image (3.36)

In the above equation, i stands for an assayed metal and j for a process unit.

The calculation of these best estimates requires the calculation of the derivatives of S for each of the unknowns, that is, the non-measured solid mass flowrates. The best estimates are those unknown variable values for which the derivatives are all together equal to zero:

δSδWK=0 (3.37)

image (3.37)

It is worth noting that with the obtained best estimates the mass conservation equations are not rigorously verified; essentially, the node imbalances are only minimized. By experience, the estimates obtained by this method are as good as the measured values used for the calculation are themselves (Example 3.10).

Example 3.10

For the example data set in Table ex 3.9, apply the Node Imbalance Minimization method to estimate the solid split value and compare it with the 2-product formula estimates.

Solution

In this example there is one process unit and three metals assayed; hence, there are four mass conservation equations and hence four imbalances. The criterion S can be written as:

S=WF2+WC2+WT22WFWC2WFWT+2WCWT+WF2i=13xFi2+WC2i=13xCi2+WT2i=13xTi22WFWCi=13xFixCi2WFWTi=13xFixTi+2WCWTi=13xCixTi (3.38)

image (3.38)

where i stands for the assayed metal.

The derivatives of S for WC and WT are then:

δSδWC=2(WCWF+WT+WCi=13xCi2WFi=13xFixCi+WTi=13xCixTi)=0 (3.39)

image (3.39)

δSδWT=2(WTWF+WC+WTi=13xTi2WFi=13xFixTi+WCi=13xCixTi)=0 (3.40)

image (3.40)

Now, the solid split value can be obtained from the above 2 equations:

WTWF=(1+i=13xFixTi)(1+i=13xCi2)(1+i=13xFixCi)(1+i=13xCixTi)(1+i=13xTi2)(1+i=13xCi2)(1+i=13xCixTi)2 (3.41)

image (3.41)

and WCWF=1WTWFimage

Using the data set of Table ex 3.9, Table ex 3.10 is derived.

Table ex 3.10

Node Imbalance Example

 WTWF(%)image WCWF(%)image
Using Zn, Cu and Fe 80.8 19.1
Using Zn and Cu 93.1 6.87
Using Zn only 93.3 6.67

Image

The results of Table ex 3.10 show that the method is sensitive to the accuracy of the metal assays. Indeed, the Fe assay is bad as demonstrated by the application of the 2-product formula (see Table 3.2).

Limitations

The node imbalance minimization method solves two issues with the n-product formula:

• It handles data in excess

• It works for more than one process unit at a time

However, the method exhibits 2 limitations:

• It is sensitive to measurement errors as can be seen in Example 3.10

• It only provides estimates of the non-measured solid flowrates; measured flowrate values and assay values are not adjusted

An extension of the method aims to decrease the influence of bad measurements. The criterion node imbalance terms are weighted according to the presence or absence of gross measurement errors.

3.6.3 Two-step Least Squares Minimization

The 2-step least squares minimization method proposes a way to adjust and correct metal assays for the measurement errors that affect them (Wiegel, 1972; Mular, 1979). It is a compromise between the limitations of the node imbalance minimization method (Section 3.6.2) and the complexity of the generalized least squares minimization method (Section 3.6.4).

In the node imbalance method, the deviations to the mass conservation of solids and of metals are considered all together. The obtained estimated flowrate values minimize the deviations to all the mass conservation equations.

In the 2-step least squares method, flowrate values that rigorously verify the mass conservation equations of solids and, at the time, minimize the deviations to the mass conservation equations of metals, are first estimated. In a second step, corrected values of metal assays that rigorously verify the mass conservation equations of metals are estimated.

To achieve, taking Eqs. (3.23) and (3.24) as an example, the mass conservation equations are re-written as:

WTWF=1WCWF (3.42)

image (3.42)

xF=WCWFxC+(1WCWF)xT (3.43)

image (3.43)

and the node imbalance equations for metals become:

xFWCWFxC(1WCWF)xT=Ix (3.44)

image (3.44)

or:

(xFxT)WCWF(xCxT)=Ix (3.45)

image (3.45)

The general problem can now be stated as: find the best estimate of WCWFimage which minimizes the sum of squared imbalances Ix for all nodes and assayed metals:

S=ji(ixji)2 (3.46)

image (3.46)

By comparison with the node imbalance minimization method, the criterion only contains the node imbalances for metals and the search variables are a set of independent relative solid flowrates. Similar to the node imbalance minimization method, the calculation of the best estimates of the independent relative solid flowrates requires the calculation of the derivatives of S for each of the unknowns, that is, the independent relative solid flowrates. The best estimates are those values for which the derivatives are all together equal to zero. Due to the way the node imbalances for metals are written (Eq. (3.45)), the mass conservation of relative solid flowrates is rigorously verified for the estimated relative solid flowrate values.

Having determined WCWFimage, assuming the feed solid flowrate WFimage is measured, it is possible to estimate WC and WT. All these flowrate values rigorously verify their mass conservation equations.

In a second step, it is possible to adjust the metal assay measured values (ximage) to values (xˆ)image that verify their own mass conservation equations as well. Let us call rx the required adjustment for each metal assay value, then it follows that:

rx=xxˆ (3.47)

image (3.47)

and since:

(xˆFxˆT)WCWF(xˆCxˆT)=0 (3.48)

image (3.48)

it follows by difference with Eq. (3.45) that:

Ix(rxFrxT)WCWF(rxCrxT)=0 (3.49)

image (3.49)

The problem is now to find the values of rx which minimize the following least squares criterion:

S=ij(rxji)2 (3.50)

image (3.50)

under the equality constraint K of Eq. (3.49). In Eq. (3.50), i stands for the number of streams (or samples) and j for the number of assayed metals. The problem is best solved using the Lagrange technique where the criterion becomes:

L=ij(rxji)2+kλkKk (3.51)

image (3.51)

where k is the number of nodes (or mass conservation equations), and λk is the Lagrange coefficient of equality constraint Kk.

The problem is solved by calculating the derivatives of L for each of the unknown, that is, rxjiimage and λi. Criterion L is minimal when all its derivatives are equal to zero, which leads to a set of equations to be solved. Solving the set of equations provides the best estimates of the rxjiimage values, that is, the adjustments to the measured assay values which will make the adjusted assay values coherent from a mass conservation point of view (Example 3.11).

Example 3.11

For the example data set of Table ex 3.9 apply the 2-step least squares minimization method to estimate the solid split value and compare it with the 2-product formula and node imbalance estimates. Calculate the adjusted metal assay values and compare with the measured values.

Solution

In the present case, Eq. (3.46) is:

S=ji(Ixji)2=j=13[(xFjxTj)WCWF(xCjxTj)]2

image

where j stands for Zn, Cu and Fe. Deriving S for the unknown wC=WCWFimage gives:

δSδwC=2j=13[(xFjxTj)wC(xCjxTj)](xCjxTj)=0

image

The solution of which is:

wC=WCWF=j=13(xFjxTj)(xCjxTj)j=13(xCjxTj)2

image

Using the data set of Table ex 3.9, Table ex 3.11a is derived.

Table ex 3.11a

Two-step Least Squares Minimization Example

 WTWF(%)image WCWF(%)image
Using Zn, Cu and Fe 93.4 6.57
Using Zn and Cu 93.3 6.67
Using Zn only 93.3 6.67

Image

The results of Table ex 3.11a show that the method is much less sensitive to the accuracy of the metal assays than the results of the node imbalance method shown in Table ex 3.10. Having determined the solid split, the next step is to adjust the metal assay values to make them coherent with the calculated solid split and coherent from a mass conservation point of view.

In the present case, Eq. (3.51) is:

L=i=13(rxFi)2+(rxCi)2+(rxTi)2+i=13λi[Ixi(rxFirxTi)wC(rxCirxTi)]

image

and the derivatives for each variable are:

δLdrxFi=2rxFiλi=0

image

δLdrxFi=2rxFi+λiwC=0

image

δLdrxTi=2rxFi+λi(1wC)=0

image

δLdλi=Ixi(rxFirxTi)wC(rxCirxTi)=0

image

Solving the system of equations leads to:

rxFi=Ixi1+wC2+(1wC)2

image

rxCi=IxiwC1+wC2+(1wC)2

image

rxTi=Ixi(1wC)1+wC2+(1wC)2

image

and the numerical solution is:

 Ixiimage Adjustments rxjiimage Adjusted values xˆjiimage
  Feed Concentrate Tailings Feed Concentrate Tailings
Zn 0.049885 0.049885 0.049885 0.049885 3.903426 52.0717466 0.514828
Cu −0.01418 −0.01418 −0.01418 −0.01418 0.167552 0.65950362 0.132944
Fe −1.62385 −1.62385 −1.62385 −1.62385 12.43504 14.6131448 12.28181

Image

Table ex 3.11b shows a comparison of the measured assays (Table ex 3.9) with the adjusted ones where adjustment=(measured – adjusted)/measured as a percent. Note the concentrate stream assays are barely adjusted by comparison with the other 2 stream assays.

Table ex 3.11b

Adjustment Distribution (%)

 Feed Concentrate Tailings
Zn 0.68 0.00 −5.07
Cu −4.72 0.08 5.04
Fe −7.48 0.39 6.17

Image

Limitations

The 2-step least squares minimization method has the advantage over the previous methods of providing adjusted assay values that verify the mass conservation equations. The method remains simple and can easily be programmed (e.g., using Excel Solver) for simple flow diagrams.

Although presenting significant improvements over the previous methods, the 2-step method is not mathematically optimal since the mass flowrates are estimated from measured and therefore erroneous metal assay values. Measurement errors are directly propagated to the estimated flowrate values. In other words, the reliability of the estimated flowrate values depends strongly on the reliability of the measured assay values. The equations could, however, be modified to include weighting factors with the objective of decreasing the influence of poor assays.

3.6.4 Generalized Least Squares Minimization

The generalized least squares minimization method proposes a way to estimate flowrate values and adjust metal assay values in a single step, all together at the same time (Smith and Ichiyen, 1973; Hodouin and Everell, 1980). There is also no limitation on the number of mass conservation equations.

The mass conservation equations are written for the theoretical values of the process variables. The process variable measured values carry measurement errors and hence do not verify the mass conservation equations. It is assumed that the measured values are unbiased, uncorrelated to the other measured values and belong to Normal distributions, N(µ,σ). The balances can be expressed as follows:

WF*=WC*+WT* (3.52)

image (3.52)

WF*xF*=WC*xC*+WT*xT* (3.53)

image (3.53)

where * denotes the theoretical value of the variable. Since the theoretical values are not known, the objective is to find the best estimates of the theoretical values. With the statistical assumptions made, the maximum likelihood estimates are those which minimize the following generalized least square criterion:

S=i(WjWˆj)2σWj2+ji(xijxˆij)2σxij2 (3.54)

image (3.54)

where the ^ denotes the best estimate value and σ2 is the variance of the measured value.

It is evident that the best estimates must obey the mass conservation equations while minimizing the generalized least squares criterion S. Therefore, the problem consists in minimizing a least squares criterion under a set of equality constraints, the mass conservation equations. Such a problem is solved by minimizing a Lagrangian of the form:

L=i(WjWˆj)2σWj2+ji(xijxˆij)2σxij2+kλkKk (3.55)

image (3.55)

where λk are the Lagrange coefficients, and Kk the mass conservation equations.

As seen previously, the problem is solved by calculating the derivatives of L for each of the unknowns, that is, the Wˆimage, xˆimage and the λ. Criterion L is minimal when all its derivatives are equal to zero, which leads to a set of equations to be solved. Solving the set of equations provides the best estimate values Wˆimage and xˆimage, which verify the mass conservation equations. It is important to note that even non-measured variables can be estimated, since they appear in the mass conservation equations of the Lagrangian, as long as there is the necessary redundancy.

In Eq. (3.55), the variance of the measured values, σ2, are weighting factors. An accurate measured value being associated with low variance cannot be markedly adjusted without significantly impacting the whole criterion value. In contrast, a bad measured value associated to a high standard deviation can be significantly adjusted without impacting the whole criterion value. Hence, adjustments, small or large, will be made in accordance to the confidence we have in the measured value whenever possible considering the mass conservation constraints that must always be verified by the adjusted values (Example 3.12).

Example 3.12

For the example data set of Table ex 3.9 apply the generalized least squares minimization method to estimate the solid split value and compare it with the 2-step least squares minimization estimate. Calculate the adjusted metal assay values and compare with the measured values.

Solution

In the present case, the Lagrangian is:

L=j=13i=13(xijxˆij)2σxij2+i=13λiKi

image

There is no measured mass flowrate, there are 3 stream samples with 3 metal assays each and there are 3 mass conservation equations:

xˆFi=WˆCWˆFxˆCi+(1WˆCWˆF)xˆTi

image

where i stands for Zn, Cu, or Fe, and the problem variables are: λi, xˆijimage, and wˆC=WˆCWˆFimage.

The Lagrangian derivatives are:

δLdwˆC=i=13(xˆCixˆTi)=0

image

δLdλi=xˆFiwˆCxˆCi(1wˆC)xˆTi=0

image

δLdxˆFi=2(xFixˆFi)σxFi2+λi

image

δLdxˆCi=2(xCixˆCi)σxCi2λiwC=0

image

δLdxˆTi=2(xTixˆTi)σxTi2λi(1wC)=0

image

where i stands for Zn, Cu or Fe, for a total of 13 equations and 13 unknowns. It can be shown that once the value of wˆCimage is known then there is an analytical solution for xˆijimage:

xˆFi=xFixFixCiwˆCxTi(1wˆC)1σxFi2+wˆC2σxCi2+(1wˆC)2σxTi2×1σxFi2

image

xˆCi=xCi+xFixCiwˆCxTi(1wˆC)1σxFi2+wˆC2σxCi2+(1wˆC)2σxTi2×wˆCσxCi2

image

xˆTi=xTi+xFixCiwˆCxTi(1wˆC)1σxFi2+wˆC2σxCi2+(1wˆC)2σxTi2×1wˆCσxTi2

image

where i stands for Zn, Cu or Fe.

Therefore, it is only required to find the best value of wˆCimage and the algorithm consists in:

1. Assuming an initial value for wˆCimage

2. Calculating the xˆijimage values

3. Calculating the criterion S of Eq. (3.54)

4. Iterating on the value of wˆCimage to minimize the value of S

The solution is given in Table ex 3.12a for variance values all equal to 1 for simplicity. The initial solid split wˆCimage has been estimated at 6.67% (based on result in Example 3.11).

Table ex 3.12a

Generalized Least Squares Minimization Example

 Feed Concentrate Tailings
Zn 3.89 52.1 0.527
Cu 0.167 0.660 0.133
Fe 12.4 14.6 12.3

Image

Table ex 3.12b shows a comparison of the measured assays with the adjusted ones. The obtained results are similar to the ones obtained with the previous method.

Table ex 3.12b

Adjustment Distribution (%)

 Feed Concentrate Tailings
Zn (%) 1.00 0.00 −7.47
Cu (%) −4.64 0.07 4.96
Fe (%) −7.47 0.38 6.17

Image

While Example 3.12 presents the calculations in detail for a simple case, there exists a more elegant and generic mathematical solution that can be programmed. The solution uses matrices and requires knowledge in matrix algebra.

3.6.5 Mass Balance Models

The formulation of the generalized least squares minimization method enables resolving complex mass balance problems. The Lagrange criterion consists of 2 types of terms: the weighted adjustments and the mass conservation constraints. The criterion can easily be extended to include various types of measurements and mass conservation equations.

The whole set of mass conservation equations that apply to a given data set and mass balance problem is called the mass balance model. Depending on the performed measurements and sample analyses, the following types of mass conservation equations may typically apply in mineral processing plants:

• Conservation of slurry mass flowrates

• Conservation of solid phase mass flowrates

• Conservation of liquid phase mass flowrates (for a water balance, for instance)

• Conservation of solid to liquid ratios

• Conservation of components in the solid phase (e.g., particle size, liberation class)

• Conservation of components in the liquid phase (in leaching plants, for instance)

There are also additional mass conservation constraints such as:

• The completeness constraint of size or density distributions (the sum of all mass fractions must be equal to 1 for each size distribution)

• Metal assays by size fraction of size distributions

• The coherency constraints between the reconstituted metal assays from assay-by-sizes and the sample metal assays

All these constraint types can be handled and processed by the Lagrangian of the generalized least squares minimization algorithm. As mentioned, all the necessary mass conservation equations for a given data set constitute the mass balance model.

It is convenient to define the mass balance model using the concept of networks. Indeed, the structure and the number of required mass conservation equations depend on the process flow diagram and the measurement types. Therefore, it is convenient to define a network type by process variable type, knowing that the structure of mass conservation equations for solid flowrates (Eq. (3.23)) is different from that of metal assays (Eq. (3.24)), for example.

Not only does a network type mean a mass conservation equation type, but it also expresses the flow of the mass of interest (solids, liquids, metals…).

For the 2-product process unit of Figure 3.33, assuming 3 metals (Zn, Cu, and Fe) have been assayed on each stream, then Figure 3.35 shows the 2 networks that can be developed, one for each equation type (Eqs. (3.23) and (3.24)).

image
Figure 3.35 Mass conservation networks for the flow diagram of Figure 3.33: (a) For solids, and (b) For metal (in this case they are the same).

For the more complex flow diagram of Figure 3.34, assuming again 3 metals (Zn, Cu, and Fe) have been assayed on each of the 6 streams, then Figure 3.36 shows the 2 networks that can be developed.

image
Figure 3.36 Mass conservation networks for the flow diagram of Figure 3.34: (a) For solids, and (b) For metal (again, they are the same in this case).

Assuming a fourth metal, Au for example, has been assayed on the main Feed, main Tailings and main Concentrate streams only, then a third network should be developed for Au as shown in Figure 3.37.

image
Figure 3.37 Mass conservation networks for Au in the flow diagram of Figure 3.34.

Networks are conveniently represented using a matrix (Cutting, 1976). In a network matrix, an incoming stream is represented with a+1, an outgoing stream with a–1 and any other stream with a 0. For the networks of Figure 3.35 and Figure 3.36, the network matrices are, respectively:

M=[111] (3.56)

image (3.56)

and

M=[111000010110001011]fornodeAfornodeBfornodeC (3.57)

image (3.57)

Each row represents a node; each column stands for a stream. Equations (3.56) and (3.57) expressed in a matrix form become:

MWW=0 (3.58)

image (3.58)

MxW¯xXi=0 (3.59)

image (3.59)

where MW and Mx are the network matrices for the solid network and the metal assay network, W is the column matrix of solid mass flowrates, W¯ximage the diagonal matrix of the solid mass flowrates used in the metal assay network, and Xi the column matrix of metal assay i on each stream of the metal assay network (i stands for Zn, Cu and Fe) (Example 3.13).

Example 3.13

A survey campaign is being designed around the grinding circuit in Figure ex 3.13. For the location of flow meters and samplers, and for the analysis types performed on the samples, develop the mass conservation networks for each data type and mass conservation equation type that apply.

image
Figure ex 3.13 Grinding circuit showing sampling points.

Solution

With the slurry feed flowrate to the grinding circuit being measured, water addition stream flowrates being measured and % solids in slurry being measured, the slurry and water mass balances can be calculated and associated measured values adjusted. The mass conservation equations that apply obey to the Slurry and Water Mass Conservation Networks are given in Figure ex 3.13-S.

image
Figure ex 3.13-S The networks corresponding to Figure ex 3.13

The mass conservation equations for the dry solid obey the same network except water addition streams are removed.

Mass size fractions are not conserved through grinding devices (SAG mill and ball mills). Therefore there is no mass conservation equation for size fractions around the grinding devices and consequently there is no node for grinding devices in mass size fraction networks. The same reasoning applies to Cu-by-size fractions.

The (overall) Cu on stream samples is conserved through grinding. Since Cu has been assayed only on the SAG mill feed and the cyclone overflows feeding the flotation circuit, the mass conservation network for Cu on stream samples consists of only one node.

Assuming eleven sieves have been used for measuring the size distributions, then the total system of mass conservation equations consist in:

• Seven mass conservation equations for slurry flowrate variables

• Seven mass conservation equations for % solids variables

• Seven mass conservation equations for water flowrate variables

• Seven mass conservation equations for solid flowrate variables

• Forty eight (4 nodes×12 mass fractions) mass conservation equations for size fractions

• Forty eight mass conservation equations for Cu-by-size fractions

• One mass conservation equation for Cu

for a total of 135 equations and 316 variables.

3.6.6 Error Models

In the generalized least square minimization method, weighting factors prevent large adjustments of trusted measurements and, on the contrary, facilitate large adjustments of poorly measured process variables. Setting the weighting factors is the main challenge of the method (Almasy and Mah, 1984; Chen et al., 1997; Darouach et al., 1989; Keller et al., 1992; Narasimhan and Jordache, 2000; Blanchette et al., 2011).

Assuming a measure obeys to a Normal distribution, the distribution variance is a measure of the confidence we have in the measure itself. Hence, a measure we trust exhibits a small variance, and the variance increases when the confidence decreases. The variance should not be confused with the process variable variation, which also includes the process variation: the variance is the representation of the measurement error only.

Measurement errors come from various sources and are of different kinds. Theories and practical guidelines have been developed to understand the phenomenon and hence render possible the minimization of such errors as discussed in Section 3.2. A brief recap of measurement errors as they pertain to mass balancing will help drive the message home.

Errors fall under two main categories: systematic errors or biases, and random errors. Systematic errors are difficult to detect and identify. First, they can only be suspected over time by nature and definition. Second, a source of comparison must be available. That is achievable to some extent through data reconciliation (Berton and Hodouin, 2000). A bias can be suspected if a measure over time is systematically adjusted to a lower (or higher) value. Once a bias is detected and identified, the bias source and the bias itself should be eliminated. Remember, the generalized least square method is a statistical method for data reconciliation by mass balancing and bias is a deterministic error, not a statistical variable, and is therefore a non-desired disturbance in the statistical data processing.

Measurement errors can also be characterized by their amplitude, small and gross errors, and by the frequency distribution of the measure signal over time: low frequency, high frequency or cyclical. Although there is no definition for a gross error, gross errors are statistically barely probable. As such, if a gross error is suspected it should be corrected before the statistical adjustment is performed. Typically, gross errors have a human source or result from a malfunctioning instrument.

Measurement errors originate from two main sources: sampling errors, and analysis errors. Sampling errors mainly result from the heterogeneity of the material to sample and the difficulty in collecting a representative sample; analysis errors result from the difficulty to analyze sample compositions. The lack of efficiency of the instruments used for measuring process variables also contributes to measurement errors. Analysis errors can be measured in the laboratory and therefore the variance of the analysis error is estimable. Sampling errors are tricky to estimate and one can only typically classify collected samples from the easiest to the more difficult to collect. Assuming systematic and gross errors have been eliminated, then purely random measurement errors can be modeled using 2 functions: one that represents the sampling error and the other one that represents the analysis error. The easiest model is the so-called multiplicative error model where the analysis error variance (σAE2)image is multiplied by a factor (kSE) representing the sample representativeness or sampling error: the higher the factor value, the less representative the sample.

σME2=kSEσAE2 (3.60)

image (3.60)

The multiplicative error model can serve as a basis to develop more advanced error models.

3.6.7 Sensitivity Analysis

The remaining unknown, once a mass balance is obtained for a given set of measured process variables and measurement errors, is: how trustable are the mass balance results or is it possible to determine a confidence interval for each variable estimated by mass balance? That is the objective of the sensitivity analysis.

Assuming systematic and gross errors have been eliminated, and measurement errors are random and obey to Normal distributions, there are two main ways to determine the variance of the estimated variable values (stream flowrates and their composition): Monte Carlo simulation, and error propagation calculation. Each method has its own advantages and disadvantages; each method provides an estimate of the variance of the adjusted or estimated variables.

In the Monte Carlo simulation approach (Laguitton, 1985), data sets are generated by disturbing the mass balance results according to the assumed error model. Each new data set is statistically reconciled and the statistical properties of the reconciled data sets determined.

In the error propagation approach (Flament et al., 1986; Hodouin et al., 1989), the equations through which measurement errors have to be propagated are complex and therefore it is preferable to linearize the equations around a stationary point. Obviously, the obtained mass balance is a good and valid stationary point. The set of linearized equations is then solved to determine the variable variances. This method enables the calculation of covariance values, a valuable feature for calculating confidence intervals around key performance indicators, for instance.

From a set of measured process variables and their variances, therefore, it is possible to determine a set of reconciled values and their variances:

{(Wj,σWj2)(xij,σxij2)}{(Wˆj,σWˆj2)(xˆij,σxˆij2)} (3.61)

image (3.61)

It can be demonstrated that the variance of the adjusted values is less than the variance of the measured values. While the variance of the measured values contributes the most to the variance of the adjusted values, the reduction in the variance values results from the statistical information provided. The data redundancy and the topology of the mass conservation networks are the main contributors to the statistical content of the information provided for the mass balance calculation.

3.6.8 Estimability and Redundancy Analysis

We have seen with the n-product formula that n–1 components need be analyzed on the n+1 streams around the process unit to estimate the n product stream flowrates, assuming the feed stream flowrate is known (or taken as unity to give relative flowrates). The n-product formula is the solution to a set of n equations with n unknowns.

In such a case, there are just enough data to calculate the unknowns: no excess of data, that is, no data redundancy, and the unknowns are mathematically estimable. A definition of redundancy is therefore:

A measure is redundant if the variable value remains estimable should the measured value become unavailable.

It follows from the definition that the measured value is itself an estimate of the variable value. Other estimates can be obtained by calculation using other measured variables and mass conservation equations.

With the n-product formula, determining data redundancy and estimability is easy. In complex mass balances, determining data redundancy and estimability is quite tricky (Frew, 1983; Lachance and Flament, 2011). Estimability and redundancy cannot be determined just by the number of equations for the number of unknowns. It is not unusual to observe global redundancy with local lack of estimability (Example 3.14).

Example 3.14

In the following mass balance problem (Figure ex 3.14), determine the estimability and redundancy of each variable entering the mass balance considering the Feed stream mass flowrate is measured and:

1. One metal has been analyzed on each sample;

2. Two metals have been analyzed on each sample.

image
Figure ex 3.14 The circuit.

Solution

When one metal is analyzed on each of the 6 streams:

There are 4 equations (2 for the mass flowrates and 2 for the metal assay) and 5 unknowns. The system is globally underdetermined since there are not enough equations for the number of unknowns. Around node A, the 2-product formula can be applied and therefore the mass flowrates of unit A products can be estimated. However, around node B, a 3-product formula cannot be applied since only one metal assay has been performed. Unit B product stream flowrates are not estimable. Metal assays are not redundant and cannot be adjusted by statistical data reconciliation.

When two metals are analyzed on each of the 6 streams:

There are a total of 6 equations and 5 unknowns. The system is therefore globally redundant and estimable a priori. Around node A, the 2-product formula can be applied with each metal assay and therefore the two product stream mass flowrates are estimable and the metal analyses on the 3 streams around unit A are redundant and adjustable. Around node B, the 3-product formula can be applied and therefore the 3 product stream mass flowrates are estimable now that the mass flowrate of the feed stream to unit B is known. However, the metal analyses are not redundant and not adjustable.

What are the factors influencing estimability and redundancy? Obviously, from Example 3.14, the number of measured variables is a strong factor, but it is not the only one. The network topology is also a factor. In the case of Example 3.14, if one of the product streams of unit B was recycled to unit A, then, with two metals analyzed on each stream, all the metal analyses would be redundant and adjustable.

It is worth noting that analyzing additional stream components does not necessarily increase redundancy and enable estimability. A component that is not separated in the process unit has the same or almost same concentration in each stream sample. The associated mass conservation equation is then similar (collinear) to the mass conservation of solids preventing the whole set of equations to be solved.

Estimability and redundancy analyses are complex analyses that are better performed with mathematical algorithm. However, the mathematics is too complex to be presented here.

3.6.9 Mass Balancing Computer Programs

It might be tempting to use spreadsheets to compute mass balances. Indeed, spreadsheets offer most of the features required to easily develop and solve a mass balance problem. However, spreadsheets are error prone, expensive to troubleshoot and maintain over time, and become very quickly limited to fulfill the needs and requirements of complex mass balances and their statistics (Panko, 2008).

There exist several providers of computer programs for mass balance calculations for a variety of prices. While all the features of an advanced solution may not be required for a given type of application (research, process survey, modeling and simulation project, on-line mass balancing for automatic control needs, rigorous production and metallurgical accounting), a good computer program should offer:

• An easy way to define, list and select mass conservation equations

• An easy way to define error models

• Visualization and validation tools to quickly detect gross errors

• A sensitivity analysis tool

• An estimability and redundancy analysis tool

• Statistical tools to validate the error models against the adjustments made

Furthermore, depending on the mass balance types and application, the program should provide:

• Support for complete analyses (e.g., size distributions)

• Support for assay-by-size fractions

• Connectivity to external systems (archiving databases, Laboratory Information Management Systems) for automated data acquisition

• Automated detection and removal of gross errors

• Automated configuration of mass conservation models

• A reporting tool

• A dedicated database to support ancillary requirements

• Trend analysis

• Multi-mass balance capabilities

Additional information can be found in Crowe (1996), Romagnoli and Sánchez (1999) and Morrison (2008).

3.6.10 Metallurgical Balance Statement

One principal output of the sampling, assaying, mass balancing/data reconciliation exercise is the metallurgical balance, a statement of performance over a given period: a shift, a day, a week, etc. Assuming that the reconciled data in Example 3.12 correspond to a day when throughput was 25,650 t (measured by weightometer and corrected for moisture), the metallurgical balance would look something like in Table 3.3.

Table 3.3

Metallurgical Balance for XY Zn Concentrator for Day Z

Product Wt (t) Grade (%) Distribution (%)
Zn Cu Cu Zn
Feed 25,650.0 3.89 0.17 100 100
Zn conc 1,674.1 52.07 0.66 87.35 25.71
Tailings 23,975.9 0.53 0.13 12.65 74.29

Image

The calculations in the Wt (weight) column is from the solid (mass) split using Eq. (3.25). It should be evident using the reconciled data in Table ex 3.12a that regardless of which metal assay is selected to perform the calculation that the result for solid split is the same (including using Fe), giving WC/WF=0.065. (In checking you will note that it is important to retain a large number of significant figures to avoid “rounding errors”, rounding at the finish to create the values given in Table 3.3.) From the solid split the tonnage to concentrate is calculated by multiplying by the feed solids flowrate (i.e., 25,650×0.065=1,674.1), which, in turn, gives the tonnage to tailings. The recovery to concentrate is given using Eq. (3.27), and thus the recovery (loss) to tails is known; together recovery and loss are referred to as “distribution” in the table. The information in the table would be used, for instance, to compare performance between time periods, and to calculate Net Smelter Return (Chapter 1).

3.7 Example Mass Balance Calculations

Mass balancing based on metal assays has been described, and used to illustrate data reconciliation. As noted, components other than metal (or mineral) can be used for mass balancing. The use of particle size and per cent solids is illustrated; this is done without the associated data reconciliation, but remember, this is always necessary for accurate work.

3.7.1 Use of Particle Size

Many units, such as hydrocyclones and gravity separators, produce a degree of size separation and the particle size data can be used for mass balancing (Example 3.15).

Example 3.15

The streams around the hydrocyclone in the circuit (Figure ex 3.15) were sampled and dried for particle size analysis with the results in Table ex 3.15. Determine the solids split to underflow and the circulating load.

image
Figure ex 3.15 Grinding circuit, ball mill closed with a hydrocyclone.

Table ex 3.15

Size Distribution Data Obtained on Cyclone Feed, Overflow and Underflow

Size* (μm)CFCOFCUFf-ou-o(f–o)/(u–o)
+5927.690.0112.847.6812.830.599
−592+4194.690.347.184.356.840.636
−419+2966.680.79.965.989.260.646
−296+2097.032.619.324.426.710.659
−209+14811.297.6313.63.665.970.613
−148+10513.6213.5514.910.071.360.051
−105+7411.3911.5210.42−0.13−1.10.118
−74+539.81157.19−5.19−7.810.665
−53+375.659.373.32−3.72−6.050.615
−3722.1539.2711.26−17.12−28.010.611
 100100100   

Image

Solution

Solids split: We can set up the calculation in the following way, first recognizing the two balances:

Solids: F=O+Uimage

Particle size: Ff=Oo+Uuimage

where F, O, U refer to solids (dry) flowrate of feed, overflow and underflow, and f, o, u refer to weight fraction (or %) in each size class. Since the question asks for solids split to underflow, U/F, we can use the solids balance to substitute for O (=FU) in the particle size balance. After gathering terms and re-arranging we arrive at:

UF=(fo)(uo)

image

While this may seem that we could simply apply the two-product formula, it is always advisable to set up the solution starting from the basic balances. Attempting to remember formulae is an invitation to error.

The estimates of U/F derived from each size class are included in the table. This is another example where errors in the data produce uncertainty in the mass balance. Note that the two estimates that are far from the others, for the 105 μm and 74 μm size fractions, correspond to situations where there is little difference in data between the three streams, making the calculation numerically unstable (if all three were equal no solution can be found).

By re-arranging we can write:

(fo)=UF(uo)

image

This has the form of a linear equation passing through the origin (y=mx) where the slope is solids split U/F. Figure ex 3.15-S shows the resulting plot and includes the solids split resulting from linear regression using the function in Excel.

image
Figure ex 3.15-S Plot of (f-o) vs. (u-o) to determine solids split.

The plot agrees with expectation and the solids split is therefore:

UF=0.616

image

Circulating load: While this is encountered and defined later (e.g., Chapter 7), it is sufficient here to note it is given by the flowrate returning the ball mill (U) divided by the fresh feed rate to the circuit (N); that is, and observing that N=O (at steady state), we can write:

CL=UO

image

We could approach the problem in the same way as we did to solve for U/F (by substituting for F in this case). More directly CL is given by noting that:

U=0.616FandthusO=0.384F

image

giving:

UO=0.6160.384=1.60

image

Thus the circulating load is 1.60, often quoted as a percentage, 160%. We will see that this is typical value for a closed ball mill-cyclone circuit (Chapters 7 and 9).

Example 3.15 is an example of node imbalance minimization; it provides, for example, the initial value for the generalized least squares minimization. This graphical approach can be used whenever there is “excess” component data; in Example 3.9 it could have been used.

Example 3.15 uses the cyclone as the node. A second node is the sump: this is an example of 2 inputs (fresh feed and ball mill discharge) and one output (cyclone feed). This gives another mass balance (Example 3.16).

Example 3.16

At the same time as the streams around the cyclone were sampled, the streams around the sump were also sampled, and sized. The measurement results for two size classes, +592 μm and −37 μm are given in the Table ex 3.16.

Table ex 3.16

Raw (unadjusted) Size Distribution Data for New Feed, Ball Mill Discharge, and Cyclone Feed

 n b f n-f f-b (n–f)/(f–b)
592 19.86 2.16 7.69 12.17 5.53 2.201
−37 19.84 20.27 22.15 −2.31 1.88 −1.229

Image

Determine the circulating load.

Solution

The mass balance equations for the sump are:

Solids: N+B=Fimage

Particle size: Nn+Bb=Ffimage

and CL=BNimage

Combining equations to eliminate F we deduce that:

CL=nffb

image

The estimates of CL are included in the table. Clearly, there is error, the result for −37 μm even being negative. After reconciliation the adjusted data and mass balance are given below:

59219.421.448.2711.156.831.633
−3720.6621.6121.25−0.59−0.361.639

Image

Clearly now the data do agree more closely with the result in Example 3.15. By experience, mass balancing around the sump is less reliable than around the cyclone.

In Chapter 9 we return to this grinding circuit example using adjusted data to determine the cyclone partition curve.

3.7.2 Use of Percent Solids

A unit that gives a large difference in %solids between streams is a thickener (Examples 3.17); another is the hydrocyclone (Example 3.18).

Example 3.17

Feed to a thickener is 30% solids by weight and the underflow is 65% by weight. Calculate the recovery of water to the overflow Rw/o assuming no solids report to the overflow.

Solution

We can approach the problem in two ways: 1, starting with a balance on the slurry, and 2, using dilution ratios. Using F, O, U to represent feed, overflow and underflow flowrates, f, o, u to represent % solids (as a fraction), and subscripts sl, s and w to represent slurry, solids, and water we can write the appropriate balances.

Balances based on slurry: We can write three balances:

Slurry: Fsl=Osl+Uslimage

Solids: Fslfs=Oslos+Uslusimage

Water: Fsl(1fs)=Osl(1os)+Usl(1us)image

and thus water recovery will be given by:

Rw/o=OslFsl1os1fsOslFsl×11fs

image

Since Usl does not appear in the recovery equation it can be substituted by (FslOsl) from the slurry balance into the water balance (this is effectively combining the two equations). In this manner the ratio Osl/Fsl is obtained from the water balance equation:

OslFsl=usfsus

image

and thus Rw/o is determined:

Rw/o=usfsus(1fs)

image

Balances based on dilution ratios: Using dilution ratios (Section 3.4) the water mass flowrate can be obtained, as illustrated using the feed stream as an example:

Feed solids: Fs=Fslfsimage

Feed water: Fw=Fsl(1fs)image

From the ratio: Fw=Fs1fsfsimage

Analogous expressions for water flowrate in the other streams can be written. It will be noticed that the expression for Ow in this example is not determinable (as Os and os are zero). Thus the Rw/o is determined from:

Rw/o=1UwFw

image

After substituting by the dilution ratios and re-arranging we obtain:

Rw/o=usfsus(1fs)

image

Thus the two approaches give the same result, as must be the case.

Solving we obtain:

Rw/o=0.650.300.65×(10.3)Rw/o=0.769or77%

image

Example 3.18

Referring to Example 3.15, at the same time % solids data were determined on each stream: feed, 55%; overflow, 41%; and underflow, 70%. Determine the split of solids to underflow and the split of water to the underflow.

Solution

The problem could be set up starting with the slurry balance, but it lends itself to the dilution ratio approach. Using the same symbolism as in Example 3.17 the dilution ratios are:

Fw=Fs100fsfs=Fsfs'=Fs(0.818)Ow=Os100osos=Osos'=Os(1.439)Uw=Us100usus=Usus'=Us(0.429)

image

Solids split: The two balances are:

Solids: Fs=Os+Usimage

Water: Fsfs'=Osos'+Usus'image

giving:

Solids split: Rs/u=UsFsimage

By combining equations to eliminate Os we derive:

Rs/u=fs'os'us'os'=0.8181.4390.4291.439Rs/u=0.614(or61.4%)

image

This compares well with the estimate in Example 3.15, which is often not the case using raw, that is, unadjusted data; the % solids data could be included in mass balance/data reconciliation, and must be if use is to be made of stream values derived from those data.

Water split: The water split is:

Rw/u=UwFw=Usus'Fsfs'Rw/u=0.614×0.4290.818Rw/u=0.322(or32.2%)

image

The water split to underflow, also referred to as water recovery to underflow, is an important parameter in modeling the performance of a cyclone (Chapter 9).

3.7.3 Illustration of Sensitivity of Recovery Calculation

See Example 3.19.

Example 3.19

A concentrator treats a feed of 2.0% metal. Compare the 95% confidence interval on the recovery for the following two conditions:

a. Producing concentrate grading 40% metal and a tailings of 0.3% metal.

b. Producing concentrate grading 2.2% metal and a tailings of 1.3% metal.

Take that the relative standard deviation on the metal assays is 5% in both cases.

Solution a)

Recovery:

From Eq. (3.27): R=85.6%

Standard deviation on R:

Solving for VR (Eq. (3.33)) gives:

VR=0.14492(394.023Vf+0.001806Vc+16044.44Vt)

image

To insert the values of Vf, Vc, and Vt the relative standard deviations must be changed to absolute values, σ (remembering the Eq. (3.33) is written in terms of fractions):

σf=5% of 2.0%=0.1% (or 0.001);

σc=5% of 40%=2% (or 0.02);

and σt=5% of 0.3%=0.015% (or 0.00015)

Substituting gives (recalling that V2):

VR=0.14492 (0.000394+7.225E–10+0.000361)

VR=0.00011

That is: σR=0.01046 (or 1.05%)

The standard deviation on the recovery is ±1.05% or an approximate 95% confidence interval of 2×1.05% or 2.1%. The recovery is best reported, therefore, as:

R=85.6±2.1%

image

Solution b)

Recovery:

From Eq. (3.27): R=85.6% (i.e., the same)

Standard deviation of R:

Solving for VR (Eq. (3.33)):

VR=3.11675(0.2025Vf+0.10124Vc+0.02367Vt)

image

To insert the values of Vf, Vc, and Vt the relative standard deviations must be changed to absolute values, σ:

σf=5% of 2.0%=0.1% (or 0.001);

σc=5% of 2.2%=0.11% (or 0.0011);

and σt=5% of 1.3%=0.065% (or 0.00065)

Substituting gives (recalling that V2):

VR=3.11675(2.025E07+1.225E07+1E08)VR=0.01044

image

That is:

σR=0.102(or10.2%)

image

And the 95% confidence interval is:

R=85.6±20.4%

image

The much higher uncertainty in the Example 3.19 (b) compared to a) again illustrates the greater uncertainty in the value of the recovery when the metal (or any component) is not well separated.

Example 3.19 raises a question as to how to estimate the variance. In this case a relative standard deviation (standard deviation divided by the mean) was taken and assumed equal for all assays. The standard deviation could be broken down into that associated with sampling and that associated with assaying, or some other “error model”. Regardless, there is some questioning of the common use of relative standard deviation as the absolute standard deviation decreases as the assay value decreases and some would argue this is not realistic. Alternative error models which avoid this have been suggested (Morrison, 2010), but none appear to be universally accepted.

References

1. Aldrich C, et al. On-line monitoring and control of froth flotation systems with machine vision: a review. Int J Miner Process. 2010;96(1-4):1–13.

2. Almasy GA, Mah RSH. Estimation of measurement error variances from process data. Ind Eng Chem Process Des Dev. 1984;23(4):779–784.

3. AMIRA, 2007. P754: Metal Accounting - Code of Practice and Guidelines, Release 3.

4. Anon., 2011. Using x-ray diffraction to solve mineral processing problems – CSIRO Process Science and Engineering, Oct 2011.

5. Apelt TA, et al. Inferential measurement of SAG mill parameters II: state estimation. Miner Eng. 2002;15(12):1043–1053.

6. Arena T, McTiernan J. On-belt analysis at Sepon copper operation. Proc Metallurgical Plant Design and Operating Strategies (MetPlant 2011) Perth, WA, Australia: AusIMM; 2011; pp. 527–535.

7. Bauer M, Craig IK. Economic assessment of advanced process control – A survey and framework. J Process Control. 2008;18(1):2–18.

8. Bauer M, et al. A profit index for assessing the benefits of process control. Ind Eng Chem Res. 2007;46(17):5614–5623.

9. Bearman RA, Milne RW. Expert systems: opportunities in the minerals industry. Miner Eng. 1992;5(10-12):1307–1323.

10. Bergeron A, Lee DJ. A practical approach to on-stream analysis. Min Mag. 1983;Oct.:257–261.

11. Berton A, Hodouin D. Statistical detection of gross error in material balance calculations. In: Herbst JA, ed. Control 2000: Mineral and Metallurgical Processing. Littleton, CO, USA: SME; 2000:23–33. (Chapter 3).

12. Blanchette G, et al. Estimation errors for gold inventory in large carbon-in-leach tanks. Proc 43th Annual Meeting of the Canadian Mineral Processors Conf Ottawa, ON, Canada: CIM; 2011; pp. 477–490.

13. Box GEP, Draper NR. Evolutionary Operation: A Statistical Method for Process Improvement Classics Edn. New York, NY, USA: John Wiley and Sons; 1998.

14. Boyd, R., 2005. An innovative and practical approach to sampling of slurries for metallurgical accounting. <www.thermo.com.cn/Resources/201007/27101513615.doc>.

15. Braden TF, et al. On-line composition analysis of mineral slurries. In: Littleton, CO, USA: SME; 2002:2020–2047. Mular AL, ed. Mineral Processing Plant Design, Practice and Control. vol. 2.

16. Brdys M, Tatjewski P. Iterative Algorithms for Multilayer Optimizing Control Imperial College Press 2005.

17. Burgos R, Concha F. Further development of software for the design and simulation of industrial thickeners. Chem Eng J. 2005;111(2-3):135–144.

18. Cao S, Rhinehart RR. An efficient method for on-line identification of steady state. J Process Control. 1995;5(6):363–374.

19. Caro D. Automation Network Selection: A Reference Manual second ed. Publisher: ISA (The International Society of Automation) 2009.

20. Chang JW, Bruno SJ. Process control system in the mining industry. World Min. 1983;36(May):37–41.

21. Chen J, et al. Robust estimation of measurement error variance/covariance from process sampling data. Comput Chem Eng. 1997;21(6):593–600.

22. Clarkson, C.J., 1986. The potential for automation and process control in coal preparation. Norrie, A.W., Turner, D.R. (Eds.), Proc. Automation for Mineral Resource Development: 1st IFAC Symposium, Brisbane, Queenland, Australia, pp. 247–252.

23. Cortes, G., et al., 2008. Rougher flotation multivariable predictive control: concentrador A-1 Division Codelco Norte. Kuyvenhoven, R., et al. (Eds.), Proc. Inertnation Mineral Processing Conf., Santiago, Chile, pp. 327–337.

24. Crowe CM. Data reconciliation – progress and challenges. J Process Control. 1996;6(2-3):89–98.

25. Cutting GW. Estimation of interlocking mass balances on complex mineral beneficiation plants. Int J Miner Process. 1976;3(3):207–218.

26. Darouach, M., et al., 1989. Maximum likelihood estimator of measurement error variances in data reconciliation, Proc. Advanced Information Processing in Automatic Control (AIPAC'89), IFAC, Nancy, France, pp. 109–112.

27. Duarte M, et al. Grinding operation optimization of the CODELCO-Andina concentrator plant. Miner Eng. 1998;11(12):1119–1142.

28. Duchesne C. Multivariate image analysis in mineral processing. In: Sbárbaro D, del Villar R, eds. Advanced Control and Supervision of Mineral Processing Plants. London: Springer; 2010:85–142. (Chapter 3).

29. Edwards R, et al. Strategies for instrumentation and control of grinding circuits. In: Littleton, CO, USA: SME; 2002:2130–2151. Mular AL, ed. Mineral Processing Plant Design, Practice and Control. vol. 2.

30. Finch, J.A., Matwijenko, O., 1977. Individual mineral behaviour in a closed-grinding circuit. CIM Bull., 70(788), 164–172.

31. Flament, F., et al., 1986. Propagation of measurement errors in mass balance calculation of mineral processing data, Proc. 19th International Symposium on Application of Computers and Operations Research (APCOM’86). Mineral Processing Applications I, pp. 553–562 (Chapter 50).

32. Flintoff B. Introduction to process control. In: Littleton, CO, USA: SME; 2002:2051–2065. Mular AL, ed. Mineral Processing Plant Design, Practice and Control. vol. 2.

33. Flintoff B, et al. Justifying control projects. In: Halbe DN, ed. Plant Operators. Littleton, CO, USA: SME; 1991:45–53.

34. Flintoff B, et al. Innovations in comminution instrumentation and control. Mineral Processing and Extractive Metallurgy: 100 Years of Innovation Englewood, CO, USA: SME; 2014; pp. 91–116.

35. François-Bongarçon D, Gy P. Critical aspects of sampling in mills and plants: a guide to understanding sampling audits. J S Afr Inst Min Metall. 2002a;102(8):481–484.

36. François-Bongarçon D, Gy P. The most common error in applying ‘Gy’s Formula’ in the theory of mineral sampling, and the history of the liberation factor. J S Afr Inst Min Metall. 2002b;102(8):475–480.

37. Frew JA. Computer-aided design of sampling schemes. Int J Miner Process. 1983;11(4):255–265.

38. Geelhoed B. Is Gy’s formula for the fundamental sampling error accurate? Experimental evidence. Miner Eng. 2011;24(2):169–173.

39. González GD. Soft sensing. In: Sbárbaro D, del Villar R, eds. Advanced Control and Supervision of Mineral Processing Plants. London Limited, UK: Springer-Verlag; 2010:143–212. (Chapter 4).

40. González GD, et al. Local models for soft-sensors in a rougher flotation bank. Miner Eng. 2003;16(5):441–453.

41. Gy, P.M., 1967. L’Echantillonage des Minerais en Vrac (Sampling of Particulate Materials). Vol. 1. Revue de l’Industrie Minerale, Special, No. St. Etienne, France.

42. Gy PM. Sampling of Particulate Materials: Theory and Practice second ed. Amsterdam: Elsevier; 1982.

43. Hales LB, Ynchausti RA. Neural networks and their use in the prediction of SAG mill power. In: Kawatra SK, ed. Comminution - Theory and Practice. Littleton, CO, USA: SME; 1992:495–504. (Chapter 36).

44. Hales LB, Ynchausti RA. Expert systems in comminution – cases studies. In: Kawatra SK, ed. Comminution Practices. Littleton, CO, USA: SME; 1997:57–68. (Chapter 8).

45. Harris CA, Meech JA. Fuzzy logic: a potential control technique for mineral processing. CIM Bull. 1987;80(905):51–59.

46. Herbst JA, Bascur OA. Mineral processing control in the 1980s – realities and dreams. In: Herbst JA, ed. Control ‘84 Mineral/Metallurgical Processing Mineral Processing Control in the 1980s. New York, NY, USA: SME; 1984:197–215. (Chapter 21).

47. Herbst JA, Bascur OA. Alternative control strategies for coal flotation. Miner Metall Process. 1985;2:1–9.

48. Herbst, J.A., Pate, W.T., 1996. On-line estimation of charge volumes in semiautogenous and autogenous grinding mills, Proc. 2nd International Conference on Autogenous and Semiautogenous Grinding Technology, Vol. 2, Vancouver, BC, Canada, pp. 669–686.

49. Hodouin D. Methods for automatic control, observation and optimization in mineral processing plants. J Process Control. 2011;21(2):211–225.

50. Hodouin D, Everell MD. A hierarchical procedure for adjustment and material balancing of mineral processes data. Int J Miner Process. 1980;7(2):91–116.

51. Hodouin D, Najim K. Adaptive control in mineral processing. CIM Bull. 1992;85(965):70–78.

52. Hodouin D, et al. Reliability of material balance calculations a sensitivity approach. Miner Eng. 1989;2(2):157–169.

53. Holdsworth M, et al. Optimising concentrate production at the Green’s Creek Mine. Proc SME Annual Meeting, Preprint 02-063 Phoenix, AZ, USA: SME; 2002; pp. 1–10.

54. Holmes RJ. Sampling methods: problems and solutions. In: Malhotra D, ed. Evaluation and Optimization of Metallurgical Performance. Littleton, CO, USA: SME; 1991:157–167. (Chapter 16).

55. Holmes RJ. Correct sampling and measurement—The foundation of accurate metallurgical accounting. Chemom Intell Lab Syst. 2004;74(1):71–83.

56. Hulthén E, Evertsson CM. Real-time algorithm for cone crusher control with two variables. Miner Eng. 2011;24(9):987–994.

57. Janse Van Vuuren MJ, et al. Detecting changes in the operational states of hydrocyclones. Miner Eng. 2011;24(14):1532–1544.

58. Jansen W. Applying best practices in metal production accounting and reconciliation. Eng Min J (E&MJ). 2009;211(10):68–71.

59. Johnson NW. In: Greet CJ, ed. Existing Methods for Process Analysis Flotation Plant Optimization: A Metallurgical Guide to Identifying and Solving Problems in Flotation Plants. AusIMM 2010:35–63.

60. Jonas R. Advanced control for mineral processing: better than expert systems. Proc 40th Annual Mineral Processors Operators Conf Ottawa, ON, Canada: CIM; 2008; pp. 421–443.

61. Kawatra, S.K., 1985. The development and plant trials of an ash analyzer for control of a coal flotation circuit, Proc. 15th International Mineral Processing Cong., Cannes, France, pp. 176–188.

62. Kawatra SK, Cooper HR. On-stream composition analysis of mineral slurries. In: Mular AL, Anderson MA, eds. Design and Installation of Concentration and Dewatering Circuits. Littleton, CO, USA: SME Inc.; 1986:641–654. (Chapter 41).

63. Keller JY, et al. Analytical estimator of measurement error variances in data reconciliation. Comput Chem Eng. 1992;16(3):185–188.

64. Kistner M, et al. Monitoring of mineral processing systems by using textural image analysis. Miner Eng. 2013;52:169–177.

65. Klimpel, A., 1979. Estimation of weight ratios given component make-up analyses of streams, AIME Annual Meeting, New Orleans, LA, USA, Preprint No. 79-24.

66. Ko Y-D, Shang H. A neural network-based soft sensor for particle size distribution using image analysis. Powder Technol. 2011;212(2):359–366.

67. Krstev, B., Golomeov, B., 1998. Use of evolutionary planning for optimisation of the galena flotation process, Proc. 27th International Symposium on Computer Applications in the Minerals Industries (APCOM), London, UK, pp. 487–492.

68. Lachance L, Flament F. A new ERA tool for state of the art metallurgical balance calculations. Proc 43th Annual Meeting of the Canadian Mineral Processors Conf Ottawa, ON., Canada: CIM; 2011; pp. 331–346.

69. Lachance L, et al. Implementing best practices of metal accounting at the Strathcona Mill. Proc 44th Annual Canadian Mineral Processors Operators Conf Ottawa, ON, Canada: CIM; 2012; pp. 375–386.

70. Laguitton D. The SPOC Manual: Simulated Processing of Ore and Coal Canada Centre for Mineral and Energy Technology, Department of Energy, Mines and Resources Ottawa, ON, Canada: Canadian Government Publishing Centre; 1985.

71. Laguitton, D., Leung, J., 1989. Advances in expert system applications in mineral processing. Dobby, G.S., Rao, R. (Eds.), Proc. International Symposium on Processing of Complex Ores, Halifax, NS, Canada, pp. 565–574.

72. Leskinen T, et al. Performance of on-stream analysers at Outokumpu concentrators, Finland. CIM Bull. 1973;66(Feb.):37–47.

73. Lessard J, et al. Developments of ore sorting and its impact on mineral processing economics. Miner Eng. 2014;65:88–97.

74. Lestage R, et al. Constrained real-time optimization of a grinding circuit using steady-state linear programming supervisory control. Powder Technol. 2002;124(3):254–263.

75. Ljung, L., 1999. System Identification: Theory for the User, second ed., Englewood Cliffs, NJ, USA.

76. Lyman GJ. Application of Gy’s sampling theory to coal: a simplified explanation and illustration of some basic aspects. Int J Miner Process. 1986;17(1-2):1–22.

77. Lyman G, et al. Bias testing of cross belt samplers. J S Afr Inst Min Metall. 2010;110:289–298.

78. Lynch AJ. Mineral Crushing and Grinding Circuits: Their Simulation, Optimisation, Design and Control Amsterdam; New York, NY, USA: Elsevier Scientific Pub. Co; 1977.

79. Maldonado M, et al. On-line estimation of frother concentration for flotation processes. Can Metall Q. 2010;49(4):435–446.

80. Matthews, D., du Toit, T., 2011. Real-time online analysis of iron ore, validation of material stockpiles and roll out for overall elemental balance as observed in the Khumani Iron Ore Mine, South Africa, Proc. Iron Ore Conference, Perth, WA, Australia, pp. 297–305.

81. McKee DJ, Thornton AJ. Emerging automatic control approaches in mineral processing. In: Netherlands: Springer; 1986:117–132. Wills BA, Barley RW, eds. Mineral Processing at a Crossroads – Problems and Prospects. Vol. 117 NATO ASI Series.

82. McIvor RE, Finch JA. A guide to interfacing of plant grinding and flotation operations. Miner Eng. 1991;4(1):9–23.

83. Merks JW. Sampling and Weighing of Bulk Solids Clausthal-Zellerfeld, Germany: Trans Tech Publications; 1985.

84. Merks JW. Sampling in mineral processing. In: Littleton, CO, USA: SME; 2002:37–62. Mular AL, ed. Mineral Processing Plant Design, Practice and Control. vol. 1.

85. Merks JW. Spatial dependence in material sampling. In: Geelhoed B, ed. Approaches in Material Sampling. Amsterdam: IOS Press; 2010:105–149. (Chapter 6).

86. Minnitt RCA, et al. Part 1: understanding the components of the fundamental sampling error: a key to good sampling practice. J S Afr Inst Min Metall. 2007;107(8):505–511.

87. Minnitt RCA. Sampling in the South African minerals industry. J S Afr Inst Min Metall. 2014;114(1):63–81.

88. Morari M, Zafiriou E. Robust Process Control Englewood Cliffs, NJ, USA: Prentice Hall; 1989.

89. Morrison, R.D., 2008. An Introduction to Metal Balancing and Reconciliation, JKMRC Monograph Series in Mining and Mineral Processing, No.4, The Univerisity of Queensland, Australia.

90. Morrison RD. In: Greet CJ, ed. Mass Balancing Flotation Data, Flotation Plant Optimisation. Carlton, Victoria, Australia: AusIMM; 2010:65–81. Spectrum Number 16.

91. Mular AL. Data adjustment procedures for mass balances. In: Weiss A, ed. Computer Methods for the 80’s in the Minerals Industry. New York, NY, USA: SME; 1979:843–849. (Chapter 4).

92. Mular AL. Modelling, simulation and optimization of mineral processing circuits. In: Sastry KVS, Fuerstenau MC, eds. Challenges in Mineral Processing. Littleton, CO, USA: SME; 1989:323–349. (Chapter 19).

93. Napier-Munn TJ, Lynch AJ. The modelling and computer simulation of mineral treatment processes - current status and future trends. Miner Eng. 1992;5(2):143–167.

94. Narasimhan S, Jordache C. Data Reconciliation and Gross Error Detection: An Intelligent Use of Process Data Gulf Publishing Company 2000.

95. Nuñez, E., et al., 2009. Self-optimizing grinding control for maximizing throughput while maintaining cyclone overflow specifications, Proc. 41th Annual Meeting of the Canadian Mineral Processors Conf., Ottawa, ON, Canada, pp. 541–555.

96. Osorio D, et al. Assessment of expert fuzzy controllers for conventional flotation plants. Miner Eng. 1999;12(11):1327–1338.

97. Panko RR. What we know about spreadsheet errors. J Organ End User Comput (JOEUC). 2008;10(2):15–21.

98. Patel, M., 2014. On-belt elemental analysis of lead-zinc ores using prompt gamma neutron activation analysis, Proc. 27th International Mineral Processing Congr. (IMPC 2014), Santiago, C1717, Chile, 1–9 (Chapter 17).

99. Pitard FF. Pierre Gy’s Sampling Theory and Sampling Practice: Heterogeneity, Sampling Correctness, and Statistical Process Control second ed. CRC Press 1993.

100. Rajamani K, Hales LB. Developments in adaptive control and its potential in the minerals industry. Miner Metall Process. 1987;4(2):18–24.

101. Ramasamy M, et al. Control of ball mill grinding circuit using model predictive control scheme. J Process Control. 2005;15(3):273–283.

102. Reid KJ, et al. A survey of material balance computer packages in the mineral industry. In: Johnson TB, Barnes RJ, eds. Proc 17th International Symposium(APCOM)-Application of Computers and Operations Research. New York, NY, USA: SME; 1982:41–62. (Chapter 6).

103. Romagnoli JA, Sánchez MC. Data Processing and Reconciliation for Chemical Process Operations Process Systems Engineering. vol. 2 Academic Press 1999.

104. Ruel M. Advanced control decision tree. Proc 46th Annual Mineral of the Canadian Mineral Processors Conf Ottawa, ON, Canada: CIM; 2014; pp. 179–187.

105. Saensunon B, et al. A combined 57Fe-Mössbauer and X-ray diffraction study of the ilmenite reduction process in commercial rotary kiln. Int J Miner Process. 2008;86(1-4):26–32.

106. Salama AIA, et al. Coal preparation process control. CIM Bull. 1985;78(881):59–64.

107. Sbarbaro D, et al. A multi-input-single-output Smith predictor for feeders control in SAG grinding plants. IEEE Trans Control Syst Technol. 2005;13(6):1069–1075.

108. Sbarbaro D, et al. Adaptive soft-sensors for on-line particle size estimation in wet grinding circuits. Control Eng Pract. 2008;16(2):171–178.

109. Seborg DE, et al. Process Dynamics and Control third ed. USA: John Wiley and Sons; 2010.

110. Shean BJ, Cilliers JJ. A review of froth flotation control. Int J Miner Process. 2011;100(3-4):57–71.

111. Shoenbrunn F, Toronto T. Advanced thickener control. Advanced Process Control Applications for Industry Workshop Canada: IEEE Industry Applications Society, Vancouver BC; 1999; 83-86.

112. Skogestad S. Simple analytic rules for model reduction and PID controller tuning. J Process Control. 2003;13(4):291–309.

113. Smith HW, Ichiyen NM. Computer adjustment of metallurgical balances. CIM Bull. 1973;66(Sep.):97–100.

114. Steyn CW, Sandrock C. Benefits of optimisation and model predictive control on a fully autogeneous mill with variable speed. Miner Eng. 2013;53:113–123.

115. Thornton AJ. Cautious adaptive control of an industrial flotation circuit. Miner Eng. 1991;4(12):1227–1242.

116. Thwaites P. Process control in metallurgical plants – From an Xstrata perspective. Ann Rev Control. 2007;31(2):221–239.

117. Tipman R, et al. Mass balances in mill metallurgical operations. Proc 10th Annual Meeting of the Canadian Mineral Processors Conf Ottawa, ON, Canada: CIM; 1978; pp. 336–356.

118. Toop A, et al. Advances in in-stream analysis. In: Jones MJ, Gill P, eds. Mineral Processing and Extractive Metallurgy: Papers Presented at the International Conference Mineral Processing and Extractive Metallurgy. Kunming, Yunnan, China: IMM and Chinese Society of Metals; 1984:187–194.

119. Ulsoy AG, Sastry KVS. Principal developments in the automatic control of mineral processing systems. CIM Bull. 1981;74(836):43–53.

120. Van der Spuy, D.V., et al., 2003. An on-line expert for mineral processing plants, Proc. 22th International Mineral Processing Cong., (IMPC), Paper 368. Cape Town, South Africa, pp. 223–231.

121. Visman J. Discussion 3: a general theory of sampling. J Mater. 1972;7(3):345–350.

122. Wiegel RL. Advances in mineral processing material balances. Can Metall Q. 1972;11(2):413–424.

123. Yianatos J, et al. Modelling and simulation of rougher flotation circuits. Int J Miner Process. 2012;112-113:63–70.

124. Ziegler JG, Nichols NB. Optimum settings for automatic controllers. Trans ASME AIMME. 1942;64:759–768.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
3.137.183.210