Download as pdf or txt
Download as pdf or txt
You are on page 1of 96

Part 1.

Theory

PART 1. THEORY

INTRODUCTION

Part 1 constitutes the theoretical core of the thesis and has several
objectives. It summarizes the respective works of Pierre M. Gy and C.O.
Ingamells that are central to create pathways to appropriate industrial
standards. But, it goes beyond by addressing several important issues that
constantly become a hurdle if not properly addressed:

 Appropriate, commonly accepted notations must be used.


 The concept of Data Quality Objectives must be clearly
introduced. Indeed, not much of interest can be created to help
people when they do not have well-defined objectives.
 The critical difference between error and uncertainty must be
discussed, so statisticians are clear on the reasons why the word
error was favored in the TOS.

Then, with these objectives in mind, several case studies are presented to
make the reader more familiar with the author’s objective with the thesis.

25
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

26
Part 1. Theory

CHAPTER 1

HISTORICAL CONCEPT OF PIERRE M. GY’S WORK

1.1 SUMMARY

The purpose of this chapter is strictly to prepare the ground for a more in-
depth analysis of Dr. Pierre M Gy’s work that is relevant to the subject
matter of this thesis. It is of course beyond the suggested thesis to go into
all details of the TOS; it is however its mission to prepare the ground for
pragmatic applications.

Gy’s work is a logical and intelligent synthesis with the mission of


satisfying unity and coherence in his theoretical approach and also
creating effective guidelines to be used in the industry under some
conditions often complex and not necessarily well defined. In his first
official publication in 1967 23, it was the first time that such a synthesis
was presented to the general public. In his preface for the same
publication G. Matheron said “If the Ratio Occidentalis, from the very
first time when it appropriated for itself the entire world, could change
all ancient ways of living and thinking, then completely transform the
image of the planet on which we prosper, it is bizarre and very hard to
understand the reasons why sampling techniques would escape, all alone,
to this imperial order.” It was indeed hard to believe that it took so much
time to create the basis of the Theory of Sampling so critically important
in all our current activities.

Gy also said “A good practical analysis is the one that relies on practice
but at the same time always remains ahead of it.” We shall make sure
this is the philosophy we follow throughout the journey of presenting the
TOS to the world.

1.2 HISTORICAL CONCEPT OF HETEROGENEITY

Regardless of what the constituent of interest is in any material to be


sampled, it carries a certain amount of heterogeneity. Such typical
amount of heterogeneity is called a structural property of the material. If
we limit our analysis to zero-dimensional movable lots, three kinds of
heterogeneity are identified:

27
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

Constitution Heterogeneity CH L of the lot L: it consists of typical


differences between individual fragments, and mixing has no effect on
these differences. Let’s call a i the content of a constituent of interest in
any fragment Fi of the lot; a L the average content of the lot; M i the
mass of any fragment of the lot; M i the average mass of fragments in
the lot; M L the mass of the lot; N F the number of fragments in the lot;
hi the heterogeneity carried by any fragment of the lot.

hi 
ai  aL   M i  NF
ai  aL   M i [1.1]
aL Mi aL ML

The mean of the heterogeneity carried by one fragment is zero:

hi
mhi    0 [1.2]
i NF

The variance of the heterogeneity carried by all fragments in the lot is


defined as Constitution Heterogeneity, and the following formula is
easily demonstrated.

ai  a L 2 M i2
CH L  s 2 hi  
1
NF
 hi2  N F 
i i a L2

M L2
[1.3]

This formula is a cornerstone of Gy’s theory, has no assumptions, and no


approximations. It is the basic building block of all other more pragmatic
formulas created to help practitioners calculate the variance of the
Fundamental Sampling Error.

Distribution Heterogeneity DH L of the lot L: it consists of typical


differences between groups of fragments (i.e., increments), and mixing
has an effect on these differences. Let’s call a n the content of a
constituent of interest in any group of fragments G n of the lot; a L the
average content of the lot; M n the mass of any group of fragments of the
lot; M L the mass of the lot; N G the number of groups of fragments in the
lot (i.e., for a given size of the increments, the number of increments that
28
Part 1. Theory

it would take to empty the lot); hn the heterogeneity carried by any group
of fragments of the lot.

(a n  a L ) M n (a  a L ) M n
hn    NG n  [1.4]
aL ML aL ML

The mean of the heterogeneity carried by one group of fragments is zero:

hn
mhn    0 [1.5]
n NG

The variance of the heterogeneity carried by all groups of fragments in


the lot is defined as Distribution Heterogeneity, and the following
formula is easily demonstrated.

DH L  s 2 hn  
1 an  aL 2  M n2
NG
 hn2  N G 
n n a L2 M L2
[1.6]

This formula is another cornerstone of Gy’s theory, has no assumptions,


and no approximations. It is the basic building block of another
pragmatic formula to help practitioners minimize in a preventive way the
negative effects of segregation in a sampling protocol.

Constitution Heterogeneity CH n of a group of fragments G n : it consists


of typical differences between individual fragments, and mixing has no
effect on these differences. Let’s call a nj the content of a constituent of
interest in any fragment Fnj of the group of fragments; a n the average
content of the group of fragments; M nj the mass of any fragment of the
group of fragments; M nj the average mass of fragments in the group of
fragments; M G the mass of the group of fragments; N n the number of
fragments in the group of fragments; hnj the heterogeneity carried by any
fragment within the group of fragments.

29
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

hnj 
anj  aL  M nj

a  aL   M nj
 N n nj [1.7]
aL M nj aL MG

The mean of the heterogeneity carried by one fragment is zero:

mhnj   
hnj
0 [1.8]
nj Nn

The variance of the heterogeneity carried by all fragments in the group


of fragments is defined as Constitution Heterogeneity within a group of
fragments, and the following formula is easily demonstrated.

a  aG 
2
M nj2
CH n  s hnj  
1
h  Nn  
2 2 nj
nj [1.9]
Nn nj nj aG2 M G2

1.3 DEFINITION OF DATA QUALITY OBJECTIVES

Before going any further it is important to define what a “good sample”


is. How can we define what this is if we don’t have well defined
objectives about the mission for that sample? It is appropriate to spend
some time understanding the concept of Data Quality Objectives (i.e.,
DQO). It affects everything we do. DQO provides a valuable planning
tool to manage data collection activities, during exploration, ore reserve
evaluations, feasibility, ore grade control, processing, trading and
environmental monitoring. The major goals of the DQO process can be
set as follows:

1. State and clearly define the project under study with its
objectives, so that the focus of an audit will be unambiguous
(e.g., local geology, mineralogy, density, hardness, crushing and
grinding properties, amenability of the ore to processing, metal
contents, existing database, preliminary Geostatistics, old
sampling protocols used and new ones, old assaying methods
used and new ones, old accuracy and precision tests, economic
expectations, etc…).
2. Identify the decision to be made after the necessary data is
collected (e.g., is the project economically feasible for various
selected metal cutoff grades, and for the crushing, grinding, and
selected process?).
30
Part 1. Theory

3. Then, formulate a decision statement, which can be made after


an audit is completed (e.g., we need to control sampling,
subsampling, and assaying accuracy and precision better for a
good identification and quantification of the ore reserves, and
also for a clear definition of the comminution process and the
amenability of the ore to processing).
4. Identify the informational input that is required to reach the
decision statement (e.g., how and where data will be collected,
and how they will be interpreted).
5. Define the spatial and temporal boundaries covered by the
decision statement (e.g., definition of geological units,
monitoring each drilling machine, monitoring core splitting,
subsampling procedures, assaying methods, etc…).
6. Establish realistic goals for levels of accuracy and precision for
each constituent of interest in the deposit or in the process.
Taking an example from the copper industry: some minor
constituents in ore may lead to unacceptable levels of impurities
in copper concentrate; the representative sampling of such
impurities is as important as the sampling of major constituents.
7. Optimize our strategy to obtain the necessary data, including the
human resources requirements, necessary equipment, and
necessary working environment, and have a comprehensive
Quality Assurance and Quality Control program.
8. Use appropriate software to make the best of existing data using
variograms, relative difference plots, control charts, scatter plots,
etc...
9. Define a chain of command for compulsory action and
accountability procedure for each of the above eight points.

For DQO to be a success, the input from several technical disciplines is


needed (e.g., sampling, analysis, geology, metallurgy, trading and good
statistics). For example, it should be clearly understood that ore reserve
estimations highly depend on the accurate determination of the physical
and chemical characteristics of materials, and process amenability. The
same kind of approach is necessary to successfully optimize a processing
plant, to satisfy a client or to comply with regulatory specifications. As a
result, it is important to carefully integrate DQO into all sampling
activities we perform in the field, at the plants, during trading and at the
laboratory. It is equally important to integrate DQO into statistical
studies. Management must be creative to select these iterative DQO
steps. The word “iterative” should be emphasized since the objective is

31
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

to achieve continuous improvement: especially, let’s make sure the same


mistakes are not repeated too often.

“To call in the sampling expert or statistician after the necessary


database has been acquired without clearly defined data quality
objectives may be no more than asking him to perform a post mortem
examination - he may only be able to say what the database died of”.
(R.A. Fisher; quote modified for the purpose of sampling)

1.4 NECESSARY PRECAUTIONS TO BE TAKEN WHEN


WORKING WITH RATIOS

When the estimation of a ratio becomes an important issue, well defined


DQO become critically important. Someone may wonder why such an
issue is included in this thesis and at this early stage; it is because the
information used for a denominator and a numerator may often be
obtained through sampling, subsampling and assaying, and all of these
operations are sources of errors or at least certain levels of uncertainty.
The ratio of two random variables may lead to difficulties if preventive
precautions are not taken. Such examples are not rare:

ASk
The a Sk  ratio used in the Theory of Sampling (see definitions
M Sk
at equation [2.5]). The statistical properties of this ratio are an issue in the
early development stage of the TOS. For Cu/S in the copper industry,
Fe/SiO2 in the iron ore industry, ratio between biological components in
GMO research and monitoring, and many other cases, the problem is the
same:

1. Under no circumstances should the numerator, or the


denominator, or both the numerator and the denominator be
affected by a Poisson process (i.e., very poor precision).
2. Under no circumstance should the numerator, or the
denominator, or both the numerator and the denominator be
affected by problems of sampling incorrectness (i.e., sampling
that is not equiprobabilistic).

Actually, more often than not, for the measurement of such ratio to be of
interest the levels of accuracy and precision should be very stringent.

32
Part 1. Theory

1.5 FROM ERRORS TO UNCERTAINTY

The following comments are important because experience from


attending many conferences around the world reveals that many people
are constantly shifting from the word error to the word uncertainty
without a clear vision of what is the subtle difference between the two
words. The same applies to a large swath of the scientific literature, in
which these two concepts unfortunately are often used synonymously
which is a scientific flaw of the first order.

Error: the difference between an observed or calculated value and a true


value; variation in measurements, calculations, or observations of a
quantity due to mistakes or to uncontrollable factors.

It is the word “mistake” that bothers statisticians.

Uncertainty: lack of sureness about someone or something; something


that is not known beyond doubt; something not constant.

Historically, statisticians prefer the word uncertainty. The following


words should be carefully remembered in all discussions involving the
creation of sampling protocols and their practical implementations.

Gy (1967) stated: “With the exception of homogeneous materials, which


only exist in theory, the sampling of particulate materials is always an
aleatory operation. There is always an uncertainty, regardless of how
small it is, between the true, unknown content aL of the lot L and the true,
unknown content aS of the sample S. A vocabulary difficulty needs to be
mentioned: tradition has established the word error as common practice,
though it implies a mistake that could have been prevented, while
statisticians prefer the word uncertainty which implies no responsibility.
However, in practice, as demonstrated in the Theory of Sampling, there
are both sampling errors, and sampling uncertainties. Sampling errors
can easily be preventatively minimized, while sampling uncertainty for a
pre-selected sampling protocol is inevitable. For the sake of simplicity,
because the word uncertainty is not strong enough, the word error has
been selected as current usage in the Theory of Sampling, making it very
clear it does not necessarily imply a sense of culpability.” 23

Gy’s choice was especially justified for Increment Delimitation Error,


Increment Extraction Error, Increment Weighting Error and Increment
Preparation Errors, because indeed, the magnitude of these errors is
33
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

dictated by the ignorance, unwillingness or negligence of operators,


managers, and manufacturers to make these errors negligible by
following the rules of Sampling Correctness listed in the Theory of
Sampling. For these errors, the word uncertainty would be totally
inappropriate. Therefore, in any project, if management is due diligent,
the word error should not exist and only uncertainties remain; the
problem is we are living in a world very far from perfect where the TOS
is not yet mandatory knowledge for everyone in the business of creating
an important database.

1.6 THE UNIVERSALITY OF THE THEORY OF SAMPLING

From mining, to chemicals, pharmaceuticals, environmental issues,


research on genetically modified organisms, food chain, etc… the
Theory of Sampling is universal, contrary to the opinion of many who
are not familiar with the subtleties of Dr. Pierre Gy’s work.

1.7 STANDARDIZING NOTATIONS FOR THE THEORY OF


SAMPLING

During recent years notations used in the English version of the Theory
of Sampling were modified, in agreement with its founder. Table 1.1.
lists a few important examples. The ones marked with (*) are still under
some kind of debate.

Notations must be unambiguous and time-stable. The change of notations


after many thousand sampling practitioners have been trained is
equivalent to a fire in one’s home; it is an unfortunate catastrophe.
Therefore, it should not take place too often.

1.8 CREATION OF A SAMPLING PROTOCOL

Knowledge of Constitution Heterogeneity and Distribution


Heterogeneity allow us to select a sampling protocol adequate for a given
constituent of interest. Such protocol is not necessarily adequate for
another constituent in the same lot. A protocol is a recipe showing the
step by step requirements for sample mass that would minimize the
Fundamental Sampling Error and minimize the effects of segregation; it
also suggests appropriate sampling frequency to properly integrate local,
non-random variability. A protocol is the flow sheet of what needs to be
done, yet it does not provide the sample until it is implemented.

34
Part 1. Theory

Table 1.1. List of important old and new notations

Errors Old notation New notation


Fundamental Sampling Error FE FSE
Grouping and Segregation Error GE GSE
Increment Delimitation Error DE IDE
Increment Extraction Error EE IEE
Increment Weighting Error WE IWE
Increment Preparation Errors PE IPE
Short-range Process Integration QE1 or IE1 PIE1 (*)
Error
Long-range Process Integration QE2 or IE2 PIE2 (*)
Error
Periodic Process Integration Error QE3 or IE3 PIE3 (*)
In Situ Nugget Effect NE INE

(*) PIE is here suggested by Pitard and Esbensen jointly(1), based on


extensive discussions at the World Conference on Sampling and
Blending and elsewhere, being aware that it may be subject to changes
after further deliberations with Dr. Pierre M. Gy. The word “Process”
must be understood in a broad sense along any chronology (time- ,
process- , rank-order chronology); it also corresponds to the case of an
elongated 1-dimensional body in the standard parlance of TOS 1-4,12,14 ;
see also books 1 & 3 by the author. The old notation QE stands for
Quality Fluctuation Error and IE stands for Integration Error, depending
on time of publication and documents.

1.9 IMPLEMENTATION OF A SAMPLING PROTOCOL

The practical implementation of the sampling protocol requires the use of


some sampling equipment. Depending how these tools are designed,
built, maintained, used and cleaned a new family of sampling errors are
likely to take place. These new errors are the nightmare of sampling as
they are all bias generators. If sampling stops from being
equiprobabilistic, everything falls apart.

1
This new terminology is presented simultaneously in Esbensen &
Mortensen (2009): “Process Sampling (Theory of Sampling) – the
missing link in Process Analytical Technology (PAT)”. in Bakeev, K.
(Ed.): Process Analytical Technology. 2.nd Ed. Blackwell Publications
(2009).
35
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

1.10 PRESERVING THE INTEGRITY OF CONSTITUENTS

Between sampling stages, something may be done to the increment


making up a sample: it can be crushed, ground, pulverized, screened,
dried, filtered, or transported from one location to another. All these non-
selective operations are referred to as a family of Preparation Errors;
there can be many Preparation Errors affecting a sample or an increment.
These errors are also dangerous as they are bias generators.

1.11 DEFINITION OF SAMPLING CORRECTNESS

Sampling Correctness refers to a certain number of conditions that need


to be respected in order to minimize the negative effect of all bias
generators in sampling. They are classified into four families:

o The Increment Delimitation Error IDE: all parts of the lot must
be given exactly the same probability P of becoming part of the
sample. For this condition to be fulfilled successfully the
sampling tool must be designed according to the basic isotropic
module of observation; it is a sphere if the lot is considered 3-
dimensional; it is a cylinder if the lot is considered 2-
dimensional; it is a slice of constant thickness if the lot is
considered 1-dimensional.
o The Increment Extraction Error IEE: the sampling tool coming
into contact with the material to be taken must respect the rule of
the center of gravity of fragments, or respect the rebounding rule.
Basically, under no circumstances should the sampling tool
become selective on what it is taking, otherwise P is no longer
constant. If the increment is not representative of all size
fractions after recovery, then the sample will not be
representative. This error could have been called the Increment
Recovery Error.
o The Increment Weighting Error IWE: Each increment must
represent a certain portion of the lot, and each increment must
have a mass proportional to the mass of that portion. A sampling
system must be reasonably proportional in order to maintain P
constant.
o The Increment Preparation Errors IPE: These errors can be
numerous for a single preparation stage. A single preparation
stage can generate contamination, losses and alteration of the
physical and chemical components, making it impossible for P to
remain constant.
36
Part 1. Theory

The common property of these four sampling errors is that their mean is
not zero (P ≠ constant) unless very stringent precautions are taken.
Therefore, these errors do not introduce statistical uncertainties. They are
errors from manufacturers, engineering firms, and practitioners. They
make the sampling operation useless, and they should be of great concern
for standards committees on sampling.

1.12 THE NECESSITY OF SAMPLING CORRECTNESS

All practitioners must clearly understand that if the principles of


sampling correctness are not respected, all the hard work they do using
sampling protocols will be ruined. In this thesis it is assumed these
dangerous bias generators have been taken care of. Each of these bias
generators is potent for future, valuable theses. For more information on
sampling correctness refer to chapters 6, 7 and 9.

1.11.1 Implications of sampling correctness for the industry

The well-hidden cost of the non-compliance to the principles of sampling


correctness is a financial catastrophe of huge proportion, and too often
ignored. Examples by Carrasco should not be taken lightly 26 (e.g., a loss
of two billion US dollars over 20 years in a large copper floatation plant
because of an incorrect sampling system for the final tails; a loss of US
$134,000,000 over 10 years because of an incorrect sampling protocol
followed by an incorrect implementation for open pit blasthole
sampling). Many other cases could be found in other industries, like the
food chain industry. In most cases, the problem is generated by unaware
practitioners who thought non-probabilistic sampling systems could do
the job properly. It is of the utmost importance for the industry to set
standards for sampling correctness in such a way that they are non-
negotiable conditions to achieve due diligence in any project.

1.11. 2 The logical pathway to appropriate industrial standards

With sampling errors such as the Fundamental Sampling Error FSE, the
Grouping and Segregation Error GSE, the space or time interpolation
errors addressed in Geostatistics and Chronostatistics, there is room for
compromises; any practitioner doing his home work properly, with well-
defined DQO, can decide about a preselected amount of uncertainty in
the data that will be later generated with the samples that he collected.
However, with the bias generators listed above (i.e., IDE, IEE, IWE, and

37
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

IPE) there is no room for compromise; the practical implementation of


the sampling protocol is either correct or incorrect by structure, and there
is no middle way. Being very clear on this issue, now it is possible to
concentrate on FSE and GSE as presented by Gy, Visman and Ingamells.

38
Part 1. Theory

CHAPTER 2

THE WORK OF PIERRE M. GY TO BE LINKED TO


C.O. INGAMELLS’ WORK

2.1 SUMMARY

The objective of this thesis is not to rewrite Gy’s work, which was briefly
summarized in Chapter 1. However, it is necessary to identify the work
that can be linked to Visman and Ingamells’ works and also the relevant
and pragmatic tools that should be used by sampling practitioners for
many applications in the industry.

2.2 FROM THE CONTINUOUS MODEL TO THE DISCRETE


MODEL

Geostatistics and Chronostatistics look at lots of particulate materials


using a continuous model, with no emphasis on the discrete fragments. In
the Theory of Sampling this is best accomplished using the semi-
variogram. The emphasis of this thesis is to look at discrete fragments,
looking at a lot with a “magnifying lens”, which is in line with Visman
and Ingamells’ works. As a result, in the following material presented in
Part 1 the variability identified by the continuous model is not the
objective of the analysis performed in the main body of this thesis.
However, the continuous model will be explored in Chapter 10 because of
its critically important applications in the industry for management to
discover structural problems affecting their operations.

The discrete model further discussed in Part 1 applies to ores, vegetables,


fruits, cereals, chemicals, wastes, and any of constituents such as
minerals, solids in pulps or slurries, foreign materials in cereals, particle
size distribution, moisture content, etc… It has a very wide range of
applications. For the sake of simplification in the following work we will
refer only to fragments, groups of fragments, and constituent of interest
contents as already defined for equations [1.3], [1.6] and [1.9].

39
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

2.2.1 Distribution of the frequency  i of fragments Fi in the lot

For a large number of trials K with replacement (not to be confused with


the sampling constant K introduced later in equation [2.55]), each
fragment is either selected with a probability Pi or not selected with a
probability 1  Pi  , which is typical of a binomial distribution
characterized by:

the mean: mi   K  Pi [2.1]

the variance: s 2 i   K  Pi 1  Pi  [2.2]

2.2.2 Distribution of the frequency  n of group of fragments Fn in


the lot

In a similar way, for a large number of trials Z with replacement, each


group of fragments is either selected with a probability Pn or not selected
with a probability 1  Pn  , which is typical of a binomial distribution
characterized by:

the mean: m n   Z  Pn [2.3]

the variance: s 2  n   Z  Pn 1  Pn  [2.4]

2.2.3 Properties of the Sampling Error SE

The purpose of our analysis is to study the distribution of the content of


any constituent of interest a Sk in a given sample S k obtained after one
selection step of rank k with k = 1, 2, 3, …, K. Obviously, the quotient of
the mass ASk of any constituent of interest in the sample, divided by the
mass M Sk of the sample is the content a Sk . It is important to recall
that this ratio may be the ratio of two random variables.

40
Part 1. Theory

ASk
a Sk  [2.5]
M Sk

The relative sampling error SE is generated by the substitution of the


sample S k to the original lot L:

a Sk  a L
SE  [2.6]
aL

2.2.4 Relationship between the properties of SE and those of aSk

The true unknown content aL of the constituent of interest being a


constant, the distribution law of the sampling error SE is the same as the
one of the sample content a Sk . The moments of SE are:

ma Sk   a L
mSE   [2.7]
aL

s 2 a Sk 
s 2 SE   [2.8]
a L2

2.2.5 Distribution law of aSk

A difficulty may arise from the fact that a Sk is the ratio of two random
variables that may itself follow a law that is difficult to identify. However,
the distribution law of a Sk tends towards a normal law when both the
numerator and the denominator are distributed according to a normal law,
and when the relative standard deviation of the denominator is small when
compared to one24,25.

In many cases it is relatively easy to make sure these conditions are


respected and this is exactly what the objective of Gy’s Theory of
Sampling is all about. Without giving any number a magic property, and
to prevent endless discussions, it has been shown in multiple cases by Gy,
Visman and Ingamells that when the relative standard deviation of the
sampling error SE starts to exceed 16% the above assumption starts to
degenerate. It is critically important to recall this limitation when using
41
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

Gy’s formulas, as many practitioners around the world are not aware of it,
and therefore misuse the work of a great man.

Experience proves that we often enter into that doubtful domain when we
sample for the determination of trace amounts for a given constituent of
interest (e.g., trace constituents in high purity materials, in pharmaceutical
products, in chemicals, in the environment, in wastes, precious metals in
mineral deposits, diamonds, etc…); this is exactly why Chapters 3, 4, 8
and 11 are so important.

Now, the difficult question is: what should we do when either the
denominator or the numerator of the ratio illustrated in equation [2.5] is
affected by a Poisson process? The answer to this question, or at least
partially and in a pragmatic way, is offered by Ingamells’ work in a very
attractive way. This is precisely why it is important to link Ingamells’
work to the Theory of Sampling, so we can fully measure the stunning
effects of disobeying the preventive principles taught in the TOS.

2.2.6 Moments of the content aSk

The subscripts of the probability P can be i , n, or nj according to earlier


definitions. Let’s define:

N the mean of the population of fragments N Sk in the sample S k with:

N  mN Sk    Pi [2.9]
i

M the mean of the population of masses M Sk of the sample S k with:

M  mM Sk    M i Pi [2.10]


i

A the mean of the population of masses ASk of the constituent of interest


in the sample S k with:

A  m ASk    Ai Pi   ai M i  Pi [2.11]


i i

a the following ratio:

42
Part 1. Theory

A m ASk 
a  [2.12]
M mM Sk 

Let’s define the following auxiliary variables:

x the relative deviation of the mass M Sk with its mean M which can be
written:

M Sk  M
x [2.13]
M
which can be rearranged:

M Sk  1  x M [2.14]

y the relative deviation of the mass ASk of the constituent of interest in


the sample S k with its mean A :

ASk  1  y A [2.15]

z the relative deviation of the content a Sk of the constituent of interest


with its mean a :

a Sk  1  z a [2.16]

Therefore we can write:

a Sk 
ASk

1  y A  1  y  a  1  z a [2.17]
M Sk 1  x M 1  x 

By dividing both sides by a we obtain:

1 y
1 z  [2.18]
1 x

43
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

The difficulty is that (1+x) is a random variable. But, for the conditions of
normality of a Sk we find that x should be small in comparison to 1, and
in practice it is often the case. This suggests we can develop the ratio
1
as a Mac-Laurin series:
1 x
1
 1  x  x 2  x 3  x 4  ...  x n for x 1 [2.19]
1 x


1  z  1  y  1  x  x 2  x 3  x 4  ...  [2.20]

  
z   y  x   x 2  xy  x 3  x 2 y  ...  [2.21]

The absolute value of z is also smaller than 1.

The conclusion is obvious: in relation [2.21] the term ( x 2  xy ) is much



smaller than the term  y  x  ; similarly the term x  x y is much
3 2

smaller than the term ( x  xy ) .
2

Gy demonstrated that the first approximation z1   y  x  is sufficient in


many cases, except for trace constituents if special preventive conditions
are not taken23. Then, going back to relation [2.17] we can write:

aSk 1  1  z1 a  1  y  xa [2.22]

2.2.7 Moments of aSk and SE

Using the first approximation, Gy demonstrated in 1967 and again in 1975


the following results in which subscripts could be i, n, or nj according to
earlier definitions. The full demonstration is long and tedious, and it was
kindly checked by G. Matheron (the master checking the master!).

a  M  P i i i
ma   a  i
[2.23]
M  P
Sk 1
i i
i

44
Part 1. Theory

a  aL
mSE 1  [2.24]
aL

 a 
 a M i2  Pi 1  Pi 
2
i
s 2 a Sk 1  i
2
[2.25]
 
 M i  Pi 
 i 

s 2 a Sk 1
s SE 1 
2
[2.26]
a L2

2.2.8 Properties of the sampling error SE; connection with correct


sampling

The conclusion of their work is unambiguous: for sampling to be correct


the sampling probability P for all units, either fragments or groups of
fragments must remain constant, at least to a first order approximation:

Pi = P = constant regardless of i [2.27]


Pn = P = constant regardless of n [2.28]
Pnj = P = constant regardless of nj [2.29]

It should be clearly understood that a truly equiprobabilistic sample exists


only if each fragment is collected one by one at random. In practice this is
rarely the case. Gy demonstrated that taking groups of fragments (i.e.,
increments) can lead to nearly enough equiprobability as long as random
systematic or stratified random sampling is used23. In the absence of
cycle, if increments are numerous enough, systematic sampling can also
lead to nearly enough equiprobability. However, in all cases, the
collection and preservation of these increments must be correct with
respect to IDE, IEE, IWE and IPE further explored in Chapters 6, 7 and 9.

Transposing this property into relations 2.23 to 2.26, with subscripts being
either i, n, or nj, we obtain:

a  M i i
ma   i
 aL [2.30]
M
Sk 1
i
i

45
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

m[SE ]1  0 [2.31]

 a  a L   M i2
2

1 P i
s 2 [a Sk ]1   i
2
[2.32]
P  
 M i 
 i 

1  P  ai  a L  M i2 
2

s SE 1 
2
  2 [2.33]
P i  a L2 ML 

Then, transposing these properties to the Fundamental Sampling Error and


to the Grouping and Segregation Error becomes possible.

2.3 THE FUNDAMENTAL SAMPLING ERROR

m( FSE )  0 in a first order approximation

and taking relation [1.3] into relation [2.33]:

1  P  ai  a L  M i2  1  P
2

s FSE  
2
  2 CH L [2.34]
P i  a L2 M L  P  N Fi

We don’t want to be in a position to have to count the number of


fragments in the lot. It is easy to overcome this difficulty by multiplying
CHL by a constant factor such as the average weight of a fragment which
ML
is by definition M i  . Then we have to define a new term called the
N Fi
Constant Factor of Constitution Heterogeneity IHL also called Invariant of
Heterogeneity by some authors.

ai  a L 2 M i2
IH L    [2.35]
i a L2 ML

Therefore, the following important relation can be obtained:

46
Part 1. Theory

1 P
s 2 FSE   IH L [2.36]
PML
From this point several pragmatic formulas can be derived which have
their own domain of application and limitations. For the record: there is
no such thing as the Pierre Gy’s magic formula as is the perception given
by many people around the world who are not familiar with the subtleties
of his valuable work.

For the sake of simplicity, the notation for the variance of Fundamental
2
Sampling Error will be s FSE .

2.3.1 A pragmatic formula for a parametric approach

The theoretical foundation of a quick approach and its limitations is the


most appropriate start. Let’s begin with the approximate formula used
from the development of Gy’s parametric approach (See referred textbook
#1 in Introduction and Historical Synopsis)27:

  v  M L 
   a  aL  M L 
2

 
IH L  X  Y  
   [2.37]
 ML a 2
M 
 
 
L L

This simplified equation is based on two important assumptions:

1. Experience shows that the content a of a constituent of interest


usually varies much more from one density fraction  to the
next than from one size fraction  to the next; therefore all the
values of a obtained in a size-density heterogeneity
experiment may be replaced by the average content a  of the
corresponding density fraction L . This assumption is almost
always true.
2. The study of a large number of real cases shows that in a size-
M L
density heterogeneity experiment the proportions usually
M L
vary little from one density fraction to the next; therefore, we

47
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

M L
may assume that all values can be replaced by their
M L
M L
average . This assumption may become debatable if the
ML
constituent of interest becomes liberated, and if delayed
comminution of the same constituent becomes a possibility.

The X term is relative to the size fractions and leads to the shape factor
and particle size distribution factor, while the term Y, relative to the
density fractions is the one of interest in our present analysis. We know Y
can be expressed as follows:

Y    
a   aL  M L
2

[2.38]
aL2  M L

2.3.1.1 The mineralogical factor

Y reaches a maximum when the constituent of interest is completely


liberated. This maximum is defined as the mineralogical factor c. If the
material is made of only two liberated constituents, for example the
gangue and the mineral of interest, then the density fraction containing
the pure liberated mineral has a density  M , while the gangue fraction
has a density  g . Then, it follows that for the fraction containing the
mineral of interest the content of the mineral is a  aM  1 ; then for
MM
the gangue fraction a  ag  0 ; also the ratio  aL and the ratio
ML
Mg
 1  aL  .
ML

Transposing these values in equation [2.38]:

1  aL  M M
2
0  aL 2 M g
Y  c  M  g
2
aL  M L 2
aL  M L
After simplifications we obtain:

48
Part 1. Theory

c  M
1  aL 2   1  a  [2.39]
g L
aL

2.3.1.2 The liberation factor

Let’s make several hypotheses which must be kept in mind in order to


understand the limitations of the following recommended methods. Let’s
also assume we are interested in the copper content of a lot to be
sampled.

First hypothesis: Following an analytical investigation, we suppose that


the maximum copper content amax of the coarsest fragments of the lot is
known.

Second hypothesis: We suppose that all size fractions have roughly the
same copper content aL, or at least they are within the same order of
magnitude.

Third hypothesis: We suppose that inside each fraction all of the copper
is located in a sub fraction of copper content amax , density  R , and
relative weight

M a
 L , [2.40]
M L a max

while the remainder of the size fraction of density  g , and relative


 aL 
weight 1    does not contain any copper or very little.
 a max 

Then we can rewrite equations [2.38] and [2.39] as follows:

Y  R
amax  aL  aL
2
 g
0  aL  
2
1 
aL 
 [2.41]
2 2
aL amax aL  amax 

which leads after simplification to:

49
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

a   a 
Y   R  max  1   R   g  L  1 [2.42]
 aL   amax 

a max being usually much larger than a L , the second term may become
negligible, therefore:

a 
Y   R  max  1 [2.43]
 aL 

It would be convenient to suppress the factor  R which is difficult to


estimate. Let’s call V1 the volume occupied by copper and V2 the
volume of the gangue. The average density  of their mixture can be
expressed as follows:

V1   M  V2   g
 [2.44]
V1  V2

aL
with  M being the density of the copper mineral. But V1  and
M
V2 
1  aL  , then:
g
M   g M   g
 or  aL   g   M 1  aL  [2.45]
aL   g   M 1  aL  

By transposing [2.45] into the general equation [2.39] to calculate the


mineralogical factor and after simplifications we obtain:

1  aL  M   g
c [2.46]
aL  

therefore:

50
Part 1. Theory

 amax 
 R  1
Y
   aL 
c 1  aL  M   g
[2.47]

aL  


amax  aL  R   [2.48]
1  aL  M   g
In practice we also know that  M   R     g therefore:

R  
1
M   g

We finally obtain the very practical formula:

a max  a L
 [2.49]
1  aL
a max and a L should be expressed as a proportion of the copper mineral
content (i.e., as part of one), and not as a metal content.

2.3.2 Recommended method #1: Determination of a max for each size


fraction of a typical granulometric distribution

1. Collect a large composite sample representing a single geological


unit, from leftover half core samples (e.g., fifty 6-kg samples).
2. Dry the composite.
3. Crush the composite to d = 2.54 cm. The definition of d is the size of
a screen retaining no more than 5% of the material by weight.
4. Screen the entire composite using the following screens: 2.54 cm,
1.25 cm, 0.635 cm, 0.335 cm, 0.17 cm, 0.085 cm, 0.0 425cm, and
0.0212 cm. Below 0.0425 cm the method becomes very awkward but
it can be done.
5. Wash, dry, and weigh each size fraction.
6. Spread the size fraction between 2.54 cm and 1.25 cm on a clean
surface.

51
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

7. Using a portable X-ray machine select 10 fragments showing the


highest copper content. Using a microscope, identify the main copper
mineral to calculate the mineral content. Crush and grind them in a
mortar, then assay them for their copper content. You obtain 10
copper results. Look at the distribution of the 10 results. Calculate
the average of the 10 results and call the average an estimate of a max
for d = 2.09 cm. Using formula [2.49] calculate  for d = 2.09 cm.
8. Spread the size fraction between 1.25 cm and 0.635 cm on a clean
surface.
9. Using a portable X-ray machine select 10 fragments showing the
highest copper content. Crush and grind them in a mortar, then assay
them for their copper content. You obtain 10 copper results. Look at
the distribution of the 10 results. Calculate the average of the 10
results and call the average an estimate of a max for d = 1.05 cm.
Using formula [2.49] calculate  for d = 1.05 cm.
10. Repeat the same process for the other size fractions between 0.635
cm and 0.335 cm, between 0.335 cm and 0.17 cm.
11. For the smaller size fraction, identify 10 zones where the copper
content is high with the X-ray machine. At each of these zones
collect a spoonful of fragments. Look at them under the microscope
and estimate a max using proportion standard references from the
mineralogist when you spot a fragment with high copper content.
a max will be the average of your observation from the 10 spoonful
sub-samples you collected from each respective size fraction. Then,
you can calculate  using formula [2.49] for each respective size
fraction.

2.3.3 Recommended method #2: Determination of a max for the top


size fraction of a typical comminution stage

This method is longer but may be more accurate because of the limitation
of hypothesis #2 under different conditions of comminution.

1. Collect a large composite sample representing a single geological


unit, from leftover half core samples (e.g., fifty 6-kg samples).
2. Dry the composite.
3. Crush the composite to d = 2.54 cm.
4. Split the composite into 7 sublots.
5. Screen one sublot using 2.54 cm and 1.25 cm screens
52
Part 1. Theory

6. Wash, dry, and weigh the size fraction.


7. Spread the size fraction between 2.54 cm and 1.25 cm on a clean
surface.
8. Using a portable X-ray machine select 10 fragments showing the
highest copper content. Using a microscope, identify the main copper
mineral to calculate the mineral content. Crush and grind them in a
mortar, then assay them for their copper content. You obtain 10
copper results. Look at the distribution of the 10 results. Calculate
the average of the 10 results and call the average an estimate of a max
for d = 2.09 cm. Using formula [2.49] calculate  for d = 2.09 cm.
9. Crush the second sublot to d = 1.25 cm
10. Screen the sublot using 1.25 cm and 0.635 cm screens
11. Wash, dry, and weigh the size fraction.
12. Spread the size fraction between 1.25 cm and 0.635 cm on a clean
surface.
13. Using a portable X-ray machine select 10 fragments showing the
highest copper content. Crush and grind them in a mortar, then assay
them for their copper content. You obtain 10 copper results. Look at
the distribution of the 10 results. Calculate the average of the 10
results and call the average an estimate of a max for d = 1.05 cm.
Using formula [2.49] calculate  for d = 1.05 cm.
14. Repeat the same process for the other size fractions between 0.635
cm and 0.335 cm, between 0.335 cm and 0.17 cm, by crushing
another sublot appropriately each time.
15. For the smaller size fraction, after crushing a sublot appropriately,
identify 10 zones where the copper content is high with the X-ray
machine. At each of these zones collect a spoonful of fragments.
Look at them under the microscope and estimate a max using
proportion standard references from the mineralogist. a max will be
the average of your observations from the 10 spoonful sub-samples
you collected from each respective size fraction. Then, you can
calculate  using formula [2.49] for each respective size fraction.

2.3.4 The more accurate but more complex approach: theoretical


principle

We know from the Theory of Sampling the variance of the Fundamental


Sampling Error can be estimated using the following formula:

53
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

 1 1 
2
s FSE     f  g  c    d 3 [2.50]
 MS ML 

We also know all factors except  are relatively easy to estimate.


Therefore, this would suggest the use of the following formula to
calculate  for each stage of comminution:
2
s FSE
 [2.51]
 1 1 
   f  g  c  d 3
 S
M M L 

The method suggested highly resembles a typical Heterogeneity Test,


except that it should be repeated for several stages of comminution.

Important remark

In Gy’s earlier literature the Constant Factor of Constitution


Heterogeneity IHL was written as follows:

IH L  f  g  c    d 3  C  d 3 [2.52]

The problem with this presentation was that C, which is the product of
four factors, must be calculated every time the value of d changes since
the liberation factor varies rapidly with the value of d. As a result, in the
new literature it became a tradition, for practicality, to summarize the
value of IHL as follows:

IH L  f  g  c    d 3  K  d x [2.53]

In this new presentation, the liberation factor is assumed to follow an


empirical model such as:

r
d 
   [2.54]
d 
where d  is defined as the liberation size of the constituent of interest. In
many cases when the constituent of interest is a single mineral, the

54
Part 1. Theory

exponent r is not far away from 0.5. But, as clearly shown by Gy 23 for
the liberation of ash in coals, and further demonstrated by François-
Bongarçon 18,40-41 for gold, r is not necessarily anywhere close to 0.5,
especially when the constituent of interest is located in various minerals.
Under such new conditions, equation [2-50] should be approximated as
follows:
 1 1 
2
sFSE   K  d
x
[2.55]
MS ML 

where:

K  f  g  c  d  
r
[2.56]
and

x  3 r [2.57]

K and r then become the key factors to quantify in various experiments


described by François-Bongarçon. The author of this thesis favors the
approach using a max for the determination of the liberation factor, in
which case x = 3; nevertheless, François-Bongarçon’s approach proved
to be extremely useful in the mining industry.

2.3.5 Recommended method #3: Determination of IH L  f  g  c    d 3


for each size fraction of a typical granulometric distribution

1. Collect a large composite sample representing a single geological


unit, from leftover half core samples (e.g., fifty 6-kg samples).
2. Dry the composite.
3. Crush the composite to d = 2.54 cm.
4. Screen the entire composite using the following screens: 2.54 cm,
1.25 cm, 0.635 cm, 0.335 cm, 0.17 cm, 0.085 cm, 0.0425cm, and
0.0212 cm. Below 0.0425 cm the method becomes very awkward but
it can be done.
5. Wash, dry, and weigh each size fraction.
6. Spread the size fraction between 2.54 cm and 1.25 cm on a clean
surface.
7. Collect 50 samples made of 50 fragments collected one by one at
stratified random.

55
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

8. Weigh each sample: the average weight of these samples is highly


relevant in equation [2.51]. Crush and pulverize each sample. Assay
each sample for copper. Convert the copper contents into mineral
contents using the main copper mineral. You obtain 50 copper
mineral results. Look at the distribution of the 50 results. Calculate
 using formula [2.51] for d = 2.09 cm.
9. Spread the size fraction between 1.25 cm and 0.635 cm on a clean
surface.
10. Collect 50 samples made of 50 fragments collected one by one at
stratified random.
11. Weigh each sample: the average weight of these samples is highly
relevant in equation [2.51]. Crush and pulverize each sample. Assay
each sample for copper. Convert the copper contents into mineral
contents using the main copper mineral. You obtain 50 copper
mineral results. Look at the distribution of the 50 results. Calculate
 using formula [2.51] for d = 1.05 cm.
12. Repeat the same process for the other size fractions between 0.635
cm and 0.335 cm, between 0.335 cm and 0.17 cm.
13. For smaller size fractions, use 50 one-gram samples made of ten
stratified random 0.1-gram increments. Then, calculate  using
formula [2.51] for each respective size fraction.

2.3.6 Example of method #3 applied to a gold deposit

In any gold project, it is essential to perform a Heterogeneity Test in


order to calculate sampling nomographs and optimize sampling
protocols. The test should be performed on each type of major
mineralization. A preliminary mineralogical study reveals that the test
can be done on a 340-kg composite, using samples from the top size
fraction, made of p = 50 fragments each, mainly because the gold is
found to be relatively finely disseminated within a quartz matrix. The
following case is a real case, not necessarily appropriate for every gold
deposit.

1. The composite is made of material from 50 different locations within


a single type of mineralization. Dry the composite overnight, at
110oC. After drying the composite weighs about 340 kg. Crush the
entire composite to roughly –2.00 cm using a clean jaw crusher with
opening adjusted to about 2 cm.
2. Screen the lot through 1.25 cm, 0.63 cm, 0.335 cm, 0.17 cm, 0.085
cm, 0.0425 cm, and 0.0212 cm screens.

56
Part 1. Theory

3. Weigh each size fraction and record these weights.


4. Spread the –1.25 cm +0.63 cm fraction on a clean surface: the
Heterogeneity Test is performed on this fraction, where d = 1.05 cm.
5. From this fraction, collect 100 samples. Each sample is made of pi
fragments selected one by one at random, up to nearest 50 grams.
Number these samples from 1 to 100, weigh each of them, and
record values for pi. Results for the sample weights are illustrated in
figure 2.1, where the arithmetic average is M S  50.2 g . Results for
the number of fragments are illustrated in figure 2.2, where the
arithmetic average is p = 51.7 fragments.
6. Pulverize each sample directly in an enclosed ring and puck
pulverizer to about 95% minus 106 micron (Do not use a dusty plate
pulverizer which is known to smear gold too much).
7. Assay each sample for gold by fire assay and gravimetric finish,
using the entire sample. Atomic Absorption finish is not
recommended for the test, since the most relevant assays are the ones
showing high gold contents. Results are summarized in table 2.1 and
plotted in figure 2.3. The arithmetic average of the 100 fire assays is
98.4 g/t and the relative, dimensionless variance is a close estimate
2
of s FSE  0.4278 .
8. Crush all size fractions, and what is left from the –1.25 cm +0.63 cm
fraction, to 95% minus 0.30 cm, and collect a split about 1000 grams
from each size fraction. If the coarsest size fraction has less than
1000 grams, use the entire fraction. Pulverize each 1000-gram split
to 95% -106μm. Perform a screen fire assay using the entire sample
and a 150μm screen. Weigh and assay the entire +150μm fraction,
and perform two 50-gram fire assays on the -150μ fraction. Report
all weights and assays, which are very relevant for the interpretation
of the test. Results are summarized in table 2.2.
9. From any rejects from the test, prepare a 10000-gram sample. Screen
the entire composite sample on a 212-μm screen. There is always a
substantial amount of material that does not grind well. Wash this
coarse material. Separate the heavy minerals by panning. With the
heavy concentrate, prepare polished sections and find the exact
nature of the material that does not comminute well.

57
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

Figure 2.1. Plot of the 100 sample weights

58
Part 1. Theory

Figure 2.2. Plot of the number of fragments in the 100 samples

2.3.6.1 Advantages of the recommended approach

1. The test accumulates an enormous amount of information at a


reasonable cost.
2. The Fundamental Sampling Error for the coarsest fragments (i.e.,
about 1 cm for blasthole chips, reverse circulation chips, or the
product of a jaw crusher) is well known. It is indeed the primary
sampling stage in ore grade control and exploration that is likely to
introduce a large error.
3. The experiment uses 1-cm fragments, therefore the sample weight
and K are the only contributors to the variance of the Fundamental
Error, d3 being one or very near. This gives an accurate estimate of
the sampling constant K.
4. Differences in gold content between various size fractions are well
documented. For example, if the -212μm material contains 5 times
more gold than the coarse fragments, then any loss of fines during
the implementation of the sampling protocol would introduce large

59
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

Increment Extraction Errors and Increment Preparation Errors, which


are all devastating non-constant bias generators.
5. Screen fire assays performed on 1000-g pulps give valuable
information about the behavior of gold particles after the material
has been pulverized (i.e., delayed comminution).
6. Polished sections can provide the necessary information to correct
the calculated nomograph if judged necessary.
7. The variance of the Analytical Error is usually negligible in the
variance between the 100 hand-picked samples made of p fragments.
8. The histogram of the 100 assays performed on p 1-cm fragments can
provide a good estimate of the proportion of gold easy to sample
(i.e., the Ingamells’ low background gold content).
9. Since the p 1-cm fragments are collected one by one at random, at
least in many cases, there is no Grouping and Segregation Error
included in the estimation of the sampling constant K.

2.3.6.2 Disadvantages of the recommended approach

1. The one hundred samples made of p 1-cm fragments often generate a


few outliers. Some people see this as a problem. Actually it
translates the behavior of the gold quite well, and could even lead to
a good modeling of the Poisson distribution using the histogram of
the 100 assays. One issue is very welcome, and always true, these
outliers minimize the negative effect of the Analytical Error.
2. The test is performed on a calibrated size fraction, and is not
representative of the grade in the total sample. This leads to a particle
size distribution factor about 0.5 instead of 0.25, but we know that in
our calculations. If the gold content of the finer size fractions is
higher than the tested fraction, then the mineralogical factor used in
the tested fraction would be slightly conservative, therefore the
sampling constant K would also be conservative. In all cases, we
know exactly where we stand.
3. The entire test is performed on a single composite. But, material to
be sampled at the mine and especially at the mill is often made of
many areas as well. At least, we confine the composite to a single
type of mineralization.

60
Part 1. Theory

Table 2.1. Summary of results from the Heterogeneity Test


Sample Sample p g/t Sample Sample p g/t
number weight fragments gold number weight fragments gold
(g) (g)
1 49.11 52 43.4 51 50.37 47 94.06
2 50.75 55 52.48 52 50.88 56 176.61
3 50.7 44 111.99 53 49.26 56 64.75
4 49.74 53 81.83 54 50.09 49 64.79
5 49.94 51 116.38 55 50.83 51 140.45
6 50.62 55 44.94 56 50.1 57 91.22
7 50.7 40 43.98 57 50.71 47 131.88
8 50.43 48 87.7 58 49.66 62 87.21
9 49.98 43 91.42 59 49.67 51 160.85
10 51.51 60 28.55 60 50.42 53 141.2
11 50.85 50 98.48 61 49.49 54 48.44
12 51.26 49 71.04 62 50.79 52 106.51
13 50.59 52 182.62 63 50.34 54 52.76
14 49.98 49 61.74 64 50.85 54 57.56
15 50.35 56 81.11 65 50.62 59 129.92
16 50.31 57 40.72 66 49.43 47 69.45
17 50.79 50 61.02 67 50.49 46 77.85
18 49.91 56 95.98 68 49.98 48 214.77
19 49.8 52 82.68 69 49.71 48 114.74
20 50.35 53 98.52 70 49.59 54 36.89
21 49.45 49 40.07 71 50.13 49 92.01
22 49.85 56 307.63 72 49.28 56 52.07
23 50.62 47 112.75 73 50.33 45 158.51
24 50.66 51 159.47 74 49.65 64 176.58
25 49.94 47 122.51 75 50.13 51 69.97
26 49.66 49 84.91 76 49.51 49 48.51
27 49.58 47 104.76 77 49.66 48 91.42
28 49.91 53 205.61 78 50.58 52 73.33
29 49.73 50 83.61 79 49.53 44 80.77
30 50.37 50 31.6 80 49.34 48 55.88
31 50.22 50 449.37 81 50.34 49 87.9
32 50.62 49 98.93 82 49.47 51 130.81
33 50.42 48 50.94 83 50.5 47 59.34
34 50.54 54 88.82 84 50.03 57 61.81
35 50.25 50 137.12 85 49.18 51 36.07
36 50.78 60 97.15 86 50.4 58 43.64
37 49.78 49 108.22 87 50.58 48 208.05
38 50.48 60 85.9 88 49.77 61 57.49
39 49.93 48 62.08 89 50.93 49 48.3
40 50.76 52 110.07 90 50.41 55 58.86
41 50.66 48 60.16 91 49.65 56 127.87

61
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

42 49.72 55 260.15 92 49.8 54 30.2


43 50.79 49 55.26 93 49.89 57 29.14
44 50.69 55 55.57 94 50.33 56 62.08
45 49.75 56 36.38 95 49.74 48 102.94
46 50.32 53 61.29 96 50.28 52 40.21
47 49.14 53 123.2 97 50.76 45 275.1
48 49.47 47 130.78 98 50.03 55 85.84
49 50.23 50 158.13 99 50.03 52 41.21
50 50.72 53 118.3 100 50.32 57 117.86

Figure 2.3. Plot of the 100 fire assays for the Heterogeneity Test

2.3.6.3 Calculation of the sampling constant K

Because the fraction from which the 100 subsets were collected was made
of a very large number of fragments, formula [2.50] can simplify to:

62
Part 1. Theory

2
s FSE MS
C [2.58]
dx
with

C  f  g c

We know dx is very close to 1, therefore the following estimated value of


K is quite accurate regardless of x:

0.4278  50.2
C  18.6 g / cc
1.053

 
K  C d  18.6 1.050.5  19.1

It is interesting to notice that if we had used the average number p = 51.7


fragments instead of the average mass of the sample MS = 50.2g, an
average density of 2.7, a fragment shape factor about 0.5 for this kind of
material, and a fragment size distribution factor about 0.75 which is
appropriate for fragments between two consecutive screen openings of a
given series, we would have obtained almost the same result:

0.4278  51.7  2.7  0.5  0.75


K  19.8
1.0530.5
2.3.6.4 Calculation of the exponent x

Clearly, results from all tests described above do not provide the
necessary information to accurately calculate the exponent x in formula
[2.58]. There is no doubt it should be x = 3, unless the liberation factor is
modeled as a function of d itself. The following solutions, suggested many
times by Gy’s literature, are:
0.5
d 
 Use the approximate, empirical formula      using the
d 
liberation size d  for the liberation factor, unless better information,
based on solid metallurgical or mineralogical facts is available, or
 define a liberation curve, including the liberation size, using valuable
information from mineral processing experiments performed to
optimize the process, or, if not available yet,
63
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

 define a liberation curve, including the liberation size, using necessary


observations and tests from a logical, comprehensive, and complete
mineralogical study, which should be performed anyway for the
feasibility study of any new project.

2.3.6.5 Calculation of a sampling nomograph

Using the following equation it becomes easy to calculate a sampling


nomograph like the one illustrated in figure 2.4, test the validity of an
existing protocol, and suggest a better one if judged necessary.

 1 1 
2
s FSE   K  d
2.5

MS ML  [2.59]

As clearly shown on the nomograph, the tested protocol (i.e., a 5-kg


sample collected from a 750-kg blasthole pile, crushed to 95% minus
0.17 cm, then split to a 250-g sample, then pulverized to 95% minus 150
µm, and finally fire assayed using a 30-g analytical subsample) is a safe
protocol with respect to FSE.

2.3.6.6 Checking the validity of the calculated value for K

A size fraction analysis was performed, the results of which are shown in
table 2.2. The gold content varies by a factor 2 between size fractions. In
this particular case, the variation is not severe enough to worry about a
significant change in the value of the mineralogical factor c, therefore the
calculated value for K should be quite reliable, and slightly conservative.
It may not always be that way. However, observed differences between
size fractions should definitely alert us to potential, severe problems with
delimitation, extraction, and preparation biases if the selected sampling
equipment, and the way it is used, are not correct.

64
Part 1. Theory

Table 2.2. Testing the gold grade changes between fragment size
fractions

Size % Sample Weight (g) of Gold content Gold Weighted


fraction Split weight +150 µm (g/t) of +150 content (g/t) average
in cm (g) sent to lab µm of –150 µm gold
content
(g/t)

+1.25 10 1087 70.38 197.72 76.56 84.40

-1.25+0.63 5 2624 18.29 1475.94 89.90 99.56

-0.63+0.33 5 827 28.04 436.83 85.97 97.87

-0.33+0.17 5 778 31.94 305.81 78.47 87.80

-0.17+0.08 5 584 24.04 372.54 77.24 89.40

-0.08+0.04 5 418 20.11 283.25 71.32 81.52

-0.04+0.02 5 332 16.27 306.41 81.27 92.30

-0.0212 5 447 21.44 227.22 174.51 177.04

65
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

Figure 2.4. Testing an existing sampling protocol on the nomograph

2.3.7 Recommended method #4: Determination of IH L  f  g  c    d 3


for the top size fraction of a typical comminution stage

This method is longer but may be more accurate because of the limitation
of hypothesis #2 under different conditions of comminution.

1. Collect a large composite sample representing a single geological


unit, from leftover half core samples (e.g., fifty 6-kg samples).
2. Dry the composite.
3. Crush the composite to d = 2.54 cm.
4. Split the composite into 7 sublots.
5. Screen one sublot using 2.54-cm and 1.25-cm screens
6. Wash, dry, and weigh the size fraction.
7. Spread the size fraction between 2.54 cm and 1.25 cm on a clean
surface.
8. Collect 50 samples made of 50 fragments collected one by one at
stratified random.

66
Part 1. Theory

9. Weigh each sample: the average weight of these samples is highly


relevant in equation [2.51]. Crush and pulverize each sample. Assay
each sample for copper. Convert the copper contents into mineral
contents using the main copper mineral. We obtain 50 copper
mineral results. Look at the distribution of the 50 results. Calculate
 using formula [2.51] for d = 2.09 cm.
10. Crush the second sublot to d = 1.25 cm
11. Screen the sublot using 1.25-cm and 0.635-cm screens
12. Wash, dry, and weigh the size fraction.
13. Spread the size fraction between 1.25 cm and 0.635 cm on a clean
surface.
14. Collect 50 samples made of 50 fragments collected one by one at
stratified random.
15. Weigh each sample: the average weight of these samples is highly
relevant in equation [2.51]. Crush and pulverize each sample. Assay
each sample for copper. Convert the copper contents into mineral
contents using the main copper mineral. We obtain 50 copper
mineral results. Look at the distribution of the 50 results. Calculate
 using formula [2.51] for d = 1.05 cm.
16. Repeat the same process for the other possible stages of
comminution between 0.635 cm and 0.335 cm, between 0.335 cm
and 0.17 cm.
17. For smaller values of d, use 50 one-gram samples made of ten
stratified random 0.1-g increments. Then, we can calculate  using
formula [2.51] for each respective size fraction.

2.3.8 Cases where several phases of the mineral of interest are


present

Nature is often more complex than ideal models we create to represent it.
It is not rare that the copper ore has several origins, even within a single
geological unit. In such cases we may wonder what happens to the
mineralogical factor and the liberation factor. It happens that these
phases that are completely distinct may be completely liberated relative
to each other. In some cases these phases are partially associated to one
another. But, such association usually exists only at the coarse scale. As
the size d diminishes for finer stages of comminution these phases
usually liberate relative to one another long before the copper minerals
themselves liberate.

67
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

In such cases it is advisable to take as a reference the liberation of the


various phases instead of the liberation of the copper minerals
themselves. The determination of the mineralogical factor is easy in both
cases, while the estimation of the liberation factor is easier when
considering the liberation of the various phases, as usually  is easier to
determine accurately as it comes closer to 1.

However, to do this for  it is necessary to modify the definition of the


mineralogical factor c. This can be done by replacing the copper mineral
A and its gangue by the various copper mineral phases present. In
practice we can often assume that there is one copper mineral phase that
is rich and plays the role of the copper mineral A and another copper
mineral phase that is poor and plays the role of the gangue.

If a1 and a 2 , 1 and  2 are respectively the copper mineral contents


and the densities of the phases rich and poor, it is possible to give the
mineralogical factor the more general form as follows:

c
a1  aL aL  a2  a  a   a  a   [2.60]
a 2 a1  a2 
1 L 1 L 2 2

If we make a1 = 1 and a 2 = 0 we come back to the usual general case.

Depending on which type of copper ore that is used to determine the


liberation curves, we may want to consider such a transformation for the
calculation of c prior to the calculation of  .

2.3.9 A pragmatic formula for liberated materials

2.3.9.1 A precious metal example: cases where gold is liberated

In cases where most of the gold reports to size fractions above 80 µm it is


likely that the fine pulverization performed with laboratory mills will
liberate the gold. In this category we may also include alluvial gold.
Assumption 1 made earlier for the derivation of equation [2.37] is still
valid, while assumption 2 may show a weakness. Therefore, we may
rewrite equation [2.35] as follows:

68
Part 1. Theory

IH L   f d    3
a   aL  M L
2

[2.61]
  aL2  M L

By definition we know the gold is liberated, therefore the shape factor is a


function of the density, thus:

IH L   d  f   
3
a   aL  M L
2

[2.62]
  aL2  M L

Let’s develop this relation for the density class ρg of the gangue and for
the density class ρAu of the gold. Let’s also call 1 the infinitesimal gold
content of the gangue and 2 the infinitesimal weight of gold in the
gangue:


IH L   d  f g  g
3 1  aL 2  2
 f Au  Au
a Au  aL  M LAu 
2

 [2.63]
  aL2  M L aL2  M L 

Obviously, the first term of the sum is negligible when compared to the
second one. Furthermore, aAu = 1 by definition, and aL is usually very
small. Therefore [2.63] simplifies as follows:

  M 
IH L   d3  f Au Au2 LAu  [2.64]
  aL  M L 

By definition:

M LAu

 ML
 aL
[2.65]

Changing the sign sum for the particle size distribution factor gAu
equivalent and for aL, we obtain the very useful simplified formula:

 Au
IH L  f Au  g Au  d Au
3
 [2.66]
aL

69
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

If fAu = 0.2, gAu = 0.25, and ρAu = 16 (in practice, native gold often alloys
with some other metals), useful sampling nomographs can be calculated
with the following formula:

 1 1  0.8 3
2
s FSE    d Au [2.67]
 M S M L  aL

Using this formula, convenient sampling nomographs can be calculated


like the one illustrated in figure 2.5. A gold deposit shows gold particles
all way up to 700m in a given type of mineralization. After pulverizing
the material very fine, most of the gold is liberated or very near liberation.
We may wonder what would be the correct sample weight to perform
assays with s FSE = 0.15 (i.e., a tolerance of 15% relative). Assuming the
large sample is screened on a 150-m screen, and the + 150-m fraction
is fire assayed entirely we may wonder how much of the - 150-m
fraction should be used for fire assay. The expected gold content at the
cutoff grade, where the maximum precision is required for ore grade
control, is 1.2g/t. A 10163-g sample is needed for the metallic screen
assay, and a 100-g sample from the –150-m fraction needs to be fire
assayed. Assuming the mill feed is sampled after the grinding circuit,
providing one 8-hour shift sample, if only a 30-g or 50-g fire assay is
performed on that sample, the estimated gold content would be totally
ludicrous. Managers, manufacturers of sampling equipment, engineering
firms, and laboratory supervisors have the duty to understand, detect, and
correct such problem.

Similar derivations could be performed for other minerals of interest, and


lead to their own version of equation [2.67].

70
Part 1. Theory

Figure 2.5. Nomograph for liberated gold for 700µm and 150µm gold
particles, a grade of 1.2 g/t gold, and a targeted sFSE of 15% relative

2.3.10 A pragmatic formula for cases where a mineral is associated


with another major mineral

Many minerals such as galena, sphalerite, chalcopyrite, pyrite, chromite,


molybdenite, etc… often contain some precious metals such as gold,
platinum, palladium, rhenium or others. For example, for a molybdenite
ore, we can assume that the rhenium grade varies little inside the host
mineral. In fact, we may consider it relatively constant. Under these
conditions, which need to be verified by a mineralogist, the rhenium
content becomes somewhat irrelevant as far as sampling for rhenium is
concerned. We should concentrate our attention on the mineral content
alone (e.g. molybdenite). Then, we come back to the general cases
initiated by formulas [2.50] and [2.51].

71
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

2.3.11 A pragmatic formula for a size distribution analysis

For many industries the determination of the particle size distribution of a


material is often one of the most important criteria for the following
reasons:

 the assessment of the quality of a product,


 the assessment of the efficiency of a process,
 the compliance with the clause of a commercial contract,
 the assessment of the accuracy of a sampling system, and
 the assessment of the representativeness of a sample.

Let’s define the necessary notations in this section:

LC a size fraction of interest


a Lc the proportion of LC in the lot L
M Lc the mass of the size fraction of interest
N Lc the number of fragments in LC
Fj a fragment of LC
M j the mass of the fragment F j
a Fj the critical content of F j
FLc the average fragment of the size fraction LC
M FLc the mass of FLc

Now it is possible to rearrange the general equation [2.35] to calculate


IHL:

within LC ai  a Fj  1
outside LC ai  0
therefore:

IH L  
1  aLc 2  M 2j  M i2 1  a Lc  M j  a Lc
2 2
M i2
j
2
a Lc ML
i M j a 2
 
M Lc
 i M
L Lc L
[2.68]

72
Part 1. Theory

Now, let’s introduce the auxiliary parameters that have the dimension of a
mass:

M i2
XL   [2.69]
i ML

M 2j
X Lc   [2.70]
j M Lc

Now, assuming there is only one size fraction of main interest:

IH L 
1  a Lc 
2

X Lc  X L [2.71]
a Lc

2
a Lc is usually very small and we can eliminate it in a first approximation:

IH L 
1  2aLc  X  XL [2.72]
Lc
a Lc

X L and X Lc should be expressed in practical terms because Mi and Mj


are not known. It is relatively easy to do this when we have at our disposal
an existing fragment size distribution that has been performed in the past
on a similar material. It is sufficient to substitute Fi and Fj of any given
size fraction by the average fragment FLx of that size fraction.

Then, more notations are needed:

N Lx the number of fragments in any class Lx


M FLx the mass of the average fragment FLx
a Lx the proportion of Lx in the lot L.

By definition:

M Lx
M FLx  [2.73]
N Lx
73
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

Changing M i for the average fragments:

M i2 N  M FLx  M FLx M  M Lx
XL     Lx   FLx [2.74]
i ML x ML x ML

M Lx
But, M L  then,
a Lx
X L   M FLx  a Lx [2.75]
x
In the same way:

M 2j N Lc  M FLc  M FLc M Lc  M FLc


X Lc      M FLc [2.76]
j M Lc M Lc M Lc

IH L 
1  2aLc  M   M FLx  a Lx [2.77]
FLc
a Lc x

All terms in this formula can either be calculated or estimated when we


have at our disposal a rough idea of the actual size distribution. Proceed
by iteration if necessary.

1. By direct measurement of the mass of a given number of


fragments selected at random within the size fraction.
2. By calculation using the volume V, shape factor f, and density ρ of
fragments, with the following two formulas:

M FLc  VFLc    f    d FLc


3
[2.78]
M FLx  VFLx    f    d FLx
3
[2.79]

d FLc and in a same manner d FLx can be estimated using the upper and
lower opening of two screens defining a size fraction, with the following
formula:

74
Part 1. Theory

d FLc  3
UpperOpening 3  LowerOpening 3 [2.80]
2

Finally, IHL can take its final, complete form:

 1  3 
IH L  f     2 d FLc   d FLx
3
 aLx  [2.81]
 aLc  x 

Now we may calculate the variance of the Fundamental Sampling Error:

 1 1   1  3 
2
sFSE    f     2 d FLc   d FLx
3
 aLx  [2.82]
MS ML   aLc  x 

This formula can often be simplified for many applications:

 If M L  10M S
 If d FLc is not much different from d
 If a Lc is small, then

f   1  3
2
sFSE    2 d FLc [2.83]
MS  aLc 

Remember the limitations: If the coarse fragments represent a large


proportion of the lot, for example more than 20%, the term
x
3
d FLx  a Lx should be kept. Furthermore, if a Lc is becoming too large

(e.g., >0.30) then the approximation taken in formula [2.72] is no longer


valid: these are rare cases.

2.3.12 A pragmatic formula using the size of the constituent of


interest

Often, especially for trace constituents, it is difficult and impractical to


determine the liberation factor with sufficient accuracy, and this makes
formula [2.55] vulnerable. Enormous literature has been written on this
subject, the best one by Dr. Dominique François-Bongarçon18,40.
75
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

However, it is not a must to use the conventional, favorite approach


suggested by equation [2.55]. The following suggestion is pragmatic,
accurate, and falls in line with Ingamells’ approach; it is summarized in
the three following statements:

1. Use the approach suggested by equation [2.67] that was used for
gold; it can be generalized to many other components of interest
as shown by equation [2.79].
2. Verify that the sample mass suggested by the generalized version
of equation [2.67] is compatible with the mass necessary to
represent all size fractions in the lot by using equation [2.83].
3. The largest required sample mass for a pre-selected precision,
obtained by equation [2.83] (i.e., using d) and equation [2.79]
(i.e., using d INT defined below) necessarily takes priority on
deciding what the sampling protocol should be.

Generalization of equation [2.67] by defining new notations:

f INT the shape factor of the constituent of interest


g INT the particle size distribution factor of the constituent of interest
 INT the density of the constituent of interest
d INT the maximum size of the constituent of interest particle, liberated or
not, or cluster of such particles contained in a single fragment of the
surrounding matrix; d INT is defined as the size of a screen that would
retain no more than 5% by weight of all the particles of the constituent of
interest.

Thus, we obtain the very useful simplified formula:

 INT
IH L  f INT  g INT  d INT
3
 [2.84]
aL

Useful sampling nomographs can be calculated with the following


formula:

 1 1  f INT  g INT   INT  d INT


3
2
s FSE    [2.85]
MS ML  aL

76
Part 1. Theory

The great advantage of this approach is its accuracy and the easiness to
collect the relevant and necessary information through microscopic
observations.

2.3.13 A pragmatic formula to estimate the In Situ Nugget Effect

A similar formula like [2.79] could be derived addressing the necessary


mass of the observation support when drilling a mineral deposit. At this
stage, there are no fragments, therefore no FSE. Nevertheless, at very
short range (e.g., few centimeters or few meters), there is a certain amount
of heterogeneity related to the size of the particles of the constituent of
interest, how these particles cluster, and how much sample mass is
collected as we drill. The author chose to call this new source of
uncertainty, the In Situ Nugget Effect. A thorough study of the In Situ
Nugget Effect is presented in chapter 8.

2.3.14 The author’s pragmatic opinion from long, extensive


experience

Most of the time practitioners are using the TOS improperly. Then, as
results do not fit their theoretical approach, they often conclude Gy’s
work is far from being universal, and they too quickly dismiss his superb
work. The view of Gy’s work is too concentrated on FSE, and
furthermore Gy’s work on FSE is understood only in a very simplistic,
naïve way. One of the main objectives of this thesis is to correct this
misunderstanding by creating about ten different approaches. It would
then take very little effort for any sampling practitioner to quickly find
which approach is appropriate for his or her particular application.

2.3.15 A guideline for the acceptable standard deviation of FSE

A direct consequence of DQO introduced earlier, is to decide what would


be the appropriate relative standard deviation s SFE expressed in %. There
as well, practitioners follow vague guidelines using s SFE = 0.10 (i.e., a
tolerance of 10% relative). It does not work that way, as not only they
should satisfy their well defined DQO, but they should also make sure
they remain well within the domain of normal statistics. The following
recommendations are suggested:

77
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

1. Under no circumstances should s SFE be larger than 16%, as


above this number we slowly leave the domain of normality and
slowly enter the domain of Poisson processes. All above
pragmatic formulas are not valid for Poisson processes.
2. For early exploration and grade control of precious metals, for the
determination of trace constituents, and for environmental
sampling it is suggested to make sure s SFE is smaller than 16%.
3. For early exploration and grade control of base metals, and for
process control in general, it is suggested to make sure s SFE is
smaller than 10%.
4. For material balance or metallurgical accounting in general, it is
suggested to make sure s SFE is smaller than 5 to 7%.
5. To sample important constituents in commercial sampling in
order to determine the settlement price of commodities, very
stringent precision is required, and it is suggested to make sure
s SFE is smaller than 1 to 2%.

2.4 THE GROUPING AND SEGREGATION ERROR

m(GSE)  0 in a first order approximation

and taking relation [1.6] into relation [2.33]:

1  P  an  a L  M n2  1  P
2

s 2 GSE     a 2  M 2   P  N DH L [2.86]
P n  L L  Gn

2.4.1 Relationship between Constitution and Distribution


Heterogeneities

Variances from random variables are additive. DH L from equation [1.6]


and CH n from equation [1.9] are variances and additive:

CH L  DHL  CH n [2.87]

Therefore:

78
Part 1. Theory

CH L  DH L  0 [2.88]

It becomes clear that the maximum of DH L is CH L itself when the


groups of fragments are made of only one fragment.

DH L max  CH L [2.89]

The minimum of DH L is defined as an absolute minimum DH L min


under which it is impossible to homogenize the material further. It is an
inaccessible limit at which each group of the lot could be regarded as a
random sample. In practice, because it is much easier to induce
segregation than to correct for it, this minimum is never zero. It is actually
a random variable the mean of which can be related to CH L as follows
when the number of groups and fragments are both large numbers:

meanDH L min 
NG
CH L [2.90]
NF

2.4.2 Definition of the Grouping Factor

A Grouping Factor Y is defined as the ratio of the number of fragments to


the number of groups, hence a measure of the group size.

1 NG
 [2.91]
Y NF
NG and NF are often extremely large numbers, and they are also very
different, therefore Y is often a large number itself. As a result, for
convenience, we can transpose 1+Y to Y for the denominator, so it can
never be zero. Then:

1
mean[ DH L ]min  CH L [2.92]
1 Y

Therefore the practical domain of DHL is somewhere between DH L min


and CHL:

79
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

1 1 Y
CH L  DH L  CH L  CH L [2.93]
1 Y 1 Y

If NF = NG, or if each increment could be made of only one random


fragment, then Y = 0 and therefore DH L min  CH L . On the other hand
if the size of the group is such that it is the entire lot, then Y = +∞ .

2.4.3 Definition of the Segregation Factor

Most of the time we must deal with a certain amount of natural


Distribution Heterogeneity DHL. The domain of the variable factor is now
known, anywhere between 1 and 1+Y. We chose to write it ZY, where Z is
defined as a Segregation Factor. It follows:

DH L 
1 Y  Z a  a  M 2
CH L  1  Y  Z  G CH L  1  Y  Z N G  i 2 L 2 i
N
2

1 Y NF i aL  M L
[2.94]

This formula shows that only three factors are responsible for the
magnitude of Distribution Heterogeneity.

 CHL that is responsible for FSE, and we already know how to


minimize FSE.
 Y that depends on the size of the extracted increments; the
smaller the size of the increments the smaller Y is.
 Z that depends on the amount of transient segregation that is
taking place in the material to be sampled.

The Grouping and Segregation Error GSE is completely defined by its


mean and variance:

meanGSE   0 [2.95]

2
sGSE  YZs FSE
2
[2.96]

If the variance of GSE is the product of 3 factors, this would suggest that
the cancellation of only one factor could eliminate GSE.

80
Part 1. Theory

 It is not possible to cancel the variance of FSE unless the sample


is the entire lot, which is not the objective of sampling. However,
it should be minimized and we know how to do this.
 It is not possible to cancel Y unless we select a sample by
collecting 1-fragment increments at random one at a time. This is
not practical however it was done in recommended method #3 for
the experimental determination of IHL. In a routine sampling
protocol, the right strategy is to collect as small and as many
increments as practically possible so we can drastically minimize
Y; this proves to be by far the most effective way to minimize the
variance of GSE.
 It is not possible to cancel Z which is the result of transient
segregation. All homogenizing processes have their weaknesses
and are often wishful thinking processes; this proves to be the
most ineffective way to minimize the variance of GSE.

2.5 PREVENTIVE WISDOM

From practical experience with many minerals and other components of


interest Gy’s formulas are wonderful preventive tools and prove to work
extremely well as long as it is possible to keep the standard deviation of
the Fundamental Sampling Error below 16% relative (i.e., sFSE = 0.16).
However with some minor constituents and especially trace constituents it
is not always possible to collect a sample mass that is sufficient to fulfill
these preventive requirements. Then, the works of Visman and Ingamells
may allow us to take action in reverse, which consists of looking at the
empirical variance from a series of replicate samples. Of course, by doing
so, it is not rare to still end up in the domain of Poisson processes that
may have to be modeled in order to have an estimate of what the
consequences are. This is also especially true for the In Situ Nugget
Effect.

In any event, the simple act of replicating a series of primary samples and
process each in an identical fashion through the entire sampling-analysis
pathway, results in an extremely easy estimate of the empirical total
sampling-cum-analysis error, the relative magnitude of which (relative to
the average of the collected samples), allows an objective, quantitative
measure of how far above 16% the particular sampling procedure is.
Deviations above 100% are not uncommon with sampling operations
which have not been contemplated in the light of TOS. It is obvious that
the bias-generating sampling errors are the first items on any

81
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

improvement agenda (IDE, IEE, IWE and IPE). Only after documented
successful elimination of these errors can the focus meaningfully be
translated to combating GSE especially and lastly FSE.

82
Part 1. Theory

CHAPTER 3

THE WORKS OF VISMAN AND INGAMELLS


RELEVANT TO GY’S WORK

3.1 SCOPE

The objective of this thesis is not to compare or debate the works of


Visman 5,6 and Ingamells 8-11 with Gy’s work, as there is no doubt such a
document would be futile because Gy’s achievement is well established as
a flawless Theory of Sampling. As a result, some of the long derivations
exposed in Visman and Ingamells’ literature are of no interest. Rather, it
is important to emphasize the work that may be beneficial and also
provide relevant, harmonious additions in some areas, and there are many
possibilities. Such additions may enhance our capabilities to predict
sampling difficulties by using a stronger strategy and also to design better
sampling experiments that would allow us to further understand the
heterogeneity of minor and trace constituents.

3.2 AN INTRODUCTION TO POISSON PROCESSES

Amounts of minor and trace constituents are the key words in many
industries where their accurate determination is of the utmost importance.
The Theory of Sampling would be incomplete without an influx of what
we can do when Poisson processes are nearly inevitable; for example
quantifying trace amounts of constituents in pharmaceutical products, in
high purity materials, in the environment, in genetically modified
organisms, in precious metals exploration and many other cases. This is
where the work of Ingamells is priceless; his entire work is based on
Poisson statistics. The author should make the emphasis very clear,
because it is crucial: without a good understanding of Poisson processes
there is no possible in-depth understanding of the TOS because too many
subtleties become elusive, and this has escaped the attention of most
sampling practitioners around the world and it needs to be corrected.

3.2.1 A simple, pragmatic observation

Figure 3.1 represents the position of a few gold particles in a piece of


diamond core drilling. At first sight there is no obvious pattern, and the
particles seem distributed at random within that small volume. We could

83
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

easily imagine many other examples. A Poisson process is a model of


such phenomenon, which uses the theory of probability to describe this
sort of highly random behavior; a Poisson process takes place when the
probability of including one or several gold particles in the sample is a
small probability. Some people may object to this model, as if we look at
the same volume within a larger area, trends or local clusters with a
logical pattern may slowly appear. The point is that a sample is too often a
miniature amount of material and within such small scale the observed
feature, such as why this particle is there rather than here or why is it
bigger than the other one, is a property of statistical independence; and
yes, in practice this assumption of independence is satisfied only
approximately. The Poisson model is only the simplest and most random
possible model to describe a phenomenon where the collected sample is
obviously one or several orders of magnitude too small to contain a
sufficient, statistically significant number of gold particles.

Figure 3.1. Simplistic illustration of a few gold particles within a piece of


HQ-diameter diamond core drilling sample

3.2.2 The Poisson Model: the law of small probabilities

The Poisson model is a limit case of the binomial model where the
proportion p of the constituent of interest is very small (e.g., fraction of
1%, ppm or ppb), while the proportion q = 1-p of the material surrounding
the constituent of interest is practically 1. We will expand on this concept
84
Part 1. Theory

in Part 2, Chapter 11. Experience shows that such constituent may occur
as rare, tiny grains, relatively pure but they don’t have to be, and they may
or may not be liberated. As the sample becomes too small, the probability
of having one grain or a sufficient amount of them in one selected sample
diminishes drastically; furthermore, and this is where it becomes
confusing to many, when one grain is present, the estimator a S of
a L becomes so high that it is often considered as an outlier by the
inexperienced practitioner while it is the most important finding that
should indeed raise his attention. It is rather amusing that people get so
upset by a high value when they don’t pay attention to the many that are
under-estimated; people are easily taken prisoner of paradigms, such as
better-known normal or log-normal distributions.

Some people may object to Poisson statistics to model sampling errors


when the constituent of interest is not liberated, as they think in terms of
discrete particles. Liberation does not have to occur for Poisson processes
to take place that are solely the result of the sample mass being too small
by one or several orders of magnitude. The remark can also be generalized
to the In Situ Nugget Effect when the basic module of observation is too
small by one or several orders of magnitude in order to represent the true,
unknown grade at any given point; this is further explored in chapter 8.
Then, when using the Poisson model it is irrelevant to argue about a
sample mass that should be 1500g for example rather than 1000g; we may
look at why the sample mass is 1000g when it should be 100000g; we no
longer have to look at too many details, but we must focus on the
important ones that can make a staggering difference. Experience proves
again and again that in such a search, it is the maximum size of the
nuggets, clusters, or nearly pure constituent grains, liberated or not, that
can make a tremendous difference.

Let’s call Px  r  the probability of r low-frequency isolated coarse


grains, liberated or not, appearing in a sample, and  is the average
number of these grains per sample.

r
P x  r   e  [3.1]
r!

with r = 0, 1, 2, 3,…

85
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

The variance of the Poisson distribution is   n  p  q  n  p since q is


close to 1. The mean value of the Poisson distribution is   n  p .

3.2.3 Shape of a Poisson distribution

It takes a value of   6 for the Poisson distribution to start to vaguely


look like a normal distribution. When   1 there is the same probability
of having r = 0 or 1. When  becomes much smaller than one, the mode
of the distribution becomes skewed toward low grades in a drastic way
(i.e., most of the time we completely miss what we are looking for.).

3.2.4 A simple, but useful exercise

Ten 1000-g samples collected from a large lot, and ground to minus 500-
µm, were submitted to a gravity concentration for the coarse gold. The
number of coarse gold particles found in each sample was
2,1,0,0,2,1,3,0,1, and 2 respectively.

1. Calculate the average number of coarse gold particles per sample:

  2  1  2  1  3  1  2 / 10  1.2

2. Calculate the probability of occurrence of 0, 1, 2, 3, 4, 5, and 6


coarse gold particles in any 1000-g sample selected at random
from the same lot.

Using equation [3.1] the following results are obtained:

 1.20 
Px  0  2.718
1.2
   0.301
 0 ! 
1.2  1.2  
1
Px  1  2.718    0.361
 1! 
1.2  1.2  
2
Px  2  2.718    0.217
 2 ! 
1.2  1.2  
3
Px  3  2.718    0.087
 3 ! 
86
Part 1. Theory

 1.24 
Px  4  2.718 
1.2
  0.026
 4! 
1.2  1.2  
5
Px  5  2.718    0.006
 5! 
1.2  1.2  
6
Px  6  2.718    0.001
 6! 

3. If one of these coarse gold particles on average makes a


contribution of about 0.5 g/t, estimate the variability of the typical
1000-gram samples using the above distribution.

30.1% of the samples would show 0 g/t gold


36.1% ,, 0.5 g/t
21.7% ,, 1.0 g/t
8.7% ,, 1.5 g/t
2.6% ,, 2.0 g/t
0.6% ,, 2.5 g/t
0.1% ,, 3.0 g/t

It is interesting to notice that   1.2 , therefore there is a much greater


chance to underestimate the gold content than to overestimate it. This has
huge implications in sampling, especially in above-cutoff grade control
when too few samples are used to make an economic selection.

3.2.5 Additivity of Poisson Processes

Someone may think this is really a bad case; it is nothing when compared
to what follows. When samples taken contain discrete grains of the
constituent of interest, and they are sub-sampled in such a way that the
sub-samples also contain discrete grains of reduced size, a double Poisson
distribution of the assay values is likely to take place. Actually, this
phenomenon is very common in sampling, and people may work for years
in that case and never see a thing. The most important feature of the
Poisson distribution is indeed its additivity.
If each primary sample contains a limited average number  of
constituent of interest grains, and the sub-samples they generate also
contain a limited average number  of reduced constituent of interest
grains, the distribution of assay values is doubly Poisson.
87
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

 r  e   n  e  
Px  r , y  n  P[ x  r ]P[ y  n]   [3.2]
r! n!
x+y has distribution P   .

Because of the extreme importance of this case, more information is given


in Part 2, Chapter 11.

3.2.6 Programming factorials

Difficulties may arise in programming factorials. The use of the


improved Stirling approximation is convenient to use:

 6n  1
n
n
n!   [3.3]
e 3

3.3 VISMAN’S SAMPLING EQUATION

Visman’s sampling constants A and B are useful when the mass to be


sampled is not well-mixed (i.e., mineral grains are not thoroughly random
distributed throughout the lot to be sampled). The sampling equation is:

s2 A B
S 2
  [3.4]
N MS N

where S is the uncertainty in the average of N assays on samples of


MS
individual mass M S  .
N

This equation, on a smaller scale of course, has almost the same


significance as a similar, well-known equation suggested by D.G. Krige in
the early days of Geostatistics49:

v v 2V ' 


 2    2    
V  V '  V  [3.5]

where the variance of samples from volume v taken in the ore deposit of
volume V is the sum of the variance of samples from volume v taken in a

88
Part 1. Theory

given ore block of volume V’ and the variance of the ore blocks V’ in the
corresponding deposit of volume V. The first term can be considered in a
first approximation as a random term where the volume v taken from V’
can make a difference. The second term can be considered as a non-
random term where the quantity of well-defined blocks V’ taken from V
make the difference.

Immediately, for the sake of consistency we should define and translate


all these new terms into their equivalents in Gy’s TOS.

A is the Visman homogeneity constant. It is the Gy’s Constant Factor of


Constitution Heterogeneity IHL multiplied by the square of the average
content of the lot, with aL expressed in appropriate units, or as a
proportion as part of 1. So, from now on let’s remain cautious as in
Gy’s theory contents, such as aL, are expressed as a proportion, part
of one. There is nothing wrong in expressing the aL content in either
%, g/t, ppm or ppb as long as these units are properly kept in mind
where appropriate when calculating a relative, dimensionless
variance or a term like IHL.

A  IH L  aL2 [3.6]

B is the Visman segregation constant. It should be well understood it is a


constant only under temporary, specific conditions, as segregation is a
transient phenomenon changing all the time. It is related to the variance of
2
Gy’s Grouping and Segregation Error S GSE as follows:
A
B  s2   sGSE
2
 a L2  N  s SE
2
 a L2 [3.7]
M S
2
N being the number of increments collected in a sample, and s SE being
the segregation variance if only one increment was collected. N can also
consist of different samples collected in a lot. The weakness of Visman
and Ingamells approaches is not to emphasize the irrefutable fact that the
2
variance s SE is a transient phenomenon that can change all the time as
the lot is transferred from a given place to another; as a result it was
concluded in Gy’s theory that the quantification of this variance is
somewhat an exercise in futility, however understanding its behavior can
lead to beneficial strategies.

89
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

3.4 VISMAN’S EXPERIMENT

Valuable experiments can be made to quantify the Constant Factor of


Constitution Heterogeneity and also the variability due to local
segregation. Such segregation may be measured on a small scale (i.e.,
within a few kilograms of particulate material) or on a larger scale (i.e.,
within a mineral deposit). Visman suggested that two series of samples
properly subsampled and assayed (i.e., subsampling and assaying
uncertainties must be of the second order of magnitude) could be
collected; one series is using small samples of mass M S1 and the other
series is using large samples of mass M S 2 . As a result variances s12 from
2
small samples and s 2 from large samples are obtained, and from [2.50],
[2.90], [3.4], [3.6] and [3.7], and with N = 1, we obtain the following
relationships.

From the series of small samples, assumed to be collected at stratified


random as suggested in figure 3.2, condition with respect to equations
[2.27-2.29]:

s12  1 1 
2
 s FSE
2
1  sGSE1  
2
  IH L  s FSE1 Z  Y
2
[3.8]
aL  S1
M M L 

Assuming the mass of the lot is at least ten times larger than the mass of
the sample, [3.8] can be simplified to the following form:

IH L  a L2 IH L  a L2  Z  Y
s12  ( s FSE
2
1  sGSE1 )a L 
2 2
 [3.9]
M S1 M S1

In a similar way, from the series of large samples, as shown in figure 3.3,
we obtain:
IH L  a L2 IH L  a L2  Z  Y
s22  ( s FSE
2
2  sGSE 2 )a L 
2 2
 [3.10]
M S2 M S2

IH L  a L2
1  Z  Y   IH L  aL 1  Z  Y   1  Z  Y  M S 2  IH L  aL  M S1  IH L  aL
2 2 2
s12  s22 
M S1 M S2 M S1  M S 2
[3.11]

90
Part 1. Theory

IH L M S 2  M S1 a L2
s12  s 22  1  Z  Y  [3.12]
M S1  M S 2

1  Z  Y IH L  aL2  s1 


 s22 M S1  M S 2 
2
[3.13]
M S 2  M S1

If the material is not segregated then Z=0, or if increments are made of


one random fragment selected at random then Y=0, we obtain the well-
known, valuable Visman’s relationship:

IH L 
s
2
1 
 s22 M S1  M S 2 
[3.14]
a L2 M S 2  M S1 

3.4.1 Discussion about the experimental variances

Several scenarios may develop during such an experiment; let’s discuss


them briefly.

3.4.1.1 Case #1: s1  s2 and it is a large variance.


2 2

This case is an indication that the sample mass is not the most important
issue. The constituent of interest is segregated on a large scale and
collecting many small samples is better than few large samples.

3.4.1.2 Case #2: s1  s2 and it is a small variance.


2 2

This case is an indication that the constituent of interest carries very little
heterogeneity on a small scale and on a large scale. In such rare case, it is
not difficult to perform good sampling with just a few small samples.

91
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

Figure 3.2. A first series of small samples collected at stratified random

Figure 3.3. A second series of large samples collected at stratified random

3.4.1.3 Case #3: s1  s2


2 2

This case is an indication that the sample mass is playing an important


role. Figure 3.4 illustrates a real case with isolated iron salt impurity

92
Part 1. Theory

grains in an ammonium paratungstate pure product. A Poisson’s process


affects the small samples.

Small sample

Large sample

Isolated iron impurity


particle

Figure 3.4. A case where a Poisson process affects the small samples

It is recommended to select an optimum sample mass M Sopt that will be


defined later in equation [3.41], which is a compromise between the
sample mass that is necessary to minimize the variance of the
Fundamental Sampling Error FSE and the total number of samples N that
is necessary to account for the segregation in the lot to sample. Three
cases may develop:

1. If M Sopt  M S1 then both series of samples are valid.


2. If M S1  M Sopt  M S 2 then only the large samples are valid.
3. If M Sopt M S 2 then both series of samples are invalid; we must
iterate the sampling effort using a sample mass M Sopt .

93
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

3.4.1.4 Case #4: s12  s22

This case is puzzling to many people. It is a very important case that


should raise suspicion. The content of the constituent of interest may be
much larger than usually recognized, but it is difficult to prove this
because samples are too small by one or several orders of magnitude: a
Poisson process has been taking place for both series of samples.
Figure 3.5 shows a real case where such a process took place when
sampling gold.

Coarse gold particle

Figure 3.5. A case where a Poisson process affects both series of samples

3.4.2 Case where no large samples are available

It is not always possible to collect and assay large samples. In such cases
it is possible to create artificial large samples by compositing assays from
small samples. Obviously such an approach is a simulation of reality and
94
Part 1. Theory

the test is somewhat less powerful, though it is still possible to retrieve


valuable information about the kinds of heterogeneity carried by a given
constituent in the lot to be sampled.

If Q is the compositing factor, we may write:

IH L 
s
2
1 
 s22 M S1  Q  M S1 

s 2
1 
 s22 Q  M S21
[3.15]
a L2 QM S1  M S1   1
a L2  M S1  Q1  
 Q

IH L 
s
2
1 
 s 22 M S1
[3.16]
 1
a L2 1  
 Q

3.5 INGAMELLS’ MOST PROBABLE RESULT

When samples are of a much too small mass and contain only a few
isolated large grains of the constituent of interest, liberated or not, the
distribution of these grains in a large number of samples is given by the
Poisson probability function (See derivation in Chapter 11):

 r
P x  r   e  [3.17]
r!

with r = 0, 1, 2, 3,… and where Px  r  is the probability of r low-


frequency isolated coarse grains appearing in a sample, liberated or not,
and  is the average number of these grains per sample. It is important to
note that the highest probability occurs when r is less than  , except
when  is an integer when Px     Px    1 . Therefore, the most
probable number of low frequency grains is always less than the average
number. With real samples, in which there is usually a spectrum of grain
sizes, the most probable assay is always less than the true value. If we call
the most probable result of an assay  , then:

  a L  f   [3.18]

95
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

where f   is a function of the average number of low-frequency grains;


this function must approach zero as  increases and must approach
aL  a H  as  decreases. We define a H as a relatively homogeneous
low background content easy to sample. So, as  becomes large   a L ,
and as  becomes small   a H .

Ingamells and Switzer found that the following semi-empirical function


meets these requirements 9:

aL  aH aL  aH
f    and   aL  [3.19]
2  1 2  1

Let’s define a few terms from Ingamells:

Let’s call a D the proportion of the constituent of interest that is difficult


to sample, then by definition:

aD  aL  aH [3.20]

The total mass of the constituent of interest in the sample is M S  a L .


The mass of the constituent of interest in the proportion easy to sample is
M S  aH .
The mass of the constituent of interest in the proportion difficult to sample
is M S  a D .
From the binomial model it follows that:

aL  p  aH  q  aD  1  qaH  q  aD  aH  aH  q  aD  q  aH  qaD  aH 

Then:

a L  a H  qa D  a L 
a  aH
q L [3.21]
aD  aH

Then  can be calculated as follows:

96
Part 1. Theory

aL  aH M S
  [3.22]
aD  aH d M3  M

But, from Ingamells and Pitard 8:


A IH L aL2
d M3  M   [3.23]
aL  aH aD  aH  aL  aH aD  aH 
Then, by transposing in [3.22] and simplifying:


aL  aH  M S
2
[3.24]
IH L  a L2

By transposing in [3.19], and assuming the material is not segregated (i.e.,


Z = 0):

  aL 
aL  aH  
 2

a L 2a L  a H  M S  IH L  a L2  a L  a H IH L  a L2
2a L  a H  M S IH L  a L2 2a L  a H  M S  IH L  a L2
2 2


IH L  a L2 IH L  a L2

2a L a L  a H  M S  a H  IH L  a L2
2

 [3.25]
2a L  a H  M S  IH L  a L2
2

3.6 INGAMELLS’ GANGUE CONCENTRATION

Many of the equations developed so far depend on the estimation of the


more homogeneous low background content easy to sample a H which is
equivalent on saying that we need to estimate a D the proportion of the
constituent of interest that is difficult to sample. With real data the
estimation of a H is difficult. However, it is relatively easy to make such
calculation by using the most probable result of an assay  by looking at
the modes  1 and  2 or the harmonic means h1 and h2 of the data
distribution of the small and large samples. The harmonic mean is
calculated as follows:

97
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

N
hi  [3.26]
1
i x
i

where N is the number of samples.

From equation [3.25] we may write:


2a L a L  a H  M S1  a H  IH L  a L2
2

h1   1 
2a L  a H  M S1  IH L  a L2
2
[3.27]

2a L a L  a H  M S 2  a H  IH L  a L2
2

h2   2 
2a L  a H  M S 2  IH L  a L2
2
[3.28]

whence

2 1 a L  a H  M S1   1  IH L  a L2  2a L a L  a H  M S1  a H  IH L  a L2
2 2

2M S1 aL  aH  aL   1   IH L  aL2  1  aH 


2
[3.29]

2 2 a L  a H  M S 2   2  IH L  a L2  2a L a L  a H  M S 2  a H  IH L  a L2
2 2

2M S 2 aL  aH  aL   2   IH L  aL2  2  aH 


2
[3.30]

2M S1 aL  aH  aL   1 aL   2 M S 2  IH L  aL2  1  aH aL   2 M S 2


2

[3.31]

2M S 2 aL  aH  aL   2 aL   1 M S1  IH L  aL2  2  aH aL   1 M S1


2

[3.32]

Subtracting [3.32] from [3.31] gives

IH L  aL2  1  aH aL   2 M S 2  IH L  aL2  2  aH aL   1 M S1

98
Part 1. Theory

M S 2 aL   1   1   2  aL  aH   2  aH   M S1 aL   2   1   2  aL  aH   1  aH 

 M S 2  a L  a H  M S 2   2  a H  M S1  a L  aH  M S1   1  a H  M S1  a L   2  M S1   1   2  M S 2  a L   1  M S 2   1   2

aH aL  M S 2  M S 2   2  aL  M S1  M S1   1    1  M S 2 aL   2    2  M S1 aL   1 

 1  M S 2 a L   2    2  M S1 a L   1 
aH  [3.33]
M S 2 a L   2   M S1 a L   1 

3.6.1 Discussion about the low background content

The low background content that is easy to sample a H can be a


parameter of great interest. It is a variable of its own in a mineral deposit,
or in a high purity product, or in the environment. Such variability should
be the object of a separate thesis as it can have geometallurgical and
economic implications in a project. For example, if a deposit is difficult to
sample for its gold content, it would be critically important to find out
what proportion of the gold is difficult to sample. If it is 2/3 of the gold,
the geologist may have a serious problem. If it is only 5% of the gold, it
may not be a problem. Also, in a flotation process for molybdenum, if 1/3
of the molybdenum that is easy to sample cannot be liberated at the mill
then the molybdenum recovery is likely to be poor, which may have
serious economic consequences. On the other hand if only 5% of the
molybdenum cannot be liberated it may not be too much of a problem.

3.7 INGAMELLS’ OPTIMUM SAMPLE MASS

The Ingamells’ optimum sample mass can be described as a compromise


between the necessary sample mass to minimize the variance of FSE and
the necessary number of samples N to minimize the effect of local
segregation on the overall estimation.

When calculating IHL, sample mass alone is an issue. When minimizing


the effect of segregation, the number of increments in a composite sample
or the number of samples collected over a certain target is an issue.
Therefore, the uncertainty of a selected sampling protocol depends both
on the total weight of samples taken and on their number. There is
evidently an optimum individual increment mass or an optimum
99
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

individual sample mass which will give the most information at the lowest
cost.

The cost of a sampling campaign may be broken into cost related to the
total number of samples taken and cost related to the total mass of
samples taken, plus a fixed cost related to logistics.

$  P  M S  Q  N  F  P  N  M S  Q  N  F [3.34]

Where P is the cost per gram of sample, Q is the cost per sample such as
sub-sampling and assaying, and F is the fixed cost related to local
logistics.

From equations [3.4], [3.6], and [3.7] we may write:

IH L  a L2 s SE
2
 a L2 IH L  a L2 sSE
2
 a L2  M S* IH L  a L2  sSE
2
 a L2  M S
S2     
N  M S N N  M S N  M S N  M S

Solving for N:

IH L  a L2  s SE
2
 a L2  M S
N [3.35]
S 2  M S

Transposing [3.35] in [3.34]:

$

P IH L  a L2  s SE
2

 
 a L2  M S Q IH L  a L2  s SE
2
 a L2  M S
F

S2 S 2  M S
[3.36]

P  IH L  a L2 P  s SE
2
 a L2  M S Q  IH L  a L2 Q  s SE
2
 a L2
$    F
S2 S2 S 2  M S S2
[3.37]

Differentiating [3.37] with respect to M S :

100
Part 1. Theory

d$ P  sSE
2
 aL2 IH L  aL2  Q
  [3.38]
dM S S2 M S2  S 2

d$
The minimum cost is achieved when  0 , i.e., when
dM S

P  s SE
2
 a L2 IH L  a L2  Q IH L  Q
 or M S2 
S2 M S2  S 2 P  s SE
2

IH L  Q
M S  [3.39]
2
s SE P

Substituting [3.39] in [3.37] gives the minimum cost for any desired
variance S2:

P  IH L  a L2 P  s SE  a L IH L  Q Q  IH L  a L sSE  P Q  sSE  a L2
2 2 2 2 2
$ min     F
S2 S 2 sSE 2
P S 2 IH L  Q S2

P  IH L  a L2 a L IH L  Q  sSE  P a L IH L  Q  sSE  P Q  sSE  a L2


2 2 2 2 2
$ min     F
S2 S2 S2 S2

P  IH L  a L2 2a L2 Q  s SE
2
 a L2
$ min   2 IH L  Q  s SE
2
P  F
S2 S S2

P  IH L  a L2 Q  s SE
2
 a L2 2a L2 IH L Q
$ min    2  F [3.40]
S2 S2 S 2
s SE P

In a sampling campaign, if the sample mass is chosen such that


Q IH L
 2 , the campaign will be optimized by taking samples with an
P s SE
optimum mass M Sopt :

101
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

IH L
M Sopt  2
[3.41]
s SE

3.8 INGAMELLS’ MINIMUM SAMPLE MASS

There is a Minimum Sample Mass M S min to include at least one particle


of the constituent of interest about 50% of the time in the collected sample
which happens when r =1 in equation [3.1]; Ingamells and Pitard 8 show
that it can be calculated as follows:

IH L  a L2
M S min  [3.42]
aL  aH 2
For replicate samples to provide a normally distributed population the
recommended sample mass M Srec should be at least 6 times larger than
M S min . Gy suggests a more stringent condition in order to make sure that
there is no Poisson skewness affecting the database, with a recommended
mass about 40 times the mass M S min . At this point there is an important
issue to address.

3.9 THE LINK WITH GY’S PREVENTIVE SUGGESTIONS

In a subtle way Ingamells’ work confirms that all Gy’s suggested


equations to calculate the variance of the Fundamental Sampling Error
should be used in the domain of normal distributions. As a result it would
be a much better strategy to calculate the mass of a sample for a given
allotted variance for the Fundamental Sampling Error; this would
eliminate the vast misuse of the suggested formulas in domains where
they do not belong as clearly addressed in Gy’s theory concerning the
properties of ratios of random variables like the one shown in equation
[2.5].

102
Part 1. Theory

3.10 NECESSARY VARIANCES TO CONSTRUCT A


MEANINGFUL SAMPLING DIAGRAM

In the following equation it is assumed the mass of the sample is always


second order of magnitude relative to the mass of the lot. More complex
formulas could be obtained taking into account the mass of the lot.

3.10.1 The variance of the Fundamental Sampling Error

It is convenient to express the standard deviation of the variance of FSE


with the proper units, which is directly derived from equation [3.6].

IH L  a L2
s FSE a L  [3.43]
MS

2
3.10.2 The variance s opt taking into account the Optimum Sample
Mass

From equations [3.4], [3.6], and [3.7] we may write:

IH L  a L2 s SE
2
 a L2
s 2
opt   [3.44]
MS N
But, from equation [3.41] we know that:

IH L
2
s SE 
M Sopt

It follows, with M S  N  M Sopt

IH L  a L2 IH L  a L2
s 2
  [3.45]
N  M Sopt N  M Sopt
opt

2 IH L  a L2
sopt  [3.46]
MS

103
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

3.10.3 The variance of a single assay s N2 1

From equation [3.4] making N =1 it readily follows:

IH L  a L2
s N 1   s SE
2
 a L2 [3.47]
Ms

This standard deviation gives an idea about the segregation uncertainty


affecting a single sample as the sample mass increases. A high value of
s N 1 is an indication that many samples need to be collected in order to
assess the variability across the lot to be sampled; it is a valuable indicator
for selecting the right sampling strategy (i.e., many small samples versus
few large samples).

104
Part 1. Theory

CHAPTER 4

CASE STUDIES

4.1 EXAMPLE OF A SAMPLING DIAGRAM IN A NICKEL-


COBALT LATERITIC DEPOSIT

A lateritic nickel-cobalt deposit is drilled and assayed for its cobalt


content which is an important by-product for the project. The drilling
technique uses HQ-diameter tubing with a tungsten carbide drill-head, and
the drilling is performed dry. Each sample is 1-meter long and weighs
about 7500 grams. Some holes shows very little cobalt but some others
show attractive grade intercepts. Geologists were tempted to believe there
were some areas much richer in cobalt than others. Table 4.1 shows 12 of
these consecutive holes (there were much more). Each hole shows 12
consecutive samples (there were much more).

A mining test performed in the same area shown on table 1 followed by


pilot plant pressure sulfuric leach reveals that the cobalt content was
almost the same everywhere and slightly higher than expected.
Retrospectively, looking at these existing exploration data, we may find
out what went wrong.

4.1.1 Compositing horizontally

Q = 12

We obtain 12 horizontal composites.

4.1.2 Calculation of the Low Background Content for Cobalt

The harmonic means are used with formula [3.26]. Figure 4.1 shows the
histogram of increasing cobalt contents, and the calculated value for the
Low Background Content a H and a value of 0.03% is used in the
calculation of the sampling diagrams.

105
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

Table 4.1. 12 holes in a lateritic nickel-cobalt deposit. The horizontal


axis shows consecutive holes, while the vertical axis shows consecutive
1-meter samples. Assays are expressed in % cobalt.

0.03 0.10 1.07 0.64 0.34 0.14 0.09 0.16 0.21 0.20 0.28 0.22

0.07 0.20 0.16 0.24 0.20 0.24 0.25 0.36 0.73 2.42 0.81 0.53

0.02 0.02 0.03 0.41 0.31 0.46 0.29 0.33 0.28 0.41 0.35 0.11

0.09 0.04 0.04 0.03 0.09 0.08 0.09 0.12 0.50 0.28 0.09 0.47

0.02 0.03 0.05 0.28 0.23 0.33 1.01 0.17 0.10 0.07 0.03 0.08

0.11 0.22 0.21 0.24 0.21 0.20 0.20 0.20 0.21 0.18 0.14 0.13

0.05 0.04 0.04 0.03 0.03 0.04 0.04 0.03 0.05 0.10 0.16 0.12

0.02 0.02 0.01 0.03 0.01 0.02 0.06 0.05 0.08 0.17 0.35 0.28

0.02 0.02 0.03 0.03 0.05 0.03 0.02 0.03 0.03 0.08 0.09 0.05

0.02 0.02 0.03 0.02 0.08 0.14 0.12 0.30 1.34 1.04 0.50 0.27

0.02 0.02 0.02 0.02 0.02 0.02 0.04 0.07 0.12 0.16 0.30 0.43

0.20 0.26 0.17 0.12 0.12 0.10 0.22 0.23 0.27 0.29 0.22 0.18

4.1.3 Calculation of the Most Probable Result

The most probable result of an assay  is calculated using formula


[3.25] and shown in figure 4.2. We may observe that a single 7500-g
sample gives a most probable result halfway between the estimated
overall average content and the estimated Low Background cobalt
content. It takes the averaging of 73 samples to eliminate the Poisson
skewness.

106
Part 1. Theory

Comment about the Ingamells’ nomenclature: we could use the TOS


nomenclature for unknown values (i.e.,  , a L , a H ) or the
corresponding Ingamells’ estimated values (i.e., Y , X , L).

Figure 4.1. Calculation of the Low Background Content.


L is Ingamells’ nomenclature used in the software.
aH  L  0.03%

107
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

Figure 4.2. Calculation of the Most Probable Result  ,


illustrated as the blue line. The low background content is the red line.
The overall arithmetic average is the orange line.

4.1.4 Calculation of standard deviation of the Fundamental Sampling


Error

The calculation of the standard deviation of the Fundamental Sampling


Error is performed using equation [3.43], and it is shown in figure 4.3.
Using equation [3.42] we may observe that the calculated Minimum
Sample Mass M S min and the Recommended Sample Mass M Srec are
24 kg and 144 kg respectively, which is much larger than the basic
drilling support set at 7.5 kg. Furthermore, following the guideline of the

108
Part 1. Theory

TOS it would be unwise to collect a sample with less than a 16% relative
for the standard deviation of the FSE which leads to the collection of a
683-kg sample (91 x 7.5 kg). Of course, such sample can only be
obtained by averaging neighboring samples until the required mass is
obtained. Nevertheless, if no such averaging is performed, the Poisson
process will most certainly create very unfortunate illusions, and this is
exactly what happened during the exploration and preparation of the
geological block model.

Figure 4.3. Calculation of the standard deviation s FSE a L


,
illustrated as the dotted green line.
Ingamells’ nomenclature is used in the software.
s FSE a L  S

109
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

4.1.5 Calculation of the standard deviation sopt taking into account


the Optimum Sample Mass

The calculation of the standard deviation sopt is performed using


equation [3.46], and it is shown in figure 4.4. As suggested in the
graphic, the significance of sopt is meaningless below the Optimum
Sample Mass, which is about 59 kg when using equation [3.41]. In other
words, it is practically impossible to draw a logical geological block
model using the information from 7.5-kg samples, unless a massive field
moving average is used, which of course, in turn results in the loss of
small-scale definition.

Figure 4.4. Calculation of the standard deviation s0 pt


illustrated as the purple lines. Ingamells’ nomenclature is used in the
software.
sopt  S _ v

110
Part 1. Theory

4.1.6 Calculation of the standard deviation of a single assay s N 1

The calculation of the standard deviation s N 1 is performed using


equation [3.47], and it is shown in figure 4.5. The spread of s N 1 as
the sample mass becomes very large gives an indication of the amount of
large-scale segregation that was present in the field in the area under
study. This domain is where Geostatistics should perform very well.
Now the Ingamells’ sampling diagram is complete, and as presented it is
reasonably consistent with the TOS.

Figure 4.5. Calculation of the standard deviation s N 1


illustrated as the brown lines.
Ingamells’ nomenclature is used in the software: s N 1  S p  v

111
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

4.1.7 Compositing vertically

Q = 12

We obtain 12 vertical composites.

Steps 2 through 6 are repeated and a new, complete sampling diagram is


obtained and illustrated in figure 4.6. Observations are substantially
different from what is observed in the horizontal diagram illustrated in
figure 4.5. Results are summarized in table 4.2.

Figure 4.6. Complete sampling diagram by compositing vertically

The main conclusions and recommendations are:

 The 1-meter HQ-diameter support with a mass of 7.5 Kg is much


too small. It is recommended to drill with larger diameter.
 It is absolutely necessary to composite 7.5-Kg individual
samples until the recommended sample mass is reached.
112
Part 1. Theory

 There is less large-scale segregation as a function of depth than


as a function of lateral distance. Because holes are 25 meters
apart, this conclusion was predictable.
 A massive Poisson process is affecting the database therefore
many areas are unfairly looking unattractive that may result in an
underestimation of the ore reserves. Collect samples with
optimum sample mass.
 A few blocks may appear to have extremely high cobalt contents
when it is not true. Collect samples with optimum sample mass.

Table 4.2. Summary of observations from vertical and horizontal


sampling diagrams
Parameters Vertical Horizontal
compositing compositing

aH 0.03% 0.03%

M S min 26.2 kg 24 kg

M Srec 157 kg 144 kg

M Sopt 141 kg 59 kg

Number of 7.5-Kg
to composite for 100 91
uncertainty = 16%

s N 1
From 0.13 to 0.28% Co From 0.09 to
for very large samples 0.31% Co

113
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

4.2 THE DOUBLE POISSON DISTRIBUTION: CASE OF A


MOLYBDENUM DEPOSIT

A molybdenum ore contains one molybdenite grain per 6.394 kg of ore on


average. Each molybdenite grain weighs 19.1 g. The average molybdenite
content is 0.3%. In addition, the ore contains 0.1% of finely divided
molybdenite evenly distributed throughout the mass. 10-kg samples are
taken. The average sample contains 1.564 grains of molybdenite.
The distribution of the true grades of 1000 samples is Poisson, assuming
the molybdenite is randomly distributed in a given geological unit.

The probabilities are shown in table 4.3.

Table 4.3. Probability of a Mo grade with  = 1.564 for 1000 samples

True grade (wt.%) Number of grains of Poisson Probability


molybdenite
0.100 0 209
0.291 1 327
0.482 2 256
0.673 3 133
0.864 4 52
1.055 5 16
1.246 6 4
1.437 7 1
1.628 8 0

Each 10-kg sample is reduced by crushing in such a way that each


molybdenite grain is broken into 10 tiny pieces. The crushed material is
mixed and 1/20 is split out. The subsample weight is 500g. The average
number of grains of molybdenite per subsample is 0.782. Of 1000
samples, the 209 that contained no coarse grain of molybdenite yield 209
subsamples containing no grain of molybdenite. The 327 samples that
contained one coarse grain of molybdenite yield subsamples containing 1,
2, 3, … molybdenite grains, as do samples containing 2, 3, 4, … grains.
The overall probabilities are shown in table 4.4.

114
Part 1. Theory

Table 4.4. Combined Probabilities of a Mo grade with  = 1.564 for


primary samples and Mo grade with  = 0.782 for secondary subsamples

True grade Number of Poisson Total


(wt.%) grains of Probability Probabilities
molybdenite
0.100 0 209 539
0.291 1 327 256
0.482 2 256 124
0.673 3 133 52
0.864 4 52 18
1.055 5 16 6
1.246 6 4 2
1.437 7 1 1
1.628 8 0 0

Results in table 4.4 shows that there are about 800 chances of finding a
result below the true 0.300% grade while there are only 200 chances of
finding a result above. This example shows how a Poisson process or a
double Poisson process can lead to a massive under-estimation of the ore
reserves, and this is exactly what happened in this particular molybdenum
deposit.

Lessons to be learned for fitting statistical models

At one time, scientists became convinced that the Gaussian distribution


was universally applicable, and an overwhelming majority of applications
of statistical theory are based on this distribution.

A common error has been to reject “outliers” that cannot be made to fit
the Gaussian model or some modification of it as the popular lognormal
model.

The tendency, from some geostatisticians, has been to make the data fit a
preconceived model instead of searching for a model that fit the data.

It is now apparent that outliers are often the most important data points in
a given data set. The recommendation would be to make an effort to make
a clear difference between outliers and unexplained extreme values, and in
such endeavor Poisson statistics can greatly help.
115
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

116
Part 1. Theory

CHAPTER 5

FROM LINKS TO A SAMPLING STRATEGY

5.1 SUMMARY

The link between Gy and Ingamells is straightforward. In a first


preventive approach, Gy offers many tools to optimize sampling
protocols and to implement them in a correct and logical way. In a
second approach, Ingamells offers ways to quantify and identify the
sources of variability in an existing database. Furthermore, when no
preventive approach as suggested by Gy has been implemented, the
study of existing data as suggested by Ingamells would most certainly
demonstrate how devastating such an adventurous shortcoming can be.
In other words, by using the work of Gy and Ingamells simultaneously,
any auditor can proceed with confidence as the involved company is due
diligent in its practices.

5.2 DISCUSSIONS, CONCLUSIONS AND RECOMMENDATIONS


FOR FUTURE WORK

TOS stands on a very solid theoretical foundation and its applications are
universal. However, as clearly shown in this thesis, sampling practitioners
using the TOS must be careful that they do not enter a domain of Poisson
processes that very quickly becomes dangerous and too often
misunderstood. In such a domain, TOS can only tell the practitioner that
the database that will be generated is nearly useless. Therefore, this thesis
suggests the following due diligence strategy:

5.2.1 The wisdom of prevention for due diligence

Pierre Gy’s Theory of Sampling is totally preventive in nature. TOS


instigates a categorical imperative: sampling processes, sampling
procedures and sampling equipment must be unbiased. There is absolutely
no room for any compromise in this matter. One could call this Pierre
Gy’s paradigm.

Step #1
Always make sure the implementation of the sampling protocol that is
going to be used is flawless with respect to the fundamental principles of
sampling correctness, as suggested by equations [2.27], [2.28] and [2.29].

117
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

This is best achieved by eliminating biases generated by IDE, IEE, IWE,


and IPE. If this condition is not fulfilled, sampling becomes an exercise in
futility. Much information on this subject is given in Chapters 6, 7, and 9.
Nobody can achieve due diligence in sampling without an unconditional
commitment to respect the rules of sampling correctness; nobody can
ignore the economic implications of this statement.

Step #2
Perform a thorough microscopic investigation of the material to be
sampled in order to quantify the appropriate sample mass with respect to
well-defined Data Quality Objectives. Many Gy’s approaches to minimize
the variance of FSE are detailed in this thesis and provide a basis to start a
project correctly.

Step #3
Minimize the variance of GSE by taking the three preventive actions in
order to minimize the three factors involved in this variance as suggested
by equation [2.90].

Step #4
Optimize field or plant sampling frequency using either Geostatistics or
Chronostatistics as suggested in chapter 10. The TOS covers that subject
to a great extent.

Step #5
Using the generated existing data according to steps 1-4, perform a
verification that due diligence was applied. This can be done using the
information given in chapters 8 and 10 and the reliable literature available
on Geostatistics.

Step #6
Using the generated existing data according to steps 1-4, calculate the low
background content a H easy to sample of the constituent of interest. This
information may have important geometallurcical and economic
implications.

118
Part 1. Theory

5.2.2 Difficult cases where good prevention is, in appearance, not


realistic

Step #1
Always respect the fundamental rules of sampling correctness as
explained earlier. This step is not negotiable.

Step #2
Perform a thorough microscopic investigation of the material to be
sampled in order to quantify the appropriate sample mass with respect to
well-defined Data Quality Objectives. Proceed with a Visman’s
experiment, calculate the Ingamells’ parameters, and draw an informative
Ingamells’ sampling diagram.

Step #3
Executive managers must review the information and make a feasibility
study to allocate much more money on a wiser and necessary approach
using Gy’s requirement to minimize the variance of FSE. Someone may
say: why not start with this in the first place? The answer is an economic
one and the Visman’s experiment and Ingamells’ sampling diagram
provide all the necessary information to perform the necessary feasibility
study.

5.2.3 The after-the-fact non compliance with due diligence

Step #1
Following an audit, as suggested in Chapter 12, make a full assessment of
due diligence with respect to sampling correctness. If non compliance is
found, everything stops right there as the generated data is not statistically
valid and totally useless. If compliance with respect to sampling
correctness is found, proceed with the following step.

Step #2
Using existing data proceed with the Visman’s experiment and calculate
the Ingamells’ sampling diagram. Find out what is wrong with the
existing data and outline the consequences.

Step #3
Executive managers must review the information and make a feasibility
study to allocate much more money on a wiser and necessary approach
using Gy’s requirement to minimize the variance of FSE and comply with
sampling correctness if this has been found as the major issue.
119
Pierre Gy’s Theory of Sampling and C.O. Ingamells’ Poisson Process Approach

5.3 VISMAN AND INGAMELLS’ WORKS HELP JUSTIFY AN


AUGMENTED GY’S APPROACH.

The suggestion of the above strategy is simple: use Visman’s experiments


and Ingamells’ sampling diagrams as a selling tool to justify a wiser
approach as suggested by the existing TOS. In 30 years of experience as a
consultant there have been many occasions when an important company
executive would say, “give me the facts, then I will find the necessary
funds to fix the problems.” Using existing data, it is very powerful to use
the approach suggested in this thesis in order to demonstrate the absolute
necessity to always be preventive using TOS. For more information on the
economic consequences of no diligence with correct sampling it is
beneficial to read Carrasco case studies 26.

120

You might also like