after regrinding may also be required
to achieve the necessary pulp potential value, allowing uptake of collector prior to flotation
of the target sulfide mineral. The quantity of air or oxygen that needs to be supplied is a
property of the ore that can be determined in oxygen demand tests (see Chapter 7); this needs
to be known, because this property of iron sulfide minerals varies from deposit to deposit.
ACKNOWLEDGEMENT
The important contribution of Suzanne Munro (Mineralurgy Pty Ltd/Mineralis) in preparation
of diagrams and of other aspects for the content in this appendix is acknowledged with
gratitude.
REFERENCES
Healy, T W, 1984. Pulp chemistry, surface chemistry, and flotation, in Principles of Mineral Flotation – The
Wark Symposium (eds: M H Jones and J T Woodcock), pp 43–56 (The Australasian Institute of Mining
and Metallurgy: Melbourne).
Hiskey, J B and Schlitt, W J, 1982. Aqueous oxidation of pyrite, in Proceedings SME–SPE International
Solution Mining Symposium: Interfacing Technologies in Solution Mining (eds: W J Schlitt and J B Hiskey),
pp 55–74 (American Institute of Mining, Metallurgical, and Petroleum Engineers: Englewood).
Hu, Y, Sun, W and Wang, D, 2009. Electrochemistry of Flotation of Sulphide Minerals (Tsinghua University
Press and Springer).
Payant, R, Rosenblum, F, Nesset, J E and Finch, J A, 2012. Galvanic interaction in self-heating of
sulphide mixtures, CIM Journal, 3(2):169–177.
Trahar, W J, 1984. The influence of pulp potential in sulphide flotation, in Principles of Mineral Flotation
– The Wark Symposium (eds: M H Jones and J T Woodcock), pp 117–135 (The Australasian Institute of
Mining and Metallurgy: Melbourne).
Vogel, A I, 1962. A Text-book of Quantitative Inorganic Analysis, third edition (Longmans).
Wills, B A and Finch, J A, 2016. Mineral Processing Technology, eighth edition (Elsevier/B-H).
Witika, L K and Dobias, B, 1995. Electrochemistry of sulfide minerals, in Flotation Science and Engineering
(ed: K A Matis), pp 157–177 (Marcel Dekker, Inc).
CHAPTER 9
Surface chemical characterisation
for flotation plant optimisation
Alan Buckley*
ABSTRACT
The principal surface chemical characterisation techniques of relevance to the optimisation
of flotation plants will be described in this chapter. The main emphasis is on conventional
X-ray photoelectron spectroscopy (XPS), the most widely applied spectroscopy for obtaining
ex situ surface chemical analyses. The complementary information that can be obtained by
synchrotron XPS and by X-ray absorption spectroscopy is also described. Other important
ex situ characterisation techniques covered include static secondary ion mass spectrometry
and to a lesser extent, vibrational spectroscopy. General considerations for the preparation of
specimens for the ex situ techniques are discussed, including the relative merits of investigating
single particles rather than pulp specimens. Off-stream, but in situ (wet), surface chemical
characterisation techniques are outlined.
INTRODUCTION
As discussed in preceding chapters, the floatability of a particle in a flotation cell depends
on its hydrophobicity, and hence on its surface chemical composition. The separation of a
valuable ore component from gangue, or the selective flotation of one valuable component
relative to others, normally requires one component to be significantly more hydrophobic
than the other constituents (or the converse for reverse flotation). Thus, it can be important
to know the surface chemical composition of not only the component to be concentrated,
but also the other components or gangue that should report to the tails. Of course, flotation
has been an industrial process for more than a century, and for much of this period, surface
chemical characterisation of the particles in flotation pulps has not been carried out or even
feasible. Advances in flotation technology have usually arisen from empirical developments
rather than as a result of surface chemical knowledge. However, when a flotation process fails
to deliver the required recovery or grade, or when there is some other need to understand
flotation behaviour, then surface characterisation becomes not an option but a necessity.
* This chapter was revised by Alan Buckley, who sadly passed away in 2017, prior to the publication of this edition.
Most flotation plants process base metal ores, and most base metal ores comprise sulfide
minerals that can sometimes be partially oxidised. Accordingly, while emphasis will be
placed on sulfide minerals in this chapter, oxide minerals will be covered sufficiently to
illustrate the main surface characterisation differences for the two mineral types. Quite apart
from ‘industrial minerals’, oxide minerals can be important sources of metals as most rare
earth and actinide elements occur in ores predominantly in the form of oxides, carbonates or
phosphates. There is also an emphasis in this chapter on the surface chemical characterisation
of mineral specimens prepared in the laboratory expressly to help optimise generic flotation
plant circuits rather than on pulp samples to identify and solve plant-specific problems. The
latter approach will be covered briefly here but addressed more fully in the next chapter
where some case studies are described in detail. Regardless of whether the surface chemical
composition to be determined is that of mineral particles from a flotation plant or of a single
piece of mineral conditioned in a laboratory, the underlying principles of any characterisation
technique and its data interpretation will be the same.
The surface chemical composition of relevance in flotation systems is usually that of mineral
particles either bearing organic adsorbates, especially collectors, or being free of organic
adsorbates but altered by interaction with oxygen, added reagents other than collectors
such as activators, or normal constituents of flotation pulps such as metal ions. In the first
category, adsorbed collector coverage can range from the submonolayer of chemisorbed thiol
that is just sufficient to impart hydrophobicity to an unoxidised sulfide mineral particle, to a
multilayer of metal thiolate and/or of dithiolate. In the second category, the altered surface
can vary from one involving only the outermost atomic layer of the mineral lattice, such as
the interaction of Cu ions with pyrite (eg Wang, Forssberg and Bolin, 1989), to one involving
several atomic layers, such as the interaction of Cu ions with sphalerite (eg Finkelstein, 1997).
Thus, in both categories, the altered surface layer to be chemically characterised can vary in
thickness from one atomic layer (~0.2 nm) to several or even many atomic layers (typically
1–5 nm), notwithstanding the fact that it is only the outermost atomic layer that will have a
significant influence on the hydrophobicity. Characterisation of the former sometimes calls
for extreme surface sensitivity at the limit of available technology, while examination of the
latter is manageable using conventional laboratory surface analysis techniques such as X-ray
photoelectron spectroscopy (XPS). The advanced approaches, which usually require access
to major infrastructure such as a synchrotron, have become well-established for flotation
research in recent years, and are now at the stage that they should be considered for flotation
problem-solving when appropriate.
This chapter is concerned with the surface chemical nature of mineral particles, but readers
should be reminded that surface physical aspects, including microstructure and particle
shape, can also influence floatability. The chemical and physical aspects are often inextricably
linked when the crystal structure is particularly anisotropic. For example, the basal planes of
molybdenite have been shown, as anticipated, to have lower reactivity than the edges (Buckley
et al, 1995), and an analogous reactivity difference is expected for covellite. In these cases, the
particle size will influence the ratio of edges to basal planes, and hence the composition of
surfaces predominantly encountered by the bubbles in a flotation cell.
small inclusions of other phases or weathering products (although that in itself may be of
relevance in the flotation of some minerals). On the other hand, the main difficulty involved
in using abraded surfaces is the additional complexity introduced by the need to interpret
the observed behaviour of altered surfaces. Indeed, in order to obtain meaningful data, there
is sometimes the need to fracture surfaces under the flotation-related aqueous phase (eg
collector solution).
are also used, especially when Auger electron peaks interfere with photoelectron peaks
ejected by Al X-rays. The electron energy analyser is usually of hemispherical design with
~150 mm mean radius preceded by an electron lens and with a solid-state detector typically
consisting of either a number (usually three to ten) of individual channel electron multipliers,
or a multichannel plate. The specimen is inserted into the spectrometer via a load lock and
subsequently transferred to the analysis chamber where it is mounted on a manipulator of
some kind. An electron spectrum is normally obtained by maintaining a fixed pass energy
(eg 20 eV) for the hemispherical analyser, and incrementing a potential to retard the kinetic
energy of the emitted electrons to this pass energy value.
An X-ray photoelectron spectrum is a plot of the intensity of detected electrons as a function
of their kinetic energy. In most cases, it is the binding energy of the electrons from each
core level that is of primary interest; consequently, it is traditional to show the spectrum
abscissa as binding energy increasing from right to left, rather than kinetic energy (the actual
independent variable) increasing from left to right. The counter-current relationship between
kinetic energy and binding energy exists because a photoelectron binding energy, measured
relative to the Fermi level, is simply the exciting photon energy (usually 1486.6 eV) minus
the photoelectron kinetic energy minus the spectrometer work function (typically ~4 eV).
A survey electron spectrum excited by 1486.6 eV photons from a chalcocite fracture surface
prepared in air is shown in Figure 1 as a typical example of a wide scan.
In the spectrum shown in Figure 1, photoelectron peaks from the Cu 2s, Cu 2p, Cu 3s,
Cu 3p, S 2s and S 2p core levels and from the valence band are observed. A peak from S 1s
electrons is not observed because their binding energy (~2470 eV) is greater than the photon
Cu 2p
Cu LMM
Cu 3p
S 2p
Cu 3s
C 1s
S 2s
O 1s
FIG 1 – Survey electron spectrum from a chalcocite fracture surface prepared in air.
energy. There are two well-resolved Cu 2p peaks, Cu 2p1/2 and Cu 2p3/2, because the spin-orbit
splitting (~20 eV) is much greater than the available resolution. The corresponding splitting
of the S 2p1/2,3/2 doublet is not obvious in the wide scan because it is only 1.19 eV, but it is
evident in high resolution narrow scans. By contrast, the Cu 3p splitting (2.47 eV) is of the
same order of magnitude as the experimental linewidth. The expected relative intensity of
the 2p doublet components is [2×(1/2)+1]:[2×(3/2)+1] = 1:2. The Cu 2p peaks are more intense
than the Cu 3p because the photoionisation cross-section at a photon energy of 1487 eV for
the former is significantly greater than for the latter, notwithstanding the fact that there are
six 2p and six 3p electrons in each Cu atom. Peaks from Cu LMM Auger electrons are also
observed, the most intense Cu L3M4,5M4,5 peak appearing at an ‘effective binding energy’ of
569 eV or, more correctly, at a true kinetic energy near 1487 - 569 = 918 eV. Auger electrons
are not ejected directly by photons; rather, they are emitted as part of one of the processes by
which core electron vacancies are filled, the other process being X-ray emission (fluorescence).
Thus, Auger electrons have a kinetic energy that is independent of the energy of the photons
that created the core electron vacancies by photoemission. Auger electron emission is more
probable than X-ray fluorescence for low atomic number elements. Cu L3M4,5M4,5 Auger
electrons are produced when a vacancy in the L3 (2p3/2) level is filled by an electron from the
M4,5 (3d3/2,5/2) levels and another electron from the M4,5 levels is ejected. It will be shown later
that these X-ray excited Auger electrons can be extremely useful for mineral systems; however,
it should be noted that ‘Auger electron spectroscopy’ (AES) usually denotes Auger electron
emission brought about by the creation of core level vacancies by means of an electron beam,
as distinct from the photon beam in XAES. Electron excited AES is practically useless for
surface chemical characterisation of minerals because of the concomitant chemical alteration
brought about by the primary electron beam, although it can be of use for surface elemental
analysis. Finally, in Figure 1 it can be seen that small peaks from C 1s and O 1s electrons are
discernible, the C and O most probably adsorbed from the atmosphere after the mineral had
been fractured. The ubiquitous nature of C and O at the surface of air-exposed surfaces can
prove to be a nuisance when attempts are made to detect collector adsorption by monitoring
photoelectron peaks from those elements. On the other hand, the adventitious C can be useful
as an indicator of any charging of the specimen surface layer. The C 1s binding energy of the
hydrocarbon peak from such contamination should be close to 285.0 eV relative to the Au 4f7/2
binding energy of 84.0 eV for Au metal used in calibration.
A wide scan electron spectrum, then, can reveal which elements are present at the surface,
and their relative concentrations within the depth analysed (typically 1–5 nm), provided
those surface concentrations are greater than the detection limit, which is of the order of one
atomic per cent. The elements present can be identified because the various core electron
binding energies and Auger electron kinetic energies are essentially unique for each element.
Hydrogen has no closed shell electrons, and consequently photoelectrons from H cannot
be identified as arising from that element. However, the great strength of XPS (and XAES)
arises from the fact that the core electron binding energies (and Auger kinetic energies) for a
particular element do vary to a small extent to reflect different chemical environments of that
element. In most cases, this ‘chemical shift’ is large enough to allow adequate discrimination
of different oxidation states, with the binding energy usually increasing with increasing
oxidation state or ‘positive charge’, but not so large that the origin of a photoelectron peak
becomes uncertain. Thus, a survey scan is usually no more than an overview of the chemical
composition of a surface; a detailed surface chemical characterisation necessitates the
determination of a series of narrow, high resolution scans from which the position of each
peak to ±0.05 eV, its intensity (area above background), and its shape (including linewidth)
can be obtained. In many cases, an electron spectrum will consist of a partially resolved
superposition of several components, and curve fitting will be required to extract the binding
energy, intensity and shape for each component.
The most important narrow scan for the surface characterisation of sulfide minerals is that
covering the S 2p region. The S 2p spectrum from a fracture surface of the semiconductor
chalcocite is shown in Figure 2 as an example. Although there are 12 S sites (and 24 Cu sites)
in the Cu2S unit cell, the S 2p spectrum is an uncomplicated doublet with no obvious line
broadening arising from the distribution of electronic environments in the lattice. The 2p3/2
binding energy of 161.55 eV is within the range observed for most monosulfide minerals,
which is from 160.6 eV for galena to 161.7 eV for millerite. By contrast, the S 2p spectrum from
a chalcocite surface abraded in air and subsequently washed with water (Figure 2) does not
appear to comprise one uncomplicated doublet. It is certainly different from the spectrum for
the fracture surface, most probably because of the superposition of a second doublet from
an altered, S-rich (Cu-deficient) environment at a binding energy near 162.3 eV arising from
loss of Cu during the abrasion and washing. Similar, atypically high S 2p3/2 binding energies
are observed for unaltered heazlewoodite, Ni3S2 (162.35 eV) and for the five-coordinate S
in pentlandite, (Ni,Fe,Co)9S8 (162.2 eV). The corresponding values for the disulfides such as
pyrite and vaesite (NiS2) are normally in the range 162.4 to 162.7 eV. In between those values
for the disulfides and that for elemental sulfur, S8 (163.6 eV), are the 2p3/2 binding energies for
the oligosulfides (Buckley et al, 1988) and severely metal-deficient sulfide lattice environments
(Buckley and Walker, 1988). The values for S-O species such as sulfite (~166.5 eV) and sulfate
(~169 eV) are significantly higher than those for S with a zero or negative oxidation state.
The S 2p spectrum from essentially unaltered covellite (Figure 3) is much more complicated
than that from chalcocite for at least two reasons. Firstly, there are two quite different
S environments in the unit cell of covellite (CuS but often represented as Cu2S.CuS2), a
monosulfide giving rise to a doublet with a 2p3/2 binding energy of 161.6 eV, and a disulfide
giving rise to a doublet of twice the intensity of the former at 162.4 eV. The monosulfide
2p1/2 component is buried underneath the disulfide 2p3/2 component. Secondly, covellite
has metallic conductivity; consequently, the lineshape of individual components will be
asymmetric, with a ‘tail’ on the high binding energy side arising from the high density of
S 2p
Abraded
Fractured
states at the Fermi level (Doniach and Sunjic, 1970). XPS data analysis software packages
usually have at least one asymmetric lineshape that can be used to help fit such spectra. To
further complicate an already complicated situation, there may also be broad contributions to
the spectrum at 2–3 eV higher, binding energy from each principal component arising from
energy loss processes. Fortunately, such energy loss contributions are usually of low intensity
relative to the principal photoelectron peaks, but their possible presence must always be kept
in mind when subtleties are introduced into spectral interpretation. From a practical point
of view, the best way to approach this complexity is to interpret changes in spectra resulting
from a flotation-related treatment.
Analysis depths
The analysis depths associated with S 2p and Cu 2p electrons are quite different. For excitation
by 1486.6 eV X-rays, S 2p photoelectrons would have a kinetic energy (KE) of about 1487 - 162
= 1325 eV while the corresponding Cu 2p value would be 1487 - 932 = 555 eV. The IMFP for
electrons in solids increases approximately in proportion to √KE beyond a minimum at about
50 eV, so the Cu 2p photoelectrons would be significantly more surface sensitive than the S
2p. While this is of no consequence for a fairly thick surface layer that is uniform throughout
its depth, it can be of crucial importance for non-uniform surface layers, for example in a
comparison of the Cu activation of pyrite and sphalerite. The analysis depth associated with
A
A
B
B
C
C
D
D
E E
FIG 4 – Cu 2p and Cu LMM spectra from (A) copper metal; (B) chalcocite; (C) covellite; (D) bornite and (E) chalcopyrite.
Cu 3p photoelectrons, which have a binding energy near 77 eV, is closer to that for S 2p
electrons. Thus, the analysis depth for conventional XPS varies across the binding energy
range, the greatest surface sensitivity being at the highest binding energy (lowest kinetic
energy) end.
An appreciation of the magnitude of the analysis depth and its variation with photoelectron
energy is fundamental to the meaningful application of XPS to flotation-related systems. As
discussed in the introduction, for flotation it is the chemical composition of the outermost
atomic layers that is of most relevance, whereas the analysis depth for XPS varies from
about 1 nm to 5 nm. At best, then, XPS has adequate surface sensitivity, but away from
that maximum sensitivity, the contribution to a spectrum from the outermost atomic layers
becomes a minor one.
It is difficult to provide a precise value for the analysis depth. The intensity of photoelectrons
escaping from a surface through a distance x in a specimen is proportional to (1 - e-x/λ), where
λ is the attenuation length (which takes into account elastic as well as inelastic scattering).
Thus, 63 per cent of the intensity would arise from x = λ, 86 per cent would arise from x =
2λ, and 95 per cent would arise from x = 3λ. In practice, a thickness of 3λ is usually taken as
the thickness analysed, but as explained below, the relationship between that thickness and
the surface layer depth analysed depends on the photoelectron take-off angle (θ). For 1 keV
electrons, λ is of the order of 1–2 nm.
can become positively charged, even when a sample is in good electrical contact with the
spectrometer specimen stage, leading not only to photoelectron peak shifting to lower kinetic
energy and hence higher apparent binding energy, but also to peak broadening. However,
when an adsorbate or altered layer is either thicker than a few nm or for some other reason in
poor electrical contact with a highly conducting mineral substrate, even those surface layers
can become sufficiently charged to noticeably affect photoelectron peak energies. With poorly
conducting minerals, and with thick altered or adsorbate layers, a small flux of low energy
(<5 eV) electrons from a flood-gun is used to ‘compensate’ for the positive charge. Although
some spectrometers now have highly effective, in-lens flood-guns, they are operated in a way
that slightly over-compensates, rather than exactly compensates, for the positive charging.
Because the specimen is usually electrically isolated when the flood-gun is in use, the surface
layers become uniformly negatively charged and hence all photoelectron peaks become
narrower and are shifted by up to the flood-gun electron energy to lower apparent binding
energy. The precise energy shift is deduced from the observed position of a peak of known
binding energy (such as the C 1s peak from aliphatic hydro¬carbons), and other binding
energies adjusted assuming that all photoelectron peaks will be shifted by the same energy
(which is a reasonable but not entirely correct approximation).
h h
e-
e-
63% of
analysis
depth
FIG 5 – Schematic illustration of the effect of electron take-off angle (Θ) on X-ray photoelectron spectroscopy (XPS) analysis depth.
detected electrons will have passed furthest through the specimen for a given depth below
the surface, and the analysis depth will be minimised. However, only if the specimen surface
is a relatively smooth single particle one will lowering the electron take-off angle lead to a
marked enhancement of surface sensitivity. In practice, the minimum electron take-off angle
is ~15°, but even at that angle, the analysis depth is four times lower than for 90°. For fine
particulate specimens, reducing the take-off angle has a much smaller effect.
The other way to enhance the surface sensitivity of XPS is to lower the photon energy so that
the photoelectron kinetic energy is reduced. If there weren’t a need to also minimise the X-ray
source linewidth and complexity, reducing the photon energy would simply be a matter
of replacing the usual Al or Mg anode in the X-ray source with an anode made of another
suitable metal and utilising an X-ray line other than Kα, such as the 192.3 eV Mζ line from
Mo, but most alternative sources would require the development of a new monochromator
(Krause, 1971). Rather than go down that path, it is preferable to use a synchrotron as a source
of X-rays, as in addition to making SXPS possible, the high flux of photons of tunable energy
and aligned polarisation allows angle-resolved NEXAFS spectroscopy to be carried out. The
application of SXPS to flotation-related mineral systems is covered in a later section.
A B
408 404 400 396 392 908 912 916 920 924 928
Binding Energy (eV) Kinetic Energy (eV)
C D
408 404 400 396 392 908 912 916 920 924 928
Binding Energy (eV) Kinetic Energy (eV)
FIG 6 – (A, C) N 1s and (B, D) Cu L3M4,5M4,5 spectra from an abraded chalcocite surface conditioned 10 min in
0.01 M IPETC at pH ~6 obtained with the specimen earthed and: (A, B) flood-gun off; (C, D) flood-gun on.
A B
C D
FIG 7 – N 1s spectra from a cubanite fracture surface conditioned 10 min in ~0.02 M hydroxamate at pH 9.5 with specimen: (A–C)
earthed; (D) floating, and with: (A) flood-gun off, at the outset; (B) flood-gun off, at end of initial spectral suite; (C, D) flood-gun on.
good for mineral studies, and one covering 100 to 2500 eV (which also encompasses the S
K-edge near 2470 eV) would be ideal. In the soft X-ray range, the photon energy is normally
selected by means of a grating monochromator. It is not unusual for more than one grating to
be required to span the full energy range, and for the X-ray linewidth and flux to vary across
the energy range covered by each grating. Thus, photon energy tunability comes at the cost
of resolution varying with energy, but the available flux is usually sufficient to maximise
resolution when necessary. The X-ray ‘spot’ size on the specimen depends on the beam-line,
but it is rarely bigger than 3 × 0.5 mm and is typically 0.5 × 0.1 mm. Because of the high
flux, specimen charging can be a problem with poorly-conducting materials (such as pure
sphalerite), even when a low energy electron flood-gun for charge compensation is employed.
Also, because of the high flux, the possibility for beam damage to the specimen is greater than
in conventional XPS, but in practice, it is more likely that the high flux will be used to improve
resolution (by narrowing slits, for example) than to reduce data acquisition time.
Synchrotron X-ray beams are horizontal, therefore an end-station suitable for SXPS and
NEXAFS spectroscopy is essentially a high-quality XPS instrument with the electron energy
analyser oriented so that its lens is horizontal, with an optimally positioned NEXAFS partial
yield detector, with the specimen mounted on a five-axis manipulator, and with the specimen
handling facilities necessary for investigating mineral fracture surfaces bearing adsorbed
collector species. The end-station on the soft X-ray spectroscopy beam-line at the Australian
Synchrotron is shown in Figure 8. That end-station, which was designed to facilitate minerals
processing research, allows liquid nitrogen cooling in the entry lock, preparation chamber
and analyser chamber to retain any moderately high vapour pressure reaction products on a
specimen surface, has the entry lock within a glove box that can be flushed with an inert gas,
and has a double-blade cleaver for the preparation of mineral fracture surfaces under UHV.
Specimen transfer between the entry lock and various chambers is via a rotary distribution
mechanism. End-stations enabling minerals processing research are available on soft X-ray
beam-lines at most synchrotrons.
FIG 8 – The end-station on the soft X-ray spectroscopy beam-line at the Australian Synchrotron.
A B
C D
FIG 9 – S 2p spectra for Cu-activated sphalerite determined at photon energies of (A) 800 eV; (B) 400 eV; (C) 260 eV and (D) 210 eV.
IV
S2p
II
III
hv = 480 eV
hv = 350 eV
hv = 240 eV
hv = 210 eV
FIG 10 – S 2p spectra for 2-mercaptobenzothiazole adsorbed on galena determined at different photon energies (Szargan et al, 1997).
the cross-section ratio is 4.4:1, so that now the S 2p spectrum is much more intense than the
Pb 4f, and it can be difficult to extract chemical information from the Pb 4f peaks. However,
at a photon energy of 250 eV, the cross-section ratio is close to 1:1 while both S 2p and Pb 4f
photoelectrons remain quite surface sensitive.
A major advantage of the ability to tune the photon energy is that X-ray absorption
spectroscopy (XAS) becomes feasible. For surface science and/or for the soft X-ray region, it is
the near-edge structure in the X-ray absorption spectrum that is of interest.
Pb 4f
S 2p
FIG 11 – S 2p/Pb 4f spectrum from galena determined at a photon energy of 1487 eV.
of NEXAFS spectroscopy in flotation research, the technique has been deemed too specialised
to be included in this Metallurgical Guide, and interested readers are encouraged to refer to
relevant journal reports (eg Goh et al, 2006a, 2008).
is called ToF-SIMS. The technique is a UHV-based one, but surprisingly, not all spectrometers
have the facility to cool the specimen before and during analysis. In ToF-SIMS, the primary
ion source is operated in pulsed mode, and a typical repetition rate of 5 kHz gives a mass
spectral range to m/z 1440 with resolution of at least 0.05.
There are several general comments that should be made about static SIMS. The first is that
the primary ion beam usually ejects more neutral species than ions; consequently, there is
no necessary relationship between the abundance of an ion peak at a particular m/z and the
surface concentration of the species giving rise to that peak. Some ions are more stable than
others, and for a given specimen, the intensities of peaks from more stable ions are expected
to be greater than the intensities of peaks from less stable ions. In particular, some species
are more stable as positive ions than negative, or vice versa, therefore it is almost always
better to determine both positive and negative secondary ion mass spectra. For an organic
adsorbate such as a collector, it is possible that only collector fragment ions will be detected
rather than the parent ion (of mass equal to the complete collector radical). In practice, and
with good reason, it is assumed that negligible interaction of ions occurs between ejection and
detection, so that if a substantial ion of a species is detected, then that species is most probably
present at the specimen surface as a discrete entity or as a fragment of a larger entity. The
exception to this assumption is the loss or gain of hydrogen. Hydrogen is often a common
constituent of the vacuum in the vicinity of a material under primary ion bombardment, and
it is not unusual to observe both positive and negative ions that have an m/z one or two atomic
mass units (amu or Da) greater than expected. These observations are not artefacts of an
incorrectly calibrated m/z scale, but it is important that the mass scale is calibrated across as
much of the range as possible with ions of unambiguous m/z. Indeed, mass scale calibration is
a critical aspect of ToF-SIMS when fragment identification is required, and when calibration
difficulties are exacerbated by mineral specimens being rough at the µm level. Most organic
fragment ions of unambiguous identity, such as CH3+, C2H5+, C3H7+, CH⁻ or C2H⁻, have low
mass values, in which case extensive extrapolation is required in order to determine the mass
values of large ions. Elemental ions such as Cl⁻ or K+ would normally be available too, but
it is generally accepted that if molecular ions are of primary interest, atomic ions are best
avoided in the calibration procedure, and it is presumed that the same advice would apply
to organometallic secondary ions. ToF-SIMS mass calibration difficulties have been discussed
by Green, Gilmore and Seah (2006).
As explained below, the unambiguous identification of some ions can require high m/z
resolution, however, for most of the ions encountered in systems related to mineral flotation,
an m/z resolution of 0.05 is usually sufficient. The identification of metal-containing ions is
often facilitated by the diagnostic stable isotopic ratio that is reflected in the cracking pattern.
For example, as can be seen in Figure 12, Cu-containing ions give rise to doublets with
components of intensity ratio approximately 2:1 separated by 2 amu reflecting the 63Cu to 65Cu
ratio of 69:31, and Ag-containing ions produce doublets with components of approximately
equal intensity separated by 2 amu reflecting the 107Ag to 109Ag ratio of 52:48. Ions containing
a single Pb give rise to triplets of intensity ratio ~1:1:2 separated by 1 amu because the 206Pb to
207
Pb to 208Pb ratio is 24:23:52. The natural abundance of the 204Pb isotope is only 1.48 per cent.
Identification of a peak at m/z 28, on the other hand, would require a mass resolution better
than 0.05, as CO would be at m/z 27.995 and C2H4 would be at m/z 28.031.
Secondary ion mass spectra are usually tedious to interpret, but they can contain chemical
bonding information, and some of the findings from static SIMS data for sulfide mineral/
collector systems have been surprising. For example, it is evident from Figure 12 that Cu(DTP)2–
S _
A DTP– Cu(DTP)H Cu(DTP)2
3.5 0.08 0.10
3.0 0.06
2.5
2.0 0.04 0.05
1.5
1.0 0.02
0.5
0.0 0.00 0.00
185 190 248 250
Cu2S – DTP
0.06 0.10
1.2
1.0
0.8 0.04
0.6 0.05
0.4 0.02
0.2
63 79 95 111 0.0 0.00 0.00
185 190 248 250 426 428 430 432 434 435 436 440
B 0.008
Ag(DTP)H– 0.05
Ag(DTP)2–
0.006 0.04
0.004 0.03
DTP– 0.002
0.02
0.01
185 0.000
Ag2S – DTP
0.00
292 294 472 474 476 478 480 482
0.06 0.04
0.04 0.03
0.02
0.02
0.01
0.00 0.00
292 294 472 474 476 478 480 482
FIG 12 – Negative secondary ion mass spectra for monolayer (upper spectrum) and multilayer
(lower spectrum) diethyl dithiophosphate (DTP) adsorbed on (A) chalcocite and (B) silver sulfide.
ions were detected for multilayer diethyl dithiophosphate (DTP) adsorbed on chalcocite, but
not for monolayer coverage. This observation suggested that the multilayer species might
either be, or contain, Cu(DTP)2, which was not expected as XPS data had indicated that the
multilayer species was (CuDTP)n. The data from both techniques could only be rationalised
if the Cu(DTP)2 ions arose from contiguous CuDTP molecules adsorbed at the surface, or if
the multilayer was predominantly CuDTP but Cu(DTP)2 was a minor component. The Cu L3-
edge NEXAFS spectrum is particularly sensitive to the presence of Cu(II) (Pattrick et al, 1997;
Goh et al, 2006b), and a small Cu(II) peak was discernible in the Cu L3-edge NEXAFS spectrum
for multilayer DTP on chalcocite, an observation consistent with the presence of a small
concentration of Cu(DTP)2 (Goh et al, 2006c). The situation for DTP adsorbed on silver sulfide
was similar, and fortuitously because of the presence of the low concentration of Cu or Ag
dithiolate in the multilayer, it was concluded that differentiation of monolayer and multilayer
coverage by static SIMS alone would be expected for all thiol collector/sulfide systems.
small organic fragments are generated in the central impact region, but immediately outside
this impact region, more extensive fragments that might have undergone some structural
rearrangement are produced. Further away from the impact region, the energy available for
rearrangement is lower, so that larger and minimally rearranged fragment ions (and even
parent ions) might be expected. On this basis, it might be anticipated that the larger secondary
ions should better reflect species present at the surface prior to primary ion impact and hence
be diagnostic. However, the larger the secondary ion, the greater the need to extrapolate
the mass range calibration, which means that the more relevant a secondary ion, the more
uncertain would its measured m/z value be. Extended mass range calibration is particularly
difficult for mineral specimens that are rough at the µm level, which is the case for all but
polished single piece specimens. Thus, while static SIMS can complement XPS for determining
the chemical bonding arrangement of a collector adsorbed at a mineral surface, it is unlikely
to be able to establish the bonding independently of XPS. For a multiphase mineral, on the
other hand, static SIMS can provide surface atomic information with high sensitivity that can
reveal to which mineral phase a collector is chemisorbed.
concentrate and tail fractions overlapped to some extent for all samples investigated. Although
oxidation products such as FeO and FeOOH contributed to the variation in chalcopyrite
surface chemistry, they were not correlated with the contact angle. To date, a similar approach
to contact angle determination has not been applied to other valuable sulfide minerals or to
pyrite, though fundamental work on the quartz-silane system (Brito e Abreu and Skinner,
2012) suggests the methodology could be extended to different collector-mineral systems.
usually determined with a laser source (eg of wavelength 785 or 1064 nm) and a cooled solid-
state detector. The spectroscopy is rendered surface sensitive by suitably ‘roughening’ the
surface of a metal substrate. FT-SERS is now a well-established technique for the examination
of adsorbed layers, including self-assembled monolayers, at electrochemically roughened Ag,
Au and Cu surfaces, and it was used in conjunction with XPS in a flotation-related context
to verify that xanthate chemisorbed on Ag without decomposition (Buckley et al, 1997). By
contrast, the value of SERS spectroscopy for investigating alteration of, and adsorption on,
sulfide mineral surfaces is only starting to be realised. Promising results have been reported
by Hope and co-workers on evaporative ‘decoration’ with gold of an MBT layer adsorbed on
chalcopyrite to induce SERS (Hope et al, 2006), and also on gold decoration of chalcopyrite
prior to interaction with MBT (Buckley et al, 2006). In the latter case, chemisorption of MBT on
the Au decoration was found to be not a problem, as adsorption on the mineral sulfide occurred
at a potential lower than for the Au. However, it cannot always be assumed that collector
adsorption on the Au nanoparticles will not occur for pre-decorated mineral surfaces. While
there is usually sufficient spectral resolution to distinguish scattering peaks from collector
chemisorbed on Au from that chemisorbed on a mineral, for example the C-S stretch in MBT
chemisorbed on Au is at ~710 nm whereas for MBT chemisorbed on the Cu in a Cu sulfide
mineral it is at ~717 nm (Woods, Hope and Watling, 2000), it might not be clear whether a
multilayer species such as (MBT)2 was adsorbed on the Au decoration or on the mineral itself.
Nevertheless, a significant strength of SERS spectroscopy is its spatial resolution, the laser
spot size being typically 5 µm in diameter, and as discussed in a later section, another major
strength is its applicability as an in situ characterisation technique.
organic adsorbates. In general, the in situ techniques are used less frequently than the ex situ
techniques for flotation plant problem-solving, and hence are described here in less detail.
Electrochemical techniques
The application of electrochemical techniques in flotation research has been described
by numerous authors, including Chander and Gebhardt (1989). The most widely used
electrochemical technique is voltammetry, especially when rotating ring disc electrode and
photoelectrochemical measurements are included, but less common transient response
techniques such as AC impedance spectroscopy and chronoamperometry can be informative
for some systems. The application of linear potential sweep (uni-directional) or cyclic (bi-
directional) voltammetry for investigating the oxidation of sulfide minerals, and the interaction
of sulfide minerals with collectors and other surface modifiers such as hydrosulfide ions,
has been described by Woods in the section on Electrochemical Investigative Techniques in
Chapter 8 and elsewhere in more detail (Woods, 1996). In cyclic voltammetry, the variation of
the current that flows as the potential applied to the mineral electrode is varied in a cyclical
scan reveals the potential at which a reaction (eg chemisorption) occurs at the electrode, and
the extent of that reaction, but it does not unambiguously establish the identity of the reaction
products. The reaction products are deduced from the potential at which the reaction occurs,
the effect of pH on that potential, the inferred reactions of the products on the reverse scan,
and, if a rotating ring disc electrode arrangement is in use, the effect of rotating the electrode
to disperse soluble products. One weakness in this procedure is that inferred reactions are
usually based on bulk rather than surface reaction potentials. Thus, voltammetry is most
powerful when used as an in situ adjunct to an ex situ technique, such as XPS, that does
identify reaction products more directly (Buckley and Woods, 1993, 1995).
It should also be noted that in cyclic voltammetry, electrode reactions are driven at sweep
rates of the order of 10 mV.s⁻1, therefore it is conceivable that a reaction that takes place during
a potential scan may be different from the reaction that takes place spontaneously but slowly
at open circuit. Also, it is possible that a reaction that takes place under the steady-state
conditions of multiple scan cyclic voltammetry might be different from the corresponding
reaction that occurs during the first potential scan. Indeed, when the oxidation behaviour of
pyrite and pyrrhotite fracture surfaces was investigated by means of cyclic voltammetry, it
was found that, commencing at the rest potential, the first and second positive-going scans
for pyrite were similar, whereas they were different for pyrrhotite (Buckley, Hamilton and
Woods, 1988). However, additional information concerning the formation of a metal-deficient
lattice during initial oxidation was obtained from interpretation of those observations.
Voltammetric investigations of the interaction of sulfide minerals with flotation collectors
have revealed a pre-wave on the positive-going scan at a potential below the reversible value
for the formation of the metal thiolate for numerous sulfide/collector systems. In each case,
the charge associated with the prewave corresponds to that expected for the charge-transfer
adsorption of a monolayer of the collector (Buckley, Hope and Woods, 2003).
Infrared spectroscopy
Water is a strong absorber of IR radiation, therefore the in situ application of IR spectroscopy
is not straightforward. Nevertheless, mineral sulfide/collector systems have been investigated
by ERS under potential control in specially designed cells (eg Bozkurt et al, 1996) that enable
the mineral electrode to be positioned close to (within 100 μm of) the cell window during
the recording of single reflection spectra. Calcium fluoride is one cell window material that
has been used. However, in most of this work, as discussed in the consideration of ex situ
techniques, the requirements of a moderately large area (~10 × 10 mm) and highly polished
mineral specimen surface are disadvantages.
As an alternative to ERS, internal reflection spectroscopy, or ATR, has been used with a
finely ground mineral in carbon paste working-electrode pressed against a Ge ATR crystal
(eg Leppinen, Basilio and Yoon, 1989). These measurements have been extremely informative
in revealing the identity of the adsorbed species for particular Eh/pH conditions, and it is
conceivable that they could be of use in flotation plant problem-solving applications.
Raman/SERS spectroscopy
Raman scattering spectroscopy has some advantages over IR spectroscopy in that water is a
weak (but not negligible) Raman scatterer, and glass cells can be used. In situ SERS spectroscopic
investigations of the interaction of flotation collectors with electrochemically-roughened Au,
Ag and Cu metal surfaces have been carried out under potential control (Hope et al, 2006).
Of more relevance for investigating flotation plant problems, the demonstration that Au
decoration of a chalcopyrite surface to impart a SERS response could be carried out prior
to interaction with MBT solution (Buckley et al, 2006) raised the possibility of determining
in situ SERS spectra of other sulfide mineral/collector systems. More recently, in situ and
ex situ SERS spectroscopy has been used to investigate the interaction at different potentials
of n-octanohydroxamate collector with sulfide minerals (such as pyrite) pre-decorated with
Au (Parker et al, 2012), and the adsorption of thiol collectors on a range of pre-decorated Cu
and Cu-Fe sulfide minerals.
CONCLUDING REMARKS
The purpose of this chapter was to assist a minerals processing practitioner to more fully
appreciate the potential role of surface chemical characterisation in optimising flotation plant
performance, and to better understand any surface analytical data obtained for that reason.
Such data might have been sought to address a particular recovery or grade issue encountered
in a flotation plant. A typical problem might have been a concentrate grade that was lower
than expected, in which case surface chemical information relating to possible explanations
such as whether or not a gangue mineral was inadvertently becoming activated, a target
mineral surface was being altered by physical contact with a non-target mineral, or the
collector used was insufficiently selective, might have been sought. The problem might have
been a recovery that was lower than expected, in which case surface analytical data providing
information on collector coverage, adsorbed collector species, mineral surface oxidation, or
co-adsorption of hydrophilic species might have been pursued. It has been explained in this
chapter that most plant performance concerns can be investigated by surface characterisation
of either pulp samples or mineral pieces appropriately conditioned in a laboratory.
A reader may have gained the impression that insufficient attention was given to some
surface chemical characterisation techniques, and excessive emphasis given to others. In one
sense this is true, but it must be kept in mind that this chapter addressed only those surface
chemical characterisation techniques of most use for optimising flotation plants. It is not
through any bias of the author, who has very little involvement in the day-to-day activities
of a surface analysis facility, that the ex situ techniques of XPS and to a lesser extent ToF-
SIMS have become the almost universal starting point for surface chemical problem-solving.
These spectroscopies have a successful track record, and accordingly, the emphasis in this
chapter has been on them. The synchrotron-based spectroscopies of SXPS and XAS deliver on
information needed to complement that from XPS and static SIMS, especially the information
from SXPS, which has the surface sensitivity essential for the identification of some flotation
problems. The application of SXPS and XAS can only increase as more synchrotron soft X-ray
beam-lines become available in countries with vibrant minerals processing industries.
ACKNOWLEDGEMENTS
The author is grateful to Prof Rüdiger Szargan, Leipzig University, for Figure 10, Prof Bill
Skinner, University of South Australia, for the end-station photograph in Figure 8, and Dr Bill
Gong, UNSW Analytical Centre, for assistance in obtaining XPS and 69Ga SIMS data.
REFERENCES
Benninghoven, A, 1994. Chemical analysis of inorganic and organic surfaces and thin films by static
time-of-flight secondary ion mass spectrometry (TOF-SIMS), Angewandte Chemie International Edition
England, 33:1023–1043.
Bozkurt, V, Brienne, S H R, Xu, Z, Rao, S R, Butler, I S and Finch, J A, 1996. Development of in situ
external reflection spectroscopy technique for adsorbed films, Minerals Engineering, 9:351–355.
Briggs, D and Seah, M P, 1990a. Practical surface analysis, Auger and X-ray Photoelectron Spectroscopy,
second edition, volume 1, 657 p (Wiley: Chichester).
Briggs, D and Seah, M P, 1990b. Practical Surface Analysis, Ion and Neutral Spectroscopy, second edition,
volume 2, 756 p (Wiley: Chichester).
Brito e Abreu, S, Brien, C and Skinner, W, 2010. ToF-SIMS as a new method to determine the contact
angle of mineral surfaces, Langmuir, 26:8122–8130.
Brito e Abreu, S and Skinner, W, 2011. ToF-SIMS-derived hydrophobicity in DTP flotation of
chalcopyrite: contact angle distributions in flotation streams, International Journal of Mineral Processing,
98:35–41.
Brito e Abreu, S and Skinner, W, 2012. Determination of contact angles, silane coverage, and
hydrophobicity heterogeneity of methylated quartz surfaces using ToF-SIMS, Langmuir, 28:7360–7367.
Buckley, A N, 1994. A survey of the application of X-ray photoelectron spectroscopy to flotation
research, Colloids and Surfaces A, 93:159–172.
Buckley, A N, Hamilton, I C and Woods, R, 1988. Studies of the surface oxidation of pyrite and pyrrhotite
using X-ray photoelectron spectroscopy and linear potential sweep voltammetry, in Proceedings
International Symposium Electrochemistry in Mineral and Metal Processing II (eds: P E Richardson and
R Woods), pp 234–246 (Electrochemical Society: Pennington, NJ).
Buckley, A N, Hope, G A, Lee, K C, Petrovic, E A and Woods, R, 2014. Adsorption of O-isopropyl-N-
ethyl thionocrabamate on Cu sulfide ore minerals, Minerals Engineering, 69:120–132.
Buckley, A N, Hope, G A and Parker, G K, 2014. A spectroscopic investigation of the interaction of
n-octanohydroxamate collector with Cu–Fe sulfide minerals, Minerals Engineering, 64:23–37.
Buckley, A N, Hope, G A and Woods, R, 2003. Metals from sulfide minerals: the role of adsorption of
organic reagents in processing technologies, in Solid–Liquid Interfaces, Topics in Applied Physics (eds:
K Wandelt and S Thurgate), 85:61–96 (Springer: Berlin).
Buckley, A N, Laajalehto, K, Suoninen, E and Woods, R, 1995. Flotation-related surface reactivity of
molybdenite, in Processing of Hydrophobic Minerals and Fine Coal (eds: J S Laskowski and G W Poling),
pp 29–42 (Canadian Institute of Mining, Metallurgy and Petroleum: Montréal).
Buckley, A N and Parker, G K, 2013. Adsorption of n-octanohydroxamate collector on iron oxides,
International Journal of Mineral Processing, 121:70–89.
Woods, R, 1996. Chemisorption of thiols on metals and metal sulfides, in Modern Aspects of Electrochemistry
(eds: J O M Bockris, B E Conway and R E White), volume 29, pp 401–453 (Plenum: New York).
Woods, R, Hope, G A and Watling, K, 2000. A SERS spectroelectrochemical investigation of the
interaction of 2-mercaptobenzothiazole with copper, silver and gold surfaces, Journal of Applied
Electrochemistry, 30:1209–1222.
Zachwieja, J B, McCarron, J J, Walker, G W and Buckley, A N, 1989. Correlation between the surface
composition and collectorless flotation of chalcopyrite, Journal of Colloid and Interface Science,
132:462–468.
ABSTRACT
The application of surface analysis techniques to solving problems in flotation plants can
provide tremendous insights into why things have happened. It is also important to recognise
that surface analysis is not a ‘magic bullet’, but data that complements information collected
from plant metallurgical surveys, recovery-by-size analysis and pulp chemical surveys. This
appendix aims to complement the information provided in Chapter 9 with some practical
examples.
INTRODUCTION
Three examples are provided where surface analysis techniques have been used in tandem
with pulp chemistry and classical metallurgical tests to identify the root cause of the problem.
This has led (potentially) to a more focused solution.
120 10.0
100
80
40
20
pH
8.0
0
-20
-40 7.0
-60 Grinding Copper rougher Lead rougher circuit
-80 circuit circuit
-100 6.0
er
s
ed
f
il
il
il
il
il
il
il
il
f
o/
o/
u/
ta
ta
ta
ta
ta
ta
ta
ta
ee
at
fe
e
e
en
E
C
G
F
rf
w
on
on
04
D
01
01
04
04
01
04
04
he
s
re
04
cl
es
cl
Circuit position
sc
ug
cy
cy
oc
ro
SM
y
y
Pr
ar
ar
u
D
nd
C
im
Pr
co
Eh Dissolved oxygen pH
Se
FIG A1 – Pulp chemical measurements (Eh, pH and dissolved oxygen) through the grinding and copper/lead rougher flotation circuits.
occurring. Lime was added to conditioning tank 04A, and the pH was gradually reduced
through the lead rougher circuit. This corresponded to an increase in Eh and was explained in
terms of water equilibria1.
The Eh-pH changes through the circuit can be more readily observed in Figure A2. (Note:
The diagonal lines in this graph are parallel to the water-oxygen line.) Changes that are
approximately parallel to the water-oxygen line, for example the grinding circuit (points 1 to 3,
in Figure A2) and the lead rougher circuit (points 8 to 13, in Figure A2), can be ascribed to the
maintenance of water equilibria. However, if the changes are approximately perpendicular to
the water-oxygen line, for example the first two cells of the copper rougher circuit (points 4
to 6, in Figure A2), then oxidative reactions are the cause. Thus, the Eh-pH data suggested that
oxidation was occurring in the copper rougher circuit.
60
11
40 12 10
1
13 6
20 1. DSM screen u/s
7
2. Primary cyclone o/f
5 9
Eh, mV (SHE)
-100
7.0 8.0 9.0 10.0
pH
FIG A2 – Eh-pH position of the pulp at each stage within the circuits.
Note: the diagonal lines in this graph are parallel to the water-oxygen line.
The dissolved oxygen profile (Figure A1) indicated that the oxygen levels were low through
the grinding circuit, particularly after tower milling. Once the pulp entered the copper rougher
flotation circuit, the dissolved oxygen levels began to increase, but remained comparatively
low, implying that oxygen was being consumed from the pulp, presumably as oxidation
occurred. The dissolved oxygen concentration reached saturation values in the lead rougher
circuit, and remained relatively constant after cell 04A. Thus, oxygen consumption was low
in the lead circuit and suggests that oxidation may be substantially completed prior to lead
rougher flotation.
EDTA was used in diagnostic leach tests to selectively dissolve the oxidised species from
the surfaces of sulfides to determine the level of oxidation of the various species present in
the system. The EDTA extraction data, collected using the method described by Rumball and
Richmond (1996), are presented in Figure A3.
A small increase in EDTA extractable lead occurred across the primary grinding circuit
(Figure A3) indicating that galena oxidation was reasonably limited in this part of the
circuit. However, the EDTA extractable lead levels increased dramatically across the copper
rougher section, from 2.8 to 5.2 per cent for approximately the same lead head grade. This
implied that significant galena oxidation occurred in the copper rougher circuit. The EDTA
extractable lead profile through the lead rougher showed a substantial increase in oxidised
lead. However, as the lead head grade decreased through lead rougher flotation, these data
suggest that the galena remaining in the lead rougher tailing was in a more oxidised state.
The lead EDTA extraction data therefore indicated that the majority of the galena oxidation
occurs through the copper rougher circuit. Further, the galena that reported to the lead
rougher tailing was significantly more oxidised than the galena in the lead rougher feed (ie
4.9 versus 10.4 per cent). These observations are the same as those made by Rumball and
Richmond (1996).
The EDTA extractable zinc remained comparatively low throughout the circuit (Figure A3).
The EDTA extractable iron (Figure A3) increased across the primary grinding circuit as contact
with the grinding media increased. There was a further small increase in EDTA extractable
12.0 1.0
10.0
0.8
% EDTA Zn and Fe
8.0
% EDTA Pb
0.6
6.0
0.4
4.0
0.2
2.0
ed
d
f
il
il
il
il
il
il
il
il
f
o/
o/
u/
ta
ta
ta
ta
ta
ta
ta
ta
e
fe
fe
ne
e
en
E
C
G
on
er
04
D
01
01
04
04
01
04
04
lo
re
04
h
cl
Circuit position
sc
ug
cy
cy
ro
SM
y
y
ar
ar
u
D
nd
C
im
Pr
co
Se
Pb Zn Fe
iron through the copper rougher circuit, which may be associated with the oxidation of pyrite
in this part of the concentrator. The per cent oxidised iron then remained relatively constant
across the lead rougher circuit. It is known that the oxidation products of pyrite, for example,
FeOOH (Hiskey and Schlitt, 1982) and higher iron oxides are less soluble in EDTA (Blesa,
Rueda and Grassi, 1985) than iron hydroxides, and therefore are not as readily extracted with
EDTA. Hence, it can be claimed that the pyrite surfaces on reaching the lead rougher circuit
were coated with FeOOH and higher iron oxide species, which resulted in the passivation of
the pyrite. The lower solubilities of these species in EDTA caused the EDTA extractable iron
concentration to remain approximately the same through this part of the circuit.
Solution chemistry
A major technical issue at the Hellyer concentrator was the low galena recovery (42 per cent)
to final lead concentrate (Richmond, 1993), which is considerably lower than that achieved
by most other lead/zinc operations within Australia. So, a comparison of the lead rougher
concentrate and tailing was made.
Water analysis
The results of the water analysis performed on the lead rougher concentrate and tailing
slurries are presented in Table A1, and the corresponding X-ray fluorescence (XRF) analysis
of the solids contained in the two process streams are provided in Table A2.
The water-soluble lead, zinc, iron and magnesium concentrations were comparatively
low (less than 20 parts per million) in both the concentrate and tailing samples. Calcium
was present at significantly higher concentrations, and was marginally higher in the tailing
sample compared with the concentrate. The high calcium ion content can be attributed to the
addition of lime to the circuit as the pH modifier. These measurements compare favourably
with those made by Woodcock and Jones (1970) and Grano, Ralston and Johnson (1988), who
reported that the levels of soluble lead, zinc and iron in process streams were generally less
than 20 parts per million.
TABLE A1
Cation and anion analysis of water extracted from the lead rougher concentrate and tailing process streams.
Extracted species Pb rougher concentrate Pb rougher tailing
Cations
Pb2+ 8.2 ppm 1.9 ppm
Zn2+ <0.1 ppm <0.1 ppm
Fe2+/Fe3+ <0.1 ppm <0.1 ppm
Mg 2+
13.2 ppm 16.6 ppm
Ca2+ 310 ppm 410 ppm
Anions
OH- <0.1 ppm <0.1 ppm
CO32- <0.1 ppm <0.1 ppm
HCO3 -
<0.1 ppm <0.1 ppm
CN -
32.2 ppm 46.8 ppm
SO4 2-
310 ppm 402 ppm
SxOy2- 10.6 ppm 70.1 ppm
TABLE A2
X-ray fluorescence assays of the lead rougher concentrate and tailing process streams.
Stream Elemental composition (%)
Ag Pb Zn Cu Fe S Ba As Sb
(ppm)
Pb rougher concentrate 551 52.54 6.72 0.24 12.83 25.74 0.20 0.99 0.18
Pb rougher tailing 69 2.90 12.41 0.15 26.01 35.55 0.85 1.99 0.09
The anion analysis showed that hydroxide, carbonate and bicarbonate were below the
detection limit in both streams. Cyanide ions were detected in both process streams, with
marginally higher levels of cyanide detected in the tailing compared with the concentrate. The
sulfate and sulfoxy (possibly thiosulfate) levels observed in the tailing stream were slightly
greater than those recorded for the concentrate (Table A1). Although the cation analysis did
not provide a clear indication of the chemical state of the pulp, the anion data, in particular
the variations in sulfoxy species, did suggest that the tailing sample contained more oxidised
species than the concentrate.
EDTA analysis
It has been recognised by numerous workers (Shannon and Trahar, 1986; Grano, Ralston
and Johnson, 1988; Rumball and Richmond, 1996) that EDTA has the ability to solubilise
the oxidation products of sulfide minerals. However, there is still considerable debate
surrounding the solubility of sulfide minerals in EDTA solutions. Wang and Forssberg (1990)
suggest that in the presence of excess EDTA, after the ‘hydroxides’ have been removed (which
may leave a sulfur-rich surface), the EDTA continues to leach the bulk sulfide (galena, pyrite,
arsenopyrite). However, it was further suggested that the dissolution of sulfide with a sulfur-
rich surface would be slow. Pang and Chander (1992) support this view, for pyrite at least.
Thus, assuming that the dissolution of sulfide minerals with a sulfur-rich surface by EDTA is
slow, it is reasonable to assume that the measured concentrations of EDTA extractable ions
relate directly to the oxidation of the constituent minerals present in the pulp provided the
leach time is not excessive.
Greet and Smart (2002) describe experimental work conducted to determine the solubility
of galena and its oxidation products in EDTA. The results can be summarised as:
•• The experiments conducted demonstrated that EDTA readily solubilises lead hydroxide,
lead carbonate, lead sulfate and lead hydroxy-carbonate, all typical galena oxidation
products. Further, a 0.1 molar EDTA solution will solubilise galena oxidation products
rapidly at low concentrations, ie within five minutes. Purging the pulp with argon during
the extraction process significantly reduces the possibility of continuing oxidation during
leaching.
•• EDTA did not extract lead from galena. Continuing lead extraction with time is due to
continuous oxidation of galena.
•• A modified EDTA extraction technique, based on these observations was developed. The
new test employs argon as the purging gas, and a leach time of five minutes. Experiments
were conducted comparing the new method against other commonly used techniques.
The results of these tests suggest that the new method produces significantly lower levels
of EDTA extractable lead than the other techniques. These differences were attributed to
variations in the oxygen content of the pulp during leaching.
The results of EDTA extractions performed on the lead rougher concentrate and tailing
process streams are contained in Table A3.
The concentrations of EDTA extractable lead were similar for both process streams; however,
the significance of this data becomes apparent only when the metal content of the process
stream is considered (Kant, Rao and Finch, 1994; Rumball and Richmond, 1996). In this
example, the lead rougher concentrate assayed 52.5 per cent lead, while the lead rougher
tailing contained 2.9 per cent lead (Table A2). When these grade differences are taken into
TABLE A3
Cation and anion analysis of ethylene diaminetetraacetic acid (EDTA) solutions used in diagnostic
leach tests performed on the lead rougher concentrate and tailing process streams.
Extracted species Pb rougher concentrate Pb rougher tailing
Cations
Pb2+ 2918 ppm 2445 ppm
Zn 2+
106 ppm 184 ppm
Fe2+/Fe3+ 867 ppm 1394 ppm
Mg2+ 42 ppm 116 ppm
Ca 2+
1293 ppm 10 853 ppm
Anions
OH- <0.1 ppm <0.1 ppm
CO3 2-
<0.1 ppm <0.1 ppm
HCO3- <0.1 ppm <0.1 ppm
CN -
26.5 ppm 43.3 ppm
SO4 2-
494 ppm 660 ppm
SxOy2- 24.4 ppm 66 ppm
TABLE A4
The percentage of ethylene diaminetetraacetic acid (EDTA) extractable lead,
zinc and iron for the lead rougher concentrate and tailing process streams.
Metal ion Pb rougher concentrate (%) Pb rougher tailing (%)
Pb 0.78 15.88
Zn 0.22 0.28
Fe 0.96 1.01
account when calculating the percentage of EDTA extractable lead (or per cent oxidised lead),
significant differences are observed (Table A4):
Mass of Pb in EDTA solution
% EDTA extractable Pb = # 100
Mass of Pb in solids
Using the above equation, the percentage of EDTA extractable lead, zinc and iron for the
lead rougher concentrate and tailing process streams were determined and are presented in
Table A4.
The percentage of EDTA extractable (oxidised) lead in the tailing as a proportion of lead in
the tailing stream was about 20 times greater than that observed in the concentrate (that is,
15.9 per cent versus 0.8 per cent). This suggested that approximately 16 per cent of the lead
in the tailing sample was present as galena oxidation products, and it is proposed that the
galena reporting to the tailing is heavily oxidised compared to the galena in the concentrate.
While differences were observed in the concentrations of zinc and iron, between the lead
rougher concentrate and tailing (Table A3), the percentages of EDTA extractable zinc and
iron were determined to be similar in both process streams, with a small tendency for more
extractable (oxidised) species in the tailing.
Analysis of the anion species contained within the EDTA solutions indicated that the
sulfate content of the tailing was approximately 30 per cent greater than that reported for the
concentrate (Table A3). Again, sulfoxy species (probably thiosulfate) were observed to be in
higher concentrations in the tailing than the concentrate. Gardener and Woods (1979) showed
that in alkaline solutions galena oxidation produced both sulfate and thiosulfate, whereas
pyrite oxidation produced only sulfate (Hamilton and Woods, 1986). Thus, the presence of
higher levels of sulfoxy species in the tailing would suggest that the galena in this stream
was more extensively oxidised than the concentrate. These observations also corroborate the
comments made in relation to the cation analysis presented above.
Both the water and EDTA analysis indicated that the galena in the tailing process stream
was more heavily oxidised than the galena in the concentrate.
Surface analysis
A number of surface analysis techniques were used during this project to examine the surface
composition of particles from various flotation streams. These are listed below.
Sample preparation
In general, the same sample preparation method was used for each of the techniques
detailed below. Slurry from the process stream of interest was collected in plastic phials,
which were purged with nitrogen for two minutes and then sealed. The sealed phial was
then ‘snap frozen’ in liquid nitrogen and stored in a freezer until analysis could be arranged
(Smart, 1991).
Samples for examination were introduced into each instrument in much the same way. The
prepared slurry sample2 was carefully spread over the sample stub and placed into the fore-
chamber of the instrument. As this chamber was evacuated, the solution phase of the slurry
sample was evaporated off. Once the pressure in the fore-chamber was approximately the
same as that of the main chamber the sample was introduced and the analysis performed.
Several surface analysis techniques were used during this investigation to determine the
differences in surface composition of galena found in the lead rougher concentrate and
tailing process streams. X-ray photoelectron spectroscopy (XPS) was used initially. The XPS
analysis provided information about the composition of the surface, and the chemical states
of the various species detected. This technique examines an area containing a large number
of particles, and yields the average surface composition. That is, the XPS analysis is not particle
specific. Therefore, the XPS analysis did not give data regarding the surface compositional
differences between individual galena grains from the concentrate and tailing, so other surface
analysis techniques were employed. Scanning Auger microscopy (SAM) and time-of-flight
secondary ion mass spectroscopy (ToF-SIMS) were used to examine individual particles from
the concentrate and tailing process streams.
2. In some instances, the pulp samples were pre-treated by successive sedimentation/decantation steps, with the solution being
replaced with nitrogen-purged demineralised water. This process was used to remove fine (or slimes) particles, which, as the
solution was evaporated off under vacuum, may deposit on to the surfaces of the minerals of interest and inadvertently interfere
with the surface analysis. This process also removed soluble species (ie sulfates) from solution that may precipitate out of solution
as the solution was evaporated off under vacuum.
FIG A4 – Secondary electron images of lead rougher concentrate: (A) ‘as received’;
and (B) deslimed (using six cycles of sedimentation/decantation/centrifuged).
TABLE A5
Atomic concentration data for elements detected on the surfaces of particles from the lead rougher concentrate
and tailing in the ‘as received’ and deslimed condition, using X-ray photoelectron spectroscopy (XPS).
Element Lead rougher concentrate Lead rougher tailing
‘As received’ Deslimed 'As received' Deslimed
Original Etched a
Original Etched Original Etched Original Etched
C 28.62 9.56 33.89 10.04 19.78 3.65 16.09 4.12
O 45.71 52.44 37.89 36.24 55.98 62.38 57.48 63.19
S 10.16 12.42 9.16 15.72 7.01 8.14 4.26 5.43
Pb 3.29 5.05 5.73 10.29 0.92 0.83 1.56 1.38
Zn 0.88 1.95 2.00 6.98 1.16 2.57 1.87 3.25
Cu 0.27 5.33 0.07 0.00 0.00 0.00 0.00 0.00
Fe 2.54 0.00 3.67 10.91 1.00 4.97 4.03 6.72
Al 1.69 2.65 2.81 1.99 3.32 4.30 2.99 3.99
Si 2.49 4.83 3.97 6.48 6.54 8.18 9.84 9.68
Ca 4.35 5.77 0.81 1.35 4.29 4.98 1.88 2.24
a. In most instances once data had been acquired from the original surface an etch was applied, using a 3 kV accelerated argon ion beam.
Data was then collected from the ‘fresh’ surface. At the time these experiments were conducted, it was assumed that etching
removed approximately 2.5 nm from the surface (Skinner, personal communication, 1994). The argon ion etch was applied in order to
investigate the subsurface of the minerals.
Comparing the data for the concentrate with the tailing showed that the surface carbon
concentration was significantly greater for the concentrate sample (Table A5). Presumably,
this was because of the presence of xanthate on the surfaces of the galena particles in this
stream.
The surface oxygen concentration was considerably lower for the concentrate compared with
the tailing (Table A5). This suggested that the tailing contained a higher proportion of oxidised
species. However, if the aluminium and silicon detected on the surface are assumed to be in
the expected forms of alumina (Al2O3) and silica (SiO2), respectively, their contributions to
the surface oxygen concentration can be determined and subtracted from the original oxygen
content. Further, part of the calcium present may be as calcium sulfate. From the sulfur spectra
(Figure A5), it is known that the desliming treatment removed the sulfate from the mineral
surfaces; therefore, it can be surmised that the calcium remaining after desliming was not
present as calcium sulfate. Thus, by subtracting the calcium content of the deslimed sample from
that of the ‘as received’ sample, the concentration of the calcium associated with sulfate can be
estimated, and the associated oxygen content determined. If this value is added to the oxygen
concentrations calculated for the aluminium and silicon surface concentrations, the surface
oxygen concentration attributable to the non-sulfide gangue component can be established.
Subtracting this component from the original surface oxygen content will yield the oxygen
associated with the sulfide component of the surface. Table A6 summarises these calculations.
(1)
(1) Pb rougher concentrate - "as received"
Arbitrary units
(3)
M1-xS
FeSO4
(4) PbSO4
PbX2
ZnS
FeS2 PbS
176.0 174.0 172.0 170.0 168.0 166.0 164.0 162.0 160.0 158.0
Binding energy, eV
FIG A5 – Sulfur 2p spectra for lead rougher concentrate and tailing process streams.
TABLE A6
Summary of calculations to determine the oxygen concentration associated with the sulfide mineral
component of the surface for the lead rougher concentrate and tailing process streams.
Lead rougher concentrate Lead rougher tailing
Atomic concentration of O (%) Atomic concentration of O (%)
Original O concentration 45.71 55.98
O associated with Al (as Al2O3) 1.69 × 3/2 = 2.54 3.32 × 3/2 = 4.98
O associated with Si (as SiO2) 2.49 × 2/1 = 4.98 6.54 × 2/1 = 13.08
O associated with Ca (as CaSO4) (4.35 - 0.81) × 4/1 = 14.16 (4.29 - 1.88) × 4/1 = 9.64
Total non-sulfide gangue O 21.68 27.70
Sulfide minerals O 24.03 28.28
This analysis suggested that the non-sulfide gangue surface oxygen component for the
concentrate sample was dominated by the high calcium sulfate contribution. In the tailing
sample, the alumina and silica were the major components of the non-sulfide gangue
oxygen concentration, and this can be directly attributed to the higher levels of alumina and
silica gangue minerals in this process stream. The oxygen concentration associated with the
surfaces of sulfide minerals was greater in the tailing (28.28 per cent), than that calculated
for the concentrate (24.03 per cent). This suggested that there was more oxygen associated
with the sulfide minerals in the tailing, which implied that the tailing sulfides were more
oxidised than those appearing in the concentrate. It is recognised that this interpretation
relies on uncertain assumptions based on the Si and Al speciation, and can only indicate
relative differences.
The sulfur concentration was significantly greater for the concentrate sample compared with
the tailing (Table A5), suggesting greater exposure of sulfide mineral surfaces (in particular,
galena) in the concentrate. This indicated that the tailing sample was more oxidised (that
is, less sulfur present). This assertion is further supported by an examination of the sulfur
surface to bulk ratio (the bulk solids assays are provided in Table A2), which show that the
concentrate ratio was 0.39, almost double that of the tailing (0.20).
The surface concentrations of lead, zinc and iron were generally higher for the concentrate
sample compared with the tailing (Table A5).
Argon ion etching generally increased the surface content of sulfur, lead, zinc and iron in
both cases. However, the percentage increase in the surface concentration of these elements
was usually greater for the concentrate sample, indicating that the oxidation layers were
comparatively thin on particles contained in the concentrate compared to those in the tailing.
The XPS spectra for sulfur and lead appear in Figures A6 and A7, respectively. These
figures contain the spectra for both the ‘as received’ and deslimed samples of lead rougher
concentrate and tailing. The spectra have been charge corrected to allow better comparison.
The lead rougher concentrates in the original spectra were charge shifted by 0.8 eV for the ‘as
received’ sample, and 0.6 eV for the deslimed sample. The lead rougher tailing samples were
charge shifted by a greater amount, that is 2.2 eV for the ‘as received’ sample, and 3.7 eV for
the deslimed sample. Charge shifting indicates that the surfaces being examined have some
insulating characteristics. In this case, the tailing sample was charge shifted considerably more
than the concentrate, which suggested that the particles in the tailing were more insulating
than those in the concentrate. The insulating characteristics observed in the tailing sample can
be attributed to both the higher proportion of insulating, non-sulfide gangue present, and the
propensity for the sulfide minerals to be more oxidised.
As the system under investigation is a mixture of many minerals, the assignment of specific
binding energies corresponding to known species is difficult. For example, the uncharged S
(2p3/2) signal for galena occurs at 160.4 eV (Buckley and Woods, 1984), 161.6 eV for sphalerite
(Buckley, Wouterlood and Woods, 1989), and at 161.1 eV for pyrite (Buckley and Woods,
1987). Therefore, the surface sulfide peak develops into a relatively broad emission region
accounting for the various sulfide species present (for example, galena, sphalerite, pyrite,
chalcopyrite, arsenopyrite etc).
It is immediately apparent from Figure A6 that there were differences in the spectra between
the ‘as received’ and deslimed samples. The ‘as received’ samples (for both the concentrate
Arbitrary units
(2) Pb rougher concentrate - "as received",
etched surface
(3) Pb rougher tailing - "as received", original
(2) surface
(4) Pb rougher tailing - "as received", etched
surface
(3)
(4)
FeSO4 M1-xS
PbSO4 PbX2 PbS
ZnS FeS2
176.0 174.0 172.0 170.0 168.0 166.0 164.0 162.0 160.0 158.0
Binding energy, eV
FIG A6 – Sulfur 2p spectra for the lead rougher concentrate and tailing process
streams comparing the original and etched surfaces of the ‘as received’ samples.
and tailing) exhibited a sulfate signal at about 168 eV (Buckley and Woods, 1984, 1987)3,
compared with the sulfate-free spectra of the deslimed samples. The variations in the sulfide
peak (nominally 160 to 165 eV) were only slight, with an increase in the intensity of the sulfide
signal and small reductions in the contribution made by the high binding energy shoulders.
Hence, the desliming procedure resulted in the removal of sulfate, but left the remainder of
the surface sulfur species reasonably intact.
A comparison between the lead rougher concentrate and tailing spectra revealed that
significant differences existed between the two process streams. For the ‘as received’ samples,
it is apparent that the sulfate present on the surfaces of particles from the concentrate was
considerably less than that for the tailing. This is illustrated by examining the sulfide:sulfate
peak height ratio. The ratio for the concentrate was 1.6, while the ratio for the tailing was
1.2. When these surfaces were etched, the sulfate present on the concentrate was reduced
considerably, indicating that the sulfate species formed only a thin oxidation layer on the
surface of these particles. Conversely, the sulfate signal remained strong for the tailing sample
after etching, suggesting that the oxidation layers were relatively thick (Figure A6).
Examination of the sulfide region of the S(2p) spectra revealed that the signal for both
concentrate and tailing samples had strong sulfide signals corresponding to galena, at
160.4 eV (Buckley and Woods, 1984), and lead xanthate at 162.1 eV (Laajalehto et al, 1991).
A relatively small high binding energy shoulder, at 163 eV, represented the metal deficient
sulfide species (Fornasiero et al, 1994). The tailing sample also exhibited a very broad sulfide
signal, which contained galena at 160.4 eV (Buckley and Woods, 1984), sphalerite at 161.6 eV
(Buckley, Wouterlood and Woods, 1989), pyrite at 161.1 eV (Buckley and Woods, 1987), and
lead xanthate at 162.1 eV (Laajalehto et al, 1991).
3. The sulfate signal, for both the lead rougher concentrate and tailing samples, appears to occur at a binding energy considerably
greater than that reported by Buckley and Woods (1984, 1987), ie for galena 167.9 eV, and for pyrite 168.5 eV. This is due to the
insulating properties of sulfate species. That is, the sulfate signal is charge shifted to higher binding energies than expected.
The difference between the ‘as received’ and deslimed lead spectra for the concentrate
sample were only slight (Figure A7). However, the tailing spectra appeared to be very
different. The third peak, at 140.0 eV, between the Pb(4f5/2) and the Pb(4f7/2) signals for
the deslimed sample, corresponded to the Zn(3s) peak. This peak interferes with the Pb(4f)
spectra particularly when the surface concentration of lead is low. It is interesting to note that
desliming the ‘as received’ tailing sample removed sufficient lead species from the pulp to
allow the Zn(3s) peak to interfere with the Pb(4f) spectra. This suggested that some of the lead
oxidation products detected in the ‘as received’ sample were present on the surface of the
minerals as precipitated species formed during the drying process, and were loosely attached
to particles, easily removed by decantation.
There were obvious differences between the concentrate and tailing spectra. The concentrate
sample displayed very sharp, narrow lead spectra, at 137.4 eV, which corresponded to galena
(Buckley and Woods, 1984). A small high binding energy shoulder at 138.4 eV was also
detected, which corresponds to lead hydroxide (Buckley and Woods, 1984). Etching removed
all traces of the lead hydroxide high bonding energy shoulder, implying that these oxidation
layers were comparatively thin. The tailing spectra, however, was broad, with a galena signal
at 137.4 eV (Buckley and Woods, 1984), lead xanthate at 138.1 eV (Laajalehto et al, 1991), lead
hydroxide at 138.4 eV (Buckley and Woods, 1984), and lead sulfate at 139.2 eV (Buckley and
Woods, 1984) detected on the surfaces. Etching did little to sharpen the lead signal; therefore,
it can be assumed that the surface layers produced by the galena oxidation products were
considerably thicker.
Unfortunately, the XPS analysis, while providing information about the surface oxidation
state of particles in the lead rougher concentrate and tailing, does not yield data about the
location of the oxidation products. That is, where do the galena oxidation products occur in
the concentrate sample? Do the lead oxidation products remain on the galena only, or do they
precipitate on to other minerals? To answer some of these questions a number of experiments
were conducted on the lead rougher concentrate and tailing samples using more particle-
specific surface analysis techniques.
(1)
Arbitrary units
PbS
(4) Zn(3s)
PbSO4
Pb(OH)2
PbX2
154.0 152.0 150.0 148.0 146.0 144.0 142.0 140.0 138.0 136.0 134.0
Binding energy, eV
FIG A7 – Lead 4f spectra for lead rougher concentrate and tailing process streams.
TABLE A7
Average atomic concentrations of elements detected on the surfaces of galena, sphalerite
and pyrite grains examined from the lead rougher concentrate and tailing process streams.
Mineral/process stream Element, atomic %
C O S Pb Zn Fe
Galena
Lead rougher concentrate 20.1 7.0 10.0 57.6 0.0 5.3
Lead rougher tailing - - - - - -
Sphalerite
Lead rougher concentrate 16.4 10.4 32.6 6.1 34.5 0.0
Lead rougher tailing - - - - - -
Pyrite
Lead rougher concentrate 23.5 5.5 43.2 0.0 0.0 27.8
Lead rougher tailing 21.7 23.6 28.2 3.0 0.0 23.4
determined, it was not possible to obtain any meaningful data about the surface composition
of the tailing sphalerite due to charging. However, the fact that the tailing sphalerite was charging
and the concentrate sphalerite did not, suggested strongly that the surface composition of sphalerite
from the two sources was significantly different.
The concentrate sphalerite composition was mainly zinc, sulfur and carbon, with minor
quantities of oxygen and lead (Table A7). No iron was detected on the surface of the sphalerite.
Again, several comments are worth noting:
•• Sulfur was close to its stoichiometric value (for zinc sulfide, 32.92 weight per cent), but
the zinc content was considerably below the expected value (67.08 weight per cent). This
suggested that the dissolution of sphalerite, in this case, was through the removal of zinc
ions from the surface, while leaving the sulfur intact. It is assumed, therefore, that the
dissolution process leaves a zinc deficient sulfide surface.
•• The surface carbon was present as both hydrocarbon contamination and residual collector.
•• The presence of lead (6.1 atomic per cent) on the surface was of great interest, and may be
responsible for the inadvertent activation of sphalerite, and subsequent recovery into the
lead rougher concentrate.
•• The oxygen present was most likely the result of lead hydroxide species on the surface.
Particles of pyrite were detected and examined in both the concentrate and tailing samples,
so a comparison was possible. Carbon, oxygen, sulfur and iron were detected on the surface
of both samples; however, lead was found on pyrite particles in the tailing only (Table A7).
There were apparent differences in the surface composition of the pyrite grains:
•• The oxygen content of the pyrite in the concentrate was significantly lower (5.5 atomic per
cent) compared to the tailing (23.6 atomic per cent). This indicated that the pyrite in the
concentrate was considerably less oxidised than that reporting to the tailing.
•• The sulfur concentration of the pyrite in the concentrate (43.2 atomic per cent) was
dramatically greater than that of the tailing pyrite (28.2 atomic per cent). This suggested
that the surfaces of the pyrite in the concentrate were sulfur rich, which was probably the
main contributor to its recovery into the concentrate.
•• Carbon and iron levels were similar for the pyrite from both sources.
TABLE A8
Surface species detected on galena from the lead rougher concentrate and tailing process streams using time-of-flight
secondary ion mass spectroscopy (ToF-SIMS). The relative intensity of each element is expressed as strong (s), medium (m)
or weak (w). Note: gallium, from the primary ion beam, and indium, from the sample mount, were also detecteda.
Stream/particle Ion
Pb Zn Cu Fe K Na Ca Si Al C2H3 C3H7
Lead rougher concentrate
1 s w m w s m s m m s s
2 s w w w s w m m w m s
3 s w w w s w w w w m m
4 s w w w w w w w w w w
Lead rougher tailing
1 s w - s s w m s w w m
2 s w - s m w w m w w w
3 s w - m m w m m w w w
4 s m - m s w s m w w w
a. During this analysis, several sphalerite grains from both the lead rougher concentrate and tailing process streams were examined. It was
interesting to note that the sphalerite surfaces in the rougher concentrate had elevated levels of lead, potassium and hydrocarbon
compared to those present in the tailing. These data, although only qualitative, would suggest that sphalerite recovery into the lead rougher
concentrate may be the result of inadvertent lead activation.
A B
Ga Ga
FIG A8 – Time-of-flight secondary ion mass spectroscopy (ToF-SIMS) chemical maps of the lead rougher concentrate:
(A) lead; and (B) copper. The bright areas represent regions of high concentration for each element (Ga – galena).
intensities were weaker for the lead rougher concentrate sample (Table A8), than the lead
rougher tailing.
The intensity data (Table A8) suggested that particles in the lead rougher concentrate
sample had a high concentration of copper associated with their surfaces, while copper was
not detected on the surfaces of particles in the lead rougher tailing sample. An examination of
the ToF-SIMS chemical map of copper for the lead rougher concentrate (Figure A8) indicated
that the copper ions were only weakly associated with the galena grains, and were more
strongly associated with other species within the sample, possibly chalcopyrite or copper
activated sphalerite.
From Table A8 it is suggested that galena particles in the lead rougher concentrate sample
had a higher concentration of collector (potassium amyl xanthate) associated with their
surfaces than particles in the lead rougher tailing sample. An examination of Figure A9,
the ToF-SIMS chemical maps for potassium (Figure A9b) and hydrocarbon (Figure A9c),
A B
Ga
Ga
C D
Ga
Ga
FIG A9 – Time-of-flight secondary ion mass spectroscopy (ToF-SIMS) chemical maps for the lead
rougher concentrate: (A) lead; (B) potassium; (C) hydrocarbon (C3H7); and (D) iron. The bright areas
represent regions of high concentration for that particular element (Ga – galena).
indicates that the highest intensities for potassium and hydrocarbon correspond to the
highest intensities for lead. This match suggested that the galena particles in the lead rougher
concentrate were strongly associated with collector. The same comparison, when made for
the lead rougher tailing sample (Figure A10), revealed that the intensity of hydrocarbon on
the surfaces of galena particles was lower than that observed for the lead rougher concentrate.
This observation indicates that the galena in the lead rougher tailing may have had insufficient
collector on its surface to produce the required level of hydrophobicity for flotation.
One possible reason for the lower collector coverage of the galena in the lead rougher tailing
is the higher concentrations of iron and calcium on the surfaces of particles in this process
stream compared with the lead rougher concentrate (Table A8). An examination of the ToF-
SIMS chemical maps for iron (Figures A9d and A10d) revealed that the galena particles
contained within the lead rougher tailing sample exhibited higher concentrations of iron on
their surfaces than the galena contained within the lead rougher concentrate sample. While
not shown here, the same observation holds true for calcium. These observations suggest
that galena in the lead rougher tailing process stream is coated with hydrophilic species. For
example, the iron occurs as ‘iron hydroxides’ emanating from the oxidation of pyrite and
the corrosion of the grinding media (high carbon, low alloy steel). It is suspected that the
hydrophilic layers coating the galena in the tailing may have been responsible for blocking the
collector, and preventing adsorption on to these particles, thereby reducing their floatability.
Summary of findings
The chemical characteristics examined showed, for both the solution (that is, free ions in
solution and EDTA analysis) and surface analysis (that is, XPS, SAM and ToF-SIMS), the
abundance of lead oxidation products was significantly greater in the lead rougher tailing
compared to that observed in the concentrate. Further, the XPS analysis indicated that the
oxidation products formed comparatively thick layers on the particles present in the lead
rougher tailing. The ToF-SIMS analysis was more particle specific, and while qualitative in
A B
Ga Ga
C D
Ga Ga
FIG A10 – Time-of-flight secondary ion mass spectroscopy (ToF-SIMS) chemical maps for the
lead rougher tailing: (A) lead; (B) potassium; (C) hydrocarbon (C3H7); and (D) iron. The bright areas
represent regions of high concentration for that particular element (Ga – galena).
nature suggested that the galena in the lead rougher concentrate was hydrophobic because of
its strong association with hydrocarbon fragments, derived from the collector. The evidence
also suggests that the galena may have been copper activated. The surface composition of
the galena in the lead rougher tailing was significantly different. The data suggested that this
galena had insufficient collector coverage to render it hydrophobic; and significant coverage
of hydrophilic species, such as ‘iron hydroxide’. Both observations would severely hamper
the galena flotation response.
A secondary observation made from both the SAM and ToF-SIMS analysis indicated
appreciable levels of lead on the surfaces of sphalerite reporting to the lead rougher concentrate.
This suggested strongly that sphalerite recovery into the lead rougher concentrate was due to
inadvertent activation by lead.
The positive changes to pulp chemistry and metallurgical response observed with the
change in grinding media in the primary tower mill work provided the impetus to commenced
examining the impact of high chrome grinding media added to the lead regrind circuit on lead
metallurgy. The initial study conducted on lead rougher concentrate, comparing forged steel
media with a variety of high chrome (above 15 per cent chrome) alloys showed forged steel
produced a superior lead grade/recovery curve, with low overall zinc losses when compared
to the high chrome alloys. That is, as the chrome content of the grinding media increased,
and the mill pulp chemical environment became more oxidising, the lead flotation rate,
concentrate grade and recovery noticeably decreased. The deterioration in lead metallurgy
appeared to be related to the increased floatability of sphalerite as the chrome content of the
grinding media increased (Kinal, Greet and Goode, 2009).
A series of laboratory tests were completed to develop an understanding of this behaviour,
and determine which alloy was best to trial in the plant.
Laboratory studies
A sample of fresh lead regrind feed from South 32’s Cannington Mine in north-western
Queensland, Australia was tested in the laboratory. The results showed forged steel produced
a superior lead grade/recovery curve, with low overall losses of zinc compared to the three
high chrome alloys tested. As the chrome content of the grinding media increased, and
the mill pulp chemistry environment became more oxidising (Table A9), the lead flotation
rate, grades and recoveries noticeably reduced. A main reason for the drop in lead flotation
performance was found to be poor selectivity for galena against sphalerite and non-sulfide
gangue minerals.
As the liberation characteristics were nominally the same for both media types, the
conclusion was drawn that it must have been the change in pulp chemistry affecting
selectivity. Surface analysis (ToF-SIMS) of the lead concentrate from a forged and high
chrome test showed that the sphalerite reporting to the lead concentrate in the high chrome
case had more surface coverage of chromium and zinc than the sphalerite particles prepared
using forged steel, which had a higher surface coverage of iron, but also lead and silver
(Figure A11).
The presence of more activating ions (silver and lead) on sphalerite surfaces for the forged
steel case is rather perplexing, as conventional wisdom would suggest that sphalerite
recoveries in this instance should be elevated. However, they are not. Sphalerite recoveries
after grinding with high chrome media were considerably greater than those observed for
forged steel grinding. This suggests that the presence of iron species on the surfaces of
sphalerite tend to control its floatability. That is, when grinding with forged steel the iron
corrosion products tend to react with the sphalerite’s surface more rapidly than the activating
species (for example, silver and lead), and thereby depress its flotation. When these corrosion
TABLE A9
Pulp chemistry data for laboratory lead cleaner tests comparing fresh
lead regrind feed ground with forged steel and high chrome media.
Media Mill discharge Flotation feed
Eh, mV (SHE) pH DO (ppm) Eh (mV) (SHE) pH DO (ppm) EDTA Fe (%)
Forged -62 7.9 0.0 129 7.6 0.0 1.4
Hi Cr 199 8.4 3.9 195 8.4 2.9 1.0
0.5
con(326)
+SIMS
0.4 con(327)
0.3
Intensity
0.2
0.1
0.0
Mg Al Si Ca Cr Fe Cu Zn FeOH Ag Ag Pb PbOH
con(326) 0.0455 0.026 0.0556 0.0373 0.0022 0.2917 0.0411 0.1287 0.0058 0.0245 0.0227 0.308 0.011
con(327) 0.0776 0.0485 0.0774 0.0599 0.0068 0.2195 0.055 0.1508 0.0041 0.0106 0.0099 0.2689 0.011
(+) Fragment
0.5
con(326)
-SIMS
0.4 con(327)
0.3
Intensity
0.2
0.1
0.0
CH O OH S SO3 EX IPX IBX AX
con(326) 0.16 0.36 0.25 0.21 0.0057 0.0011 0.0003 0.0004 0.0003
con(327) 0.17 0.38 0.27 0.17 0.0035 0.0009 0.0002 0.0004 0.0002
(+) Fragment
FIG A11 – Statistically evaluated (+) and (-) mass spectra for sphalerite in concentrates during
tests 326 (Forged) and 327 (High Cr). Confidence intervals were calculated for P = 95 per cent.
products are significantly reduced by grinding with high chrome media, the activating species
are now able to react with the surfaces of the sphalerite leading to significantly higher zinc
recoveries to lead concentrate.
Given this surface analysis knowledge, diagnostic tests were conducted on fresh lead
regrind feed ground with high chrome grinding media with different addition of iron sulfate
powder (Fe2(SO4)3) to determine if this would simulate the forged media test (Figure A12).
These data indicated that with increasing iron sulfate addition the lead grade/recovery
improved. That is, the lead concentrate grade at a specific lead recovery increased with
increasing iron sulfate addition due to better electivity for galena against sphalerite. This
suggested a low chrome alloy should restore the lead metallurgy. A low chrome ball was
sourced and tested in the laboratory.
The difference in grinding pulp chemistry between the two media types was noticeably
smaller when the low chrome media was used in place of a high chrome alloy (Table A10).
However, there was still a slight increase in pulp potential that would have an impact on
flotation performance. In fact, Figure A13 shows that the low chrome media had a beneficial
impact on lead cleaner flotation with a small increase in grade and recovery, compared to
forged steel.
80.0
70.0
60.0
Grade - Pb(%)
50.0
40.0
30.0
Forged Std
20.0
HiCr 1 g Iron Sulfate
10.0 HiCr 5 g Iron Sulfate
HiCr 10 g Iron Sulfate
0.0
0.0 20.0 40.0 60.0 80.0 100.0
Recovery - Pb(% )
FIG A12 – Lead grade/recovery curves showing the effect of increasing
the addition rate of iron sulfate powder to the high chrome grind.
TABLE A10
Pulp chemistry data for laboratory lead cleaner tests comparing fresh
lead regrind feed ground with forged steel and low chrome media.
Media Mill discharge Flotation feed
Eh, mV (SHE) pH DO (ppm) Eh (mV) (SHE) pH DO (ppm) EDTA Fe (%)
Forged -84 8.26 0.11 -110 8.34 0.00 8.26
LoCr -32 7.93 0.12 -69 8.13 0.00 7.30
60.0
55.0
50.0
Grade - Pb(%)
45.0
40.0
Plant trial
Based on these laboratory results a plant trial was arranged. Given that the first cleaner circuit
operates with two parallel grinding/flotation trains, a direct comparative trial was possible.
The pulp chemistry data indicated that the shift from cylpebs to a low chrome ball did result
in a subtle increase in the lead regrind tower mill discharge Eh, but all other pulp chemical
parameters were more or less the same. Metallurgically, the shift to a low chrome ball resulted
in an increase in lead grade through improved selectivity for galena against sphalerite, with
respect to first cleaner feed (Kinal, Greet and Goode, 2009).
Summary
The application of a surface analysis technique, in this case ToF-SIMS, helped identify the
issue of poor selectivity for galena against sphalerite in the laboratory. Knowing this valuable
piece of information led to the identification of a potential solution (ie a low chrome media),
which was tested in the laboratory before successfully transferring this result to the plant.
TABLE A11
Magotteaux Mill® discharge pulp chemistry.
Media pH Eh, mV (SHE) DO (ppm) Temp (°C) EDTA Cu (%) EDTA Fe (%)
Forged 9.9 134 0.0 35.7 0.77 0.99
HiCr 9.9 179 0.0 36.0 0.35 0.50
Ceramic 9.9 189 0.0 35.5 0.38 0.04
media corrosion, exhibited a marked decrease between the forged steel (0.99 per cent) and
the high chrome alloy (down to 0.50 per cent). The EDTA extractable iron decreased further
to 0.04 per cent when grinding with ceramic media. In essence, as the iron content of the
grinding media was reduced the EDTA extractable iron (or amount of corrosion products
generated) decreased.
The pulp chemistry data showed that moving from the electrochemically reactive forged
steel grinding media to high chrome and then ceramic resulted in the mill discharge pulp
potential becoming more oxidising and the level of corrosion (EDTA extractable iron) to
decrease.
Surface chemistry
The first surface analytical technique applied to the -38/+20 micron sample of chalcocite was
XPS. Generally, XPS is a bulk surface analysis taking an average surface composition of all
particles contained within an area of 400 × 700 microns. This is not an analysis of individual
particles. Fortunately, the sample examined was predominantly chalcocite so the analysis is
a reasonable approximation.
The atomic surface concentration of copper, sulfur, iron and oxygen are provided in
Table A12. The data show that as the iron content of the grinding media decreased the
exposure of copper and sulfur on the surfaces of the chalcocite particles increased. For
example, the atomic concentration of copper increased from 20.5 atomic per cent for forged
steel to 31.0 for ceramic media. This also corresponded to a decrease in the oxygen content
of the surface.
Remarkably, the surface iron concentration increased as the iron content of the grinding
media decreased. This may be related to changes in the surface chemistry of the trace levels
of chalcopyrite, bornite and pyrite contained in this sample of natural chalcocite. That is, as
media corrosion is reduced, moving from forged steel to high chrome to ceramic grinding
media, the mineral surfaces should be cleaner. This means that not only are the surfaces
of chalcocite cleaner (that is, increased exposure of copper and sulfur), the surfaces of
TABLE A12
Surface concentrations of copper, sulfur, iron and oxygen measured using X-ray photoelectron spectroscopy
(XPS) (normalised to remove the effects of carbon, magnesium, silica, sodium and nitrogen).
Media Atomic concentration
Cu S Fe O
Forged 20.5 15.3 4.0 60.2
HiCr 27.1 16.6 5.9 50.4
Ceramic 31.0 19.5 5.3 44.2
chalcopyrite, bornite and pyrite particles contained within the sample would also have
less contamination. Therefore, for these minerals (all iron bearing) not only are the surface
concentrations of copper and sulfur increased, so is the iron.
The XPS results only partially answer the question: are particles that are ground with an
electrochemically active grinding media surface contaminated by grinding media corrosion
products? The next level of analysis should examine individual particles to determine their
surface composition. The first such technique used was ToF-SIMS. Unfortunately, ToF-SIMS
while being extremely surface sensitive is not quantitative.
Figure A14 shows the positive SIMS results from the analysis of eight particles in the
-38/+20 micron size fraction prepared by grinding with forged steel, high chrome and ceramic
grinding media. The data does show that the surface iron content does decrease as the iron
content of the grinding media decreased. The data also shows that the particles prepared
using forged steel grinding media had a higher affinity for calcium, magnesium, aluminium
and silicon.
The ToF-SIMS analysis starts to provide evidence that corrosion products for electrochemically
active grinding media do impact the surface chemistry of particles. Unfortunately, as ToF-
SIMS is not quantitative, further work was completed using SAM. The SAM analysis allowed
the examination of individual grains of chalcocite, and obtained the surface concentration of
these particles.
Figure A15 provides an example of the SAM spectra obtained during this experiment. The
atomic concentration data collected are provided in Table A13. These data clearly show that
both the copper and sulfur surface concentration increase as the iron content of the grinding
media decreased. There is a corresponding decrease in the iron and oxygen concentrations.
0.6
0.5
ToF-SIMS counts (normalised)
0.4
0.3
0.2
0.1
0.0
Na Mg Al Si K Ca Fe Cu CuH2SO
FIG A14 – Time-of-flight secondary ion mass spectroscopy (ToF-SIMS) results for positive SIMS.
dN(E)
Fe2
O1
C1
Cu1
S1
30 130 230 330 430 530 630 730 830 930 1030
Kinetic Energy (eV)
FIG A15 – Scanning Auger microscopy spectra for chalcocite ground with forged steel.
TABLE A13
Surface concentrations of copper, sulfur, iron and oxygen measured using scanning Auger microscopy (SAM).
Media Atomic concentration
Cu S Fe O
Forged 42.7 29.1 13.5 14.7
HiCr 53.1 37.3 4.6 5.0
Ceramic 59.7 39.1 0.3 0.9
Summary
Changing the grinding media from forged steel to high chrome to ceramic when grinding
chalcocite did alter the pulp chemistry of the laboratory mill discharge to more oxidising Eh
values and lower EDTA extractable iron concentrations (ie lower media corrosion).
The XPS (bulk) surface analysis did show increased exposure of copper and sulfur as the
iron content of the grinding media decreased, which suggests that the chalcocite particles
are cleaner.
The ToF-SIMS surface analysis (individual particles) did show a decrease in the iron
concentration on the surfaces of the chalcocite particles. However, as ToF-SIMS is qualitative
these results are a good indication but further quantitative data is needed.
The SAM (individual particle) surface analysis also showed increased exposure of copper
and sulfur as the iron content of the grinding media decreased. In this analysis, it was
evident that the change in grinding media did result in a reduction in iron species on the
surface of chalcocite.
The pulp chemical data, particularly the EDTA information, is corroborated by the three
surface analysis techniques. That is, more electrochemically active grinding media (ie forged
steel) results in higher percentages of EDTA extractable iron, and reduced exposure of copper
and sulfur, but elevated iron concentration on their surfaces. Shifting to media with lower
iron contents (ie high chrome and ceramic) the EDTA extractable iron decreases as does the
concentration of iron on the particle surfaces.
REFERENCES
AMIRA, 1991. P260 – Iron sulphide interactions and their influence on sulphide mineral flotation,
Volume 4: Aberfoyle Resources Limited – Hellyer, prepared by the Particle and Surface Technology
Research Group, University of South Australia (Australian Mineral Industries Research Association:
Melbourne).
Blesa, M A, Rueda, E H and Grassi, E B, 1985. Adsorption and dissolution in the system geothite/
aqueous EDTA, Journal of Colloid and Interface Science, 106(1):243–247.
Buckley, A N and Woods, R, 1984. An X-ray photoelectron spectroscopic study of the oxidation of
galena, Applied Surface Science, 17:401–414.
Buckley, A N and Woods, R, 1987. The surface oxidation of pyrite, Applied Surface Science, 27:347–352.
Buckley, A N, Wouterlood, H J and Woods, R, 1989. The surface composition of natural sphalerite
under oxidative leaching conditions, Hydrometallurgy, 22:39–56.
Fornasiero, D, Li, F, Ralston, J and Smart, R StC, 1994. Oxidation of galena surfaces: I. X-ray
photoelectron spectroscopic and dissolution kinetic studies, Journal of Colloid and Interfacial Science,
164:333–344.
Gardner, J R and Woods, R, 1979. A study of the surface oxidation of galena using cyclic voltametry,
Journal of Electroanalytical Chemistry, 100:447–459.
Grano, S R, Ralston, J and Johnson, N W, 1988. Characterisation and treatment of heavy medium plant
slimes in the Mount Isa Mines lead/zinc concentrator, Minerals Engineering, 1:137–150.
Grano, S R, Wong, P L M, Skinner, W, Johnson, N W and Ralston, J, 1995. Detection and control
of calcium sulphate precipitation in the lead circuit of the Hilton Concentrator of Mount Isa Mines
Limited, Australia, in Proceedings XIX International Mineral Processing Congress, volume 3, pp 171–179
(Society for Mining, Metallurgy, and Exploration: Littleton).
Greet, C J, 2002. Galena flotation: Hellyer Mine case study. PhD thesis, University of South Australia.
Greet, C J and Smart, R StC, 2002. Diagnostic leaching of galena and its oxidation products with EDTA,
Minerals Engineering, 15:515–522.
Hamilton, I C and Woods, R, 1986. Investigation of electrochemical properties of lead ethyl xanthate by
linear potential sweep voltammetry, Langmuir, 2:770–773.
Hiskey, J B and Schlitt, W J, 1982. Chapter 7: Aqueous oxidation of pyrite, interfacing technologies
in solution mining, in Proceedings Second SEM-SPE International Solution Mining Symposium (eds:
W J Schlitt and J B Hiskey), pp 55–74 (American Institute of Mining, Metallurgical, and Petroleum
Engineers: New York).
Johnson, N W, 1988. Application of electrochemical concepts to four sulphide flotation separations, in
Proceedings Electrochemistry in Mineral and Metal Processing II, pp 131–149.
Kant, C, Rao, S R and Finch, J A, 1994. Distribution of surface metal ions among the products of
chalcopyrite flotation, Minerals Engineering, 7(7):905.
Kinal, J, Greet, C and Goode, I, 2009. The effect of grinding media on zinc depression in a lead cleaner
circuit, Minerals Engineering, 22:759–765.
Laajalehto, K, Nowak, P, Pominowski, A and Suoninen, E, 1991. Xanthate adsorption at PbS/aqueous
interfaces: comparison of XPS, infrared and electrochemical results, Colloids and Surfaces, 57:319–333.
Moore, J J, 1981. Chemical Metallurgy (Butterworths: London).
Natarajan, K A and Iwasaki, I, 1973. Practical implications of Eh measurements in sulphide flotation
circuits, AIME Transactions, 256:323–328.
Pang, J and Chander, S, 1992. The effect of EDTA on collectorless flotation of pyrite, in Proceedings
International Symposium on Electrochemistry in Mineral and Metal Processing III (eds: R Woods and
P E Richardson), p 221.
CHAPTER 10
Laboratory flotation testing – an essential
tool for ore characterisation
Kym Runge
ABSTRACT
The properties of the ore have a significant impact on the ultimate grade and recovery
achievable in a flotation circuit. To maximise recovery or yield, the proportion of the
valuable mineral that is floatable should be maximised. To maximise the concentrate quality
or grade, the flotation rate of the valuable mineral should be significantly higher than the
flotation rates of the gangue minerals in the ore. This is often referred to as ‘maximising the
selectivity’.
Laboratory batch flotation, in which a 1–2 kg sample of ore is floated using a standard set
of operating conditions, is the standard technique used to assess ore floatability. These tests
have been performed extensively to determine the ultimate flotation design, screen potential
reagents, determine optimum feed grind size and predict the change in performance of
different ore types. These tests have traditionally been performed using the feed ore sample
to the process, but they are increasingly being used to characterise the floatability of internal
flotation circuit streams and ‘map’ how the floatability is changing around a flotation circuit
and across different processes.
Key ore floatability parameters (flotation rate constants and proportion of mineral in
different floatability classes) can also be derived from the laboratory test flotation response.
The values of these ore floatability parameters can be evaluated and compared to rank
different ore types or reagent suites. These parameters are also suitable for predicting,
through mathematical modelling, the grade–recovery relationship achievable in a multistage
flotation circuit.
There are various graphical and modelling techniques that have been developed to interpret
laboratory batch flotation test results. In this chapter, the recommended procedure for
performing these laboratory flotation tests and the various graphical and modelling analysis
methods used to interpret these test results will be outlined.
INTRODUCTION
To achieve separation in the flotation process, those particles containing the valuable mineral
need to float and they need to float more quickly than those containing predominantly gangue
minerals. The particle size, degree of liberation of the minerals and the hydrophobicity of the
mineral surface will in large dictate a mineral’s recovery rate. Grinding and chemical addition
are weapons in the metallurgist’s armoury used to modify particle properties for a particular
ore type in an attempt to achieve maximum separation.
A vast array of techniques can be used to predict the optimum grind size and chemical
suite to use for a particular ore. Ultimately, there is always a need for small-scale testing to
validate the results of these analyses or to explore, screen or prioritise different options. This
is where laboratory batch flotation testing and the various techniques that can be used to
analyse these batch test results are useful. They are relatively easy and inexpensive to perform
in comparison to detailed liberation or chemical speciation analyses and they go some way
to replicating the process conditions that will be present in the full-scale processing plant.
The aim of this chapter is to outline how these ore characterisation tests should be performed
and demonstrate the various graphical and modelling techniques that can be used to
interpret the results produced from these tests.
different operators. When using a cell where the impeller is top mounted, reproducibility can
be improved by using two fixed-depth scrapers with rubber glued to the end of each scraper.
This rubber moves around the impeller shaft as the scraper is brought forward from the
back of the cell to the cell lip, thus enabling the entire froth phase to be removed during each
scrape. Whatever cell is used to perform the test work, it is important that a great deal of care
is taken to keep the procedure as consistent as possible.
A typical test involves washing an ore pulp sample into the laboratory cell. Make-up water
is added to achieve the desired volume. The impeller is started and set to the desired speed.
Frother, collector and depressant (if required) are then added and appropriate conditioning
time is allowed to elapse. The order by which these reagents are added is important and
should be tailored to achieve the desired surface chemistry. Depressant, for example, often
needs to be added prior to the collector to stop collector attaching to gangue mineral surfaces.
The appropriate reagent addition rates are best determined from reagent optimisation testing.
At the instant the air flowmeter is set to the appropriate air rate, the timing of the test
commences. The concentrate produced from the cell flows and/or is scraped into trays and
at designated time intervals the trays are removed and replaced. Water is continually added
to the cell to maintain the desired froth level. The test continues until the froth is barren
and no more particles are reporting to the froth in significant quantities. It is recommended
that one more concentrate be collected after this point to ensure the end point of the test can
be established.
All concentrate samples and the final tailing sample are weighed wet and dry and assayed
for the elements of interest. These results are used to calculate the rate of mineral recovery as
a function of time, usually on a cumulative basis. A typical cumulative mineral recovery curve
is depicted in Figure 2.
As an aside, the water in the concentrate versus the water in the flotation cell will be used
to estimate recovery by the entrainment mechanism. Thus, known quantities (or no) water
should be used to wash concentrate into the tray.
Samples to be tested can be collected directly from an operating flotation circuit, referred to as
‘hot samples’, or prepared in the laboratory – grinding to produce the appropriate particle size
or adding reagent to achieve the required pulp chemistry. The time between sample collection/
preparation and flotation should be minimised to prevent ageing or oxidation of particle
surfaces prior to testing.
100
Mineral Recovery (%)
80
60
40
20
0
0 2 4 6 8 10 12 14
Time (min)
FIG 2 – Typical mineral versus recovery time curve obtained from a laboratory batch test.
The weight or size of sample added to the laboratory flotation test should be chosen to
achieve the required pulp density. Pulp density is a variable that often affects batch flotation
test results and therefore comparative tests should be performed at the same pulp density.
When attempting to replicate plant performance it is recommended that one use the same
per cent solids as in the plant. When performing tests to assess or model ore floatability,
however, it is recommended that all samples be floated at a low per cent solids (that is, less than
ten per cent solids). The use of a low per cent solid minimises the risk of bubble saturation
or overloading. In tests where there is a high proportion of floatable solids performed using
a low air rate and at a high pulp density, bubble surfaces can become saturated, resulting in
low recovery rates. This condition should be avoided as the result is no longer a function of
just the ore, but the air rate used in the test as well.
The water used to dilute the samples and added during a test to maintain froth depth
should be as similar to that used (or to be used) in the plant as possible. Water properties (that
is, pH, Eh, temperature, dissolved oxygen and dissolved ions) will affect the composition of
reagent and deposition products on particle surfaces, which ultimately affects ore floatability.
Whenever possible, plant water should be used. When plant water is not available, synthetic
plant water created after an analysis of the cationic and anionic chemistry should be added.
Reagent additions may also need to be increased when using synthetic plant water because
of the lack of residual collector, which is often present in the recycled water used in a plant.
Frother (so long as it has no collecting properties) should also be added to all make-up water
in sufficient quantities to maintain a constant bubble size in the pulp and minimise bubble
breakage within the froth phase. For most frothers, a concentration of 20 ppm within the
make-up water will be sufficient for this purpose. Distinct additions of frother in the middle
of a test should be avoided as they can result in a spike in flotation recovery.
It is important that all tests are performed using the same air rate, impeller speed, froth
depth and froth scraping rate so that any change in a batch test result is a consequence of a
change in the ore characteristics rather than to any change in the environment in which it is
floated. The air rate chosen should be high to prevent bubble saturation and maintain a fluid
non-collapsing froth. Air rates of 10–18 L/min are usually appropriate for tests performed
using a 5 L cell. Impeller speed should be sufficient to keep all solids in suspension but not so
high that molecules loosely deposited on surfaces are stripped off. An impeller speed between
750 and 1000 rev/min is recommended for standard flotation tests, although higher speeds
may be required when processing very coarse material.
Ideally the froth should be removed the instant it is formed so that differences in results can
be attributed to changes in pulp flotation kinetics rather than to any change in the flotation
froth characteristics. As a practical compromise, it is recommended that 2 – 3 cm of froth
be maintained throughout the test and a scraper of sufficient depth be used to remove all
but 0.5 cm of this froth at regular time intervals. Make-up water should be added regularly to
maintain a constant froth depth as solids and water report to concentrate. A fast froth scraping
rate is desirable but one stroke every ten seconds is usually a speed that allows the execution
of other tasks (addition of water and changing of trays).
The time intervals over which each concentrate is collected should be chosen such that the
shape of the mineral recovery time curves can be well established for the particular operating
conditions chosen to do the tests. A test of a slow floating platinum ore, for instance, may require
45 minutes to an hour to reach completion and thus the first concentrate can be collected after
two minutes of flotation. A faster-floating copper porphyry flotation test, however, may finish
in five minutes and a first concentrate may need to be collected after 20 seconds.
Depending on one’s circumstances, some deviation may be required from the recommended
test conditions (for example, use of high per cent solids to obtain sufficient concentrate
samples for assay or samples left to sit for long periods). Tests should be performed to
establish whether any adverse effect arises from the proposed change.
A 100
90
80
16
70
Recovery (%)
60
8
50
40
4
30
20
10
0
0.001 0.01 0.1 1 10
60
50
Reflotation
40
30
20
10
0
0.001 0.01 0.1 1 10
-1
Flotation Rate (min )
FIG 3 – Recovery of particles with different flotation rates: (A) after four, eight and 16 minutes
of flotation, and (B) after one, two and three refloatation stages of eight minutes in duration.
(Figure 3b). This occurs because of the increase in the fast to slow ratio of particles in the
feed to subsequent stages. The net result of this improvement in separation efficiency is an
improvement in the concentrate grade or purity achievable at a particular mineral recovery.
Locked cycle batch flotation experiments provide a better insight into the expected grade
and recovery achievable from multistage processing of an ore. These tests are performed such
that they replicate, on a small scale, the full-scale flow sheet. To incorporate the effect of
recycle streams, they are often performed multiple times with the recycled streams from a
previous test added at the appropriate place in the subsequent test. Tests are performed until
the mass flow in all samples converges to a constant value. Figure 4 shows a schematic of
a locked cycle test performed to replicate a rougher/cleaner flotation circuit in which the
cleaner tail is recirculated back to the feed.
The disadvantage of these types of tests is that they are time consuming to perform and they
pertain to a single flow sheet arrangement. To investigate an alternative flow sheet or a change
in the residence time of a stage, another experiment needs to be performed.
A more time-effective option for estimating multistage circuit grade and recovery is to
model the data from a single-stage experiment (or a locked cycle test). This model can be
used to predict circuit grade and recovery achievable from different circuits mathematically.
The procedures to perform this type of analysis will be outlined later in the chapter.
New Feed
Recycled Tail
New Feed
Recycled Tail
Circuit Concentrate
40
20
0
0 2 4 6 8 10 12 14
Time (min)
FIG 5 – Cumulative mineral recovery curve with the important features denoted.
It is desirable that the valuable mineral in a sample exhibit fast flotation kinetics (steep
slope) and the proportion of non-floating mineral in the sample be minimised. The
proportion of non-floating valuable mineral is an indicator of the amount of mineral that
is either in particles too large or small for flotation, locked with non-floating particles or
with a hydrophilic surface due to insufficient collector coverage or depressant coatings.
Regardless of the reason, non-floating mineral is not recoverable in a flotation circuit and
should be minimised if circuit recovery is to be maximised.
Flotation is a separation process and thus it’s important that the valuable mineral is not only
recoverable but that it’s recoverable to a greater extent than the gangue mineral. Selectivity
curves are ideal for this type of evaluation. On these curves, the recovery of the valuable
mineral is plotted against the recovery of a gangue mineral of interest (Figure 6).
The 45 degree line on this plot represents the points at which no selectivity occurs – the
recovery of the valuable mineral is equal to the gangue mineral recovery and no separation
is achievable. The further to the right of this line the selectivity curve lies, the better. Different
selectivity curves should be created for each of the gangue minerals of interest as different
minerals often exhibit different degrees of selectivity with the valuable mineral.
The overall measure of selectivity in the batch flotation test is the grade (or purity) of the
concentrate produced. The cumulative concentrate grade versus cumulative recovery curves
(Figure 7) produced from different tests can be compared to determine those conditions that
result in the best selectivity with respect to all minerals.
These curves do not, however, give an estimate of the grade and recovery achievable in a
multiple stage flotation circuit. Locked cycle tests or modelling are better suited to this task.
These curves do give a good relative measure of selectivity. An improvement in selectivity in
these tests should result in a better grade at a particular recovery at full scale.
100
60
40
Increasing Selectivity
20
0
0 20 40 60 80 100
Valuable Mineral Recovery (% )
FIG 6 – Mineral selectivity curves with the important features denoted.
100
Improving Overall Selectivity
Cumulative Recovery to
80
Concentrate (%)
60
40
20
0
0 5 10 15 20
Cumulative Concentrate Grade (% )
FIG 7 – Cumulative recovery versus cumulative concentrate grade produced during a batch flotation experiment.
gangue recovery. This section outlines techniques that can be used to estimate the degree of
entrainment in a laboratory batch flotation test.
Entrainment is a consequence of particles suspended in the pulp phase following the
water into the froth and ultimately into the concentrate. The mass of a mineral entrained into
the concentrate is proportional to the water flow to concentrate and can be estimated using
Equation 1 (Johnson, 2005). Equation 2 is an alternative equation used to estimate entrainment.
It is a simplification of Equation 1 but is only valid at low water recovery values.
Fi = C f, i ~ i, pulp Fw (1)
where:
Fi is the flow of component i into the concentrate due to entrainment
Cf,i is the classification function for component i
ωi,pulp is the mass of component i to the mass of water in the pulp phase
Fw is the flow of water to the concentrate
The classification function (Cf,i) is a factor to account for the degree of drainage with respect
to the water flow and must be estimated to perform the calculation. It is known to fall between
zero and one; zero representing the condition of total drainage where there is no entrained
recovery of particles regardless of the water recovered, and one representing the condition
where there is no drainage and the concentration of particles in the water in the concentrate
(due to entrainment) is the same as in the pulp phase. Cf,i is known to be strongly dependent
on size with the coarser particles exhibiting a high degree of drainage (Cf,i = 0) and the fine
particles usually following the water phase (Cf,i = 1).
Thus, to calculate the recovery by entrainment in a batch test, the water to solids ratio
in the pulp, water flow to concentrate and an estimate of Cf,i is required. The water flow
to concentrate can be measured and the average water to solids ratio in each stage can be
estimated based on knowledge of the cell volume and the particle mass balance. The challenge
is to estimate the degree of drainage in the system.
If one of the minerals is known to be non-floating, an inversion of Equation 1 can be used to
calculate a Cf,i estimate for the other minerals in the ore. This, unfortunately, is not usually
the case with even the non-floating minerals locked to some extent with the valuable and
exhibiting floatability. Johnson (2005) published typical values of Cf,i for siliceous non-sulfide
gangue (Table 1). These values coupled with knowledge of the particle size distribution of a
particular sample can enable a reasonable estimate of Cf,i to be calculated.
Alternatively, Cf,i can be estimated using results from the batch test by recognising
that towards the end of an experiment, the predominant particle recovery mechanism is
entrainment. If one calculates the ratio of mineral to water in both the concentrate and
pulp and plots this value for each stage of a batch experiment (Figure 8), it usually decreases,
asymptoting at a particular value towards the end of the test. This value is an estimate of the
Cf,i value. The larger values at the beginning are because recovery is due to entrainment and
flotation in these earlier stages.
TABLE 1
Typical values of Cf,i (after Johnson, 2005).
Size fractions (μm)
-11 -16 +11 -23 +16 -33 +23 -44 +33
0.77 0.46 0.30 0.24 0.13
2.5
Gangue to Water Ratio in Concentrate
Ratio =
2.0 Gangue to Water Ratio in Pulp
1.5
Ratio
0.5
Cfi Estimate
0.0
0 5 10 15
Time (min)
FIG 8 – Cf,i calculated using Equation 1 versus time in a batch flotation test.
Once a Cf,i estimate has been obtained and the gangue entrainment flow has been calculated
(Equation 1), a cumulative entrainment recovery curve can be constructed that can be compared
to the overall recovery curve for that mineral (Figure 9). The difference in these two curves is
the proportion of the mineral recovered due to true flotation (attachment to bubbles).
70
Gangue Recovery (%)
60
50
Overall
40
Entrainment
30 Estimate
20
10
0
0 5 10 15
Time (min)
FIG 9 – Curves used to assess the predominant gangue recovery mechanism in a batch flotation test.
80 Rougher Tail
Scavenger Concentrate
60 Scavenger Tail
Cleaner Feed
Cleaner Concentrate
40
Cleaner Tail
Recleaner Feed
20
Recleaner Concentrate
Recleaner Tail
0 Cleaner Scavenger Tail
0 2 4 6 8
Time (min)
The final tailing is of special importance as it is discarded and any valuable mineral
contained in this stream is a lost opportunity. High valuable mineral recovery rates coupled
with good selectivity in a tailing stream batch test is an indication of insufficient residence
time in the circuit. Slow floating material in the final tailing, however, is often unrecoverable
in a circuit without producing an unacceptable concentrate grade. Increasing circuit residence
time to recover slow-floating material will often only result in large recirculating loads and
no improvement in overall recovery. Large quantities of slow-floating material in the tailings
is an indication that particle modification (either in terms of particle size or chemistry) is
required in the circuit to increase recovery.
Internal streams with poor valuable mineral to gangue mineral selectivity are appropriate
targets for regrinding or depressant or collector addition. Batch flotation of different streams
in a flotation circuit, combined with liberation analysis, enables the appropriate streams for
this type of modification to be delineated from those that would not benefit from this type
of processing.
Regular batch laboratory flotation test mapping during circuit surveys will provide a record
of ore floatability properties and enable the comparison of this property between different
surveys.
where:
F Mineral
Stream s
j
is the flow rate of mineral j in stream s
R Stream s, t minutes
Mineral j
is the cumulative recovery of mineral j in stream s after t minutes of flotation
in the standard laboratory batch flotation test
A B
C D
FIG 11 – Examples of nodes in a flotation circuit. (A) Flotation cell; (B) tower mill; (C) pump sump; (D) flotation circuit.
Take for example a flotation column. A column has a single feed but two product streams.
Equation 3 can be used to recombine the concentrate and tailing stream to produce a single
‘combined’ product stream result. The combined stream floatability can then be compared
to the feed stream floatability to assess whether there has been any change across
the unit. Figure 12 shows the galena recovery rate measured in a batch flotation test of the
feed, concentrate and tailing streams of an industrial column. It also shows the calculated
‘combined product’ stream which, in this case, is very similar to the column feed. It can,
therefore, be concluded that floatability does not change significantly across this unit. The
difference in galena recovery rates in the concentrate and tailing stream is attributed to the
faster-floating galena-containing particles concentrating in the concentrate streams and the
slower-floating galena-containing particles concentrating in the tailing streams. Using this
technique in a number of different flotation circuit operations, it has been shown that ore
floatability is not often affected in a full-scale flotation cell (Runge et al, 2006).
Regrinding and reagent addition are employed in a flotation circuit specifically to alter
particle floatability. The effectiveness of these processes can be assessed by a statistically
planned and analysed on-off plant trial. Alternatively, comparison of ore floatability using
nodal analysis techniques can enable a direct assessment of the effectiveness of these
particle modification operations. In Figure 13a for example, it is obvious that staged reagent
addition has resulted in a significant increase in the proportion of floatable copper in the
circuit. In Figure 13b it is obvious that regrinding has not resulted in an improvement in
the chalcopyrite/pyrite selectivity. These types of observations provide the metallurgist with
invaluable information regarding the effect of an ore modification process.
These techniques will be demonstrated in a worked example included at the end of this
chapter. For a more detailed example, the reader is referred to Runge, Franzidis and Manlapig
(2004) in which these techniques are used to assess the effect of staged reagent addition,
regrinding and lime addition on ore floatability in an industrial circuit.
100
90
FIG 12 – Galena recovery versus time for the feed, concentrate, tailing and combined
product of an industrial-scale flotation column (after Runge et al, 1997).
A
Chalcopyrite Recovery (%)
100
80
60
40
Feed to Reagent Node
20
Product of Reagent Node
0
0 2 4 6 8 10
Time (min)
B 100
Combined Feed
Pyrite Recovery (%)
80 Product
60
40
20
0
0 20 40 60 80 100
Chalcopyrite Recovery (%)
FIG 13 – Nodal analysis comparative graphs: (A) mineral recovery before and after
collector addition, (B) chalcopyrite/pyrite selectivity before and after regrinding.
Flotation is considered a first-order kinetic process, recovery being a function of the time in
the process (t) and a flotation rate constant (k). In a batch laboratory flotation environment,
recovery can be calculated using Equation 4.
R = 1 − exp(−kt) (4)
The flotation rate constant is a function of the operating conditions in the cell and also the size,
mineralogy and surface chemical speciation of the particles being floated. There are a number
of different types of particles in an ore sample and each will float with a different flotation rate
– thus an ore sample exhibits a distribution of floatabilities. There are no methods currently
available for measuring each particle’s individual rate. Instead, the problem is simplified by
assuming the form of the floatability distribution. The two methods most commonly used
involve assuming a continuous shape for the distribution (for example, rectangular distribution
(Klimpel, 1980)) or splitting a mineral into different floatability components. In this chapter,
derivation and use of multiple component parameters will be discussed. For those wanting
more information regarding shaped distributions, the reader is referred to a paper written by
Chander and Polat (1994).
Mineral recovery in a batch flotation test using multiple components can be calculated using
Equation 5.
n
R= / m (1 - exp (- k t))
i i
(5)
i=1
The rate of each component (ki) and the proportion of the mineral in that class (mi) are the
ore floatability parameters of the system that can be derived from batch laboratory flotation
test data. This involves calculating the ‘best’ set of these parameters that minimise the sum
of squares difference between the recoveries calculated using Equation 5 and that measured
in the laboratory batch flotation experiments. This difference can, optionally, be weighted by
the standard deviation of the experimental data (Equation 6). These problems are easily set
up on an Excel spreadsheet using the Solver add-in to perform the minimisation.
(R j,calculated - R j,experimental) 2
SSE = / SD j,experimental 2
(6)
j
by using information collected from laboratory batch flotation tests performed using different
streams of a circuit. A circuit by its very nature results in a redistribution of the floatability
components – fast floatability components predominantly reporting to the concentrate streams
and the slower and non-floating components predominantly reporting to the tailing streams.
If we assume that the feed consists of a number of floatability components, the other streams
are also made up of these same floatability components, just in a different ratio. Analysis
involves assuming the rate of each floatability component in the different batch tests is the
same and that the mass of each floatability component is conserved across different nodes in
the process. These constraints enable the data to parameter ratio to be significantly increased
and enable parameters of multicomponent systems to be derived with statistical confidence.
It is the author’s experience that a two floating and one non-floating model (Equation 9) is
usually sufficient to describe or model the behavior of minerals in a flotation circuit. The
measured distribution of this type of model around a circuit derived using the Northparkes
batch flotation test data depicted in Figure 10 is shown in Figure 14.
In a similar vein, laboratory batch flotation tests can be performed with multiple stages or
with recycle streams to increase the number of data points available for parameter estimation.
Thus, multiple ore floatability parameter estimation can be performed without the need to
collect samples from an operating flotation circuit.
A step-by-step worked example of ore parameter derivation from a multiple stage batch
flotation test will be provided at the end of this chapter. The reader is also referred to
85.1
2.5
2.1
0.5
0.0
0.0 0.0
11.7 0.5
2.7 0.0
80.7
0.0
1.8 11.7
0.0 2.2
0.0
4.8
Legend
1.9 6.9 Fast Floating
0.0 0.3 Slow Floating
0.0 Non Floating
FIG 14 – Mass in different copper floatability components around the Northparkes industrial
flotation circuit (based on 100 units of copper in the feed) (Runge et al, 2003).
Varadi, Runge and Franzidis (2010), who provides additional examples of model parameter
derivation from multi-stage batch laboratory flotation data.
R= kx (10)
1+kx
The flotation recovery equation can be extended to include parameters associated with
entrainment and the cell operating variables (for example, air rate, froth depth). Flotation
cell water recovery equations are also required to calculate the circuit water flow information
required for cell residence time calculations. For more information on these types of simulations
and the methods required to determine the parameters of these relationships, the reader is
referred to Harris et al, 2002.
A simplification of these more complex models can be used to assess the separability of a
particular ore in a particular flow sheet. The effect of the cell operating variables and residence
time can be lumped into a single parameter, the cell operating constant, θ (Loveday, 1966),
which typically varies between zero and 100. Flotation recovery of a component is a function
of only this parameter and its particular batch flotation test rate constant (Equation 11).
k batch i
R= (11)
1 + k batch i
Multiple simulations are performed for a particular flow sheet, with the cell operating
constants, θ, randomised in each stage. This randomisation is a representation of the range
of different cell volumes/cell operating conditions that could be used to achieve flotation
separation. It is equivalent to the flotation operator slowing or speeding up the process to
achieve the desired concentrate grade and recovery at each stage.
The resulting final concentrate grade/recovery numbers from these simulations can be
plotted. Figure 15 shows examples of the type of curves generated using this type of analysis.
These curves were generated using ore floatability constants derived from batch laboratory
experiments applied to a three-stage flotation circuit. The result is not a single grade versus
recovery relationship, as changes in the cell operating constant in one stage results in both a
change in recovery but also a change in the separation efficiency of the circuit.
These curves can be generated to compare the floatability of different ore types (Figure 15b)
or to compare the ability of different flow sheets to separate the minerals in a particular ore
sample (Figure 16).
Simulations required to create these curves can be performed on an Excel spreadsheet.
Alternatively, packages such as JKSimFloat, HSC Chemistry©, and Limn© have been designed
to enable these types of simulations to be performed rapidly in a user-friendly environment.
1. Note that this perfect mixing equation can only be used to predict flotation recovery from a single flotation cell. If multiple cells are
present in a bank, then recovery should be calculated for each cell individually and then combined to predict the overall recovery.
56
A
B 56
Final Concentrate Grade (%)
54
52
50
48
46
44
42
40
40 50 60 70 80 90 100
Circuit Recovery (%)
FIG 15 – Circuit grade and recovery simulated in a three-stage flotation circuit. (A) For a particular
ore type, (B) for the same ore subject to different reagent/regrinding schemes.
60
Final Concentrate Grade (%)
58
56
54
52
50
48 Three Stage
46 Four Stage
44
42
40
40 60 80 100
Circuit Recovery (%)
FIG 16 – Circuit grade and recovery simulated for the same ore processed in a three- and four-stage flotation circuit.
The cell scale-up factor is a complex function of the operating conditions at full scale
and those used in the batch laboratory flotation test. It will also be strongly dependent on
the ore, its particle size distribution and its effect on the full-scale froth characteristics. It is
often assumed that the batch test rate is about twice the rate at full scale (that is, C = 0.5).
In reality, the scale-up factor can change significantly with the cell operating conditions used
in both the batch test and full-scale cells and typically is between 0.05 and 1. It is the author’s
recommendation that this scale-up factor be determined by comparing batch and full-scale
kinetics in an existing concentrator or assumed to be equal to that measured in another
operation treating a similar type of ore. Scale-up numbers will usually be higher in a rougher
and cleaner bank than a scavenger bank (Runge et al, 2001, 2003). Scale up is also sensitive to
particle size with coarser particles having a lower scale-up number than fines because of low
coarse particle froth recoveries in the full-scale cells. A worked example is provided at the
end of this chapter demonstrating an example of how the scale-up factor can be determined.
The other complication when scaling the kinetics from a batch test is that the multiple
component parameters required to fully describe the floatability of an ore (that is, fast, slow
and non-floating rates and mass fractions) cannot be determined confidently from a single
test. To overcome this problem, simplified model structures such as the one floating and
one non-floating model can be used. These simplified models can be used to predict the
recovery achievable off an industrial bank of flotation cells based on a batch test of its feed.
However, prediction of the performance of an entire circuit where streams are recycled
based on results from a single batch test of a feed sample will be poor with this simplified
type of model structure.
Full-scale multiple stage circuit prediction requires the model to better incorporate the
distribution of floatabilities in the feed. There are a number of options. Parameter estimation
can be performed using locked cycle batch test data (rather than a single batch flotation
test) to enable the data to be modelled using fast, slow and non-floating components.
Alternatively, the products from a single batch test can be sized and the rate and proportion
of floatable mineral can be determined for each size class. In this approach, each size class
will need to be assigned a different scale-up constant due to the scale-up constant being
sensitive to particle size.
Once the batch test ore floatability parameters and estimates of cell scale-up numbers have
been established, simulations can be performed. For a new operation, one can estimate the
number of flotation cells of a particular volume required to achieve a particular recovery from
a flotation bank. Additional capacity is usually factored into this analysis to account for any
error and the usual increase in capacity (over design) often demanded from a flotation circuit
over time. Alternatively, for an existing plant, the response of different ore types or the effect
of changing variables such as reagent conditioning, grinding environment or particle size on
the grade and recovery achievable in the plant can be predicted.
AREAS OF APPLICATION
Laboratory batch flotation testing can be used as a diagnostic tool throughout the development
and life of a flotation circuit.
Even before a project has commenced, batch laboratory flotation of drill core samples provides
the information required to make decisions regarding a project’s viability. Throughout the
development of a flotation circuit flow sheet, batch tests are used to screen and determine the
appropriate reagent suite and dosage rates, the target grind size and an estimate of the circuit
residence time requirements.
From the day a flotation circuit is commissioned, the metallurgist strives to better understand
the process and implement changes that will improve or optimise operation. Batch laboratory
flotation testing is one of a suite of tools used in this process. New reagents are almost always
screened first in these small tests before implementation at full scale.
After implementation of a change in grind size, reagent or circuit change at industrial scale,
batch flotation tests can be used to assess the effect of the change on mineral recovery rates and
selectivity.
The ore feeding a flotation circuit is continually changing as rock is extracted from different
areas of the mine. Batch testing is a valuable tool for comparing the floatability of different
ores or predicting the metallurgical performance that will be achieved when a particular ore
is processed. Geometallurgical programs are often conducted, involving hundreds of batch
flotation tests to achieve this objective.
Flotation modelling techniques published in the literature (Harris et al, 2002; Runge et al,
2001; Coleman, Urtubia and Alexander, 2006) are using batch testing to determine ore
characterisation properties. These models can be used for flotation circuit design, flotation
circuit diagnosis and simulation of alternative operating strategies. They enable an estimation
of the overall change in grade and recovery achievable in a particular process fed an ore with
a particular mineral selectivity.
In summary, laboratory batch flotation testing is an essential tool that, when used
appropriately, helps a metallurgist design new flotation processes, develop strategies for
improving a flotation circuit and ultimately assess the effectiveness of these strategies.
ADVANTAGES AND LIMITATIONS OF USING BATCH FLOTATION TESTING FOR CIRCUIT DIAGNOSIS
Small-scale batch laboratory tests are relatively cheap and easy to perform. Almost all
concentrator laboratories are equipped with laboratory flotation cells and the associated
equipment required to prepare and process the samples produced from a test. They also
have the advantage that, to a significant degree, they replicate the flotation process that occurs
at full scale.
Batch tests also have the advantage that they enable the ore floatability to be analysed in
isolation to those effects related to circuit operation. Poor flotation recovery measured in a
circuit survey can be due to a number of different effects: poor ore floatability, insufficient
residence time, use of overly deep froth depths, to name a few. Batch laboratory tests are usually
performed using the same operating conditions and thus any change between tests can be
attributed solely to a difference in ore floatability. Batch tests performed in conjunction with
circuit surveys can help decouple ore floatability and circuit operation effects and highlight
the cause of flotation recovery loss.
What is often forgotten when analysing batch flotation test results, however, is that they are
a good measure of pulp floatability but sometimes a poor indicator of full-scale froth phase
performance. The froth in a small-scale laboratory test is only a couple of centimetres in
depth and is continually removed. In a full-scale cell the froth can be 5 cm to 50 cm in depth
and there are usually significant transportation distances between a point on the surface of the
froth and the nearest launder. The froth produced in a small-scale cell is often very different
in bubble size, texture and viscosity to that observed when the same ore is processed at full
scale. This can result in a particular reagent suite, producing a superior pulp floatability
(and thus batch test result) but poor recovery in a full-scale cell because of its effect on
movement of the froth phase.
The primary assumption associated with batch laboratory test analysis is that pulp
selectivity in the small scale is similar to what would be achieved in the full-scale flotation
machine. It’s likely this is a valid assumption in mechanical cells but less likely in
alternative flotation technologies. The low degree of turbulence in a column, for instance,
can result in the poor recovery of coarse particles, which are recovered quickly in a batch
flotation test.
Another risk during batch laboratory flotation testing is the removal of loosely deposited
surface coatings. The higher agitation energies that exist in a small-scale cell can sometimes
remove molecules that would remain in situ in the full-scale flotation environment.
These limitations, as with most diagnostic techniques, mean that those conclusions drawn
from batch laboratory testing should be verified through full-scale trials.
Much can be gained by performing batch laboratory flotation tests in conjunction with other
measurements. Batch laboratory flotation tests indicate the overall result of a change. Sizing,
mineralogical and chemical analysis tools can be used to determine the reason for a particular
result, often suggesting alternative batch laboratory experiments to trial.
CONCLUDING REMARKS
In this chapter, batch laboratory testing procedures and the modelling and graphical
techniques that can be used to analyse the results of these tests have been reviewed.
It has been demonstrated that batch laboratory flotation tests have a role to play throughout
the lifetime of a particular flotation circuit operation. They can be used during the design of
new flotation processes, to develop strategies for improving a flotation circuit and, ultimately,
to assess the effectiveness of these strategies.
They have the advantage that they are relatively cheap and easy to perform and to an extent,
replicate the full-scale flotation process. They enable an assessment of ore floatability without
the complication of changing cell operating variable effects. Effects, due to differences between
the small-scale and full-scale cell, should be kept in mind when interpreting results.
The value of laboratory batch flotation testing is greatly enhanced when performed in
parallel with other techniques that can assist with data interpretation.
REFERENCES
Chander, S and Polat, M, 1994. In quest of a more realistic flotation kinetics model, in Proceedings
IV Meeting of the Southern Hemisphere on Mineral Technology and III Latin American Congress on Froth
Flotation (eds: S Castro and J Alvarez), pp 481–500.
Coleman, R G, Urtubia, H E and Alexander, D J, 2006. A comparison of BHP Billiton’s Minera Escondida
flotation concentrators, in Proceedings 38th Annual Meeting of the Canadian Mineral Processors, pp 349–
370 (Canadian Institute of Mining, Metallurgy and Petroleum: Montreal).
Harris, M C, 1998. The use of flotation plant data to simulate flotation circuits, in Proceedings SAIMM
Mineral Processing Design School (Southern African Institute of Mining and Metallurgy: Johannesburg).
Harris, M C, Runge, K C, Whiten, W J and Morrison, R D, 2002. JKSimFloat as a practical tool for
flotation process design and optimization, in Mineral Processing Plant Design, Practice and Control,
vol 1, pp 461–478 (Society for Mining, Metallurgy, and Exploration: Littleton).
Hay, M P and Rule, C M, 2003. Supasim: a flotation plant design and analysis methodology, Minerals
Engineering, 16(11):1103–1109.
Johnson, N W, 2005. A review of the entrainment mechanism and its modelling in industrial flotation
processes, in Proceedings Centenary of Flotation Symposium (ed: G J Jameson), pp 487–496 (The
Australasian Institute of Mining and Metallurgy: Melbourne).
Kelsall, D F, 1961. Application of probability in the assessment of flotation systems, Transactions of the
Institute of Mining and Metallurgy, 70:91–204.
Klimpel, R, 1980. Selection of chemical reagents for flotation, in Mineral Processing Plant Design, pp 907–
934 (The American Institute of Mining, Metallurgical, and Petroleum Engineers: Littleton).
Loveday, B K, 1966. Analysis of froth flotation kinetics, Trans IMM, 75:C219–C225.
Runge, K C, Dunglison, M E, Manlapig, E V and Franzidis, J-P, 2001. Floatability component modelling
– a powerful tool for flotation circuit diagnosis, in Proceedings Fourth International Symposium on
Fundamentals of Minerals Processing, pp 93–107 (Canadian Institute of Mining, Metallurgy and
Petroleum: Montreal).
Runge, K C, Franzidis, J P and Manlapig, E V, 2004. Optimisation of flotation pulp selectivity utilising
nodal analysis, in Proceedings 36th Annual Meeting of the Canadian Mineral Processors, pp 575–597
(Canadian Institute of Mining, Metallurgy and Petroleum: Montreal).
Runge, K C, Franzidis, J P, Manlapig, E V and Harris, M C, 2003. Structuring a flotation model for
robust prediction of flotation circuit performance, in Proceedings 22nd IMPC Conference, Cape Town,
September.
Runge, K C, Harris, M C, Franzidis, J P and Manlapig, E V, 2006. Conservation of floatability across
processes in industrial flotation circuits, in Proceedings XXIII International Mineral Processing Congress
(eds: G Onal, N Acarkan, et al), vol 3, pp 1909–1914 (Promed Advertising Agency).
Runge, K C, Harris, M C, Frew, J A and Manlapig, E V, 1997. Floatability of streams around the Comino
Red Dog lead cleaning circuit, in Proceedings Sixth Annual Mill Operators’ Conference, pp 157–163 (The
Australasian Institute of Mining and Metallurgy: Melbourne).
Varadi, R, Runge, K C and Franzidis, J P, 2010. The rate variable batch test (RVBT): a research method
of characterising ore floatability, in Proceedings XXV IMPC, pp 2471–2480 (The Australasian Institute
of Mining and Metallurgy: Melbourne).
TABLE A1
Mineral grade of samples produced from batch flotation testing of a lead concentrate before and after grinding.
Sample Time Before After regrinding
(min) Galena Sphalerite Pyrite NSG Galena Sphalerite Pyrite NSG
Concentrate 1 0–0.5 76.71 14.08 4.02 5.20 72.80 14.86 6.77 5.58
Concentrate 2 0.5–1.5 68.66 20.80 5.73 4.81 68.54 17.67 7.66 6.13
Concentrate 3 1.5–4 48.42 31.44 9.12 11.03 57.27 25.34 9.52 7.87
Concentrate 4 4–8 25.88 45.83 16.35 11.95 28.29 44.10 15.60 12.00
Concentrate 5 8–12 18.17 50.83 21.40 9.60 15.99 50.99 19.02 14.01
Concentrate 6 12–20 12.54 51.77 23.49 12.21 12.19 50.05 20.80 16.96
Tailing 4.72 53.02 21.25 21.01 5.75 45.20 22.84 26.21
NSG – non-sulfide gangue.
100
Galena Recovery (%)
80
Rougher Concentrate
60
40 Reground Rougher
Concentrate
20
0
40 50 60 70 80
Galena Grade (%)
FIG A1 – Galena recovery versus grade achieved in batch flotation tests of the lead rougher concentrate before and after regrinding.
A 70
70
B
Rougher Concentrate
60
Pyrite Recovery (%)
50 Reground Rougher
Concentrate
40
30
20
10
0
0 20 40 60 80 100
Galena Recovery (%)
60
C
Rougher Concentrate
50
NSG Recovery (%)
Reground Rougher
40 Concentrate
30
20
10
0
0 20 40 60 80 100
Galena Recovery (%)
FIG A2 – Galena versus gangue mineral selectivity before and after regrinding of the lead rougher concentrate.
This result is primarily due to deterioration in the pyrite-galena selectivity. Galena recovery
rate has changed very little by the regrinding process, whereas there has been a significant
increase in pyrite recovery (Figure A3). There is very little difference in the galena-sphalerite
selectivity and a marginal improvement in galena-gangue selectivity.
The grinding environment often results in changes in the condition of particle surfaces – iron
hydroxide coatings, for instance, can deposit on all surfaces during grinding with mild steel
balls. Some change is occurring to the pyrite particles in this grinding system, resulting in
an increased recovery rate. Surface chemical speciation of particle surfaces before and after
grinding would almost certainly explain the reason for this observed result and maybe suggest
strategies to reverse the effect.
60
A
Rougher Concentrate
20 Reground Rougher
Concentrate
0
0 5 10 15 20 25
Time (min)
100
B
Galena Recovery (%)
80
Rougher Concentrate
60
Reground Rougher
40 Concentrate
20
0
0 5 10 15 20 25
Time (min)
FIG A3 – Mineral recovery versus time before and after regrinding of the lead rougher concentrate.
Under the current chemical conditions, however, regrinding of lead concentrate should be
discontinued.
Activator Collector
To Final Tail
Stage 1 Sump
Lime Activator
Regrind
Collector
Stage 2
To Final Concentrate
Sump
Cleaning Columns
TABLE A2
Metallurgical performance achieved during each survey.
Details of survey Concentrate grade (%) Recovery to concentrate (%)
Galena Sphalerite Pyrite NSG Galena Sphalerite Pyrite NSG
No lime added 6.2 79.7 7.0 7.2 57.8 86.2 11.4 5.5
Lime added 5.0 79.0 8.0 8.0 33.0 61.5 8.4 4.2
NSG – non-sulfide gangue.
that the sphalerite, galena and non-sulfide gangue mineral recovery rates are very similar
in the feed batch flotation tests. The pyrite recovery rate is low and thus sphalerite/pyrite
selectivity of the feed is superior in the lime survey (Figure A5). It can therefore be concluded
that the low recoveries in the lime survey are not due to poorer feed ore floatability.
Lime is supposed to improve sphalerite/pyrite selectivity by depressing pyrite in preference
to sphalerite. The change in sphalerite/pyrite selectivity in the two circuit surveys can be
assessed using nodal analysis techniques. The circuit is a node with a single feed and two
product streams. The two product stream results can be combined using Equation A1 to create
an estimate of the combined product floatability.
R Concentrate
Mineral j
# R Concentrate,
Mineral j
t
+ R Tail
Mineral j
# R Tail, t
Mineral j
R Combined Product, t
= (A1)
Mineral j 100
where:
R Concentrate
Mineral j
is the recovery of mineral j to concentrate
TABLE A3
Cumulative mineral recovery achieved in the feed, concentrate and tailing batch tests performed during the two circuit surveys.
Cumulative time Mineral cumulative recovery in feed batch test
(min) No lime addition Lime addition
Galena Sphalerite Pyrite NSG Galena Sphalerite Pyrite NSG
0.33 22.5 33.3 10.8 7.5 23.2 36.4 9.7 10.3
1 49.9 63.4 28.2 19.8 49.0 66.6 22.2 21.8
2 66.8 79.1 43.3 29.9 65.7 80.7 33.5 31.8
4 82.4 89.2 58.0 40.9 79.3 90.4 48.3 42.0
8 89.5 94.3 69.6 51.1 87.7 94.8 62.1 52.2
12 92.4 96.1 75.5 57.6 91.2 96.5 69.7 58.9
R Tailing
Mineral j
is the recovery to tailing of mineral j
R Stream s, t
Mineral j
is the cumulative recovery of mineral j in the batch test of stream s after t minutes
of flotation
The results of this calculation are shown in Table A4.
Once the combined product floatability has been calculated, feed and product mineral
recoveries can be compared. Figure A6 shows the sphalerite/pyrite selectivity in the feed
and product streams of this particular circuit. Sphalerite/pyrite selectivity in the survey when
lime is not added, improves dramatically across the circuit. It can therefore be concluded that
100
Survey - No Lime Added
40
20
0
40 50 60 70 80 90 100
Sphalerite Recovery (%)
FIG A5 – Sphalerite/pyrite selectivity measured in the feed batch tests performed during the two circuit surveys.
TABLE A4
Cumulative mineral recovery of the combined product streams from the two circuit surveys.
Cumulative time Mineral cumulative recovery in combined product
(min) No lime addition Lime addition
Galena Sphalerite Pyrite NSG Galena Sphalerite Pyrite NSG
0.33 24.3 35.7 9.6 6.8 15.6 21.2 8.2 6.3
1 53.0 71.9 25.2 13.1 38.2 49.0 20.6 14.7
2 69.5 87.4 38.6 19.4 54.7 66.6 31.8 21.4
4 80.5 94.5 51.3 28.1 68.7 79.8 43.0 27.5
8 87.4 97.0 65.6 37.5 80.4 88.2 56.2 36.3
12 89.7 97.6 70.0 42.5 85.0 90.8 62.9 41.8
NSG – non-sulfide gangue.
regrinding and the reagent addition during this survey have had a positive effect on the ability
to separate sphalerite and pyrite. In contrast, sphalerite/pyrite selectivity deteriorates in the
survey in which lime is added.
This deterioration in sphalerite/pyrite selectivity is due to a decrease in the sphalerite
recovery rate (in contrast to an increase when lime is not added to the circuit) and a small but
similar decrease in the recovery of pyrite (Figure A7).
Use of nodal analysis in this example has shown that lime addition in the circuit does not
result in appreciable improvements in sphalerite/pyrite selectivity. The deterioration in
valuable to gangue mineral selectivity will have had a significant impact on the ability to
separate minerals in this circuit. Adjustments in circuit operation would have occurred to
maintain grade but at a cost of losing zinc recovery.
As a confirmation of this conclusion, nodal analysis was also performed before and after
lime addition in this circuit (Runge, Franzidis and Manlapig, 2004). These tests show that lime
addition dramatically decreases the floatability of all minerals, but sphalerite in particular.
It thus can be concluded that lime addition is clearly not the solution to improving sphalerite/
pyrite selectivity in this circuit. Batch laboratory testing performed in conjunction with survey
work enabled a more definitive analysis of circuit operation.
40
20
0
40 50 60 70 80 90 100
Sphalerite Recovery (%)
80
Feed
60 Combined Product
40
20
0
40 50 60 70 80 90 100
Sphalerite Recovery (%)
FIG A6 – Sphalerite/pyrite selectivity measured in the feed and product of the two circuit surveys.
A 100
Feed No Lime
60
Feed Lime
40 Product No Lime
Product Lime
20
0
0 5 10 15
Time (min)
100
B
Pyrite Recovery (%)
80
60
Feed No Lime
40
Feed Lime
20 Product No Lime
Product Lime
0
0 5 10 15
Time (min)
FIG A7 – Sphalerite and pyrite recovery rates measured in the feed and product of the two circuit surveys.
Rougher Scavenger
(2 minutes) (6 minutes)
New Feed Tailing
Concentrates
FIG A8 – Representation of multiple stage batch laboratory flotation test – samples produced denoted by a red circle.
It is decided to represent the ore floatability of the system by three lead components (fast,
slow and non-floating lead) and three gangue components (fast, slow and non-floating
gangue). Three components are the maximum number that could be derived with confidence
from the experimental data available. Gangue is a term to refer to the weight in each
sample that is ‘not lead’. To determine the flotation rate and proportion of lead and gangue
in each of the floatability components, an Excel spreadsheet was set up to calculate the weight
and assay of the experimental samples based on a flotation model. Solver, an add-in tool to
Excel, was then used to determine the ore floatability parameters that minimise the difference
between the measured and calculated sample information.
TABLE A5
Multiple-stage batch test results.
Sample Solids (g) % Lead % Gangue
Cleaner concentrate 1 13.4 75.9 24.1
Cleaner concentrate 2 10.0 74.0 26.0
Cleaner concentrate 3 7.5 71.9 28.1
Cleaner concentrate 4 13.5 67.3 32.7
Cleaner scavenger concentrate 1 9.2 55.8 44.2
Cleaner scavenger concentrate 2 6.7 50.5 49.5
Cleaner scavenger concentrate 3 8.9 42.8 57.2
Cleaner scavenger concentrate 4 14.5 25.7 74.3
Cleaner scavenger tailing 19.6 7.2 92.8
Tailing 2096.7 0.9 99.1
Recalculated feed 2200.0 3.1 96.9
Tables A6, A7 and A8 show the calculation tables that were set up in Excel to perform the
computations.
The calculations are performed separately for each component. The first table for the lead
calculation contains the parameters to be derived from the fitting exercise. The values first input
into the table are an estimate. Note that the rate and mass fraction of the non-floating component
are not fitted parameters. The rate of the non-floating component is zero and the proportion of
material in the non-floating component is one minus the fraction in the other components.
The second table calculates the recovery of each component in each stage of the batch test
(Rs). Recovery is a function of the flotation rate of the component (ki) and the time of flotation
(t) (Equation A2).
In the final table, the weight of each component in each batch test sample throughout the
duration of the test is calculated. The weight in each component in the test feed is established
by multiplying the total weight of lead in the feed (an input) by the proportion of lead in
each component as specified in the first table. The weight in the concentrate of each stage is
determined by multiplying the weight in the feed to the stage by the stage recovery (Rs). The
weight in the tail of each stage is the feed minus the concentrate weight. These calculations
are performed in sequence, with either the concentrate or tailing (or both in the case of the
cleaner scavenger feed) of previous stages becoming the feed to subsequent stages.
The total weight of lead reporting to each concentrate or tailing stream in the batch test is
then calculated by summing the weights of its floatability components.
The identical calculation sequence is performed to determine the mass of gangue in each of
the batch test samples according to a set of gangue specific floatability components (Table A7).
The weight of lead and gangue calculated in each stream can be used to calculate the total
weight of solids and the lead and gangue assay of each batch test sample. This calculated
assay is then compared to the actual assay measured during the test (Table A8).
TABLE A6
Lead flotation model Excel spreadsheet.
Parameter table Fast Slow Non
Lead flotation rates (min )-1
0.932 0.201 0.000
Proportion of lead in feed in each component 0.642 0.099 0.258
TABLE A7
Gangue flotation model Excel spreadsheet.
Parameter table Fast Slow Non
Gangue flotation rates 0.654 0.099 0.000
Proportion of gangue in feed in each component 0.009 0.030 0.961
Calculation of the sum of squares error associated with the model fit.
Calculated data Experimental data Standard deviation Squared error
Batch test sample Assay Solids (g) Assay Solids (g) Assay Solids (g) Assay Solids (g)
Lead Gangue Lead Gangue Lead Gangue Lead Gangue
Feed 3.10 96.90 2202.00
Rougher concentrate 60.46 39.54 64.99
Scavenger concentrate 25.94 74.06 38.34
Tailing 0.91 99.09 2098.67 0.85 99.13 2096.7 0.19 1.00 52.4 0.09 0.00 0.00
Cleaner feed (rougher concentrate) 60.46 39.54 64.99
Cleaner concentrate 1 75.94 24.06 13.21 76.7 24.09 13.4 1.00 1.00 0.3 0.58 0.00 0.22
Cleaner concentrate 2 73.84 26.16 10.00 73.1 25.97 10.0 1.00 1.00 0.3 0.54 0.04 0.00
Cleaner concentrate 3 71.56 28.44 7.60 71.0 28.09 7.5 1.00 1.00 0.2 0.31 0.13 0.10
KYM RUNGE
Cleaner concentrate 4 66.85 33.15 13.74 65.0 32.67 13.5 1.00 1.00 0.3 3.43 0.23 0.36
Cleaner tail 35.48 64.52 20.43
Cleaner scavenger feed (SC + CT) 29.25 70.75 58.77
Cleaner scavenger concentrate 1 55.45 44.55 9.22 58.3 44.16 9.2 1.00 1.00 0.2 8.10 0.15 0.03
Cleaner scavenger concentrate 2 50.07 49.93 6.65 48.2 49.51 6.7 1.00 1.00 0.2 3.49 0.18 0.02
FLOTATION PLANT OPTIMISATION | SPECTRUM SERIES 25
Cleaner scavenger concentrate 3 42.23 57.77 8.72 41.1 57.17 8.9 1.00 1.00 0.2 1.28 0.36 0.39
Cleaner scavenger concentrate 4 24.81 75.19 14.60 24.0 74.35 14.5 1.00 1.00 0.4 0.65 0.71 0.06
Cleaner scavenger tailing 7.38 92.62 19.57 7.5 92.79 19.6 0.82 1.00 0.5 0.02 0.03 0.00
Combined cleaner concentrate 71.92 28.08 44.56
Combined cleaner scavenger concentrate 40.18 59.82 39.20
Total concentrate 57.06 42.94 83.76
Total tailings 0.97 99.03 2118.24
Sum of squares error 21.52
CHAPTER 10 | LABORATORY FLOTATION TESTING – AN ESSENTIAL TOOL FOR ORE CHARACTERISATION
To calculate the total sum of squares error associated with the system, the error of each
experimental data point must be estimated. For this exercise, the Whiten function is used to
estimate the standard deviation of the assays (Equation A3) and a relative error of 2.5 per cent
is used as the error of the solid weights.
SD = 1 if assay > 9%
SD = assay / 10 + 0.1 if assay < 9% (A3)
The squared error associated with each experimental data point is calculated using
Equation A4. The total sum of squares error is the sum of the squared errors.
(dcalculated - d experimental) 2
SE = (A4)
SD experimental 2
Once the spreadsheet has been set up in the way described above, Solver (an add-in tool in
Excel) can be used to determine the lead and gangue parameters that minimise the total sum
of squares error. For this example, Solver is able to find a set of parameters that result in a
good fit to the experimental data.
In Table A9, the parameters derived from the analysis above are compared to those
derived from a similar test performed using the ore currently processed by the plant.
The flotation rates of the different components in the two different ores are similar. The ore
does have poorer floating characteristics than the current ore. The proportion of floatable
lead is significantly lower and the proportion of floatable gangue has increased. This is an
indication that the future ore is less liberated at the grind size tested.
The flotation model that was developed to determine the ore parameters can also be used
to simulate various different batch test configurations – predicting the change in grade and
recovery that would be produced if the test was run in a different way.
To use the model, the parameters of the ore are input into the parameter table and the time
of each stage is input into the stage recovery matrix. The lead grade and lead recovery of the
combined concentrate (cleaner and cleaner scavenger concentrate) is calculated by the model.
Using the parameters derived for the current ore and the future ore, the model was used to
predict the combined concentrate lead grade and recoveries that would be achieved if
different times were used in the various stages of the batch test. Table A10 shows the different
simulations that were performed and the results. Figure A9 shows a graphical comparison of
the results derived for the two ores.
The ore floatability of the future ore, as measured in the batch flotation test, is clearly
poorer than that currently treated. The lead recovery achievable at a particular grade is
TABLE A9
Stage batch test ore floatability results.
Parameter Future ore Current ore
Fast Slow Non Fast Slow Non
Lead flotation rates 0.93 0.20 0.0 0.99 0.16 0.0
Proportion of lead in each component in feed 0.64 0.10 0.26 0.71 0.08 0.21
Gangue flotation rates 0.65 0.10 0.0 0.62 0.07 0.0
Proportion of gangue in each component in feed 0.009 0.03 0.96 0.007 0.02 0.97
TABLE A10
Multiple stage batch test simulation results.
Model Batch test stage time (min) Future ore Current ore
simulation Rougher Scavenger Cleaner Cleaner Lead Lead Lead Lead
scavenger grade (%) recovery (%) grade (%) recovery (%)
Base case 2.00 6.00 2.00 6.00 57.06 70.02 70.44 74.72
Simulation 1 2.00 6.00 0.50 1.50 66.69 55.63 77.69 61.67
Simulation 2 4.00 12.00 2.00 6.00 52.54 71.41 66.97 75.92
Simulation 3 6.00 18.00 2.00 6.00 50.52 71.78 65.05 76.32
Simulation 4 12.00 24.00 2.00 6.00 48.91 72.04 63.26 76.61
Simulation 5 12.00 24.00 4.00 12.00 43.01 73.71 57.32 78.28
Simulation 6 12.00 24.00 1.00 3.00 55.64 68.08 69.04 73.31
Simulation 7 6.00 18.00 1.00 3.00 57.03 67.83 70.34 73.08
Simulation 8 2.00 6.00 1.00 3.00 62.21 66.21 74.16 71.71
Simulation 9 2.00 6.00 0.50 1.50 66.69 55.63 77.69 61.67
Simulation 10 1.00 3.00 0.50 1.50 69.76 52.34 79.80 58.69
Simulation 11 1.00 3.00 1.00 3.00 66.11 62.97 76.87 69.06
Simulation 12 1.00 3.00 2.00 6.00 62.02 66.60 73.98 72.02
90
80
Lead Recovery (%)
70
60
50 Future Ore
40 Current Ore
30
20
10
0
0 20 40 60 80 100
Lead Grade (%)
FIG A9 – Simulated batch test lead grade and recovery of the future and current ore.
ten per cent lower in the future ore simulations than in the current ore simulations. Increased
recoveries may be possible at a finer grind. Additional test work would be required to
substantiate this hypothesis.
tests. In parallel, feed tonnage was recorded and samples were collected from the plant
(designated in Figure A10) to determine the kinetics in the rougher and scavenger bank when
processing this ore.
It was decided to use one floating and one non-floating component to describe the
floatability distribution of the ore (a simplification that is possible because it was only the
rougher and scavenger bank recoveries that were to be predicted). Solver in Excel was used
to determine the batch test flotation rate (kbatch) and the proportion of floatable lead in the feed
to the circuit (Rmax), which minimised the difference between the recoveries calculated using
Equation A5 and that measured experimentally in the batch flotation test. The rate calculated
was 1.04 min-1. Rmax or the proportion of lead in the feed that was floatable was determined to
be 92.3 per cent.
R = R max (1 - exp (- k t)) (A5)
The survey data was mass balanced to determine the solids, water and lead flow in the
sampled streams. A circuit model was then set up to predict the flow of lead to every stream
of the circuit. The circuit consists of four banks of cells, with each bank consisting of two or
three cells. All cells in a particular bank were assumed to behave in a similar manner and have
the same scale-up factor (as individual factors could not be derived as individual cells were
not sampled). Recovery, however, was calculated for each cell individually down each bank
using Equation A6.
C j kbatch x
R = mf (A6)
1 + C j kbatch x
The proportion of floatable lead feeding the circuit was defined as that fitted and
determined using the batch flotation test data. The proportion of floatable lead feeding each
individual cell was a function of the amount recovered in the preceding cells. Residence
time was estimated by dividing the total cell volume of a bank by the volumetric flow
rate measured in its feed and then dividing by the number of cells in the bank. This is
Rougher
Scavenger 1A Scavenger 1B
Scavenger 2
Grinding
Circuit
To
Tailings
To
Cleaning
FIG A10 – Eureka rougher scavenger flotation circuit denoting points at which samples were collected during a survey.
a reasonable approximation for rougher and scavengers where the flow in the feed and
tailing of each cell is similar.
A unit matrix and a stream matrix were set up in Excel to calculate the unit recoveries
and stream flows for the model (shown as Tables A11 and A12). Solver in Excel was used
to calculate the scale-up factors for each bank, which minimised the difference between the
calculated lead bank recoveries produced by the model and that measured experimentally in
the circuit survey. As the problem has zero degrees of freedom, the model is able to exactly
match the experimental data and the total sum of squares error associated with the problem
is zero.
The derived bank scale-up factors are higher for the cells in the rougher and lower for the
scavenger cells at the end of the circuit. This is usual and is a consequence of froth recoveries
in the scavengers, where there is little mass being recovered, being significantly less than
in the roughers. Thus, if additional capacity were to be added to this bank, the cells would
behave in a similar manner to scavenger 2 and have significantly lower recovery rates than
achieved in the rougher cells.
Future ore flotation recoveries can now be calculated using the developed model, assuming
that the bank scale-up numbers remain the same. This should be a reasonable assumption so
long as the ores do not have a property that would significantly affect full-scale froth recoveries.
The batch flotation tests performed using these future ores can be used to determine the batch
flotation rate and proportion of floatable lead in the feed required as input to enable these
simulations to be performed.
TABLE A11
Spreadsheet table used to calculate lead recovery based on model parameters.
Bank of circuit Model parameters Lead
Cj Res time kbatch mf recovery
(min-1/min-1) (min) (min-1) (%) (%)
Cell 1 0.63 3.3 1.04 92.3 63.25
Rougher Cell 2 0.63 3.3 1.04 81.5 55.85
Cell 3 0.63 3.3 1.04 66.0 45.26
Cell 1 0.50 3.1 1.04 51.6 31.89
Scavenger 1
Cell 2 0.50 3.1 1.04 42.0 26.00
Cell 1 0.23 3 1.04 34.9 14.65
Scavenger 1B
Cell 2 0.23 3 1.04 31.4 13.18
Cell 1 0.04 2.9 1.04 28.5 2.77
Scavenger 2
Cell 2 0.04 2.9 1.04 27.9 2.71
TABLE A12
Calculated flow of floating and non-floating lead in the streams of the rougher
and scavenger bank based on the calculated unit recoveries in Table A11.
Stream Lead flow (based on 100 units in feed) Bank recovery (%)
Floating Non-floating Total Modelled Exp SSE
Rougher feed 92.30 7.70 100.00
Rougher Cell 1 concentrate 58.38 0 58.38
Rougher Cell 1 tailing 33.92 7.70 41.62
Rougher Cell 2 concentrate 18.95 0 18.95
Rougher Cell 2 tailing 14.98 7.70 22.68
Rougher Cell 3 concentrate 6.78 0 6.78
Rougher Cell 3 tailing 8.20 7.70 15.90
Combined rougher concentrate 84.10 0 84.10 84.1 84.1 0.00
Scavenger 1A Cell 1 concentrate 2.62 0 2.62
Scavenger 1A Cell 1 tailing 5.58 7.70 13.28
Scavenger 1A Cell 2 concentrate 1.45 0 1.45
Scavenger 1A Cell 2 tailing 4.13 7.70 11.83
Scavenger 1A combined concentrate 4.07 0 4.07 49.6 49.6 0.00
Scavenger 1B Cell 1 concentrate 0.61 0 0.61
Scavenger 1B Cell 1 tailing 3.53 7.70 11.23
Scavenger 1B Cell 2 concentrate 0.46 0 0.46
Scavenger 1B Cell 2 tailing 3.06 7.70 10.76
Scavenger 1B combined concentrate 1.07 0 1.07 25.9 25.9 0.00
Scavenger 2 Cell 1 concentrate 0.08 0 0.08
Scavenger 2 Cell 1 tailing 2.98 7.70 10.68
Scavenger 2 Cell 2 concentrate 0.08 0 0.08
Scavenger 2 Cell 2 tailing 2.90 7.70 10.60
Scavenger 2 combined concentrate 0.17 0 0.17 5.4 5.4 0.00
Combined concentrates 89.40 0 89.40
Final tailing 2.90 7.70 10.60 Total SSE 0.00
CHAPTER 11
Designing and analysing plant trials
Tim Napier-Munn
ABSTRACT
This chapter deals with the principles of running plant trials. The purpose of a plant trial
is (usually) to determine whether a change in some process condition will lead to an
improvement in performance. Detecting the presence and magnitude of any change in
process performance is difficult against the inevitable background of process noise (error) if
the correct procedures are not followed. The source and management of error is discussed,
and then several desirable experimental methods are described, including the paired t-test,
the analysis of variance (ANOVA) and the randomised block design. Other designs discussed
briefly include the factorial experiment and evolutionary operation (EVOP). A formula is
given for the number of trials required for a t-test, and a procedure for eliminating outliers.
If a formal design cannot be implemented or is desired to analyse historical data, modelling
is the appropriate method, and several approaches are described. CUSUM (cumulative sum)
control charts are recommended as a powerful way of examining data collected in a time
sequence. The chapter finishes with a consideration of some of the practical aspects of doing
a plant trial.
There really is no alternative to the adoption of a formal experimental design or, if all
else fails, a rigorous analysis of the data from some other experimental activity. Experience
repeatedly shows that a systematic approach to conducting plant trials will produce the
correct answer in the minimum time at minimum cost. Any other approach is very likely
to take much longer and cost more to produce the wrong answer, or at best to produce no
answer at all.
INTRODUCTION
The great statistician R A Fisher (1938), whose work and philosophy underpin many of the
methods described in this chapter, once said:
To call in the statistician after the experiment is done may be no more than asking him to
perform a post-mortem examination. He may be able to say what the experiment died of.
Fisher was a brilliant mathematician who worked for 20 years at the famed Rothamsted
Agricultural Research Station in England, and the experimental methods that he developed
for Rothamsted’s scientists have been practical and effective, and have stood the test of
time. The problems that agricultural scientists face are very similar to those encountered by
metallurgists charged with a policy of ‘continuous improvement’ and the consequent testing
of new reagents, equipment or process circuits. The data are very noisy and the experiments
often difficult to do.
This chapter is a plea to follow Fisher’s implicit advice. You don’t necessarily have to call in a
statistician, but you do have to plan the experiments properly. Many laboratory experiments
and plant trials are conducted with the naïve view that hard work and enthusiasm are
adequate substitutes for planning, rigour and thought. The protagonists throw themselves
into an experiment in which hope triumphs over method, and the answer is optimistically
expected to emerge from the fog, like advice from the oracle. As a consequence this often
leads to no clear result or, worse still, the wrong result. The irony is that the right way to
do things is often cheaper than the wrong way, and leads to the correct result more quickly.
There is therefore no real excuse for doing a poor experiment.
In the past the Eureka concentrator was no exception. Figure 1 shows the daily recovery
from a plant trial conducted some years ago to test the efficacy of an expensive new reagent.
The average zinc recovery with the new reagent was found to be seven per cent more than
with the old. This is an astonishing improvement and duly excited the metallurgical staff
no end.
Would you, looking at this simple graph of the recovery time series, support the view that
Eureka should invest in the new reagent? I hope not. There is good visual evidence that the
recovery was climbing steadily even before the new reagent was introduced, and in fact one
could take the view that the improvement was actually impeded by the new reagent, even
though the average recovery under the new condition was much higher than that under
the old.
The key point to take from this all too familiar scenario is that, in the absence of other
information, there is no way in which any change in metallurgical performance can be unequivocally
90
85
80
75
Zn Recovery (%)
70
65
60
Reagent 1 Reagent 2
55 R = 70% R = 77%
50
45
40
0 20 40 Day 60 80
attributed to the new experimental condition. This is the fundamental flaw in such ‘before and
after’ experiments, and it cannot be avoided, except by doing some special analysis, which
is discussed further on. A much better option is to adopt a simple experimental protocol to
deal with the problem illustrated in Figure 1 and use the corresponding methods of data
analysis, rather than suffer the indignity of Fisher’s wrath and, more importantly, the waste
of time and money (and sometimes credibility) that failure brings. The good news is that
appropriate protocols and analysis methods exist that are relatively easy to implement and
so maximise the chance of a successful outcome. And once the design is chosen, the data
analysis is straightforward.
This chapter describes the most accessible experimental designs and data analysis
procedures. In most cases the analysis can be carried out using Microsoft Excel’s excellent
statistical toolbox, and reference is made to the appropriate routines. The methods can
be applied to laboratory experiments, pilot plant tests and full-scale plant trials, and no
distinction is made between these other than some discussion at the end of the chapter about
the particular practical problems of conducting plant trials. The quality system Six Sigma has
much in common with the techniques described.
This chapter is not a textbook on statistics. It is intended to introduce metallurgists to
the use of good practice in experimental design and analysis to assist process optimisation.
Discussion is confined to principles, and demonstration of the use of Excel in some of the
more widely used methods. For a more comprehensive discussion and the underlying
mathematics, the interested reader is referred to the references and further reading list,
including Napier-Munn (2014a).
It is therefore very important to conduct the trial in such a way that a firm conclusion is
reached (one way or the other) in the minimum time. The problem, as in all experiments, is
to understand and deal with the basic experimental error. In plant trials, as often as not, this
is apparent mainly as a variation of performance with time. Figure 2 shows the actual daily
variation in recovery from historical plant data. A one per cent recovery improvement may
be worth several million dollars a year in a large operation, but how is one to detect it in data
that varies over a 25 per cent range? The answer is: easily, if you do the right experiment.
1. Strictly speaking one should make a distinction between repeats, which are repeats of a particular measurement, and
replications, which are repetitions of experiments or parts of experiments. Similarly there is a distinction between repeatability,
which for example measures the variability of a particular assay method within a particular laboratory, and reproducibility, which
measures the variability between laboratories.
85
80
Metal Recovery (%)
75 70% ?
70
65 69%
60
55
0 10 20 30 40 50 60
Day Number
s=
/ (x - x)
i
2
(1)
n-1
where:
xi = data point i
x = the mean or average
n = number of data points
(n - 1) is called the degrees of freedom
(The units of s are the same as those of the original data.)
For n = 1, s is indeterminate. A single result therefore carries no information about precision,
and, one might argue, in those cases where little if anything is known about the experimental
system being studied, no information of any kind. So it is important to get into the habit of
performing replicates. A plant trial run over many days or weeks (or even months) produces
many replicates. Laboratory tests can also be done with many replicates. In assessing the
value of a particular quantity (recovery in a laboratory float test for example) replication
directly reduces variance, thus making real differences between conditions (eg reagents)
easier to detect against the background noise of experimental error.
Laboratory practice
Variations in operator, machines, batch of reagent, method, ambient conditions and many
other sources will contribute to error.
Assay
The baseline of error is nearly always an assay: chemical, mineralogical or particle size.
The repeatability of assay methods should be well characterised by the assay chemist or
metallurgist.
The plant
The plant is home to the biggest variabilities of all, usually driven by feed changes over
time, but also by poor feed distribution, shift variations, parallel circuit differences, ambient
conditions, reagent and many other factors. Figure 2 is a good example of the trends, cycles
and random noise evident in concentrator recovery over a two-month period. Figure 3 shows
a consistent bias in recovery between two parallel flotation circuits, Line 1 operating at about
1.3 per cent higher recovery than Line 2, and Figure 4 shows a statistically significant negative
impact of feed rate on recovery.
All these effects conspire to increase the variability of plant data, making it essential to adopt
the correct experimental design procedures to be able to detect differences (improvements)
that are often small compared to the background noise.
95
90
Cu Recovery (%)
85
80
75
70 Line 1
65 Line 2
60
1 3 5 7 9 11 13 15 17 19 21 23 25 27 29 31
Day
100
Cu recovery (%)
95
90
R = -0.31
85
80
200 250 300 350 400 450
Feed rate (t/h)
difference is not zero, ie there is a real difference between the two means: X 1 - X 2 ≠ 0.2 We
will only accept H1 if there is a sufficiently low risk in doing so, and it is the great power
and ingenuity of these statistical methods that allows the risk to be quantified and rational
decisions made on that basis.
Statistical tests of significance generate a probability, P, which is interpreted as the risk of
being wrong in rejecting the null hypothesis, ie concluding that there is a difference when in
fact there isn’t, which statisticians call a Type I error. 100(1 - P) is the confidence level of the
test. So, if the test returns a P-value of 0.05, we infer that there is a five per cent chance of being
wrong in concluding there is a difference, or, conversely, we are 95 per cent confident that
there is a difference. P will always exceed zero so there is always a risk of being wrong – the
only certain things in life are death and taxes!
A question which always arises is: what level of risk (or conversely, confidence) should we
choose as a hurdle rate for making decisions? This will depend on the consequences of the
decision. The lower the value of P we choose the less chance we have of making a mistake
but also the less positive decisions we will make. If we were doing a reagent trial in the plant
and the new reagent was much more expensive than the old, we would need to be more
certain of a difference before recommending the change than if there were no real downside
to the decision.
If you have nothing else to go on, the most common hurdle rate for P used by most
experimenters is 0.05, following recommendations originally given by Fisher in 1926.
However, there is no magic bullet. It would probably be unwise to choose P to be any more
than 0.1 as this gives a one in ten chance of being wrong. If you were in a casino with a
90 per cent chance of success in each game, you would never leave! But here we are dealing
with a different situation: you want to minimise the chance of failure; in fact you would prefer
never to fail, but as we know that P > 0, we want to limit the chance of failure as much as
possible, especially if the decision turns out to be career-limiting.
Now we will consider a number of different methods of making decisions about differences
in quantities of interest. All are based on the principle of the hypothesis test.
COMPARING MEANS
The two-sample t-test
This tests the difference between the averages (means) of two samples of data, which need
not be of the same size. An example is the testing of two different ore sources or two different
reagents using a standard laboratory float test. Example 1 shows a test of a new zinc collector
at Eureka, in which a single sample of ore was ground, split ten ways, and four float tests done
with the old collector (A) and six with the new (B). The null hypothesis is that the difference
in means is zero. Excel’s two-sample t-test is invoked as follows: Data > Data Analysis3 >t-test:
Two-Sample Assuming Equal Variances4 and complete the dialogue box. The result is as
shown in Example 1 (data on the left, Excel output on the right).
2. Strictly speaking, we distinguish between H1: X1 - X2 ! 0 , and H1: X1 - X2 > 0 , the first being a ‘two-sided’ test and the second
a ‘one-sided’ test. More of this later.
3. If the Data Analysis option is not visible, find Add-Ins and check the Analysis ToolPak box.
4. Excel also provides a routine for unequal variances. ‘Unequal’ means statistically different (usually determined using an F-test).
In most cases variances can be considered ‘equal’ even though they may look different. Using the ‘unequal’ routine reduces the
confidence in any observed difference.
EXAMPLE 1
Two-sample t-test for laboratory flotation test.
A B
Mean 86.5 87.65
Variance 0.366667 0.971
Observations 4 6
Pooled variance 0.744375
A B Hypothesised mean difference 0
85.8 86.5 df 8
87.1 89.0 t stat -2.06494
86.2 87.2 P(T< = t) one-tail 0.0364
86.9 86.8 t critical one-tail 1.859548
88.5 P(T< = t) two-tail 0.072799
87.9 t critical two-tail 2.306004
The important numbers to interpret are means and variances, hypothesised mean difference
and P(T<= t).
P(T<= t)
P(T<= t) is the chance (risk) of a Type I error, ie saying there is a difference when in fact there
isn’t one. A low P means that the calculated value of t is unlikely to have arisen by chance
and that the difference is therefore statistically significant. Whether it is one- or two-sided
(‘tail’) will depend on how the objective of the experiment has been defined. In general, if
we are seeking a change in a particular direction (eg an improvement) it is one-sided. If the
question is ‘are the means different?’ (in either direction) it is two-sided. As can be seen from
the Excel output, a two-sided problem has a higher risk and lower confidence than a one-
sided problem.
In this experiment the comparison is one-sided as we are seeking an improvement. P =
0.036, ie there is a 3.6 per cent chance of being wrong in rejecting H0, or alternatively, we are
96.4 per cent confident that the difference is not zero (and our best estimate of the difference
is 87.65 - 86.50 = 1.15 per cent). Is this enough? This is up to the experimenter and will depend
upon the consequences of the decision, but in most cases if P <0.05 we would conclude that the
difference is real.
A useful statistic to calculate is the confidence interval, CI, for the difference in means,
given by:
CI = ta s 1 + 1 (2)
n1 n 2
where:
tα = t-value for a two-sided P-value of α
s = square root of pooled variance (from the Excel output)
n1 = number of results in sample 1
n2 = number of results in sample 2
tα can be obtained from tables of the t-distribution, or from the Excel function TINV(α, df)
where df = degrees of freedom = n1 + n2 - 2, and α is the two-sided P-value for the chosen
confidence level. For example, if a 90 per cent two-sided confidence interval is required, α
is 0.10.
In Example 1, CI95% = ±1.28 per cent, ie the difference in mean recoveries is 1.15 ± 1.28
with 95 per cent confidence. There is no contradiction in the fact that this interval includes
zero, even though our t-test has told us that we are more than 95 per cent confident that
the difference is not zero. The t-test was a one-sided test and the confidence limits here are
two-sided.
In fact, it can often be useful to form a one-sided confidence limit. This is done by using
½ α in Equation 2 and the Excel function T.INV in place of TINV. T.INV produces a one-
sided value for the α level of confidence, preceded by a minus sign. For Example 1, the
one-sided 90 per cent confidence interval is -0.78, and we could interpret this as saying: we
are 90 per cent certain that the recovery improvement is at least 1.15 - 0.78 = 0.37 per cent
(the worst-case scenario). If a cost-benefit analysis requires the difference to be at least
one per cent to get a return on the increased cost, then one might be justified in not
proceeding with the new collector even though it was shown to be better than the old by
more than one per cent, because the risk of failure is too high.
The two-sample t-test is not always appropriate for plant trials because it implies that the
data have not been collected under one of the protocols suggested below. For example, they
may have been collected in the ‘before and after’ fashion illustrated in Figure 1. If so, even
though the t-test may show a real difference, there is no guarantee that the difference is due
to the variable being investigated rather than some other factor that was not controlled by the
experimental design.
d ! ta sd n (3)
where:
sd = standard deviation of the paired differences
n = number of pairs
tα has the same meaning as in Equation 2
Example 2
Two flotation circuits are operated in parallel. Each day, one circuit is run with the usual
collector (A) and the other with a new collector (B). The order is switched randomly between
the circuits each day to avoid any systematic differences between the circuits. The experiment
is run for 12 days, and the daily recovery is monitored. It is required to determine whether the
new collector gives a significantly improved recovery. The results are as shown in Example 2.
The mean difference is 0.88 per cent. The question (as always) is: is that difference real, or
has it arisen by chance due to experimental error? Excel provides a routine to do the analysis:
Data > Data Analysis > t-Test: Paired Two Sample for Means. Again, this is a one-tailed test as
we are seeking an improvement. P(T< = t) = 0.075. Thus, we are 92.5 per cent confident that the
difference is not zero. Is this sufficient confidence? If the new collector were more expensive,
we may wish to improve our confidence by doing more trials (see below for a formula for the
number of trials needed).
TABLE 1
Typical arrangement for paired trial in parallel circuits.
Day Circuit A Circuit B Difference
1 T N T-N
2 N T T-N
3 T N T-N
4 N T T-N
5 T N T-N
6 etc etc etc
EXAMPLE 2
Paired t-test for plant flotation trial.
Day Collector A Collector B t-test: paired two sample for means
1 84.2 85.8 A B
2 71.9 75.0 Mean 79.03333 78.15
3 77.8 76.6 Variance 22.06788 16.88818
4 80.1 82.5 Observations 12 12
5 75.6 77.7 Pearson correlation 0.90725
6 70.8 70.0 Hypothesised mean difference 0
7 81.3 85.0 df 11
8 82.6 80.6 t stat 1.544139
9 78.5 77.1 P(T< = t) one-tail 0.07541
10 79.2 81.5 t critical one-tail 1.795885
11 80.6 82.2 P(T< = t) two-tail 0.15082
12 75.2 74.4 t critical two-tail 2.200985
The power of the paired test is illustrated by using Excel’s two-sample t-test to analyse the
same data. P(T< = t) = 0.31 (not significant) and we would have to accept the null hypothesis
and thus erroneously conclude that Collector B was no better than A.
TABLE 2
Truth table for decision-making.
Decision Truth about the population
H0 true H0 false
(no difference) (true difference)
Reject null hypothesis H0 Type I error Correct decision
P=α P=1-β
Accept null hypothesis H0 Correct decision Type II error
P=1-α P=β
2 (z a + z b) 2 s 2
n= (round n up to a whole number) (4)
D2
where:
zα = normal deviate at a confidence level of α (one or two-sided)
zβ = normal deviate at a confidence level of β (always one-sided)
s = expected standard deviation
n is the size of each sample for the two-sample t-test, and the number of pairs for the paired
t-test. For small samples n should be multiplied by the ratio (f + 3)/(f + 1), where f is the
degrees of freedom: f = 2n - 2 for the two-sample t-test and n - 1 for the paired t-test. zα and zβ
are the standard normal deviates for probability levels α and β. These can be obtained from
tables of the normal distribution, or from the Excel function NORMSINV(1-P) where P is α or
β one-sided (use (1 - P)/2 for two-sided α), or from Table 3. Note that for the paired t-test, the
leading two in Equation 4 is omitted, and n = number of pairs.
After carrying out n experiments, the t-test is then conducted at a confidence level of
α (taking the usual account of whether the test is one or two-sided). If P(t) ≤ α, then the
alternative hypothesis is accepted (there is a real difference) with a probability α of being
wrong (Type I error). If P(t) >α, then the null hypothesis is accepted (no real difference of
size D) with a probability β of being wrong (Type II error). 1 - β, the probability of correctly
concluding that there is a difference, is called the power of the test.
Some textbooks provide power curves (Mason, Gunst and Hess, 1989; Montgomery and
Runger, 1999) or tables (Davies, 1967), allowing the experimenter to read off the number of
trials required from an assumption of α, β and the expected signal-to-noise ratio D/s.
The balance of α and β is important. Because the number of experiments needed to control
both types of error can be large, and increases as α and β decrease, β is sometimes chosen to
be somewhat larger than α to limit the number of tests. Thus if we choose α = 0.05 (95 per cent
confidence level) then we might choose β = 0.2 (80 per cent confidence level), since it is usually
more important to avoid concluding that there is a real difference when there really is not
TABLE 3
Values of z.
Tail α or β
0.3 0.2 0.1 0.05 0.025 0.01
One-sided 0.524 0.842 1.282 1.645 1.960 2.326
Two-sided 0.842 1.282 1.645 1.960 2.326 2.576
(Type I error, controlled by α) than to assert no difference when one really exists (Type II
error, controlled by β). In trials of new operating conditions, for example, this gives more
weight to retaining the status quo.
Example 3
Equation 4 can be used to determine the number of paired trials to be conducted in the
flotation trial of Example 2. Here, s is the standard deviation of the paired differences, sd.5 It is
desired to control the risk of incorrectly concluding that there is a process improvement
when there is not at five per cent (α = 0.05) and to have a 90 per cent chance of detecting a
recovery improvement as large as one per cent – the power of the test (β = 0.1). From previous
experiments it is believed that sd ≈ 2. How many pairs of tests should be conducted? From
Table 3, one-sided z0.05 = 1.645, one-sided z0.10 = 1.282. Substituting directly in Equation 4 with
D = 1 and s = 2, and with the correction for degrees of freedom gives n = 37 pairs.
It is easy to use Equation 4 to construct general curves relating the number of trials required
to the signal-to-noise ratio D/s for a range of confidence levels for one-sided and two-sided
tests. Figure 5 shows such curves, from which it is clear that as the signal-to-noise ratio
declines (either the error increases or the difference required to detect it decreases, or both)
the number of trials required increases rapidly.
5. The sd of the paired differences can be estimated from historical plant data by calculating the sd of 30 or more sequential pairs
of differences.
been no change in the process. If that probability is small (eg ≤ 0.05), then it is concluded that
a real change has occurred. So essentially it is another significance test. However, in this case
the probabilities are measured directly from the historical data.
Suppose that the trial period gives us ten results, with a mean recovery three per cent more
than the mean of the last ten results immediately before the change. The question, as ever,
is: ‘is this difference significant (real) or has it arisen by chance?’. This can be re-stated rather
more formally as: ‘what is the probability with which such a difference would arise, if no
change had occurred?’. This is answered directly by interrogating the reference distribution.
Suppose that the reference distribution (RD) comprises 100 sequential data points. We can
then look at the 100 - 10 + 1 = 91 sequential sets of ten results (1–10, 2–11, 3–12 etc) and see how
many times one set of ten (eg 1–10 or 2–11) is followed by another set of ten (eg 11–20 or 12–21)
with a mean recovery higher by three per cent. If this occurs rarely in the RD (eg ≤5 per cent
of times), then the trial result must be regarded as significant.
The problem with this method is the key assumption that the reference distribution is
stable and represents the constant condition of the process before the change. This is an
assumption which is often not met – the process mean moves around and it is difficult to
define the appropriate RD. The cusum chart (see below) can help to determine whether the
RD is homogenous. Also, it cannot help with the nuisance influence of covariates such as feed
grade, and it is a less efficient test than the t-test, generally requiring many more data points.
But it is free of the statistical assumptions required by many of the other methods.
Outliers
Outliers may be cynically regarded as inconvenient points, ‘wild’ results that do not appear to
belong to the rest of the data set or perhaps conflict with our prejudices. They are best identified
by plotting the data. They can often have a significant effect on hypothesis tests, mainly due
to the substantial increase in standard deviation that they cause. As a general rule, however,
outliers should not be rejected without some formal justification. For example, if it is found
that a mistake was made in collecting the data, such as a biased sample, or a miscalibrated
instrument, then the relevant point(s) can be rejected. If, however, no such reason can be
found but suspicion persists that they do not belong, a formal test can be invoked.
Several such tests have been devised, but there is no consensus amongst the statistical
community and certainly no magic bullet. Mason, Gunst and Hess (1989) recommend the
Grubbs test for a normally distributed sample whose mean is of interest. This test is more
powerful than some others and can reject more than one outlier at a time. The test statistic
is a ratio of sums of squares, (xi - x ) 2 , the SS of the data without the outlier(s) being divided
by the SS of all the data with the outlier(s), giving a ratio less than one. The column of data
should first be sorted using the Excel Data >Sort option. The outlier(s) will be at the top
or bottom of the column depending on whether they are the highest or lowest values and
how the sort has been done. The Excel function DEVSQ (array) is then used to calculate the
two sums of squares, with and without the outliers. The ratio of these is then compared
with Table A26 in Mason, Gunst and Hess (1989) (also given as Table A1.3 in Napier-Munn
(2014a)) with n = total number of data points and k = number of outliers. (The values in the
table have been multiplied by 103 to avoid the confusion of decimal points.) If the calculated
ratio is less than the tabulated value at the desired α-value then the outlier(s) should be
considered for rejection.
Example 4
In a large feasibility study, several operators in different laboratories are involved in laboratory
float tests. It is important to ensure that the operators do not produce statistically different
results. To establish whether they all give similar results, a single ore sample is divided and
each of four operators conducts a number of repeat tests on the sample. The results are shown
on the left side of Example 4. On the right is the Excel output invoked by Data > Data Analysis
> ANOVA: Single Factor. The full block of data is selected, including the labels in the first row,
and the ‘Labels’ box in the dialogue box is checked.
The summary gives the number of tests for each operator, and the corresponding mean
recoveries and variances. The ANOVA table provides the key information to judge whether
there are differences between the operators (see Example 4). This is done by comparing
the mean squares (MS) of the ‘Between Groups’ and ‘Within Groups’ variation, which
have units of variance. The ‘Between Groups’ variance is due to real differences between
the operators, and the ‘Within Groups’ variance is experimental error, measured from the
repeats done by each operator. The comparison is made by forming the ratio of the two MS
values, which is called the F-ratio (after Fisher, who invented it); the P-value then gives the
EXAMPLE 4
The one-way analysis of variance (ANOVA) in comparing operators.
ANOVA: Single factor
Op 1 Op 2 Op 3 Op 4 Summary
67.3 60.5 63.5 60.7 Groups Count Sum Average Variance
66.2 58.9 64.9 64.7 Op 1 4 261.42 65.35393 3.222062
63.3 54.1 66.4 67.3 Op 2 3 173.45 57.81624 10.82865
64.6 64.3 Op 3 3 194.78 64.92636 1.993553
Op 4 4 256.97 64.24125 7.376275
ANOVA
Source of variation SS df MS F P-value F crit
Between groups 118.557 3 39.51887 6.880095 0.00855 3.7083
Within groups 57.4394 10 5.743942
Total 175.996 13
Type I error risk. In this case P = 0.009, ie there is less than one per cent risk in saying that
the operators do indeed give different results, or conversely, we are more than 99 per cent
sure that they are different.
A glance at the summary shows that Operator 2 produced a lower mean recovery than the
other three, and therefore some mentoring or re-training might be in order.
A useful piece of additional information is the MS for ‘Within Groups’, which is the
experimental error variance based on the repeats. The square root of this value (2.4 per cent)
is the error standard deviation, which can be used in planning future experiments, eg in
Equation 4. In this case it should be calculated after Operator 2’s data have been removed.
Example 5
A trial is conducted on a concentrator with three parallel closed grinding circuits. The purpose
is to increase the fineness of grind by using an alternative cyclone geometry (a combination
of vortex finder and apex). After doing simulations and consulting the manufacturer, three
likely geometries are selected for test, plus the current arrangement. It is thought likely that
the circuits do not operate exactly the same but it is important that the conclusions apply
to all the circuits. A randomised block design is therefore run, with ‘circuit’ the blocked
variable and each of the four geometries being tested in random order once in each circuit.
A composite sample of cyclone overflow is collected for each circuit/geometry combination
and the per cent -75 μm determined.
The results are shown on the left side of Example 5. On the right is the Excel output invoked
by Data > Data Analysis > ANOVA: Two Factor Without Replication. The full block of data is
selected, including the labels in the first row and column, and the ‘Labels’ box in the dialogue
box is checked.
Now there are three sources of variation: the geometries (rows), the circuits (columns) and
error. The null hypotheses are that the four geometries give the same product size, and the
three circuits also give the same product size. The P-values for the F-tests show that both
hypotheses can be rejected: the geometries and the circuits are significantly different. The
geometry effect is significant at >95 per cent confidence (P = 0.043). We can therefore choose
the best cyclone with the finest product (from the summary – Cyclone C).
The circuit effect is significant at 93.6 per cent (P = 0.064), which is probably sufficient in a
trial of this kind to conclude that the difference is real, as suspected. Inspecting the summary
shows that Circuit 3 produces a coarser product than the other two. This bonus information
can be used to explore why this is so and perhaps correct Circuit 3 so that an additional
process improvement can be achieved.
The square root of the error MS (2.5 per cent -75 μm) is the experimental standard deviation
for the trial.
Note that if the experiment had not been blocked, ie not arranged as a formal randomised
block design, the block variation (columns) would not have been distinguishable from the
error. The consequence of this can be simulated by analysing the data as a one-way ANOVA,
as in Example 4. In this case the circuit labels are not selected as ‘data’, so the analysis simply
compares the four cyclone geometries against an inflated error variance (now including
the circuit differences). P = 0.11, which would suggest that there is insufficient evidence to
conclude that the geometries gave different product sizes.
EXAMPLE 5
The two-way analysis of variance (ANOVA) in comparing cyclone geometries.
ANOVA: Two-factor without replication
Cyclone Circuit 1 Circuit 2 Circuit 3 Summary Count Sum Average Variance
A 73.8 74.2 69.0 A 3 217 72.33333 8.373333
B 68.9 75.5 65.2 B 3 209.6 69.86667 27.22333
C 79.2 78.5 73.8 C 3 231.5 77.16667 8.623333
D 72.5 69.6 70.0 D 3 212.1 70.7 2.47
FACTORIAL DESIGNS
The factorial is a widely used and powerful experimental design, and is comprehensively
covered in the literature (eg Davies, 1967; Box, Hunter and Hunter, 1978; Mason, Gunst
and Hess, 1989; Napier-Munn, 2014a). It is the basis of the Evolutionary Operation (EVOP)
process optimisation method discussed briefly below, but is most widely used in the
laboratory environment. It will therefore not be dealt with in detail here; the reader is
referred to the literature.
The key feature of a full factorial is that each factor (variable) is investigated at every level
of every other factor, ie a measurement is made at every factor combination. The number of
experiments needed is cxn, where x is the number of levels, n is the number of factors and c
is the number of replicates of the whole experiment. If three factors are investigated at two
levels each without replication, the number of tests required is 23 = 8. A 23 factorial can be
thought of as a cube in space, with each of the three dimensions representing one of the three
factors, A, B and C. A test is run at each of the eight corners of the cube (see Figure 6).
The analysis falls into two parts: the calculation of the actual effects (how each factor affects
the response), and testing of the significance of each effect using an ANOVA. Some estimate
of error is required for the ANOVA – either a full replication of the experiment, or a number
of centre point runs at the centre of the experimental space (as shown in Figure 6), or some
other source. No specific routines are available in Excel for the analysis of factorials, though a
two-level factorial can be analysed by a simple method suitable for spreadsheets, called Yates’
analysis (Box, Hunter and Hunter, 1978; Mason, Gunst and Hess, 1989; Napier-Munn, 2014a).
The two-level factorial only detects linear relationships between factors. This is satisfactory
in many cases. However, three or more levels are required to detect non-linear effects. A
specialist statistics package is best used both for the design and analysis of the factorial and
its many variants.
The design is not limited to the same number of levels for each factor. However, a
disadvantage is that many tests are required if the situation calls for several levels and/or
several factors, the number of tests increasing exponentially with the number of factors. In
such cases subsets of the full design can be used, called fractional factorials. Special designs
such as the central composite rotatable design (CCRD) are also available. The CCRD is able to
investigate quadratic effects for the development of response surface models in fewer tests
than would be required for a full three-level factorial; the models can be used for process
optimisation (Napier-Munn, 2014a, 2014b). CCRDs are particularly useful for laboratory and
pilot plant experiments, but in recent years some have also been successfully conducted in
sections of production flotation plants (eg Venkatesan, Harris and Greyling, 2014).
The factorial has important advantages over other less formal approaches to experimentation,
including:
•• It is mathematically the most efficient way of estimating the factor effects. This is
particularly useful in collecting data for modelling (see below).
•• It is necessary to detect interaction, in which the effect of one factor depends systematically
upon the level of one or more other factors.
•• It can incorporate blocking and randomisation, which are important weapons in dealing
with error to improve the ability of the experiment to detect important effects.
•• The conclusions hold over a wide range of conditions.
where:
Y = performance variable of interest, or dependent variable (eg recovery)
X = uncontrolled variable (eg feed grade)
T = a dummy variable representing the trial condition: T = 0 for the old condition
and T = 1 for the new condition6
a0, a1, a2 are coefficients to be estimated from the data
ε represents the residual error (noise)
X and T are independent variables. The null hypothesis here is that the trial condition does
not influence recovery, ie that a2 = 0.
6. If the trial variable has several quantifiable levels (eg collector concentration) then of course it can be represented as a normal
quantitative variable.
Example 6
A plant trial was conducted to test a new flotation reagent. An appropriate protocol (eg a
paired trial) could not be conducted, so data using the new reagent were compared with a
historically equivalent period when using the plant’s usual reagent. Equation 5 is fitted to
the data by invoking the Excel regression tool: Data > Data Analysis > Regression, which
estimates the linear parameters (coefficients) in the model using the least squares method.
The Y-range in this case is the column of Recovery, and the X-range the contiguous block
of feed grade and trial condition, including the labels. Example 6 shows part of the data set
on the left, and the Excel output on the right (excluding the 95 per cent confidence limits
for clarity).
The key parts of the output can be interpreted as follows. The significance in the ANOVA
(refer Example 6) (1) is the P-value for a test of whether the whole regression is significant.
A value much less than 0.01 is desirable, and certainly less than 0.05. The trial coefficient (2)
is an estimate of the magnitude of the effect that the trial condition has had on recovery: in
this case an increase of 1.48 per cent. (Note that this will not be equal to the difference in the
mean values, which is 1.8 per cent, because it is essentially a mean weighted towards the
centre of gravity of the data.) It is essential as always to test whether this is significant, using
the corresponding standard error (3) and t-value. This is done as usual with the P-value (4),
in this case 0.046, which is significant at the 95 per cent level of confidence. (Similar P-values
are available for the grade coefficient (5) and intercept, both of which are highly significant.)
EXAMPLE 6
Regression model of a plant trial with intervention analysis.
Feed grade Trial Recovery Summary output
0.364 0 66.9 Regression statistics
0.320 0 63.7 Multiple R 0.59745195
0.453 0 66.7 R square 0.35694883
0.413 0 75.0 Adjusted R square 0.34514973
0.617 0 76.8 Standard error 3.86107058
0.503 0 82.9 Observations 112
0.500 0 73.4
0.518 0 72.6
… … … ANOVA
0.560 1 68.8 df SS MS F Significance F
0.493 1 68.1 Regression 2 901.9914438 450.996 30.2522 3.54308E-11
0.590 1 75.8 Residual 109 6 1624.957399 14.9079 1
0.697 1 73.3 Total 111 2526.948843
0.560 1 73.6
0.570 1 74.6 Coefficient Standard error t stat P-value
0.520 1 66.0 Intercept 53.4035743 2.180237461 24.4944 4E-46
0.537 1 71.0 Grade 32.4636222 4.40265655 7.37364 3.4E-11 5
0.580 1 74.7 Trial 2 1.47612985 0.731053365 3 2.01918 0.04592 4
Note that Excel returns a two-sided P-value for the coefficients. If a one-sided value is
required, as in this case (we are looking for an improvement), then use the Excel function ‘=
TDIST(t, df, 1)’. t is the value of t for the trial coefficient (2.02 in this case), df is the degrees
of freedom for the residual (6) (109 in this case), and ‘1’ implies a one-sided test (use ‘2’
for a two-sided test). In this case TDIST returns the value P(t) = 0.023; so we are more than
97 per cent confident that the new reagent gives a higher recovery than the old.
The standard error for the trial coefficient (3) can also be used to calculate one- or two-sided
confidence limits, as before. The 95 per cent two-sided confidence interval is approximately
1.96 × standard error = ±1.4 per cent. So we are 95 per cent confident that the trial conditions
produces an increase in recovery, and our best estimate of that increase is 1.5 per cent
± 1.4 per cent. (These limits are given in the Excel output but not shown in Example 6).
Example 7
Historical data for 114 values of daily feed rate and feed grade were available from two
periods running with a different flotation reagent. Trial = 0 indicates the period running with
the old reagent and Trial = 1 the period with the new (see Table 4).
Recovery is the dependent variable (Y) and feed rate, feed grade and ‘trial’ (the dummy
variable) are the independent variables (X1, X2, T). Choosing Excel’s Data > Data Analysis >
Regression will produce the basic output as shown in Example 7 (leaving out the ANOVA
for clarity).
The P-value (two-sided) for the trial coefficient (1) is 0.310, and the one-sided P-value
calculated from the function TDIST is 0.155. Neither is statistically significant, and the main
conclusion of the trial is that the new reagent has no significant effect on recovery.
TABLE 4
Data from plant trial for Example 7.
Day Feed rate (t/h) Feed grade (%) Trial Recovery (%)
1 354 1.63 1.00 90.1
2 364 1.71 1.00 90.3
3 343 1.62 1.00 90.5
4 324 1.75 1.00 91.8
5 365 1.70 1.00 89.9
6 373 1.39 1.00 87.6
… … … … …
109 352 2.06 0.00 92.8
110 353 1.87 0.00 91.4
111 337 1.95 0.00 93.1
112 343 2.05 0.00 93.3
113 328 1.93 0.00 93.0
114 334 1.86 0.00 92.1
EXAMPLE 7
Multi-variable regression model of a plant trial with intervention analysis.
Summary output
Regression statistics
Multiple R 0.81970252
R square 0.67191222
Adjusted R square 0.66296437 2
Standard error 1.06188161 3
Observations 114
However, the model is quite interesting in its own right. The model itself is:
Recovery (per cent) = 88.59 - 0.019 (feed rate) + 5.49 (feed grade) (6)
This simple process description gives us additional useful information from the trial.
An increase of 0.5 per cent in feed grade increases recovery by about 2.7 per cent, and an
increase of 50 t/h decreases recovery by about one per cent.
Excel provides statistics to judge the quality of the model. The adjusted R2 (2) is the coefficient
of multiple determination (adjusted for the degrees of freedom), which is a goodness-of-fit
criterion describing the proportion of the total data variation accounted for by the model (the
balance would be due to missing terms, model inadequacies and error). R2adj = 66 per cent is
high for plant data of this kind in which values of less than 50 per cent are not uncommon,
and suggests that the model gives reasonable predictions. This is confirmed by the standard
error (3), which is the standard deviation of the residuals, the difference between the observed
and predicted values of recovery. 1.96 × standard error gives approximately the 95 per cent
confidence interval of the predictions of the model, in this case ±2.1 per cent.
Regression models of this kind are helpful but users should be aware of their limitations
and dangers:
•• They are simplistic models and should not be over-interpreted. In particular they will not
apply to different process conditions and so should not be extrapolated.
•• The model may be biased if terms (eg other significant process variables) have been omitted.
•• Outliers should be removed where appropriate, and the regression re-run. These can
be identified by plotting observed Y (on the y-axis) versus predicted Y (on the x-axis) –
this is a useful plot anyway, giving a visual indication of the goodness-of-fit. The points
should be scattered around the 45° line, the closeness of the points to the line indicating
the quality of fit. Outliers will lie outside the general cluster of points. A more rigorous
way of rejecting them is to select Standardised Residuals from the regression dialogue box.
These are residuals divided by a standard error and form a kind of t-test. Any absolute
value greater than about 2.5 can be considered for rejection with 99 per cent (two-sided)
confidence; though don’t forget that a large data set would be expected to have one per
cent of its residuals of this value or greater by chance alone.
•• The least squares method is based on a number of assumptions. One of the most important
is that the residuals (predicted – actual values) are normally distributed; this can be visually
checked using a histogram (the Excel regression dialogue box permits you to produce a
column of residuals and predicted values). Another is that the errors are independent; this
assumption is violated when the data are serially correlated, and the time relationship
would need to be accommodated in the model (see below).
Regression analysis is a big subject and interested readers should consult appropriate
texts for further details (eg Box and Draper, 1987; Box, Hunter and Hunter, 1978; Chatfield,
1983; Draper and Smith, 1998; Montgomery and Runger, 1999; Snedecor and Cochran, 1968;
Napier-Munn, 2014a).
yt = μ + φ(yt-1 – μ) + et (7)
where:
yt = performance (eg recovery) in current time period (eg today)
yt-1 = performance in previous time period (eg yesterday)
φ = autocorrelation coefficient (0 ≤ φ ≤ 1), which measures how strongly yt depends on yt-1
μ = process mean: the value about which the process is randomly oscillating
et = error for time t: a measure of the residual error after the model is fitted
Equation 7 can be rearranged in the form of a linear expression, which can be easily fitted
by Excel’s regression tool or ‘Add Trendline’ in Chart.
Here yt is the column of daily data and yt-1 is this column displaced down by one cell. (The
first yt value and the last yt-1 value must be eliminated to produce two columns with the same
number of data points.) a0 (the intercept of the fitted line) = μ(1 – φ), and a1 (the gradient) = φ,
and so μ = a0/(1 – a1).
Now we invoke Equation 5 to perform intervention analysis on the time series model:
yt = a0 + a1(yt-1) + a2T (9)
where again:
T is the presence or absence of the trial condition
The process shown in Example 6 is then followed.
Example 8
One type of experiment that is difficult to do on a plant scale is the testing of grinding media,
which cannot be switched easily to conform to a particular experimental protocol such as a
paired t-test. In this example a study was made of the effect of two types of grinding media
on final concentrate grade from the flotation plant. A homogeneous period using the first
type of media was taken as representing that condition (T = 0). The new media was then
progressively added to the mill and a second period of data collected once the mill was fully
charged with the new media (T = 1). Three hundred and ten days of data were available in all.
The data and Excel Regression output were as shown in Example 8.
The media coefficient is -1.31 per cent and its two-sided P-value 1.8E - 06, which is highly
significant. We may therefore conclude that moving from Media 0 to Media 1 reduced the
concentrate grade by 1.3 per cent.
Note that the concentrate grade (t - 1) coefficient is 0.45. This is a measure of the dependency
of each day’s concentrate grade (t) on the previous day’s figure (t - 1), and is also highly
significant. This confirms the validity of using the AR(1) time series model.
EXAMPLE 8
Modelling a plant trial as a time series.
Media Conc grade (t-1) Conc grade (t) Summary output
0 34.34 33.27 Regression statistics
0 33.27 32.86 Multiple R 0.6092497
0 32.86 33.11 R square 0.3711852
0 33.11 30.44 Adjusted R square 0.3670887
0 30.44 29.99 Standard error 2.0754444
… … … Observations 310
1 28.50 28.39
1 28.39 28.79
1 28.79 28.50 Coefficient Standard error t stat P-value
1 28.50 27.23 Intercept 16.531569 1.535578721 10.766 3.83E-23
1 27.23 29.85 Media -1.312924 0.26970263 -4.868 1.81E-06
1 29.85 29.55 Conc grade (t-1) 0.4530328 0.050493461 8.9721 2.95E-17
A ‘cusum’ chart is a time sequence plot of the cumulative sum of the current value minus
some mean value plus the previous cusum. In symbols:
Ct = Ct-1 + Rt – µ (10)
where:
Ct = cusum at time t
Ct-1 = cusum at time t – 1
Rt = value (eg recovery) at time t
μ = some mean or target value (often the process mean over the period)
Figure 8 (top) shows a normal time series of mill P80 over an 18-month period from a gold
plant. Apart from the usual noise, there appear to be three distinct periods of prevailing
P80, a low leading to a high and then back to a low again. The bottom image in Figure 8 is
the cusum series of the data, using a mean value μ = 174 μm (the mean of the whole series).
The points of change are now very clear, and careful inspection of the chart shows that the
changes occurred on days 130 and 404. This information can now be used forensically to
explore the reason(s) for the change.
In cusum charts, the absolute value of the cusum at any point is not important. It is the
gradient of the line over a characteristic period that indicates the prevailing process mean
over that period, and changes in gradient indicate changes in process mean. A change in gradient
can be from positive to more or less positive, negative to more or less negative, or more
dramatically (as in Figure 8) from negative to positive or vice versa. If the gradient over a
given period is negative then the prevailing mean is less than μ, if positive then the prevailing
mean is more than μ, and if flat then the prevailing mean is equal to μ.
In constructing a cusum chart, therefore, the choice of μ is important, and it is usually
worth trying several to see which gives the most obvious changes in gradient. This is easy if
Equation 10 is implemented in a spreadsheet. Typical choices are the total mean of the data
(as in Figure 8, for which the cusum of the final point will always be zero), some target value
for the process being sought, or the mean for the period before a known change is made to a
process variable.
In many cases, changes in gradient can be assessed by eye. However, rigorous methods are
available for determining if the change in gradient at any given point is statistically significant
(Himmelblau, 1970; Montgomery and Runger, 1999; Napier-Munn, 2014a).
300
250
P80 (micron)
200
150
100
50
0
0 100 200 300 400 500
Day
3000
2000
Cusum of P80
1000
-1000
-2000
0 100 200 300 400 500
Day
EVOLUTIONARY OPERATION
Evolutionary Operation, or EVOP, has been widely used in some process industries and has
been advocated for mineral processing (Mular, 1971; Mular and Bull, 1973; Napier-Munn,
2014a, 2014b) but never seems to have caught on. It is not clear why this should be, as its
features suit the characteristics of concentrators rather well. EVOP is a process optimisation
procedure in which simple factorial designs are used to explore and approach the optimum
performance in a noisy process. The range of each factorial is small to prevent damaging
interference to the process. The basic experiment is replicated until effects become statistically
significant, and the design follows the contours or response surface of the process until it detects
a maximum. It is one of a broad class of optimisation methods called hill climbing techniques,
and has a number of variants including SSDEVOP: simplex self-directed EVOP. It is easy to
implement and to calculate, and is promoted as a method in which plant operators can play
an important role. It does, however, require a commitment to the process over long operating
periods. The method is covered in detail in Mular and Bull (1973) and Napier-Munn (2014a)
and is also discussed in Box, Hunter and Hunter (1978) and Himmelblau (1970).
Of these, the paired t-test is probably the most efficient if only two conditions are being
tested. The randomised block design is appropriate where more than two conditions are
being trialled. Modelling using intervention analysis will generally require more data than
the other two, and is normally used only where practicalities prevent the use of a formal
design, or in the analysis of historical data. Some comments on each follow.
Clearly it is no good running for a time (a day, week or month) with one condition, then
switching to the other condition for the same period, and then comparing mean performance
over the two periods, as was done in the experiment illustrated in Figure 1. Other factors
may have changed with time which would interfere with the comparison, most notably ore
type or grade. The error would also be high and uncontrolled. The paired trial or randomised
block design are intended to deal with this situation. In both cases, it is necessary to choose a
duration for each replicate of the experiment, eg for each member of the pair. Selection of an
appropriate time period will depend on:
1. How the plant does its metallurgical accounting and reconciliation.
2. Material residence time in the process (the period must be selected to be long compared
to mean residence time; at least three residence times are required, and preferably much
more – five to ten).
3. The period of large-scale changes and trends in process performance, such as those evident
in Figure 2. The replicate times should be short compared to this period.
Common durations for a replicate are a shift or a day. A day is preferred, as shift performance
often varies systematically over the 24-hour cycle, and a day is usually long compared with
mean residence time but short compared with long-term trends. However, the longer the
period, the longer (and more expensive) is the total duration of the experiment.
The switching order in the trial must be randomised. A problem arises if it is difficult
to switch condition, eg when a major item of equipment is to be included and excluded,
although this is often easier than it seems if a little thought is given to the problem. In such
cases the switch can be made less frequently but the trial will have to proceed for longer to
ensure a time bias does not accumulate.
In some cases identical parallel circuits are not available. Here the pairs are generated
sequentially in time. It is often considered that after switching a condition the plant requires
a full period (eg a day) to settle down, and the performance of the day following any change
should therefore be ignored. The order of each pair must also be randomised. This produces a
sequence of groups of three or four days: four days when the first day of the group is different
to the last day of the previous group, and three days when the first day is the same as the last
day of the previous group and no day is required for stabilisation. A typical sequence is as
shown below (groups numbered in sequence; N = normal and T = trial condition). The pairs
used in the analysis are underlined:
If 16 ‘pairs’ were required to detect a difference in this case, then the experiment would run
for 55 days, assuming no breaks.
A checklist
In planning a plant trial (or any experiment), work through the following checklist:
1. Define the objective(s) of the trial; write them down and be specific. Why are you doing it?
What will be your actions depending on the outcome?
2. Formulate the objectives in experimental terms. What experiments need to be conducted
to achieve the objectives?
3. What response(s) and factors are to be selected for investigation? What do we know about
the random variation of the response(s) – ie experimental error? Conduct pilot-scale tests
if necessary.
4. Over what range do we need to alter the factors? Are we interested in knowing factor inter
action? Is there any unwanted or uncontrollable variable whose effect must be eliminated?
5. Taking into account all of these considerations, plus the time, money and labour available,
and the possible effect on current operating processes, select an appropriate experimental
design.
REFERENCES
Box, G E P and Draper, N R, 1987. Empirical Model-Building and Response Surfaces (John Wiley and Sons:
New York).
Box, G E P, Hunter, W G and Hunter, J S, 1978. Statistics for Experimenters (John Wiley and Sons: New
York).
Bruey, F and Briggs, D, 1997. The REFDIST method for design and analysis of plant trials, in Proceedings
Sixth Mill Operators’ Conference, pp 205–207 (The Australasian Institute of Mining and Metallurgy:
Melbourne).
Cavender, B W, 1993. Review of statistical methods for the analysis of comparative experiments, SME
Annual Meeting, Reno, February, Preprint 93–182 (Society for Mining, Metallurgy, and Exploration:
Littleton).
Chatfield, C, 1983. Statistics for Technology, third edition (Penguin).
Davies, O L (ed), 1967. Design and Analysis of Industrial Experiments (Oliver and Boyd: London).
Draper, N R and Smith, H, 1998. Applied Regression Analysis, third edition (John Wiley and Sons: New
York).
Fisher, R A, 1926. Statistical Methods for Research Workers (Oliver and Boyd).
Fisher, R A, 1938. Presidential Address to the First Indian Statistical Congress, Sankhya, 4:14–17.
Himmelblau, D M, 1970. Process Analysis by Statistical Methods (John Wiley and Sons: New York).
Lang, A, 1910. Speech quoted in Alan L Mackay The Harvest of a Quiet Eye (1977), as reported in Chambers
Dictionary of Quotations (2005), p 488.
Mason, R L, Gunst, R F and Hess, J L, 1989. Statistical Design and Analysis of Experiments (John Wiley
and Sons: New York).
Montgomery, D C and Runger, G C, 1999. Applied Statistics and Probability for Engineers (John Wiley and
Sons: New York).
Morrison, R D (ed), 2008. An Introduction to Metal Balancing and Reconciliation (JKMRC, The University
of Queensland: Brisbane).
Mular, A L, 1971. The selection of optimization methods for mineral processes, in Proceedings Decision-
Making in the Mineral Industry, special volume 12, pp 442–450 (Canadian Institute of Mining,
Metallurgy and Petroleum: Montreal).
Mular, A L and Bull, W R, 1973. Mineral Processes: Their Analysis, Optimization and Control, course
manual, 1969–1973.
Napier-Munn, T J, 1998. Analysing plant trials by comparing recovery-grade regression lines, Minerals
Engineering, 11(10):949–958.
Napier-Munn, T J, 2008. Some practical problems in running statistically valid plant trials, and their
solution, in Proceedings Metallurgical Plant Design and Operating Strategies (MetPlant) 2008, pp 249–258
(The Australasian Institute of Mining and Metallurgy: Melbourne).
Napier-Munn, T J, 2014a. Statistical Methods for Mineral Engineers: How to Design Experiments and Analyse
Data (JKMRC, The University of Queensland: Brisbane).
Napier-Munn, T J, 2014b. Cunning solutions to process improvement, in Proceedings 12th AusIMM Mill
Operators’ Conference, pp 5–10 (The Australasian Institute of Mining and Metallurgy: Melbourne).
Napier-Munn, T J and Meyer, D H, 1999. A modified paired t-test for the analysis of plant trials with
data autocorrelated in time, Minerals Engineering, 12(9):1093–1100.
Snedecor, G W and Cochran, W G, 1968. Statistical Methods, sixth edition (Iowa State University Press:
Oxford).
Venkatesan, L, Harris, A and Greyling, M, 2014. Optimisation of air and froth depth in PGM flotation
using a CCRD design, Minerals Eng, 23:621–626.
FURTHER READING
Cochran, W G and Cox, G M, 1966. Experimental Designs, second edition (John Wiley and Sons).
Davies, O L and Goldsmith, P L (eds), 1972. Statistical Methods in Research and Production, fourth edition
(Oliver and Boyd).
Napier-Munn, T J, 1995. Detecting performance improvement in trials with time-varying mineral
processes – three case studies, Minerals Eng, 8(8):843–858.
CHAPTER 12
Project evaluation and communication
Joe Pease
ABSTRACT
This chapter is about the final stage of project approval – economic evaluation and
communication to ‘sell’ the project. It is an exciting stage and easy compared to the work you
have done to reach this point. But be careful not to underestimate it; many poor projects are
approved and many good projects rejected because of mistakes made at this stage. Economic
calculations are easy but depend entirely on the quality of data and assumptions. Meanwhile,
the approach to ‘selling’ a project is quite different from the rigorous engineering process you
needed to get to this point.
The chapter starts with a discussion on some of the simple evaluation techniques used to
assess projects. More importantly, it highlights the dangers in such techniques; that is, their
simplicity can conceal the unavoidable uncertainty in your data. The really hard work in the
project evaluation is retrieving the right data and knowing its limitations.
We then look at how to increase the chances of project approval. This is where engineering
meets marketing and psychology. The directors might be clever, hard-working people, but
they may not be metallurgists so you cannot assume they grasp the details; they need to ration
capital between an extensive list of competing projects. Because you are trained in the scientific
method you avoid quoting anything that is not backed up by hard data. Some of the groups
competing for funds are not so rigorous, but they do fantastic PowerPoint presentations.
So how do you ensure that the company considers your project fairly and makes the right
decision? Importantly, communication is an ongoing process that starts in the earliest stages
of the project. Indeed, it starts with selecting the right project to work on in the first place.
EVALUATION TECHNIQUES
This section is deliberately brief. There are many well-qualified texts that demonstrate the
theory and mechanics of project evaluation methods and risk analysis. Calculators have
built in net present value (NPV) and internal rate of return (IRR) functions and spreadsheets
are ideal for laying out an economic evaluation. These techniques can be easily researched
elsewhere. The crucial task is to use the right numbers in the evaluation. This can take years,
whereas the economic model will take just hours. Spend your time making sure you look at
the right project in the right way and collect accurate numbers. Your company will want a
financial analyst to review your project against the company standards and you’ll need to
help. Understand what is important to their analysis and make sure they understand the
basic metallurgy and limiting mineralogy behind the project. Encourage them to ask you
tough questions – better to deal with them now than at the Board Meeting. If you establish
mutual professional respect with the financial analyst you’ll have an influential ally.
NPV, $ M
15
Change in revenue
12 – eg production,
recovery, prices
Change in capital or
operating cost (lines 9
overlap)
3
1 year commissioning delay
0
-30% -20% -10% 0% 10% 20% 30%
-3 % change in assumption
FIG 1 – Sensitivity of net present value (15 per cent) to key assumptions.
Further, if the inflation rate is four per cent, you’ll need $104 next year to buy what $100 would
buy you today. In this case $104 is called ‘nominal’ dollars and $100 is ‘real’ dollars (today’s
value of the future amount). You need to account for both interest and inflation. Usually
inflation is handled by doing evaluations in ‘real’ (today’s) dollars. Your company will have its
own procedures for forecasting inflation, costs and metals prices. Just make sure all numbers
are compatible. If you are using today’s dollars, the discount factor, metal price forecasts and
operating cost forecasts all need to be compatible and expressed in today’s dollars.
As well as allowing for interest and inflation, discount rates usually add a risk factor. Projects
and future assumptions are risky. The company needs a margin on projects to allow for risk
and still leave room for profit. The risk factor will vary with the perception of risk, the need
for profit and the company’s financial situation. Often a high risk factor is used to ‘ration’
capital and resources to select only the highest performing projects.
TABLE 1
Simple financial analysis spreadsheet.
No Item Year 0 Year 1 Year 2 Year 3 Year 4 Year 5 Years 6–13 Year 14 Year 15 Note
1 Capital cost -10 000 000 1
2 Additional commisioning -500 000 2
3 Additional operating -1 800 000 -2 000 000 -2 000 000 -2 000 000 -2 000 000 etc -2 000 000 0 3
4 Ramp-up profile 81.3% 100% 100% 100% etc 100% 0% 4
5 Increased concentrate tonnes produced 12 188 15 000 15 000 15 000 etc 15 000 0 5
6 Net value of one tonne of extra concentrate 500 500 450 400 etc 400 6
7 Net sales revenue of extra concentrate 6 093 750 7 500 000 6 750 000 6 000 000 etc 6 000 000 7
8 Net salvage value (after decommissioning costs) 750 000 8
JOE PEASE
9 Net operating cash flow before tax -10 000 000 -2 300 000 -4 093 750 5 500 000 4 750 000 4 000 000 etc 4 000 000 750 000 9
10 Allowable depreciation on new equipment -1 500 000 -1 500 000 -1 500 000 -1 500 000 -1 500 000 Stops in year 7 10
11 Extra operating profit before tax -3 800 000 2 593 750 4 000 000 3 250 000 2 500 000 etc 4 000 000 750 000 11
12 Extra tax paid 1 140 000 -778 125 -1 200 000 -975 000 -750 000 etc -1 200 000 -225 000 12
13 Cash flow after tax (real dollars, undiscounted) -10 000 000 -1 160 000 3 315 625 4 300 000 3 775 000 3 250 000 etc 2 800 000 525 000 13
FLOTATION PLANT OPTIMISATION | SPECTRUM SERIES 25
14 Cumulative cash flow (after tax, undiscounted) -10 000 000 -11 160 000 -7 844 375 -3 554 375 230 625 3 480 625 etc 46 302 813 46 827 813 -
15 Payback period 4 years 14
16 At 15 per cent discount rate: discounted cash flow -10 000 000 -1 008 696 2 507 089 2 158 369 2 158 369 1 615 824 etc 395 720 64 520 15
17 NVP (15 per cent) 11 543 000 16
18 At 30 per cent discount rate discounted cash flow -10 000 000 -892 308 1 961 908 1 957 214 1 321 732 875 319 etc 71 113 10 257 17
19 NVP (30 per cent) 986 100 -
20 Internal rate of return (IRR) 33% 18
FLOTATION PLANT OPTIMISATION | SPECTRUM SERIES 25
All dollar values estimated in real dollars (ie in today’s dollar values). This example uses single years. You may have to break early years into months to capture cash flow details.
Notes:
1. Capital cost includes all items from your analysis in the Sample Financial Analysis section, including capital spares. State the accuracy of the estimate, eg ±10 per cent, ±30 per cent.
2. Budget to use extra resource and for contingencies during commissioning.
3. Includes concentrator, filtering and drying, transport, etc.
4. This example assumes 50 per cent of benefit in first three months, 75 per cent for next three months, then 100 per cent. The ramp up profile will relate to project complexity, quality of engineering, and quality
of commissioning.
$119 M
Project NPV
$39 M
$80 M
$43 M
Project
NPV
$23 M
$20 M
NPV of Net
Capital NPV of Net Capital
Revenue
Revenue
Project A Project B
FIG 2 – Which is the better project?
•• Project A costs $80 M, and has an NPV (real, after tax at 15 per cent discount rate) of $39 M,
for an IRR of 28 per cent
•• Project B costs $20 M, and has an NPV (real, after tax at 15 per cent discount rate) of $23 M,
and an IRR of 43 per cent.
Assume both proposals have been prepared professionally. Which is the better project?
NPV comparison says Project A. But consider two other factors:
•• What is the risk/return and the unavoidable uncertainty in the analysis?
•• Many projects will be competing for limited capital.
While Project B has a lower NPV, it risks less capital and achieves a higher return (IRR) on
it. Even the most professional work can’t eliminate risk – metal prices may fall, treatment
charges and exchange rates will change, the mine may not achieve its schedule. Project B may
be more robust against this uncertainty than Project A.
Further, the Board will be considering lots of projects and capital is limited. They need to
choose the highest returning projects. If they only have $80 M to spend, doing four projects
such as Project B will have a higher NPV than just doing Project A. This also spreads their risk.
Payback period
Payback period simply ranks projects by how many years it takes to pay back the investment.
Like IRR, it can be easily obtained from the NPV spreadsheet. Though it is simplistic and
favours simple short-term projects over ‘company builders’ it is nevertheless useful to
consider. A project that is paid back in three years is inherently less risky than one that takes
ten years to pay back, especially if there is metal price risk or sovereign risk.
take 18 months to execute. You estimate it will take three months after start-up to deliver full
recovery gains.
•• Estimated maintenance and operating costs such as power, water, air, reagents,
maintenance consumables, operating and maintenance labour and instrument labour.
•• Estimated a start-up profile for your project for cost increase and revenue increase. What
resources are necessary to achieve the start-up profile? These will be significant if you are
to start up quickly.
•• Considered any impacts on safety and environment. You will have done a ‘hazop’ or risk
analysis study to meet company procedures, and allowed for any mitigating actions that
resulted from this study.
So, that was the hard bit. The financial analysis is a breeze!
COMMUNICATION
There is nothing worse than a good project not gaining approved because of poor
communication and presentation. As they say, presentation isn’t everything, it is the only thing.
Actually, there is something worse – bad projects that do gain approval because of great
communication and presentation. Here, style has replaced substance and it is the Board’s
job to recognise this. Of course, it will get it wrong sometimes, and there is nothing you
can do about that. Your focus is to ensure that your numbers are right and that you will
deliver if your project is approved. Then you can add some style to your substance. Sure, the
Board likes style, but most of all it likes to back a proven winner. Develop a track record of
delivering on your promises.
Communication or presentation?
What we’ve just discussed refers to the Board presentation, but this is misleading – the
real communication work should be done long before you present to the Board. The final
presentation should be a formality; they should be convinced before you even walk into the
room. Let’s consider the different stages of communication.
•• Am I prepared to stake my reputation on the project? If not, find another project. Otherwise be
happy – your plant is running well and you can save the company’s money for important
projects elsewhere.
By answering these questions honestly you know how to justify the project. If you still
believe in the project, you will have the passion and commitment to convince others.
THE PRESENTATION
By now the presentation should be easy. You have done all the hard work answering
the questions and disarming the torpedoes. You have built champions at all levels of the
organisation. People have had a long time for the ideas to ‘soak in’. Momentum has been
built, and there is a sense that the project is inevitable. You have put yourselves in the shoes
of your boss, the senior management and the Board, and have considered what they are
interested in and what they need to know.
There are plenty of books and courses on presentation skills so I’ll avoid repeating them here.
If your project has rigour and substance and you have passion, you will get the presentation
right. Just follow some simple rules:
•• Keep it brief. These are very busy people with a lot on their mind. They do not need to know
the details of your job – they just need to do their job. Distil to the essence. Less is more.
•• Structure – tell ’em what you’re gonna tell ’em, tell ’em, then tell ’em what you just told ’em.
•• Tell a story. Try to link your data with a meaningful narrative. That is how people learnt
and remembered long before writing.
•• Less words, more pictures. Use pictures and graphs wherever possible. Look for the picture
that ‘tells a thousand words’. Dwell on it, give it time to talk. Powerful presentations can
be based on a single slide.
•• Keep each slide simple. Use the bare minimum of words. Limit them to a few brief dot points
per slide.
•• Avoid fancy effects. Choose simple colours with good contrast (keeping in mind that many
people are red–green colour blind).
•• Keep graphs simple. Eliminate all unnecessary grids, legends, titles and so on. Use the
minimum amount of ink to quickly convey the message in the data, and the minimum
number of graphs – try to seek the single graph that captures the story of your data. Tufte
(1983) has produced an excellent text on this topic, with Figures 3 and 4 demonstrating the
principle of removing unnecessary ink that he refers to as ‘chart junk’.
•• Put yourself in their shoes. They are smart people, but they are not metallurgists. You are so
familiar with mineral ‘liberation’ that you don’t even consider it jargon. Are you sure they
know what you mean?
90
80
70
60
50 East
40 West
30 North
20
10
0
1st Qtr 2nd Qtr 3rd Qtr 4th Qtr
FIG 3 – Too much ink hides the message in the data.
100
80 East
60
North
40
West
20
0
1 2 3 4
Quarter
FIG 4 – A simple graph lets the data speak clearly.
available capital, up to 100 per cent RR. The team could guarantee a 30 per cent return but
the project was too big and complex to achieve 100 per cent. This might have been overcome
except for the real barrier – the Board failing to understand what we were talking about!
REFERENCES
Buffet, W E, 2002. Chairman’s 2001 Letter to Shareholders, Berkshire Hathaway Inc. annual report,
28 February. Available from: http://www.berkshirehathaway.com/2001ar/2001letter.html [Accessed:
15 January 2016].
Tufte, E R, 1983. The Visual Display of Quantitative Data, 197 p (Graphics Press: Connecticut).
Particle size
300 microns
Key:
1. ZnS = sphalerite (grey)
2. Gangue = non-sulfide gangue
3. PbS = galena (white)
Particle size
300 microns
Key:
1. ZnS - easily liberated piece of sphalerite
2. ZnS - more difficult to liberate
3. Gangue (dark)
4. Pyrite (spherulitic)
These particles are coarse, but we do not have the grinding capacity to grind them finer.
These particles assay less that 30 per cent zinc, so we would lose money if we put them
in a concentrate.
70
65
Per cent
60
55
Sphalerite Liberation
50
1982 1983 1984 1985 1986 1987 1988 1989 1990 1991
Month/Year
FIG 6 – The graph that distils a thousand graphs.
CHAPTER 13
Operational geometallurgy
Dean David
ABSTRACT
The orebody is the only asset that a mine has at its disposal to generate revenue. Nature rarely
provides neat and consistent orebodies, so it is essential that process engineers understand
orebody variability and how this interacts with the proposed mine plans. A fundamental
difficulty in understanding the orebody is that most of the available information is geological
in nature and not readily accessed by process engineers. To make matters worse, the
metallurgical information is typically sparse and disconnected from the geology. This chapter
is an attempt to unlock some geological mysteries and provide tools for the process engineer
to link the geological and metallurgical data sets to their maximum advantage, especially
when attempting to optimise base metal flotation plants.
INTRODUCTION
Geometallurgy can be defined as the marriage of geological and metallurgical concepts in
order to provide useful inputs to process design and operational planning. The focus of
geometallurgy is entirely on process outcomes but its basis should be firmly in the geology.
Orebodies are naturally occurring phenomena with little consistency from location to
location. Orebodies situated within a few hundred metres of each other can be as geologically
different as orebodies on opposite sides of the earth. Within a single orebody it is possible
to have primary igneous rock sources, igneous intrusions, faulted zones, metamorphosed
alterations, oxidised zones, alluvial accumulations and much more. From a geological
viewpoint, the various aspects of each orebody can be defined with scientific definitions and
statistical accuracy using well-established conventions. As with any science, the terminology
used in geology is specific to the discipline and meets its own requirements.
The consequent problem for the process engineer is interpreting the science with the aim of
finding a place for the geological information within the metallurgical world. It must always
be remembered that geological data is primarily aimed at satisfying geological needs and
not process ones. Therefore, much of the available geological information may not be of any
practical use to the process engineer. The process engineer must carefully and selectively
examine the relationship between geological data and process response before committing to
a geometallurgical strategy.
DEFINITIVE ISSUES
Geometallurgy, like any study topic, requires a problem definition before embarking upon an
analysis. The process engineer must define problems and issues that, if solved, generate value
for the organisation. Geometallurgy becomes important when it is identified that the key to
understanding or solving a process problem lies in an improved understanding of the source
and nature of the ore being fed to the plant.
Examples of geologically based issues that regularly arise for process engineers in base
metal sulfide operations include:
•• inability to predict and manage semi-autogenous grinding (SAG) mill feed rate variability
•• inability to predict and manage concentrate grades
•• inability to predict and manage minor elements in concentrates
•• simultaneous control of plant feed grade and throughput rate
•• planning for throughput expansions
•• planning for new ore types or orebodies to be treated in an existing plant.
The economic importance of an issue is determined by estimating the value that could
be realised if it was eliminated or addressed successfully. The potentially redeemable
value becomes the basis for justifying the resources that need to be invested to solve the
problem. For example, stabilising an erratic SAG mill feed rate typically results in an average
throughput benefit of five per cent. A five per cent throughput benefit achieved through
operational stabilisation typically translates to a five per cent revenue improvement with
little or no real capital or operating cost penalty. Additional revenue at little or no cost equals
profit. Assuming a two-year payback is acceptable to management, it should be justifiable to
invest seven to ten per cent of annual revenue in a single solution that provides a five per cent
throughput benefit by stabilising the SAG mill feed rate. However, this is unrealistic and all
investment decisions are tempered by risk assessment. More importantly, there are always
other projects competing for attention, claiming the same slice of the profit cake and enabled
by the same source of funding. For example, on a plant having an annual revenue of A$100 M,
it may be realistic to expect management to commit A$1 M to a project with good potential to
increase SAG mill throughput by five per cent. In difficult times it may be a problem getting
A$100 000 approved to target a five per cent throughput increase.
Establishing an economic case is relatively easy compared to providing the practical
pathway to solving the problem. Often, the opportunity to apply a large sum of resources to
a problem leads to the implementation of an expensive and generic off-the-shelf solution. On
occasions, this is appropriate and successful, but in many instances it is not, and could even
have an overall negative economic impact. A greater degree of success is often achievable by
understanding the problem, both in detail and breadth, and then applying a solution fit for
the purpose at hand.
The first step to understanding the problem in detail is defining the goal. In this example
the goal is stabilised SAG mill feed rate. What are the definitions of instability and stability
in this context?
new or modified models. The first job of the process engineer is to determine which of the
available predictive methods provides real information about the ore properties (and the
associated plant responses) and which methods are ineffective.
The ultimate proof of the method is in its use. In this example the definitive question is ‘Is
it possible, using the chosen method, to successfully predict the SAG mill feed rate ahead of time for
a range of plant feed situations?’ If not, then the method is either wrong or incomplete. If the
method can successfully predict feed properties then there is probably a management issue
that is causing the feed variability, such as poor implementation of the predictive method or
poor control over ore sources and feed blending.
On the (often true) assumption that the prediction of ore properties is ineffective in practice,
it is then necessary to determine if a predictive method can be developed at all. It is here that
the process engineer enters the realm of geometallurgy.
AVAILABLE DATA
The starting point for any geometallurgical investigation is the geological model for the
orebody and especially its relationship to the mine model. The geological model contains
drill data while the mine model contains ore and waste blocks derived from the drill data.
Together, this is the biggest geometallurgical asset for the process engineer but, for some, it
can be the biggest liability in such an investigation. Two characteristics of drill data libraries
and block models make them difficult for process engineers to work with: their size and their
complexity.
Geological data libraries (drill libraries) are much larger than the databases process
engineers usually work with, such as tables of metallurgical test results. Geological databases
can contain data derived from 20 000 m of drilling during the early exploration phase of a
project and during the operational phases the database can grow to more than 200 000 m for
a modest sized operation. At the simplest level the database usually contains multi-element
assays for every metre or two metres of drill intersection. However, the database is also likely
to contain information related to lithology (rock type), alteration characteristics, colour,
mineralisation, veining, competence (rock quality designation – RQD), point load index (PLI,
a measure of rock strength), specific gravity (SG), positional survey and any number of other
measures depending on the ore type involved.
The size of such a database makes it difficult to manage and presents challenges when it
comes to isolating useful information. The software tools used to manage drill databases
are geologically based and rarely suit the requirements or experience of process engineers.
Another complicating factor in the operational stage of a project is that a large portion of
the database may be irrelevant to future operations as it represents ore that has already
been mined.
The complexity of the database is the second negative factor for process engineers. The
complexity in terms of the multiplicity of data sets has been mentioned previously. More
problematic than the number of data sets is the variety of bases on which those data sets are
arranged. For example, the assays will be determined on a regular one or two metre interval
basis but the lithology is likely to be provided as irregular intervals, the lengths of which
relate to the extent of each lithological unit. Within the geological database there may be five
or more different data set tables, each with their own individual length interval set. Further
complication can happen if the geological model has been built up by a number of successive
companies or geologists, each with their own ideas about what data is important and what is
not, or even which assay method should be used for a particular element.
To make the drill library database accessible to process engineers, it is best to use
geological software to generate a single suitable database. Packages such as GenSys can
de-survey geological data sets so that all properties can be reported on a common spatial
coordinate (ie X, Y and Z) basis rather than on a multiplicity of hole and depth bases.
The ideal outcome is for each individual record of the geometallurgical database to have
a unique spatial address and include every measured data type that has any chance of
being useful. The process engineer can then analyse, and add to, the single database using
appropriate tools such as spreadsheets.
An accessible geological data set is only useful to the process engineer when coupled with
additional information about the orebody and how it is to be delivered to the process plant
over time. In an operating mine, the economic limits of the orebody will be well understood
by the geologists, and the plant ore delivery schedule is defined by the short-term, medium-
term and long-term mine plans.
A spectrum of mine-related data will exist with relevance ranging from the imminent to
the distant future, the available mine planning data can be thought of as falling into the time-
based categories as described below:
•• Imminent – ore that is ready to be processed within the next week (and often within the
next month) and is likely to be either already blasted, ready for blasting or in stockpiles
ahead of the primary crusher. This ore is the subject of the daily production planning
processes on site. It can be well defined in terms of actual grade through assay of blasthole
drill cuttings or grade control drilling ahead of blasting. Likely comminution or flotation
performance is either based on the test result for the closest available metallurgical sample
or based on assumptions about the properties of ore types.
Factors that impede or blur the process engineer’s understanding of imminent ores include
medium-term and long-term run-of-mine (ROM) stockpiles, blending stockpiles for plant
feed, inadequate grade control drilling and analysis, lack of a blasthole sample program,
lack of blasting (ie free dig ore), multiple or numerous ore sources (typically a problem
for complicated underground mines) and a high degree of short distance variability in the
orebody (as may be seen in vein type deposits).
•• Short term – each mining operation has a short-term planning process to make decisions
about the upcoming waste and ore allocations. The basis for short-term decision-making
is usually in-fill drilling conducted during the construction and operational phases of the
project. In-fill drilling is designed to fill the gaps between the exploration drillhole data
and allow final ore/waste decisions to be made about each block in the mine model. It is
likely that mine block sizes (the volume of ore considered to have similar properties and
is assumed to be mined as a unit) will be reduced in size, compared to blocks used for
exploration or long-term planning, to better match the upcoming production schedule.
The in-fill drilling data set will certainly provide ore grade, problematic elements and, in
some cases, it may provide comminution data. The mine block properties are recalculated
using all available drill information and blocks are then designated as ore or waste based
on economic parameters.
The most significant complicating factor influencing the process engineer’s understanding
of short-term planning data is the mathematics used to derive the block properties. It is
essential that the process engineer understands the source of the data underpinning short-
term planning and understand how it has been manipulated on its way to becoming block
data. Manipulations can include truncation of high-grade spikes, data smoothing, data
kriging and erroneous data exclusions. To understand how data has been manipulated
during block modelling it is necessary to do some before and after comparisons. With the
assistance of miners and geologists, compare the short-term planning block predictions
with the raw drill data that has been used to define the block. The full extent of smoothing
is then evident and the ore variability expectations can be assessed by the process engineer.
•• Long term – the long-term mine plan is essentially an extension of the final operational
plan arising from the exploration phase. The ore included in the long-term plan is usually
based more on geological orebody models than high density drill data. The presence or
absence of ore in a particular location, its associated grade and recovery, its lithology and
alteration characteristics are all much less certain than the data available for ore that is
to be processed within the week or mined within the month. Much of the ore will not
even classify as Proved under JORC 2012 (or similar geological reporting regulations). The
long-term plan is typically updated annually, firstly to reflect new drilling that is designed
to convert Probable ore into Proved ore and secondly to reflect ore depletion as a result of
actual mining progress.
The long-term plan is most useful for gaining an understanding of the overall orebody
and mine shapes, documenting the proportions of the major ore types that are expected in
future years and understanding trends with time and location for critical measures such
as grade, impurities, hardness and competence.
When using block model data it is important to note a number of mathematical facts:
1. The properties assigned to an ore block are dependent upon the measured properties of a
few drill intersections. It is possible, and valid, that there are no drill intersections within
a particular mining block.
2. Geostatistical techniques, especially kriging, are used to estimate the distribution of
geological properties between drill holes.
3. Some ore block properties may have been calculated using complex equations that could
have an unintended outcome.
4. Some ore block properties will be dependent upon the location of the block relative to
a geological model boundary (eg above or below water table, oxidised versus fresh ore
zone, etc).
5. Some of the block properties, or even the underlying data, may have been filtered or
truncated to avoid illogical or extreme outcomes.
A good check when first encountering a block model is to look at both the model and the
underlying geological data in the same orebody location and ensure that there is a reasonable
relationship between the two. In general, the predictive methods are reliable but it is always
advisable to verify information, especially where important decisions are being made.
The weakest link in the available data for geometallurgical analysis is invariably the
metallurgical data set. It is rarely, if ever, practical to analyse the orebody metallurgically
to anywhere near the data density that is available geologically. For example, an orebody
of 100 Mt with 10 000 m of intersections identified within ore has a geological data density
of 100 samples per million tonnes of ore. Assuming there are three major lithologies that
have been tested metallurgically and each lithology has had 50 variability samples tested
for flotation properties and grindability, the metallurgical data density is only 1.5 samples
per million tonnes of ore. At 5 Mt/a plant throughput (for example) each year’s production is
effectively based on the results from 7.5 metallurgical test samples compared to reliance on
500 m of drill intersections. In some operations, each year’s production is defined by less than
one metallurgical test sample.
One ore block typically represents hours or minutes of ore and this means block data
can be useful for short-term metallurgical planning. Usually a subset of the long-term
plan block model is populated with infill drill data and then smaller mining blocks are
chosen for medium-term planning purposes. A subset of the medium-term plan can then be
populated with blasthole assay data for production grade control purposes for the imminent
production horizon. In a few instances the metallurgical database is continually updated
with preproduction testing and may find its way into the medium and imminent planning
outputs. Generally, however, the testing gap between geology and process continues to
widen as the project advances. Fortunately, the amount of available process data begins to
catch up once production begins. It is at this point that the adequacy, or otherwise, of the
metallurgical predictions is revealed.
It is important for the process engineer to become familiar with the planning processes and
the available data sources on site. Perhaps more importantly, the process engineer should
become acquainted with the owners of the various databases, the individual geologists and
mine planners. Apart from forming essential professional networks on the site there is also
the issue that databases always have quirks and skeletons, so it is best to have an expert assist
you to navigate to the data sources essential for your own problem solving.
It is not necessary, or desirable, for process engineers to become users of mining or
geological software. Usually specific data is required to solve a problem and it is preferable
to have the experts extract the necessary data from the sources so it can be analysed offline
using familiar tools such as spreadsheets. Although it may be useful to learn how to use the
specialised software, it is not normally efficient. All modern geological and mining software
packages have excellent data export and interchange capabilities that should be exploited.
In order of priority, the most important information after the positional data is the set of
chemical assays, followed by the lithology. Where there is limited comminution data from
metallurgical testing, drill core structural information (RQD, FF) and the strength data (PLI)
then become more important.
Note that the above description is typical of a database that has not been subject to
metallurgical influence. If detailed metallurgical testing has been conducted on the orebody
for geometallurgical purposes some of the following may appear either in the drill database
or in the mining block model:
•• drop weight index (DWI)
•• SAG power index (SPI)
•• Bond (or equivalent) work indices
•• liner and milling media abrasion indices (eg Ai)
•• ore abrasion index (ta)
•• mill feed F80 predictions
•• recovery predictions for products
•• concentrate grade predictions
•• SAG throughput rate (in an existing operation)
•• achievable grind P80 (with existing mills).
If such information does exist then a few checks need to be made to ensure the data is useful
in an operational sense. For example, in an existing operation it should be relatively simple to
check if the geometallurgical predictions are useful to the operators and planners. If the data
is ignored by the operators then it is either a poor predictor of performance or systems have
not been put in place to ensure that the metallurgical information is made available to those
that need it. The latter is often the case when the data has been collected during the design
period for the purpose of selecting major equipment, but no operationally useful systems
were implemented at plant start-up.
If it is found that the geometallurgical predictions do not match operational reality then
there is something wrong with either the data that is being used, the interpretation method or
the method of distributing the data across the geological databases.
Examples of inappropriate use of metallurgical data include:
•• the use of Bond ball mill work indices to predict SAG or AG mill throughput rates
•• the adoption of unverified relationships between ore measurements and operational
outcomes
•• inappropriate assignment of metallurgical ore properties to geological categories, such as
lithologies, where there is no proven linkage between the two
•• the use of predictive methods that ignore key factors, for example prediction of copper
concentrate grade based on head grade without reference to copper mineralogy.
In a modern operational environment there is usually enough data available to verify
or discredit these predictive systems. If not, then systems should be implemented and
instruments installed to allow, at a minimum, the verification of the predictive information
that is meant to drive decision-making in both the mine and the plant. The saying ‘if you
can’t measure it, you can’t manage it’ is especially true when it comes to geometallurgy.
Unfortunately, in many instances, it appears that measurement is being done but because
the measurements or the interpretation are inappropriate the consequential management
is misdirected.
•• An established relationship between plant feed Cu:S ratio and final concentrate grade
will be applied to the geologically based block model Cu and S assays. A second
prediction will be based on the mineralogical information logged by the geologists.
Inconsistencies will be flagged for potential investigation.
•• The flotation plant feed will be sampled hourly to provide shift composites for 30 days.
Shift composites will be subjected to a standard grind and flotation test. The flotation
test Cu and Au recoveries will be compared to the shift recoveries in an attempt to
develop copper and gold flotation recovery predictors. In addition, the ore source for
the shift or period will be noted. Additional flotation tests will be conducted using
freshly drilled diamond core samples to firm up the relationship with ore from known
locations and geological characteristics.
4. Implement the predictive framework.
•• The predictions can be implemented in one of two locations. The first, and most
comprehensive is in the geological database. The second, and most convenient, is in the
mining block model.
•• Adding the predictive framework to the geological model invokes considerable
complexity because the useful process outcomes of the predictive calculations will not
be needed in the geological model but in the form of short-term planning outputs from
the mine block model.
•• The geological data set is also problematic because the inherent variability is high
and the range of ore properties encountered will always be much broader than those
encountered shift by shift in plant feed and broader than the conditions represented
in the metallurgical test sample set.
•• Adding the predictions into the geological model requires that one or more of
the many data tables be expanded to include the predictive calculations. The data
table to be expanded depends upon the inputs to the calculations. For example,
if the calculations are assay based then the assay table should be expanded. If the
calculations are ore type based then it may be necessary to expand the lithology table.
•• Adding the predictive framework to the block model is simpler because the block
model is numerically simpler than the geological model (it is a single database) and
has already been manipulated to eliminate problematic data. The block model is also
mathematically driven in its construction compared to the geological model, which is
data entry driven. Virtually all information in a block model is mathematically derived
so it is a relatively simple matter to add or generate new data sets using mathematical
equations and rule sets.
•• The prediction of SAG mill feed rate will be made in the block model using modelled DWI
and BWI values.
•• It has been found that DWI is related to lithology and to depth, so each lithology will be
assigned a base DWI value that will be increased in direct proportion to depth.
•• Similarly, the BWI value has been found to be related to lithology and depth with the
exception that one lithology is considerably harder in the north than the south. This
lithology will be divided in north and south types and each have its own base BWI and
depth relationship.
•• As a first estimate the SAG feed rate for a block will be the minimum throughput rate
based on DWI versus available SAG power and BWI versus available ball mill power.
The estimate will be capped to allowable plant maximum and minimum throughput
rates.
•• Copper concentrate grade predictions will be made in two independent calculations, an
assay-based calculation and a mineralogy-based calculation.
•• The first step is to ensure both assay (Cu and S) and copper mineralogy information
has been carried through into the block model output.
•• An equation will be inserted into the block model that is a calibrated theoretical
relationship between the Cu:S ratio and final concentrate grade. A Cu concentrate
grade will be predicted for the ore block.
•• A second equation will be added that uses the stoichiometric copper proportion of each
mineral present in the block to calculate a second estimate of the Cu concentrate grade.
•• Both grade predictions will be reported separately and compared with actual outcomes.
•• Copper recovery predictions will initially be made in the block model on the basis of
constant tails grade. This relationship will be updated following the results of the planned
flotation test program.
•• Gold recovery predictions will also be made in the block model on the basis of constant
tails grade and will be subsequently updated when actual test results are available.
•• The data that needs to be carried through from the geological model to the block model
output for all these calculations to be possible includes:
•• for SAG mill feed rate prediction it is necessary to carry the lithology through and it is
also necessary to know the block location in three dimensions
•• for copper concentrate grade prediction it is necessary to carry per cent Cu, per cent S
and copper mineralogy
•• for copper and gold recovery prediction it is necessary to carry the per cent Cu and g/t
Au values.
Assuming the outcomes are mathematically valid and there is enough data to make the
predictions statistically useful, it is now possible to add the desired predictions of SAG
throughput, copper concentrate grade, copper and gold recovery to the short-term planning
outputs. Equally important is setting up a comparison to see how accurately the predictions
matched the actual production outcomes.
This is only one example of how geometallurgical predictions can be implemented. Each
site will have different prediction needs, different data sources and different planning needs.
PLANNING
The most obvious use of the predictive data is for short-, medium- and long-term metallurgical
planning. Applying the predictions to the long-term planning process should reveal future
constraints on throughput or production that will need to be accommodated. Predictions of
this nature feed into the capital planning process and also provide early warning of systemic
problems such as hardening ore or changing mineralogical mixes.
Using such tools, the process engineer can participate more fully in the long-term
planning of the overall project. For example, hardening ore without a compensating grade
change or throughput increase will lead to a revenue reduction that must be counteracted.
Counteraction measures range from plant optimisation to plant expansions or even to new
plant construction. Marketability of the concentrate is also influenced by its grade, another of
the possible predictions. A trend to lower grade concentrates may lead to early discussions
with customers, a reduction in revenue expectation, planning for increased throughput or
sourcing of new customers for the concentrate. Note that a site with a smelter as well as a
concentrator will have an even more pressing need for a concentrate grade prediction than a
site that sells its concentrate to multiple customers.
Equally important can be the educational aspect that arises from geometallurgical planning.
Provided the predictions are shown to be reliable, the mining engineers and geologists learn
what is important to the process plant and are able to contribute more effectively to the overall
site optimisation. For example, the geologists may change the way they look at the orebody
and report more ore characteristics. Alternatively, the mining engineers may introduce, or
modify, blending practices to smooth the process plant operation by avoiding the delivery of
undesirable ore packages.
Without an integrated predictive geometallurgical system it is difficult for the process
engineer to participate in an effective manner in the site planning process, beyond stating the
fundamental requirements of recovery and quality targets.
TROUBLESHOOTING
Troubleshooting is an ongoing responsibility of production process engineers. Any system
that assists in the troubleshooting process not only makes life easier but also improves the
company bottom line by solving problems faster. An integrated predictive geometallurgical
system would facilitate rapid troubleshooting by providing a consistent basis for sourcing
and analysing data. In addition, a reliable integrated system would have a significant benefit
of having established trust with the mining and geological departments and also with the site
management. Not only would it be possible to generate the necessary data to build a case for
change, but it should also be possible to get rapid agreement on implementation.
Typically, on operating sites there is no available system where data can be instantly
sourced to back-up metallurgical arguments. In these situations it is necessary to carry out
ad-hoc geometallurgical investigations to complement a particular investigation. Examples
may include tracking down periodic recovery drops, reducing concentrate grade variability
or tracking down the source of an impurity such as arsenic or fluorine.
An ad-hoc geometallurgical investigation would typically follow a similar pattern to that
used to set up an integrated system. Even though each particular investigation is unique and
could be dealt with in a rapid and possibly superficial way, it is advisable to take a little more
time to set up generic geometallurgical tools that can be used in future investigations. Begin
by tracking back as reliably as possible from the symptom of the problem to the source. For
example, high arsenic in concentrate is related to high arsenic in feed and arsenic in feed may
be the result of any one of a number of geological features, such as particular ore types or
alteration zones. Tracing back from the problem to the source allows the issue to be managed
by flagging the current geological source of the problem, identifying possible future geological
sources, checking if past instances have a similar cause and implementing a suitable solution
such as avoidance of the ore, blending to dilute the problem, or implementing a processing
solution such as depressing the arsenic minerals during flotation by the use of cyanide.
An ad-hoc investigation still requires a good level of understanding of the geological
information and a good working relationship with the geologists and mining engineers.
It also requires a full understanding of the factors that influence the ability to establish a
link between the symptom and the source. Common complications arise from the delays
and mixing caused by storage systems such as stockpiles, fine ore bins, ROM storage and
concentrate storage. These necessary devices can make it virtually impossible to isolate the
source of a problem in the mine. It may even be necessary to arrange for extended runs on a
single ore type to ensure the ore source is unequivocal during an investigation.
Other complications can arise from generic property assignment to ore types. In these
instances it is necessary to test the assumptions by real ore testing. For example, if a particular
ore type is assumed by others in the organisation to have a constant BWI value then take five
samples of that ore (as defined by the geologists) from different locations in the mine and test
them. Repeat the exercise on other important ore types. If the results support the assumption
then continue to use it. If they disprove the assumption then it is likely to be necessary to
modify site systems and re-educate people. Where the assumption is found to be invalid, and
it is necessary to rely on the BWI value in reaching a solution to the problem at hand, then the
situation must be corrected. Additional BWI tests on ore samples specific to the investigation
are easily justified.
Statements such as ‘all this ore type is the same’, or ‘there is not much variability in the
orebody’ must be tested by examination of available data or by collecting more. In most cases
these statements prove to be false and can derail investigations at the start.
EXPANSIONS
Plant expansions are special cases where geometallurgy is essential if the intended outcome is
to be achieved. Many expansions have been less than successful because the ore to be treated
did not meet the expectations of the designers. This is a direct result of a poor understanding
of the future orebody and its properties.
The first rule with expansions is to check all assumptions that drive the process design.
If the expansion is throughput driven it is essential to check that tests have been carried out
on samples representative of the ore that is to be treated by the expanded plant. In the absence
of supporting data it is not adequate to assume that because the ore blend will be 30 per cent
X and 70 per cent Y that the properties can be predicted from past experience. Take multiple
samples of X and Y from the ore zones to be treated by the expanded plant and subject them
to all the necessary tests to prove the design assumptions.
To check if this additional geometallurgical expense is necessary, it is relatively simple
to repeat the financial calculations with ten per cent lower throughput than predicted, or a
five per cent lower recovery.
An expansion decision must be based on a reasonable guarantee of a return from the
ensuing revenue stream. The same principles that are applied to exploration companies, ie
proving the presence of the ore and establishing its grade, are the first steps in justifying an
expansion. In most cases this is well established. However, in some cases the mineralisation
in question may lie in the inferred category and will need to be elevated to a higher status
before proceeding. Note that if the mineralisation is inferred, then it follows that all available
‘ore’ properties are inferred, including the status of the mineralisation as ore.
If the expansion is based on treating future ores according to established lithological
categories then check that lithology has been a useful indicator of process properties in the
past. If this has not been the case then the future will be no different to the past and it will be
essential that metallurgical tests are conducted on samples of the future ores.
When carrying out a metallurgical test program on the future ore there will be a set of
obvious tests to perform and there will be some less obvious tests. Perform the obvious tests
(comminution, flotation, abrasiveness, etc) and then carry out some speculative tests, such
as full elemental analysis or mineralogical analysis of concentrates. For reference it is best to
carry out the same tests on current production samples. These types of tests are simple and
inexpensive checks designed to identify if there are any new problems that may arise from
the future ore. If problems are confirmed or indicated by the tests then a more comprehensive
test program may be required to identify mitigation strategies. It is far less expensive to solve
these problems in design than to rectify after installation of any new capital.
Another possibility in preparing for an expansion is to conduct a plant trial on the future
ore. This is highly desirable if it is known that the future ore differs from the past ore. The
existing plant is the ideal ‘pilot plant’ facility. The ability to conduct a meaningful plant
trial is contingent upon ore access and ore volume availability. A plant trial has additional
benefits such as providing concentrate samples for customer evaluation and providing a
clearer evaluation of plant areas that are dependent upon concentrate availability, such as
regrinding, thickening and filtration. Another less obvious benefit is that the plant operators
are exposed to the future ore. It is usually the operators that notice operational effects first
and the crew should be well prepared before, and debriefed after, the trial.
At the implementation stage of an expansion project the major risk areas must be mitigated
to the satisfaction of the disciplines on site and to the satisfaction of the management team.
If the factors listed below have been established definitively, then the expansion has been
underpinned geometallurgically:
•• future ore tonnage and grade established
•• future ore metallurgical response is known
•• acceptable concentrates can be sourced from future ore
•• future ore throughput properties are known and have been used in comminution design
•• the variability of the future ore is understood.
DATA ANALYSIS
Effective and efficient analysis of geometallurgical data is essential to any geometallurgical
exercise. It is often necessary to manipulate very large databases and spreadsheets and
summarise them into meaningful sources of information. It is also necessary to correlate
properties from the geological and metallurgical realms so that trends and indicators can be
established. It is in the data sets that the important information resides, the difficult part is
extracting it. Even more difficult can be communicating this information to others.
The breadth of available geological data has been described previously. Also described has
been the necessity to develop analysis tools useful to process engineers. Unfortunately, it is
not possible to provide a one-size-fits-all solution to geometallurgical data analysis as the
geological and mining data formats are as diverse as the orebodies they describe.
The first step is to organise the data into a form that can be manipulated meaningfully. One
method (that is preferred by the author and described earlier) is to arrange all the available
geological data into a single spreadsheet table against a common set of depth intervals. In
addition, the data is de-surveyed to convert drill hole numbers and depths into 3D, spatial
coordinates (ie X, Y and Z). This is best done with the assistance of a geologist that is expert
in using the geological software and, with direction as to which data is required, the task can
be completed in hours or minutes. Once the database is available it is then possible to begin
detailed analysis.
Spreadsheet packages, such as Microsoft Excel, contain many useful statistical and data
manipulation tools. Probably the most useful tools for manipulating large and unwieldy
geometallurgical databases are the ‘Pivot Table’ tool and the ‘AutoFilter’ tool. Excel tutorials
will not be provided here, but all readers that deal with large data sets of any kind are strongly
encouraged to become familiar with these tools; they have the ability to extract data and
trends from multi-million-cell databases with a few keystrokes.
Before making use of the data it is essential that each column of data be checked for unusual
features or errors. For example, sometimes numerical data is actually present as text in the
spreadsheet rather than as numbers. Often text information has a minimum length and what
appears to be a three-letter code is three letters followed by ten spaces. Such data anomalies
confuse the analysis and the analyser. Check for data artefacts inserted by the geologists,
for example sometimes missing data might be represented by an impossible value for the
particular parameter, say 9999 or -1. Check for entry errors such as a mix of ppm values
in a per cent column or per cent values greater than 100. There are infinite possibilities
for problematic data in databases. The general rule is if unexpected results appear during
analysis then go back and check if the data that you thought you were incorporating is in
the appropriate form. If erroneous or misleading data is found, then prepare a report and
feed the information back to the custodian of the geological data so that problems can be
addressed across the site. Errors will typically influence geological interpretation to an equal
or greater extent than they will the geometallurgical interpretations.
It is usually necessary to add additional data columns to the geological databases. For
example, ore and waste are rarely defined in the geological data because these are mining
terms. A data column that classifies the intervals into simple ore and waste sets according to a
reasonable grade cut-off level is easily added. A consequent sort on ‘ore’ then removes all the
material that process engineers are not interested in. Note that in orebodies with a relatively
simple geological structure, this assignment does not have to be strictly in accordance with
mine planning discriminations as there are many factors, such as run length, that are difficult
to incorporate into a simplified geometallurgical model. It must be remembered that planning
is not done in a geometallurgical database, it is a tool for finding data relationships.
However, in banded or vein orebodies, it may be necessary to re-interpret the ore and
waste definitions to match the reality of mining limitations. For example, ensure that 1 m
waste intervals within an ore band are considered as ore because they will not be selectively
excluded. Similarly, if the minimum mining cut width is 4 m in a vein deposit then include
some waste on either side of significant ore intersections. These limitations are best discussed
with the miners before being implemented in the database. Once implemented it is essential
that a reality check is carried out to ensure that the effect of the data manipulation is consistent
whenever it is invoked. Check between ten and 20 instances through the database of any new
data manipulations.
In databases compiled by different geological teams over time there are likely to be duplicate
assay columns or lithology columns. Decisions must be made as to how this is reduced to a
single consistent column describing that property.
Another common manipulation is to round the X, Y, Z data into useful intervals. For
example, it is much more useful to know that ore is within the depth range RL 100 to RL 150
m than it is to know that the ore RL is precisely 123.27 m. This then allows the database to be
analysed by depth slices or slices in the north or easterly directions. It also allows the ore to
be examined as blocks. Blocks defined by this rounding method are usually much larger than
the mining model blocks and are best distinguished by the term ‘Metblock’.
Another useful addition to the database is to indicate those drill intersection samples that
have been utilised in metallurgical test programs. If full core has been taken for testing then
these samples are not available for any future test work. Even if only half the core has been
taken for testing it may be desirable to retain the remaining material for future geological
reference. Again, this can be tedious, but it provides an excellent basis for evaluating the
representative nature of past sampling and is a guide to selection, or non-selection of samples
for future test programs.
If bad data exists and it is not possible to correct it to something meaningful, then several
alternatives exist. One is to eliminate the offending data record or column by deletion (or shift
it to another place in the spreadsheet so it is known to have been removed). Another option
is to ‘fix’ the data, such as by averaging the corresponding values for the records above and
below. Another is to move a decimal point if it is unequivocal that the reported value is not
real but is the result of a typographical slip. Any manipulations should be noted and the
information returned to the geologists. This is usually valuable information that can lead to
improvements in the integrity of the geological database.
Once verified, the database is ready to use. Several guidelines are useful here:
1. Identify the ore that is important to the metallurgical issue being addressed.
Analysing ore that will never be mined or will be mined 15 years from now is rarely useful.
Identify, with the assistance of mining engineers, the ore that impacts on the decision or
problem at hand and then concentrate on it. One method may be to add an additional
column to the database and flag the important ore with a ’1’ or a TRUE. The data can
then be sorted on this basis. Another alternative is to assign a future year number to ore
intervals, but this is much more complicated. This sort of assignment is best done in the
block model.
2. Identify the properties that are of most importance to the metallurgical issue.
These properties then become the subject of data analysis and guide the analysis to the
important aspects of the data set. For example, in a copper flotation plant the copper grade
is obviously important but so is the sulfur grade, the arsenic grade and probably the gold
grade. Using the data tools mentioned above, it is then simple to make these measurements
the subject of the analysis.
3. Use the data tools to focus on the important ore and the important properties.
Using pivot tables, it is possible to quickly construct a table constrained to ore in the
relevant time frame and examine average Cu, Au, S and As grades, and to then identify
the number of core intervals in each category. It is also possible to view this data in plan
on the orebody by arranging it within the rounded X and Y data axes categories. The user
can then step down through the orebody by selecting an individual RL value or a set of
RL values.
Once the data is in this form, the orebody can be examined from a process perspective.
For example:
•• trends of arsenic with location, depth or ore type can be quickly extracted
•• trends in Cu:S with time can be examined
•• specific locations for extracting new metallurgical test samples can be identified and a
click will instantly drill down into the database and show the relevant drill intersections
in that particular Metblock.
It is also possible to use graphical and statistical techniques to search for correlations
between factors, for example plot %As against g/t Au on an X/Y plot to see if gold and arsenic
are related. If a relationship is suspected within a particular ore type then use the pivot table
drill down facility to extract the relevant records then plot the specific data comparison again
on the selected data.
Tools such as correlation tables can be used to search large data sets for significant linear
correlations between large numbers of pairs of data columns. Multiple linear regression can
be applied to see if a particular property is a function of a number of other properties.
Regression tools are particularly useful when investigating the relationship between the
limited metallurgical data set and the extensive geological data set. Assuming the drill
samples that were tested metallurgically have been correctly identified, it is possible to extract
the geological data for those intervals from the database, calculate the averages of the interval
properties and correlate them to the measured metallurgical properties. It may be possible to
identify geological predictors of metallurgical properties in this manner.
Ideally the metallurgical test samples will have been made up of contiguous intervals
of single lithologies; this is much more useful than single test samples being formed as
composites from across the orebody. It is especially problematic when the only available
data relates to composites formed from across the orebody and a mix of lithologies. Both
these situations make it almost impossible to establish any linkages between geological and
metallurgical properties.
One of the most rewarding outcomes for the process engineer from such an analysis is the
deeper understanding that is gained about the orebody. With this, the process engineer is
well equipped to contribute to discussions on the future of the operation and to converse
meaningfully and factually with the geologists and mining engineers on site. If nothing else
is gained, then the improved communication across the various disciplines on site is usually
worth the effort.
REFERENCES
JORC, 2012. Australasian Code for Reporting of Exploration Results, Mineral Resources and Ore Reserves
(The JORC Code) [online]. Available from: <http://www.jorc.org> (The Joint Ore Reserves Committee
of The Australasian Institute of Mining and Metallurgy, Australian Institute of Geoscientists and
Minerals Council of Australia).