University
stringclasses
19 values
Text
stringlengths
458
20.7k
ADE
Chapter 3-Quantitative Core-Log Analysis 91 3.5.1: Discussion of the statistical results and the bar diagram of minerals. In comparison with massive pyrite deposits within the Broken Hill Block [e.g. Big Hill cobaltian pyrite, Thackaringa Group (Plimer 1977)] the lack of pyrite within the investigated samples of the Western Mineralisation suggests that the orebody is originally sulphur-poor. Previous studies (Groombridge 2003; Kitchen 2001; Patchett 2003; Sproal 2001) showed that the galena and sphalerite samples of the Western Mineralisation were highly enriched in Fe. This indicates a low sulphur system for formation of the orebody of the Western Mineralisation. Plimer (1977, 2006a) claimed that the Broken Hill orebodies have been unsaturated with S and the excess Pb and Zn have contributed to the composition of silicate minerals such as gahnite, zincian garnet, zincian biotite, zincian staurolite, zincian chlorite, zincian sericite, plumbian orthoclase, native lead, dyscrasite, native silver, native lead, zincian manganese olivine, safflorite and löllingite. This study also shows that the number of samples containing gahnite and green feldspar are relatively substantial within the investigated samples of the Western Mineralisation. However, weak acid digestion of sulphides for ICP-OES analysis would not have dissolved base metal- bearing silicates and gahnite. The low volume percentage and lack of arsenopyrite within the investigated samples may be another reason for the deficiency of primary S in the Western Mineralisation. Spry, Plimer and Teale (2008, p.232) suggested the chemical reaction (3.1) as a possibility for formation of composite arsenopyrite-löllingite mineral3 in the - Broken Hill deposit. 2FeAs + 2FeS + S = 4FeAsS (3.1) 2 2 The chemical reaction (3.1) shows that the generation of arsenopyrite during metamorphism needs to consume S rather than production of S. Despite the lack of pyrite and arsenopyrite (Table 3.5), chalcopyrite was presented in a larger number of samples of the Western Mineralisation. This may indicate that the dispersal of chalcopyrite (or minerals containing Cu) is controlled by different parameters. Plimer (2006b) argued that in both the Olary and Broken Hill Domains, Cu was remobilised during the Olarian Orogeny and participated at redox boundaries such as iron 3 Löllingite forms core and arsenopyrite forms rim of the composition mineral
ADE
92 Chapter 3-Quantitative Core-Log Analysis formations or sulphide rocks. The spatial distribution of chalcopyrite and Cu will be discussed in Chapters 6 to 8. According to Table 3.6, the number of samples containing orange garnet, hedenbergite, rhodonite and red garnet is very low within the total investigated samples. It is possible that those silicate minerals were occurred naturally very low or they were unstable over a long period and thus decomposed to other silicate minerals such as pink garnet and gahnite. Mineralogy of the investigated samples of the surface drill core shows: 1. An interdigitation of lode horizons (quartz-gahnite, quartz-garnet), 2. Replacement of metapsammite by quartz-gahnite and sulphides, 3. Replacement of metapelite by garnet rocks and ore types similar to C Lode (quartz-gahnite-sphalerite-galena), 4. B Lode (quartz-hedenbergite-red garnet-sphalerite-galena), and 5. A Lode (quartz-orange garnet-rhodonite-galena-sphalerite). The above observations suggest multiphase ore deposition in the Western Mineralisation. 3.6 Evaluating the precision of the variation of galena+sphalerite against the variation of Pb+Zn The most important aspect of the quantitative core logging is how precisely the variation of galena+sphalerite is correlated with the variation of Pb+Zn within the investigated samples. The two important factors for determination of precision are reliability and validity.  Reliability Reliability refers to the reproducibility of the measurement in experimental studies and the possible capacity for detecting agreements or internal consistency within measurements. In fact, two data sets can show a high correlation coefficient but it is possible they have little agreement and internal consistency in the variation of data. In this study, reliability was measured by the following approaches (Garson 2010; Hopkins 2000a):
ADE
Chapter 3-Quantitative Core-Log Analysis 93 1. Comparing the trend variation of mean and median concentrations for Pb+Zn and galena+sphalerite within the corresponding drill cores (Section 3.6.1), 2. Comparing the COVs of Pb, Zn and Pb+Zn with the COVs of galena, sphalerite and galena+sphalerite respectively within the corresponding drill cores (Section 3.6.1), and 3. Calculation of Spearman correlation coefficient4 and Cronbach’s alpha (UCLA: ATS 2007; Yaffee 2003; Tables 3.9 and 3.10).  Validity Validity refers to the amount of correlation between the value of the measurement and its true value. In this study, the values of Pb+Zn are considered as true values and it is important to know the extent of the correlation between the estimated values of galena+sphalerite and the values of Pb+Zn. Also, the degree of validity was calculated by the Pearson correlation coefficient5 (Hopkins 2000b) for logarithmic data (Table 3.8). But as mentioned earlier a high PCC between Pb+Zn and galena+sphalerite does not mean that the trend variation of Pb+Zn in drill cores is the same as the trend variation of galena+sphalerite. A summary of statistical terms used in this section has outlined in Table 3.7. 4 SCC 5 PCC
ADE
94 Chapter 3-Quantitative Core-Log Analysis Table 3.7: Different types of correlation coefficients and their statistical terms. The correlation coefficients are normally reported as R= (a value between -1 and +1); squaring the R value makes it easier to understand. The square of the correlation coefficient multiplied by 100 [Equation (3.1)] describes the percentage of variation in one variable in related to the variation of another variable. R2 × 100 (3.1) jk Equation (3.1) is equal to percentage of variance in common between X and X . As a j k matter of routine it is the squared correlations that should be interpreted. This is because the correlation coefficient shows the existence of more co-variation between two elements than actually exists, and this problem gets worse as the correlation approaches zero. Significance level indicates how likely the correlations reported may be due to chance in the form of random sampling error. The significance level is important, if the data set is very small. A significance level of 0.05 for a correlation coefficient means that there is (1-0.05) or 95 % certainty of the possibility of being a true correlation coefficient and a significance level of 0.01 for a correlation coefficient means that there is (1-0.01) or 99 % certainty of the possibility of being a true correlation coefficient (Creative Research Systems 1982). There are both parametric and non-parametric statistical methods for measuring the correlation coefficient. In parametric statistics it is supposed that the data set comes from random data with normal distribution and that the parameters of the distribution can be inferred by parametric statistics. In contrast, non-parametric statistics make no assumptions about the randomness of the data and normal distribution and they are less sensitive to outlier effects and skewed data. PCC requires linearity of the relationship between the data and if the actual relationship is non-linear, it may change the real degree of the correlation coefficient. Skewness of distribution (pbarrett.net 2001) and outliers can affect linear relationships - this is the case for most of the geochemical data. If the distribution of the data set is log-normal, logarithmic data is preferable for the calculation of PCC (non-parametric method). SCC is a non-parametric method that can also be used for calculation of the non-linear correlation coefficient. Spearman’s approach is a form of rank order calculation based on the median between all pairs of data in a scale that is also a type of measure of reliability (Garson 2010). Cronbach's alpha is not a statistical test but it is the most common form of reliability coefficient or internal consistency based on the average correlation among the data set. It is calculated in SPSS under the function of AnalysisScaleReliability Analysis. In this study, the dialog of the Reliability Analysis was adjusted for the model of "Two- Way Random Effects", consistency results and Cronbach's alpha. In the Two-Way Random Effects model, both judges and measures effects are random. & cirtemarap tneiciffeoc noitalerroC level ecnac i f i n g i S cirtemarap-non CCP CCS ahpla s'hcabnorC
ADE
Chapter 3-Quantitative Core-Log Analysis 95 3.6.1 Comparison of the variation of Pb+Zn with variation of galena+sphalerite Comparison of Figure 3.11 with Figure 3.3 shows that there are several similar trend variations between galena and Pb and between sphalerite and Zn. However, in drill core 4031 (Figure 3.11), the mean and median volume percentage of galena is very low, although one would expect it to be higher in regard to its corresponding assay value (Pb) in this drill core (Figure 3.3). This may have occurred due to the difficulty of visual separation of galena from sphalerite and silicate minerals during the quantitative core logging. This may especially be the case, when sphalerite appears as the major ore sulphide mineral in a sample and its lustre masks the lustre of galena. This may be the general case for the quantitative core logging of other polymetallic sulphide minerals as well and it is the weakness of visual modal mineral mapping. In this case, construction of a comparative bar diagram for variation of Pb+Zn and galena+sphalerite is more reliable for the detection of possible human error. Especially, in high grade sulphide mineralised samples, the concentration of Pb+Zn is supposed to be high, but if it is considerably lower than expected, it may be a result of one or more of the following mistakes in: 1. Reporting assay data, 2. Quantification of the ore sulphide minerals, or 3. Attributing the assay data to the wrong core sample. Therefore, the comparison of bar diagrams of Pb+Zn and galena+sphalerite (Figures 3.12 and 3.13) provides a more efficient tool for detection of potential mistakes at the preliminary stage of data collection before data processing starts. This process improves the degree of reliability of the important assay data (e.g. Pb and Zn in the Western Mineralisation). Figure 3.13 shows many similarities in COVs of the following pair variables: 1. Pb+Zn and galena+sphalerite, 2. Zn and sphalerite, and 3. Pb and galena.
ADE
98 Chapter 3-Quantitative Core-Log Analysis 3.6.2 Probability plot The construction of a probability plot is a means of identification of the normal distribution of a data set. The probability plots in Figure 3.14 were constructed for 1,219 samples containing Pb+Zn and galena+sphalerite. For construction of a probability plot, the data set should be arranged in ascending order; the largest data will be plotted at a lower percentage than 100 % and this makes it possible for some future data of Pb+Zn concentrations and galena+sphalerite (vol. %) with larger values than the current largest values to be located at a higher point in the order (Hart & Hart 2010). In Figure 3.14, the estimated cumulative probabilities were calculated by the method of Median Rank i0.3 (Benard) with the formula in which " n " is equal to the number of samples and "i" n0.4 is equal to the rank-order of each value (i.e. i = 1 indicates the smallest value and i = n for the largest). More details can be found in the "help" section of the Minitab software (Minitab Inc. 2007). Figure 3.14: (a) and (b) are related to the original data of Pb+Zn and galena+sphalerite. The red curves show the experimental percentage of Pb+Zn concentrations and the volume percentage of galena+sphalerite against their respective estimated values of cumulative probability. The straight blue lines show the theoretical percentage of cumulative probability for near-normal distribution of the variables. (c) and (d) are related to the logarithmic data of Pb+Zn and galena+sphalerite.
ADE
Chapter 3-Quantitative Core-Log Analysis 99 3.6.3 The results of correlation coefficients In the Western Mineralisation, galena and sphalerite were observed only in 1,219 out of 1,849 samples containing Pb and Zn. This means that 6301 samples did not show evidence of galena or sphalerite visually but they do have low assay values for Pb or Zn. When considering this issue, PCC, SCC and alpha coefficients were calculated for the two groups of 1,219 and 1,849 samples in order to evaluate the degree of reliability and validity of the quantitative core log data. In order to calculate correlation coefficient for 1,849 samples, a very small value of 0.000,1 was considered for the value of galena+sphalerite in 630 samples that do not have any values for volume percentage of galena+sphalerite (Tables 3.8 to 3.10). Also, because some volume percentages of galena+sphalerite are equal to one and the logarithm of one is equal to zero, for calculation of PCC, the original data plus 0.000,1 were considered for log-transformation (Table 3.8). According to Figures 3.14c and 3.14d, the logarithmic data of Pb+Zn and galena+sphalerite show near-normal distributions and PCC can be applied to the logarithmic data for evaluating the degree of validity between variation of Pb+Zn and galena+sphalerite (Table 3.8). Table 3.8: The results of PCC for logarithmic (data+0.000,1). Minerals Galena+Sphalerite Galena+Sphalerite Elements Pb+Zn 0.72* 0.45* Number of samples 1219 1849 R2 % 51.8 20.2 *The significance level (2-tailed) is 0.01 Table 3.9: The results of SCC for the original data. Minerals Galena+Sphalerite Galena+Sphalerite Elements Pb+Zn 0.73* 0.78* Number of samples 1219 1849 R2 % 53.3 60.1 *The significance level (2-tailed) is 0.01 1 1,819-1,219
ADE
100 Chapter 3-Quantitative Core-Log Analysis Table 3.10: The results of Cronbach’s alpha for the original data. Cronbach's alpha Number of samples 0.814 1219 0.818 1849 3.6.4 Interpretation of the correlation coefficients  PCC PCC in Table 3.8 shows a relatively high level of correlation (0.72) between the variation of Pb+Zn and galena+sphalerite in 1,219 samples but when 1,849 samples are considered the PCC shows a low level of correlation (0.45) between the pair values of Pb+Zn and galena+sphalerite. The PCC result shows a strong validity between variations of galena+sphalerite and Pb+Zn within the 1,219 investigated samples. In Table 3.8, the result of R2 % for 1,219 samples means that the 51.8 % of variation of galena+sphalerite can be estimated by the variation of Pb+Zn and the result of R2 % for 1,849 samples means that only 20.2 % of variation of galena+sphalerite can be estimated by the variability of Pb+Zn.  SCC SCC in Table 3.9 shows a relatively high level of correlation (0.73) between the variation of Pb+Zn and galena+sphalerite for 1,219 samples. Even when 1,849 samples are considered the value of SCC increases from 0.73 to 0.78 between the pair values of Pb+Zn and galena+sphalerite. This shows high internal consistency between the variation of galena+sphalerite and Pb+Zn for two types of abundant samples.  Alpha coefficient The alpha coefficient in Table 3.10 also shows a high level of internal consistency or reliability between the variation of galena+sphalerite and Pb+Zn for two types of abundant samples. 3.7 Exploration signature of lithologies in the Broken Hill Domain In the Broken Hill Domain, garnet quartzite, garnetite, blue quartz-gahnite lode and pegmatite show obvious spatial relationships with over 400 minor deposits of Broken Hill
ADE
Chapter 3-Quantitative Core-Log Analysis 101 Type (BHT) deposits including the main Broken Hill orebody (Spry, Teale & Heimann 2003). Although, there are many competing ideas about the origin of these rock types in the Broken Hill Domain of the Curnamona Province, these rock types are considered widely as exploration guides to BHT deposits (Spry, Teale & Heimann 2003). The question is whether it is possible to evaluate quantitatively the degree of relationship of the lithologies of the Western Mineralisation with variation of Zn+Pb and whether the rock types can be judged as major controlling factors of the mineralisation. 3.7.1 Statistical results for rock types of the Western Mineralisation The results of SCCs of the investigated rock types with Pb+Zn and their descriptive statistics are given in Table 3.11 and Figure 3.15 respectively. Table 3.11 shows at a glance whether rock types vary with Pb+Zn perfectly, or nearly perfectly, and whether positively or negatively. Table 3.11: A summary of SCC results, significance levels (2-tailed) and abundance of rock types. Elements E l e m e n t s Zn + Pb Zn + Pb Rock types Rock types Quartzite lode - 0.04 Pegmatite 0.3** Sig. (2-tailed) Sig. (2-tailed) 0.772 0.000 Number of Number of 56 339 samples samples Metapsammite Blue quartz lode - 0.13* -0.006 Sig. (2-tailed) Sig. (2-tailed) 0.033 0.880 Number of Number of 271 635 samples samples Metapsammopelite Garnet quartzite - 0.43** -0.29** Sig. (2-tailed) Sig. (2-tailed) 0.000 0.000 Number of Number of 447 644 samples samples Metapelite - 0.53** Sig. (2-tailed) 0.000 Number of 129 samples  Correlation is significant at the 0.05 level (2-tailed)  Correlation is significant at the 0.01 level (2-tailed)
ADE
Chapter 3-Quantitative Core-Log Analysis 103 In Table 3.11, the variation of metapelite and metapsammopelite show a low correlation with variation of Pb+Zn in opposite direction. It is not possible to compare directly the SCC values of the rock types in Table 3.11 with each other because they were calculated for different numbers of rock samples. The variations of the quartzite lode and blue quartz lode are independent of the variation of Pb+Zn and there is no internal relationship or consistency between them (Table 3.11). 3.7.2 Discussion of the correlation coefficient of the rock types with Pb+Zn The results of SCC (Table 3.11) indicate that variations of the rock types within the investigated samples have very low to low levels of relationship with the variation of Pb+Zn. The results do not mean that the rock types are not suitable as an exploration guide to the Western Mineralisation and variogram analysis for the rock types should be calculated to understand the spatial relationship of the rock types with Pb and Zn. This will be discussed in Chapters 6 and 8. 3.8 Summary Descriptive statistics of assay data showed some similarities in the frequency distribution of Zn-Pb, Fe-S, Bi-Sb and Ag-As-Cd. There are also some similarities in the trend variation of mean and median concentrations of Pb-Ag and Zn-S in the bar diagram of surface drill cores. Drill cores 4002 and 4031 contain high mean and median concentrations of Zn, Pb and Ag. In the underground drill cores, there are some similar trend variations among the mean and median values of Zn, Pb and S. The analysed samples of drill cores 4062 and 4064 contain higher mean and median concentrations of Zn and Pb relative to those within the analysed samples of the underground drill cores. The mean and median concentrations of Zn, Pb and Fe within the surface drill cores are less than those within the underground drill cores. In contrast, the mean and median concentrations of As, Sb, Bi and S within the surface drill cores are more than those within the underground drill cores. In the bar diagrams, drill cores with similar collar locations (fanned drill holes) show significantly different mean and median concentration for some elements. This suggests that spatial anisotropic parameters control the spatial distribution and structural variation of the elements within the orebody.
ADE
CHAPTER 4 The Relationship of Magnetic Pyrrhotite with the Orebody of the Western Mineralisation 4.1 Introduction The Western Mineralisation contains pyrrhotite. It is the only magnetic mineral in the Western Mineralisation. There has been no quantitative mineralogical and statistical approach on the magnetic properties and their variation within the Broken Hill orebodies. A knowledge of the relationship between magnetic pyrrhotite and the orebody will be useful for the development of the Western Mineralisation. It would be useful if a geophysical interpretation could be integrated with geochemical information to improve ore targeting. One way in which geophysical methods could be incorporated would be via aeromagnetic surveys, which are useful to outline an orebody’s magnetic properties (Clark 1997).The resulting aeromagnetic maps can also be applied to ore mineral tracking and drill targeting at a scale suitable for exploration . However, a high-resolution aeromagnetic survey has never been conducted over the Western Mineralisation owing to the proximity of the City of Broken Hill and the large amount of magnetic noise above the Western Mineralisation (e.g. steel structures, galvanised iron buildings, pipes, railway lines and power lines). Regional magnetic surveys in the Broken Hill area have shown the stratigraphic distribution of magnetite-bearing amphibolite, felsic rocks and metasediments, minor thin discontinuous banded iron formations comprising quartz-magnetite-spessartine-fluorapatite and the Broken Hill orebody (Figure 4.1). Godber and Bishop (2006) recommended a ground magnetic survey in order to understand the detailed magnetic properties of the lithologies that contain magnetic minerals such as pyrrhotite and magnetite. The existing drill cores provide an opportunity to conduct such a survey.
ADE
Chapter 4-The Relationship of Magnetic Pyrrhotite 107 The data set was divided into two groups and they are productive samples (galena+sphalerite ≥ 1 vol. %) and barren samples (galena+sphalerite = 0). Statistical tests were performed in each group to examine the hypothesis whether the magnetic pyrrhotite is enriched significantly in the productive units within the Western Mineralisation. The correlation coefficient was calculated to determine the extent of the relationship between the variations of volume percentage of pyrrhotite, its magnetic susceptibility and the volume percentage of galena+sphalerite. Contour plots and coefficient of variation (COV) and core logging models were constructed in order to evaluate the variation of ore sulphide minerals present with the variations of pyrrhotite abundance and its magnetic susceptibility. In view of the lack of detailed magnetic surveys of the Western Mineralisation, this magnetic susceptibility investigation and the quantification of pyrrhotite enable us to understand better the subsurface pyrrhotite distribution, different types of pyrrhotite and magnetic variation within the Western Mineralisation. The spatial distribution of magnetic susceptibility and pyrrhotite are discussed in Chapters 6 and 8. 4.2 Characteristics of pyrrhotite Pyrrhotite has a variable composition of Fe S that 0 ≤ x ≤ 0.13 (De Villiers and (1-x) Liles, 2010) and has monoclinic, hexagonal and orthorhombic polytypes. In nature, pyrrhotite occurs in various superstructures (superlattice) that are represented by the axial lengths of NiAs-type unit cell of "A" and "C" that they are equal to 3.44 Å1 and 5.70 Å respectively. However, different structures of the pyrrhotite group are presented based on Fe "C" (Table 4.1). Pyrrhotite has a variable ratio (Table 4.1). The pure FeS iron rich S member of pyrrhotite is troilite (hexagonal). Arnold (1967) stated that natural pyrrhotite is commonly a mixture of monocline and hexagonal polytypes. 1 Ångstrom
ADE
108 Chapter 4-The Relationship of Magnetic Pyrrhotite Table 4.1: Atomic percentage of Fe in pyrrhotite polytypes ( from Carpenter & Bailey 1973; Clark 1997; Kontny et al. 2000; Kruse & Ericsson 1988; Yund & Hall 1969). Fe % , Superlattice Crystal system Name dimension (Chemical formula) 2C pyrrhotite, ~ 50.0, (FeS) Hexagonal 3A, 2C Troilite 4C pyrrhotite, Magnetic pyrrhotite, ~ 46.67- (Fe S ) Monoclinic 2 3 A, 2A, 4C 7 8 Ferromagnetic Weiss-type pyrrhotite ~ 47.37- (~Fe S ) Hexagonal 2A, 5C 5C pyrrhotite 9 10 ~ 47.83- (~Fe S ) Pseudohexagonal 2A, 6C 6C pyrrhotite 11 12 11C pyrrhotite ~ 47.6- (~Fe S ) Orthorhombic 2A, 2B, 11C 10 11 (a mixture of 5C and 6C) 47.447.8 Orthorhombic 2A, 2B, nC nC pyrrhotite or Monoclinic (Fe S Fe S ) 4.8n  6 9 10 11 12 Monocline pyrrhotite is the only ferromagnetic pyrrhotite with an approximate crystallographic structure of Fe S (Zapletal 1992). It has a 4C superstructure. The 7 8 magnetic property of monocline pyrrhotite is related to cation vacancies in its crystal structure. The vacancies decrease the overall crystal symmetry. Hence, the monoclinic pyrrhotite commonly contains more defects than hexagonal forms and is therefore more magnetic (Kontny et al. 2000). At temperatures above ~320° Celsius (Centigrade) pyrrhotite loses its magnetism and, under suitable oxygen fugacities, may convert to magnetite (Clark 1997; Rochette et al. 1990). The occurrences of pyrrhotite can be considered as an important indicator of redox and temperature in metamorphic, magmatic and diagenetic rocks (Rochette et al. 1990). Hexagonal and other antiferromagnetic forms of pyrrhotite produce little magnetic susceptibility in rocks (Dekkers 1988) and they cannot carry remnant magnetism (Zapletal 1992). However, monoclinic pyrrhotite is important mineral for identification of remnant magnetisation and susceptibility anisotropy during geological times (Clark 1997) and it can
ADE
Chapter 4-The Relationship of Magnetic Pyrrhotite 109 provide information regarding Palaeomagnetic characteristics, thermal evolution of the ancient magnetisation and the genesis of pyrrhotite. 4.3 Pyrrhotite in the Broken Hill Domain Scott, Both and Kissin (1977) and Bryndzia, Scott and Spry (1988) suggested that hexagonal pyrrhotite in the Broken Hill orebody is a primary iron sulphide mineral formed during high-grade metamorphism, whereas monoclinic pyrrhotite is a retrograde or post- metamorphic product of primary hexagonal pyrrhotite. This indicates that crystallography of pyrrhotite (changing from non-magnetic to magnetic pyrrhotite) was influenced by the metamorphic evolution of the Broken Hill deposit. Experimental studies of Scott, Both & Kissin (1977) showed that primary iron-rich pyrrhotite formed at high temperature within metamorphic rocks of the Broken Hill Domain and the Broken Hill orebodies. It is hexagonal pyrrhotite with exsolved troilite. During cooling metamorphosed rocks, the primary pyrrhotite inverted to hexagonal pyrrhotite + monoclinic pyrrhotite and further inverted during retrogression to monoclinic pyrrhotite + pyrite (Scott, Both & Kissin 1977, p.1415; Figure 4.2). In nature, the conversion of hexagonal pyrrhotite to monocline pyrrhotite requires losing Fe and gaining S. Scott, Both and Kissin (1977) suggested that in the Broken Hill deposit, the reaction of pyrrhotite and sphalerite during syn- to post- metamorphism events resulted in releasing Fe and S from hexagonal pyrrhotite and sphalerite respectively. During the chemical interaction and metamorphic evolution, the hexagonal pyrrhotite was converted to monocline pyrrhotite while losing Fe and gaining S from sphalerite (Scott, Both & Kissin 1977). In the more advanced oxidation stages, the loss of iron from hexagonal pyrrhotite may result in the occurrence of pyrite, marcasite and finally hematite or magnetite (Scott, Both & Kissin 1977). Plimer (1977) stated that the presence of primary pyrite within metapelites and metapsammites of several areas of the Broken Hill Domain (e.g. Stirling Vale, Pyrite Hill and Thackaringa) indicates that pyrite is stable at the maximum metamorphic grade of granulite facies and it did not convert to pyrrhotite. This suggests that the primary pyrrhotite and troilite in the Broken Hill orebody formed because there
ADE
Chapter 4-The Relationship of Magnetic Pyrrhotite 111 5. Pyrrhotite replacement by chlorite which has resulted in decomposition of pyrrhotite to marcasite. 4.3.1 Evaluation of chemical properties of pyrrhotite among the Broken Hill orebodies There are several previous sets of EMPA analyses of pyrrhotite from the Western Mineralisation and other Broken Hill orebodies (Table 4.2). There has been no study to compare and contrast the chemical variation of pyrrhotite in the Broken Hill orebodies. In this chapter, the data sets of others on pyrrhotite chemistry are used (Table 4.2). One method was the visualisation of variation of the average atomic percent of Fe and S for pyrrhotite samples from different orebodies (Figures 4.3 and 4.4). Another useful method is correspondence analysis2 that categorises the orebodies based on major and minor elements of pyrrhotite samples and visualise their inter-relationships on two-dimensional maps (Figures 4.5 to 4.7). A three-dimensional model of association of different chemical composition of pyrrhotite was interpreted by three correspondence maps2 Table 4.2: EMPA analyses of pyrrhotite from the Western Mineralisation and the Line of Lode from CML7. Reference Orebody Number of pyrrhotite samples and drill holes Kitchen (2001) The Western Mineralisation 10 samples from 4001 and 4002 Patchett (2003) The Western Mineralisation 34 samples from 4003 47 samples from C Lode and 27 samples from Sproal (2001) C Lode and 2 Lens 2 Lens A and B Lodes, 1 Lens and 1 5 a n d 16 samples from A and B Lodes Tully (2002) respectively, 5 samples from 1 Lens and 5 Kintore Pit samples from Kintore Pit Groombridge 2 and 3 Lenses 3 and 4 samples from 2 and 3 Lenses (2003) respectively and 5 samples from south of 2 Lens 2 Lens South 4.3.1.1 Variation of Fe and S in pyrrhotite samples of the Broken Hill orebodies In Figures 4.3 and 4.4, individual pyrrhotite samples were marked by red points and the blue rhombic points represent the mean atomic percent of Fe and S. In Figure 4.3, some pyrrhotite samples of B and C Lodes and the Western Mineralisation have atomic percent 2 See the explanations of Table 4.6
ADE
Chapter 4-The Relationship of Magnetic Pyrrhotite 113 4.3.1.2 Correspondence analysis for pyrrhotite samples of the Broken Hill orebodies There are many references about mathematical theory and application of correspondence analysis e.g. Adachi (2003), Benzercri (1992), Clausen (1998), Ender (2010), Greenacre (1984, 2007), SAS Institute Inc. (1999), Social Research Update (1995), Teil and Cheminee (1975), Valenchon (1982) and Van de Geer (1993a, 1993b) but the major aim of correspondence analysis in this study is to show inter-relationship of elements of pyrrhotite samples with the Broken Hill orebodies on two-dimensional correspondence maps. Correspondence analysis is an exploratory method and non- parametric technique that has no distribution assumptions, but it needs positive data values. It is a geometric method to represent the inter-relationships of the rows and columns of a two-way contingency table (Table 4.3) in a two-dimensional map (e.g. Figure 4.4) or a low-dimensional space. Table 4.3 shows categorical variables in a matrix format and in this Table, Fe and S comprise the major average atomic percent of pyrrhotite and the other 8 elements incorporate as minor element in atomic structure of pyrrhotite. There are significant differences between average values of major and minor elements in Table 4.3. These differences create difficulties for correspondence analysis because the variable with the greatest variance will produce the highest influence on the outcome. One way for avoiding this problem is performing correspondence analysis only with minor elements of pyrrhotite (all elements of Table 4.3 without Fe and S) and another way is using standardisation that the most universal method of standardisation is z-transformation. Table 4.3: A contingency table of the average percentage of elements in pyrrhotite samples within the Broken Hill orebodies. Element S Fe Cu Zn As Ag Cd Sb Pb Bi Orebody A Lode 52.38 47.3 0.008 0.011 0.085 0.006 0.009 0.007 0.005 0.18 B Lode 51.93 47.62 0.011 0.288 0.087 0.007 0.004 0.004 0.003 0.022 C Lode 52.79 46.75 0.0231 0.035 0.076 0.009 0.006 0.003 0.094 0.016 1 Lens 52.12 47.60 0.002 0.079 0.096 0.003 0.026 0.003 0.003 0.005 2 Lens 52.72 47.00 0.0535 0.074 0.089 0.011 0.004 0.003 0.001 0.007 3 Lens 52.97 46.58 0.0192 0.313 0.083 0.006 0.007 0.002 0.007 0.004 Western 52.27 47.32 0.0427 0.196 0.062 0.008 0.008 0.004 0.009 0.021 Mineralisation
ADE
114 Chapter 4-The Relationship of Magnetic Pyrrhotite (x) The z-transformation is calculated by subtraction of the mean value of each column from each value (X) of the column and the resulting value of the column is divided by the standard deviation (S) of the column [Equation. (4.1)]. x  x z  (4.1) s The z-transformation of values of Table 4.3 produces some negative values that could not be analysed by correspondence analysis. It requires the addition of a minimum constant value (e.g. 2) to all the values of Table 4.4 to change them to positive values (Table 4.5). Although, with small changes of the constant value (e.g. addition of 4 instead of 2 to values of Table 4.4), the relative distance of points3 will change inside a correspondence map but the overall inter-relationships of the points will almost remain comparable. In this case, if a great constant value is selected (e.g. 20 instead of 2), the correspondence map will change significantly. Some statistical terms are broadly used in correspondence analysis are explained in Table 4.6. 3 See the explanations of Table 4.6
ADE
116 Chapter 4-The Relationship of Magnetic Pyrrhotite Table 4.6: A summary of statistical terms used in correspondence analysis. Points Demonstration of elements and orebodies of Table 4.3 in a correspondence map. In Table 4.5, pyrrhotite samples collected from each orebody are associated with Dimension 10 element concentrations. Therefore, each orebody can be defined by variation of the 10 elements in a 10-dimensional space. Scores show the coordinates of points in a correspondence map and each point is Scores in determined by a score in relation to the scale of each dimension of a dimension correspondence map. V a r i a nce in correspondence analysis is called inertia. Inertia is "the weighted Inertia sum of chi-square distance between each profile and the mean profile" (Ender 2010, p. 1). Mass is the marginal proportion of the row and column variables that is used to weight the point profiles for calculation of point distance. So that the sum of all Mass table entries is equal to 1.0 (StatSoft Electronic Statistics Textbook 2010) and the row and column values are standardised for producing a correspondence map. For example, in Table 4.5, each column total value will be divided by the total of sum value for the columns, i.e., 140. Determination of distances between the points provides all information about all similarities among them. In correspondence analysis, the distances between Chi-square points are measured by chi-square method rather than the Euclidian distance. method The resulting distance matrix is used as entry data of the principal component analysis1. The chi-square (Friendly 1995) is a weighted profile distance, where the weight is the mass of the row or column values (Table 4.7). T h e s u m of all the eigenvalues is named the total inertia or the total variance The total inertia explained by the dimension. The total inertia is calculated by the total chi-square value (29.952 in Table 4.7) divided by the total of the sum (140 in Table 4.5). The correspondence analysis is a method for decomposing the total inertia in Correspondence order to identify a lower-dimensional space for any given points. This analysis analysis makes easier interpretation of the internal relationships of the points. Inertia of each dimension Percentage of In Table 4.8, PPI is calculated by 100. For example, Total inertia proportional the proportional inertia of the first dimension in Table 4.8 accounts for 37.82 % inertia (PPI) of the total inertia (21.39 %). Visual presentation of the contribution of points to dimensions and contribution of the dimensions to categorisation of the points and detection of their internal- relationships. Each correspondence map accounts for part of the total variation of the points within two dimensional spaces (Figures 4.5 and 4.7). Correspondence For measuring distances of row and column values, a symmetrical map normalisation was used in this study because the principal coordinate of row and column values have slightly different scale. In a correspondence map, relative distance of points from each other and the scores of points are important parameters for interpretation (Section 4.3.1.3). 1 See the explanation of the principal component analysis in Section 5.7
ADE
118 Chapter 4-The Relationship of Magnetic Pyrrhotite 4.3.1.3 Correspondence maps for pyrrhotite samples Figures 4.5 to 4.7 display symmetrical plots of the elements of pyrrhotite samples and the Broken Hill orebodies. Figure 4.5 shows 62.75 %5 of the total chemical variation of pyrrhotite samples within the Broken Hill orebodies. In Figure 4.5, dimension 1 separates the plane from its zero value into two parts, positive and negative. C Lode, 2 Lens, S, Ag and Cu lie at the extreme end of the positive scale and 1 Lens, Fe and Cd are at the extreme end of the negative scale of dimension 1. Dimension 2 separates the plane from its zero value into two parts, positive and negative. Bi, Sb and the Western Mineralisation lie at the extreme end of the negative scale and 1 and 3 Lenses, As, Cd and S lie at the extreme end of the positive scale. Both the first and second dimensions discriminates Fe-rich pyrrhotite from S-rich pyrrhotite. Figure 4.5: Correspondence map in relation to dimensions 1 and 2 for pyrrhotite. (* The Western Mineralisation) In Figure 4.6, the correspondence map of dimensions 1 and 3 jointly accounts for 52.77 % of the total chemical variation of pyrrhotite within the Broken Hill orebodies. Dimension 3 separates C Lode and Pb-rich pyrrhotite from Zn- rich pyrrhotite. 5 Calculated by PPI (dimension 1) + PPI (dimension 2) from Table 4.7
ADE
120 Chapter 4-The Relationship of Magnetic Pyrrhotite Together Figures 4.5 to 4.7 provide a three-dimensional model6 that shows 77.69 % 7 of total variation of pyrrhotite within the Broken Hill orebodies and their results are outlined below:  A Lode contains Sb-rich pyrrhotite samples,  B Lode contains Fe-rich pyrrhotite samples,  C Lode contains Pb-rich pyrrhotite samples,  1 Lens contains Cd-rich pyrrhotite samples,  2 Lens contains Cu- and Ag-rich pyrrhotite samples,  3 Lens contains S-rich pyrrhotite samples, and  The Western Mineralisation contains Bi- and Sb-rich pyrrhotite samples. 4.4 Magnetic susceptibility measurements in the Western Mineralisation In order to interpret the down-hole variation of magnetic susceptibility in conjunction with variation of volume percentage of galena, sphalerite and pyrrhotite in the Western Mineralisation, a total of 54 exploration drill cores were investigated. These magnetic measurements were taken in CBH Resources’ core shed aiming to systematically quantify the magnetic properties in 1,928 samples of the surface drill cores. The split cores comprised the sulphide mineralised areas and their enclosing non-mineralised rocks. The magnetic susceptibility is the intrinsic property that depends on the magnetic mineral content and also on the size and orientation of magnetic particles present in a sample. The magnetic susceptibility does not however depend on the strength of the geomagnetic field in which lithologies occurred. The hand held magnetic susceptibility metre model Digital Fugro GMS-2 (Figure 4.8) was used in this study. The GMS-2 is a high sensitivity portable instrument with a detection range between 1×10-5 SI to 10 SI values and a resolution of 1×10-5 SI value. Therefore, values less than 1 × 10-5 SI are outside the detection range of the GMS-2 and are automatically assigned a value of 1 × 10-5 SI. The frequency of 760 Hz in this Model indicates that this instrument is not susceptible to the effects of high conductivity in some samples. When high magnetic susceptibility was shown, the instrument was frequently zeroed to avoid drift. The magnetic survey took place in a temperature range of about 10º-15° Celsius. 6 A combination of three correspondence maps in relation to dimensions 1, 2 and 3 7 Calculated by PPI (dimension 1) + PPI (dimension 2) + PPI (dimension 3) from Table 4.7
ADE
Chapter 4-The Relationship of Magnetic Pyrrhotite 121 Magnetic susceptibility has been measured to derive the range of magnetic degrees for different rock types and the associated sulphide mineral assemblage exhibit at intervals of one metre. The magnetic susceptibility measurements should show variation of the magnetic properties and magnetic intensity from edge to the centre of the Western Mineralisation. The following magnetic susceptibility values were measured in the Western Mineralisation: 4.4.1 Average magnetic susceptibility (AMS) Each 10 cm piece of one metre interval core sample was measured three times. The AMS values of the core were determined for the one metre core sample. It was expressed as AMS observed per metre and generally reflected lithological unit. 4.4.2 Maximum magnetic susceptibility (MMS) Owing to the presence of disseminated pyrrhotite in each one metre interval core sample, MMS was measured at one point in each one metre core sample. It was expressed as MMS observed for a particular point in each sample. Figure 4.8: A Magnetic Susceptibility Metre model Digital Fugro GMS-2.
ADE
122 Chapter 4-The Relationship of Magnetic Pyrrhotite 4.5 Descriptive statistics for magnetic susceptibility, pyrrhotite, Pb+Zn and galena+sphalerite In Table 4.9, maximum values of AMS and MMS are in the range of 10-3 SI8 (e.g. 5500 × 10-5 SI = 5.5 × 10-3 SI) and are associated with paramagnetic minerals such as olivines, pyroxenes with positive susceptibility maximum of approximately 10-3 SI value (Clark 1997). The maximum AMS value of 1500 × 10-5 was measured in drill core 4010 at the depth of 256 m which contained 10 vol. % of pyrrhotite in a blue quartz lode. The MMS values of samples were up to 5500 × 10-5 or 0.055 SI. As mentioned earlier, the 1,928 investigated samples of the Western Mineralisation were divided into two groups (Table 4.10) and their statistical results are given in Table 4.11. Table 4.11 shows that the mean values of pyrrhotite, AMS and MMS within productive samples are much greater than the barren samples. This indicates that the mean volume percentage of magnetic pyrrhotite is different between the two groups of samples in Table 4.10. Statistical tests can then be used to test if the mean values of pyrrhotite and MMS of barren and productive samples are indeed coming from two distinct distributions. The differences of the two mean values of pyrrhotite and MMS between the barren and productive samples may be significant or may be by chance. In statistics, when a result is called significance, it is not possible to have occurred by chance. This issue can be evaluated by further statistical tests. 4.5.1 Statistical tests Depends on the distribution of data, statistical tests can be carried out by parametric or non-parametric analysis. In parametric approach, data is supposed to come from a normal distribution and parameters of the distribution can be inferred by parametric statistics. In contrast, the non-parametric approach has no assumption about the distribution and it is less sensitive to the effects of outliers. 8 International System of Units (metric system)
ADE
Chapter 4-The Relationship of Magnetic Pyrrhotite 125 Figure 4.10: Probability graphs for logarithmic data of MMS, AMS, pyrrhotite and galena+sphalerite. The red curves show the experimental logarithmic values of the variables against their respective estimated percentage of cumulative probability and the straight blue lines show the theoretical percentage of cumulative probability for near-normal distribution of the variables. Statistical tests give geologists a tool to make a quantitative assessment of statistical significant differences of two sample groups. If there is a significant difference between the two sample groups, it would confirm that the magnetic pyrrhotite or magnetic property can be considered as a geophysical indicator for tracking galena and sphalerite in the Western Mineralisation. The aim is to assess whether there is enough evidence to accept the null hypothesis about two sample groups or to reject the null hypothesis (alternative hypothesis). 4.5.1.1 Hypothesis of the non-parametric statistical tests 1. The null hypothesis is that the median values of AMS, MMS and pyrrhotite are the same for both barren and productive samples, and 2. Alternative hypothesis is that the median values of AMS, MMS and pyrrhotite differ between barren and productive samples.
ADE
126 Chapter 4-The Relationship of Magnetic Pyrrhotite 4.5.1.2 P value or significance level (1-tailed or 2-tailed) 1. If the P value of data < 0.01, it will indicate that the data provides statistically significant evidence of a difference between the barren group and the productive group with 99 % confidence level. 2. If the P value of data > 0.01, it will conclude that the data does not provide statistically significant evidence of a difference between the barren group and the productive groups with 99 % confidence level. In this section, only the outcomes of the non-parametric statistical tests are presented. A comprehensive description of the mechanism of the statistical tests for independent groups is in the support section of the SPSS (SPSS Inc. 2009) software package. 4.5.2 Wilcoxon Mann-Whitney U Test In the case of independent groups, the Mann-Whitney U Test (Wilcoxon, 1945) is the most widely-used significance test for comparing two independent samples based on locations of the ranks and this test can evaluate whether two independent samples relate to the same distribution (the null hypothesis) or not (the alternative hypothesis). In the Mann- Whitney Test (Tables 4.12 and 4.13), significant values (2-tailed) and Mont Carlo Sig.9 are less than 0.01. This indicates that AMS, MMS values and volume percentage of pyrrhotite differ significantly between the two groups of productive and barren in the Western Mineralisation (at the 99 % confidence level). Table 4.12: The mean rank of AMS, MMS and pyrrhotite between the barren and the productive groups using the Mann-Whitney Test. Mann-Whitney Test Group N* Mean Rank Sum of Ranks B a r r en 682 856.88 583533.5 AMS Productive 1246 1022.55 1274094.5 Barren 682 756.50 515174.5 MMS Productive 1246 1077.41 1342453.5 Barren 207 437.54 90571.5 Pyrrhotite Productive 790 515.10 406931.5 *Number of samples 9 Mont Carlo Significance
ADE
Chapter 4-The Relationship of Magnetic Pyrrhotite 127 Table 4.13: The results of the Mann-Whitney Test. Test Statistics AMS MMS Pyrrhotite Mann-Whitney U 351312.5 282953.5 69043.5 Wilcoxon W 583533.5 515174.5 90571.5 Z -6.768 -12.228 -3.543 Asymp. Sig. (2-tailed) 0.000 0.000 0.000 Monte Carlo Lower Bound 0.000* 0.000* 0.000* Sig. (2-tailed) 99 % confidence level 0.000 0.000 0.000 Upper Bound 0.000 0.000 0.001 * Based on 10,000 sampled tables with starting seed 2,000,000 4.5.3 Kolmogorov-Smirnov Z Test The Kolmogorov-Smirnov Z test (Table 4.14) detects differences in both the locations and shapes of the distributions. This test measures the maximum absolute difference between the observed cumulative distribution functions in both groups. When the differences are large enough to be significant, the two distributions are considered different. In Table 4.14, the 99 % confidence level is smaller than 0.01 in the lower and upper bound that indicate the two groups differ in either shape or location. Table 4.14: Two-Sample Kolmogorov-Smirnov Frequencies Test. Two-Sample AMS MMS Pyrrhotite Kolmogorov-Smirnov Test Most Extreme Absolute Differences 0.147 0.295 0.157 Positive 0.000 0.000 0.010 Negative -0.147 -0.295 -0.157 Kolmogorov-Smirnov Z 3.095 6.196 2.006 Asymp. Sig. (2-tailed) 0.000 0.000 0.001 Monte Carlo Sig. (2-tailed) 0.000* 0.000* 0.000* 99 % confidence Lower Bound 0.000 0.000 0.000 level Upper Bound 0.000 0.000 0.000 * Based on 10,000 sampled tables with starting seed 2,000,000 4.5.4 Moses Test of Extreme Reactions The Moses Test is used in experimental studies to assess whether the experimental variable affects subjects in either a positive or a negative direction, creating a polarising effect. The Moses Test compares extreme responses of one sample with another sample obtained from a control group (in this study, barren group). If the probability associated
ADE
128 Chapter 4-The Relationship of Magnetic Pyrrhotite with the Moses test is less than the desired significance level (in this study P < 0.01) then it can be concluded that the two samples differ. In general, the Kolmogorov-Smirnov Z Test check the middle of the distribution for differences in central tendency and do not take account of the extreme low and high values. So it is possible when two different distributions (e.g. one normal and another polarised at the two extreme) are compared with each other and show similar central tendency that might not be found to be significantly different by the Mann-Whitney U Test and the Kolmogorov-Smirnov Z Test, while the Moses Test checks the tails of the distribution for differences in extreme tendencies (Table 4.15). Table 4.15: Moses Extreme Reactions Test Frequencies at the 99 % confidence level. Moses Test Group N Barren (Control) 681 AMS Productive (Experimental) 1246 Barren (Control) 681 MMS Productive (Experimental) 1246 Barren (Control) 207 Pyrrhotite Productive (Experimental) 790 Test Statistics AMS MMS Pyrrhotite Observed Control Group Span 1418 1661 827 Sig.(1-tailed) 0.000 0.000 0.000 Trimmed Control Group Span 1212 1487 771 Sig.(1-tailed) 0.000 0.000 0.000 Outliers Trimmed from each end 34 34 10 In Table 4.15, the barren group was defined as the control group. P values or Sig.10 (1-tailed) for AMS, MMS and pyrrhotite are below 0.01, thus indicating that the two groups of barren and productive units differ significantly at 99 % confidence level. The significance levels (1-tailed) in the trimmed group are the same as the observed group and this confirms that even if the outlier samples are removed, these two groups of barren and productive areas differ with respect to their AMS, MMS and pyrrhotite values. According to the results of non-parametric statistical tests, there is enough evidence to suggest that pyrrhotite abundance and AMS and MMS values are significantly different between productive and barren samples in the Western Mineralisation. 4.6 Bar diagrams of magnetic susceptibility and pyrrhotite In bar diagram of AMS (Figure 4.11), the drill cores 4028 and 4052 have higher mean and median values relative to other drill cores and this indicates higher concentration 10 Significance level
ADE
Chapter 4-The Relationship of Magnetic Pyrrhotite 129 of magnetic minerals in these drill cores. When considering HiP value, drill cores 4010 and 4028 show maximum AMS variation. In Figure 4.11, when considering HiP value, the high variations of MMS are present within drill cores 4010, 4014 and 4031. Drill cores 4001, 4028, 4033 and 4052 reveal high mean and median values of MMS. The trends of mean variations of AMS and MMS are roughly similar (Figure 4.11). In Figure 4.11, drill cores 4028, 4033, 4045 and 4052 have high mean and median volume percentage of pyrrhotite within the surface drill cores. The greatest variation of pyrrhotite is present in drill core 4050 followed by drill core 4030. The mean and median trend variations of pyrrhotite show similarity in a few drill cores with AMS and MMS values. For example, the mean and median values of AMS, MMS and pyrrhotite are high in drill cores 4028, 4033 and 4052 thus indicating an increase of magnetic pyrrhotite abundance within the drill cores. It is evident from Figure 4.12 that every drill core shows distinct variation for AMS, MMS, specific gravity, galena+sphalerite and Pb+Zn. There are significant similarities between the following variations:  MMS and pyrrhotite, and  Pb+Zn and galena+sphalerite. In Figure 4.12, there are several similar trends of COV for AMS, MMS, specific gravity, galena+sphalerite and Pb+Zn in drill holes: 3231, 4010, 4011, 4012, 4015, 4026, 4028, 4031, 4032, 4037, 4038, 4039, 4040, 4052 and 4054.
ADE
Chapter 4-The Relationship of Magnetic Pyrrhotite 133 4.7.1 Interpretation of the correlation coefficients In Tables 4.16 and 4.17, the significance level (2-tailed) between AMS and galena+sphalerite is more than 0.01 and it implies that there is less than (1- 0.01) or 99 % certainty for chance of being a true correlation coefficient. The non-parametric correlations in Tables 4.16 and 4.17 show 64.8 % and 50.7 % correlation coefficients respectively between pyrrhotite (vol. %) and MMS (SI). In Tables 4.16 and 4.17, the result of correlation coefficient of MMS and pyrrhotite with galena+sphalerite show values close to each other. For example, in the SCC (Table 4.16), there is a 20.2 % correlation coefficient between MMS and galena+sphalerite versus 21.7 % correlation coefficient between pyrrhotite and galena+sphalerite. This indicates the relationship of high magnetic pyrrhotite with the galena and sphalerite. 4.8 Appraisal of internal relationships among magnetic susceptibility, pyrrhotite, galena+sphalerite and Pb+Zn relative to depth There are major cases of internal relationships among variation of galena+sphalerite, Pb+Zn, pyrrhotite, AMS, MMS, specific gravity relative to depth among the 54 drill cores of the Western Mineralisation (Figures 4.13 to 4.17). The results of observed internal relationships among the variables outline below: 1. The MMS values are anomalous in the presence of magnetic pyrrhotite and the trend of intensity of AMS values is partially analogous with MMS signatures, 2. The down hole variation of MMS values, pyrrhotite abundance and specific gravity values intensify within galena+sphalerite and Pb+Zn. It is unlikely that their marked variations may be directly attributed to lithology variations, 3. The range of MMS values and volume percentage of pyrrhotite vary from 2×10-3 SI to 20×10-3 SI and from 3 to12 (vol. %) respectively within galena+sphalerite. The range of specific gravity changes from 2 to 4, and 4. The MMS values and volume percentage of pyrrhotite display considerable fluctuation in places where the sample contains rich sulphide ore. In more detail, the following
ADE
134 Chapter 4-The Relationship of Magnetic Pyrrhotite compatibilities are shown among MMS values, pyrrhotite anomalies and galena+sphalerite content. 4.1. In most core logs, the commencement and termination of the signals of MMS and pyrrhotite anomalies roughly are consistent with the presence of sulphides (Figures 4.13 and 4.14), 4.2. In a few places, there is a partial shift or complete delay phase between the start and termination of one anomaly of MMS against the presence of sulphide ore. For example, in drill core 4014 (Figure 4.15), the anomaly of sulphide ore begins to appear immediately after termination of the MMS and pyrrhotite anomalies. Drill core 4005 (Figure 4.16) shows a delay phase and just part of the anomaly of MMS and the presence of sulphide ore overlap each other, and 4.3. In some drill cores, for example, 4044 (Figure 4.17), there are some depths where MMS anomalies are appeared in the absence of sulphide ore anomaly and vice versa. The displacement and absence of sulphide ore (galena+sphalerite) against MMS and pyrrhotite anomalies may be because pyrrhotite is ductile and during the Olarian Orogeny may have been remobilised out of sulphide rocks into the enclosing silicate rocks. Textures in Figures 4.13 to 4.17 are associated to all sulphide minerals including pyrrhotite. However, among the investigated samples of the Western Mineralisation, pyrrhotite was present in various macroscopic textures such as stringer, disseminated, brecciated, network, massive, crystalline and veinlets. Where adjacent to galena and sphalerite, pyrrhotite is mostly disseminated, laminated and stringer, but further away from galena and sphalerite the dominant textures are massive and network. Moreover, to some extent the fine grained pyrrhotite reveals higher magnetic susceptibility and anisotropy (changing of magnetic intensity at different directions) than pyrrhotite with massive and network textures. The quantitative core log diagrams of this chapter have been provided in supplementary files to this thesis.
ADE
140 Chapter 4-The Relationship of Magnetic Pyrrhotite 4.9 Contour plots Contour plots are a useful tool for displaying anomalies. For example, contour plots of galena+sphalerite and Pb+Zn in the parameter space of AMS and pyrrhotite (Figures 4.18a and 4.18b) show an anomaly of galena+sphalerite and Pb+Zn that appears between 2 to 12 volume percent of pyrrhotite and between 1×10-3 SI and 6×10-3 SI value of AMS. In Figure 4.18a, another anomaly occurs between 12 and 19 volume percent of pyrrhotite and measures that are higher than 4 ×10-3 SI values for AMS. In Figure 4.18c, the major anomalies of galena+sphalerite appear between 2 and 15 volume percent of pyrrhotite and between 3×10-3 SI and 20×10-3 SI of MMS values. In Figure 4.18d, the anomalies of Pb+Zn appear between volume percent of 4 and 12 of pyrrhotite and for MMS between 4×10-3 SI to 16×10-3 SI. In Figure 4.18d, another anomaly of Pb+Zn appears in the range of more than 35 volume percent pyrrhotite and the dominant MMS is between 0.01×10-3 SI and 27×10-3 SI. This second anomaly occurs in a few samples that are not covered by the major data set.
ADE
142 Chapter 4-The Relationship of Magnetic Pyrrhotite 4.10 Summary This chapter investigated variation of magnetic pyrrhotite and its relationships with galena and sphalerite and tried to understand whether magnetic susceptibility and occurrences of pyrrhotite can be used as an exploration tool for tracking ore in the Western Mineralisation. In order to interpret the down-hole variation of magnetic susceptibility in conjunction with variation of galena (vol. %), sphalerite (vol. %) and pyrrhotite (vol. %), a total 1,928 samples of 54 exploration drill cores were investigated. The data set was then divided into two groups, productive and barren, based on the quantity of summation of galena and sphalerite. The outcomes of this study show that samples containing galena or sphalerite have considerably more magnetic pyrrhotite than the samples without galena and sphalerite. Thus, an understanding of the variation of pyrrhotite abundance and the geological significance of its magnetic patterns provide a link between the mineralisation and the geometrical configuration of the orebody. Simple correspondence analysis performed in this study to show a picture of internal relationships between chemical composition of pyrrhotite samples and the Broken Hill orebodies. The result of 3D correspondence map shows that pyrrhotite samples of the Western Mineralisation contain higher Bi and Sb in comparison with the other Broken Hill orebodies. Previous collected pyrrhotite samples from the Western Mineralisation shows chemical formula of either magnetic pyrrhotite or non-magnetic pyrrhotite. The results of the COVs show significant similarities between variations of pairs of MMS and pyrrhotite, Pb+Zn and galena+sphalerite. The statistical tests of this chapter suggest that pyrrhotite abundance, AMS and MMS values are significantly difference between productive and barren samples of the Western Mineralisation. The results of quantitative core logging suggest in most cases, the commencement and termination of the signals of MMS and pyrrhotite anomalies in depth roughly is consistent with the presence of sulphide ore anomalies. However, a partial shift or complete delay between the start and termination of one anomaly of MMS against a presence of galena+sphalerite are seen in some drill cores and in some cases, there are
ADE
CHAPTER 5 Multi-Element Relationships: The Western Mineralisation 5.1 Introduction Unlike geophysical data, geochemical data is stochastic in nature and multivariate statistical analysis is in general used for their evaluation. Geochemical data often suffers from several shortcomings, such as the detection limit problem, constrained data range (Aitchison 1986; Le Maitre 1982), abnormal or multi-modal populations or strongly skewed distributions (pbarrett.net 2001) and the presence of outliers (Filzmoser, Reimann & Garrett 2005). Most classical statistics are based on assumptions that random variables are following Gaussian distributions (Reimann & Filzmoser 2000), independent and unconstrained in Euclidean real space. Multivariate statistical analysis can be used to develop geochemical exploration models and identify mineral paragenesis and mineral generations through: 1. The analysis of multi-element interactions, 2. The recognition of significant controlling factors, 3. The reduction of the geochemical dimensions, 4. The mass balance recognition, or 5. Statistical chemical equilibrium analysis There are abundant publications regarding the geochemical characteristics and the origin of the Broken Hill orebody based on interpretation of thermodynamic, sulphide melting, isotopic data, element concentrations and geological observations (e.g. Frost, Mavrogenes & Tomkins 2002; Mavrogenes et al. 2004; Phillips 1980; Phillips & Wall 1981; Ryall 1979; Spry, Plimer & Teale 2008; Stevens, Prins & Rozendaal 2005; Tomkins, Pattison & Frost 2007). However, the application of multivariate statistical analyses in relation to an orebody such as that at Broken Hill is still an area seriously under- researched. In this chapter, there is a description of a variety of bivariate and multivariate statistical analyses that were applied to detect and quantify the underlying structure of geochemical variations of sulphide minerals. The methods are useful for identification of the controlling chemical factors of sulphide ore formation in the framework of a multi- influence elemental model.
ADE
Chapter 5- Multi-Element Relationships 145 The detailed statistical analyses include: 1. Categorising chemical composition of galena and sphalerite samples of the Broken Hill orebodies in order to identify the inter-relationship between the orebodies and the chemical composition of the minerals, 2. Application of the compositional (closed) data analysis to make independent the constrained geochemical data sets created from 10 element concentrations (Section 5.3.2), 3. Calculation of correlation coefficients for pairs of element concentrations, 4. Appraisal of multi-element relationships and the underlying structures, and 5. Evaluation of the biplot of the average atomic percentage of elements in the galena and sphalerite samples in order to understand mineral generation. The statistical analysis in this study was conducted by the following software programs: 1. Minitab software for performing simple correspondence analysis, 2. SPSS software for performing principal component analysis (PCA), 3. Matlab software (The MathWorks Inc. 2006) for programming of a 3D biplot, and 4. CoDaPack3D (Thió-Henestrosa 2008) a freeware module for conducting compositional data analysis and construction of the 3D biplot. 5.2 The comparison of mineral chemistry of galena and sphalerite samples Galena (PbS) and sphalerite (ZnS) are the major sulphide minerals of the Broken Hill orebodies. The long term effect of deformation and metamorphism has caused galena and sphalerite to show different percentages of major elements (Pb in galena, Zn in sphalerite and S in both of them) and different minor substitution of Pb in sphalerite, Zn in galena, and different traces of Fe, Ag, Sb, Bi, Cd, Cu, Zn, Mn, Co and As in samples containing both minerals. In this section, the average atomic percentages of elements in galena and sphalerite samples collected by Kitchen (2001), Sproal (2001), Tully (2002), Groombridge (2003) and Patchett (2003) are compared (Tables 5.1 and 5.2). All these samples were analysed by the CAMECA 50X electron microprobe at The University of Melbourne.
ADE
Chapter 5- Multi-Element Relationships 147 The results of comparison of the Western Mineralisation with the other Broken Hill orebodies are outlined below using Tables 5.1 and 5.2: 1. The average atomic percentages of Bi, As, Fe, Sb and Zn in the galena samples of the Western Mineralisation are at least 69.4, 9, 6.75, 2.75 and 1.6 times those of the Broken Hill Lodes and Lenses respectively (Table 5.1), 2. The average atomic percentage of Pb in the sphalerite samples of the Western Mineralisation is at least 1.59 times that of the Broken Hill Lodes and Lenses (Table 5.2), and 3. The average atomic percentage of As in the sphalerite samples of the Broken Hill Lodes and Lenses is at least 1.27 times that of the Western Mineralisation (Table 5.2). 5.2.1 Correspondence analysis for galena samples within the Broken Hill orebodies The Broken Hill orebodies are categorised based on the major and minor elements of galena and sphalerite samples using simple correspondence analysis. This method was described earlier in Section 4.3.1.2. In this section similar procedures of correspondence analysis were performed on the data of Tables 5.1 and 5.2. In the data sets of Tables 5.1 and 5.2, there are significant differences between the average values of major and minor elements. This will produce significant differences between the variance of the major and minor elements that consequently will affect greatly the result of correspondence analysis. In order to solve this problem, the data sets of Tables 5.1 and 5.2 need to be standardised using a z-transformation [Equation (4.1)]. During the z-transformation of data of Table 5.1, some negative values were produced (Table 5.3) that could not be analysed by correspondence analysis, and therefore a minimum constant value of 2 was added to all the data of Table 5.3 to change them to positive values (Table 5.4). The reason for this was explained earlier in Section 4.3.1.2. The result of decomposition of the data set of Table 5.4 into six dimensions (components) is given in Table 5.5. In Table 5.4, galena samples collected from each orebody are associated with 10 element concentrations in a 10-dimensional space. The correspondence analysis decomposes the total variation of galena into a lower-dimensional space for the 10 element concentrations. This makes it easier to understand the internal relationships between the Broken Hill orebodies and the different chemical compositions of galena.
ADE
Chapter 5- Multi-Element Relationships 149 Table 5.5: The results of decomposition of Table 5.4 into six dimensions. Proportion of Percentage of proportional Percentage of Dimensions Inertia inertia inertia (PPI) cumulative inertia 1 0.0751 0.4693 46.93 46.93 2 0.0458 0.2864 28.64 75.57 3 0.0194 0.1210 12.10 87.68 4 0.0154 0.0961 9.61 97.28 5 0.0043 0.0267 2.67 99.95 6 0.0001 0.0005 0.05 100.00 Total 0.16 5.2.2 Correspondence maps for galena samples 5.2.2.1 Interpretation of Figure 5.1 Figure 5.1 shows 75.57 % 1 of the total chemical variation of galena within the Broken Hill orebodies. Figure 5.1: Correspondence map in relation to dimensions 1 and 2 for galena. (*The Western Mineralisation) 1 Dimension 1 In Figure 5.1, geochemical processes that can be inferred from dimension 1 separate the map at zero value into two parts, positive and negative. Pb- and Cd-rich galena, A, B Lodes and 1 Lens lie on the positive scale of dimension 1 and the Western Mineralisation (WestMin in map), C Lode, 3Lens, Bi-, Sb-, Fe-, Zn-, As- and S-rich galena lie on the negative scale of dimension 1. 1 Calculated by PPI (dimension 1) + PPI (dimension 2) from Table 5.5
ADE
150 Chapter 5- Multi-Element Relationships 2 Dimension 2 Dimension 2 separates 2, 3 Lenses (Pb Lenses) and Cu-, S- and Ag-rich galena from the Western Mineralisation, B, C Lodes, 1 Lens and Pb-, Bi-, Fe-, Zn- and As- rich galena samples. 3 Discriminator Dimension 1 acts as a good discriminator of Cd- and Pb- rich galena from the Western Mineralisation. Dimension 2 acts as a good separator of S-, Cu-, Ag-rich galena from the Western mineralisation 5.2.2.2 Interpretation of Figure 5.2 In Figure 5.2, the correspondence map of dimensions 1 and 3 jointly accounts for 59.03%1 of the total chemical variation of galena samples. 4 Dimension 3 Dimension 3 is a good discriminator for separating C Lode, 2 Lens, Zn- and S-rich galena from Ag-rich galena and A Lode. Figure 5.2: Correspondence map in relation to dimensions 1 and 3 for galena. 5.2.2.3 Interpretation of Figure 5.3 In Figure 5.3, the correspondence map of dimensions 2 and 3 accounts for 40.74% of the total chemical variation of galena. 1 Calculated by PPI (dimension 1) + PPI (dimension 3) from Table 5.5
ADE
Chapter 5- Multi-Element Relationships 151 Figure 5.3: Correspondence map in relation to dimensions 2 and 3 for galena. 5.2.2.4: Interpretation of a combination of Figures 5.1 to 5.3 Together Figures 5.1 to 5.3 provide a three-dimensional1 model that explains 87.68 %2 of the total chemical variation of galena within the Broken Hill orebodies and their inter-relationships are outlined below:  A Lode contains Ag-rich galena samples,  B Lode contains Pb- and Cd-rich galena samples,  C Lode contains Zn-rich galena samples,  1 Lens contains Pb-rich galena samples,  2 Lens contains S-rich galena samples,  3 Lens contains Cu-rich galena samples, and  The Western Mineralisation contains Bi-, Fe-, As- and Sb-rich galena samples. 5.2.3: Correspondence analysis for sphalerite samples The z-transformed values of Table 5.2 and the z-transformed values plus 2 for sphalerite are given in Tables 5.6 and Table 5.7 respectively. The result of decomposition of Table 5.7 into 6 dimensions is given in Table 5.8. 1A combination of three correspondence maps in relation to dimensions 1, 2 and 3 2 Calculated by PPI (dimension 1) + PPI (dimension 2) + PPI (dimension 3) from Table 5.5
ADE
Chapter 5- Multi-Element Relationships 153 Table 5.8: The results of decomposition of Table 5.7 into six dimensions. Proportion of Percentage of Dimensions Inertia PPI inertia cumulative inertia 1 0.1029 0.5535 55.35 55.35 2 0.0404 0.2175 21.75 77.10 3 0.0235 0.1266 12.66 89.76 4 0.0116 0.0626 6.26 96.02 5 0.005 0.0271 2.71 98.73 6 0.0024 0.0127 1.27 100.00 Total 0.1859 5.2.4 Correspondence maps for sphalerite samples 5.2.4.1 Interpretation of Figure 5.4 Figure 5.4 shows 77.1 % of chemical variation of sphalerite samples within the Broken Hill orebodies in relation to dimensions 1 and 2.  Dimension 1 In Figure 5.4, geochemical processes that can be inferred from dimension 1 separate Fe-, Sb- and S-rich sphalerite and 3 Lens at the extreme end of the negative scale from A Lode, 1 Lens, Zn-, Ag-, Bi-, Cd-rich sphalerite at the extreme end of the positive scale.  Dimension 2 Dimension 2, on the other hand, separates the Western Mineralisation and Pb-rich sphalerite samples from 1 Lens and As-rich sphalerite samples. Figure 5.4: Correspondence map in relation to dimensions 1 and 2 for sphalerite.
ADE
Chapter 5- Multi-Element Relationships 155 Figure 5.6: Correspondence map in relation to dimensions 2 and 3 for sphalerite. 5.2.4.4 Interpretation of a combination of Figures 5.4 to 5.6 Together Figures 5.4 to 5.6 provide a three-dimensional model that explains 89.76 %1 of the total chemical variation of sphalerite within the orebody of the Broken Hill deposit and their inter-relationships can be summarized as follows:  A Lode contains Ag-rich sphalerite samples,  B Lode contains Fe- and Sb-rich sphalerite samples,  C Lode contains As-rich sphalerite samples,  1 Lens contains Cd- , Bi- and Cu-rich sphalerite samples,  2 Lens contains Pb- and S-rich sphalerite samples,  3 Lens contains Sb- and Fe-rich sphalerite samples, and  The Western Mineralisation contains Pb-rich sphalerite samples. There is a temptation in exploration to think that the "next" Broken Hill will be identical to the main Broken Hill deposit on the Line of Lode. The result of correspondence analysis for galena and sphalerite shows that such a view is not valid. The chemical mineralogy of the Broken Hill Domain is similar to: 1 Calculated by PPI (dimension 1) + PPI (dimension 2) + PPI (dimension 3) from Table 5.8
ADE
156 Chapter 5- Multi-Element Relationships 1. The Eastern Fold Belt of the Mount Isa Inlier where IOCG deposits (e.g. Ernest Henry), 2. Iron formation Cu-Au deposits (e.g. Starra line, Osborne), 3. Co-As deposits (e.g. Mount Cobalt), 4. Pb-Ag-Zn deposits (e.g. Pegmont, Cannington), and 5. Mo-Re deposits (e.g. Merlin) Another "Broken Hill" may not even have an abundance of gahnite and garnet rocks. There may even be a Mount Isa-type Pb-Zn deposit at Broken Hill in the cover rocks. What is known about the Broken Hill Domain is that it contains the largest Zn-Pb- Ag deposit in the world, that there are thousands of minor deposits of different commodities and that the area is a major metallogenic province. This suggests that the Broken Hill Zn-Pb-Ag deposit is not an orphan. 5.3 Assay data used for multivariate statistical analysis Some core samples (1,059) of the Western Mineralisation have equal assay data for the concentration of the following ten elements: Zn, Pb, Cu, Fe, S, Ag, Cd, As, Sb and Bi. The 1,059 samples have been cut from surface drill cores 4003 to 4032, 4040 and 4042. Some of the element concentrations including Sb, As, Bi, Cd and Ag were measured based on parts per million (ppm). Their values were converted to percentages so all assay values have the same unit, hence simplifying the analysis. The data base has been provided in supplementary file to this thesis. 5.3.1 Statistical distributions of assay values The statistical distributions in terms of histograms of the raw data of the ten elements are displayed in Figure 5.7 and their corresponding probability plots are given in Figure 5.8. In Figure 5.7, all histograms except for Fe show strongly positive skewed distributions that indicate that the proportion of a low concentration of elements is much higher than that of a rich concentration of elements. The distribution of Bi and Sb show that they have an approximately constant concentration value. It is clear from Figures 5.7 and 5.8 that the data for the element concentrations does not follow normal distributions. They will have to be transformed to normal distributions first for effective bivariate analysis. The reason for this has been explained in Table 3.7.
ADE
Chapter 5- Multi-Element Relationships 159 5.3.2 Compositional data analysis Compositional data is used in subjects such as geochemistry, petroleum chemistry and environmental issues (Labus 2005). Compositional data is multivariate data defined as a vector X with components of non-negative values X ,…, X that sum up to a constant, 1 D usually 100 percent. As a consequence, elements of compositional data are not independent. For example, if one of the elements is enriched, other elements must be depleted to retain the geochemical mass balance (100%)1. Therefore, conventional statistics applied to raw closed geochemical data may produce incorrect outcomes. A sample space for D-part compositional data (e.g. D=10 for current data set), SD, is called "Simplex" (Aitchison 1986). If D=3, a ternary diagram can show clearly the different contributions of three elements for the closed data and if D=4, a tetrahedron diagram can display the different contributions of four elements for the closed data (Figure 5.9). However, there is no satisfactory way to evaluate and demonstrate the contribution of more than 4 elements in ternary and tetrahedron diagrams. Figure 5.9: Contribution of 4 element concentrations within 1,059 samples of the Western Mineralisation in a tetrahedron model. 1This problem is called closure effect (Filzmoser, Hron & Reimann 2009, p.627).
ADE
160 Chapter 5- Multi-Element Relationships Aitchison (1986) introduced logarithms of ratios (log-ratio) to convert compositional data into an unconstrained form in the real space RD. The transformed data avoids the closure effect and the transformation is appropriate for elements that are measured in percentages. The following three log-ratios are used for the transformation of data into the real space (RD): 1. Additive log-ratio (alr) transformation: value of each element (x) is divided by the value of a selected element (x ) before taking the logarithm i.e.: D xSD  YRD1  x x x  Y  alr (x)  log 1 , log 2 ,...,log D1   x x x  D D D 2. Centred log-ratio (clr) transformation: value of each element (x) is divided by the geometric mean (m ) of the data before taking the logarithm i.e.: g xSD  YRD for i 1,..., D.     x x x x Y  clr (X)  log   log 1 ,log 2 ,...,log D   m   m m m  g g g g m  D ΠD x Where m is the geometrical mean defined as . g g i1 i 3. Isometric log-ratio (ilr) transformation: the transformed vector is defined by sequential binary partition that solves the problem of data collinearity resulting from clr-transformation (Egozcue & Pawlowsky-Glahn 2005; Egozcue et al. 2003) i.e.: xSD  YRD-1 for i  1,...,D-1  i i Πi x  Y  ilr (x)   log j1 j   i 1 x   i1 
ADE
Chapter 5- Multi-Element Relationships 165 The results of Table 5.9 and 510 are outlined below. 1. Positive correlations range from 0.021(between Ag and S) to 0.955 (between Sb and Bi), 2. Negative correlations change from -0.048 (between Ag and Cu) to -0.781(between Sb and Zn), 3. Elements of Bi, Sb, As, Fe and Cu are negatively correlated with Pb, Zn, S, Ag and Cd in this orebody, 4. The degree of correlation between Cu and other elements is generally very low, 5. There are strong positive correlations between Pb and Ag and between Zn and Cd, 6. Pb and Zn both have strong negative correlation with Fe, Sb and Bi, 7. Zn shows moderate positive correlation with S while Pb shows weak correlation with S, and 8. Sb and Bi have strong positive correlations with each other and also with Fe, while Sb and Bi both have strong negative correlations with S. 9. Table 5.10 shows that 60 % of the variability of the Zn concentration is negatively correlated with the variation of Sb or Bi. According to Table 5.10, Sb and Bi are negatively correlated with 40 % and 41 % of the variability of Pb respectively. 5.4.1 Interpretation of the PCC results Silver and Cd contributed in the form of solid solution to the atomic structure of galena and sphalerite respectively in the Western Mineralisation. Low correlation coefficients of Cu with other elements indicate that enrichment or depletion of other elements did not influence the degree of variability of Cu and vice versa. It is possible Cu originated from a secondary geochemical process with a different source because the major sulphide minerals of the orebody are galena and sphalerite. The low relationship of Cu with Pb and Zn in the Western Mineralisation is similar to almost all of the Australian sedimentary exhalative Pb-Zn-Ag deposits (McArthur River, Century, Hilton, George Fisher and Cannington) that contain a Cu-content. However, the Mount Isa deposit in the north of Australia has a spatial relationship with rich Cu-bearing sulphide minerals (Large et al. 2005).
ADE
166 Chapter 5- Multi-Element Relationships 5.5 Linear multivariate regressions (LMR) The LMR allows for prediction of the behaviour or variability of individual elements such as Pb and Zn concentrations based on other elements as predictors. Moreover, the LMR shows how well a predictor element can represent the variability of Pb or Zn concentrations. The result of the LMR shows the most economical exploration model for prediction of Pb and Zn concentrations by measuring the minimum number of predictor elements. Thus, it is important to determine whether all of the predictor elements are equally important to predict the variability of Pb or Zn and if not which of them has more priority for the predication of the Pb and Zn concentrations. The method of LMR is also used in exploration of gold (Bellehumeur & Jébrak 1993). It is obvious that the number of predictor elements for Zn (or Pb) can increase if we have access to more than 9 element concentrations associated with the Western Mineralisation but as mentioned earlier only some of them may be considered as good predictor elements of Pb and Zn. The resulting predictor elements of Pb and Zn can be calculated by variogram analysis so as to understand whether the predictor elements are suitable as "pathfinder" or "indicator" elements (Levinson 1974, pp.54-55; Peters 1987, p.403) for tracking galena and sphalerite in the Western Mineralisation or similar Pb-Zn sulphide orebodies. More details about pathfinder elements are given in Section 7.3 where the spatial zonation of geochemical haloes is discussed for the Western Mineralisation. In this study, a stepwise multiple regressions analysis using SPSS was carried out to evaluate the contribution of all predictor variables for a prediction model of Pb or Zn concentrations in the Western Mineralisation. In this process, the nine predictor elements of Pb, Fe, S, Cd, Ag, Sb, Bi, As and Cu for Zn and the nine predictor elements of Zn, Fe, S, Cd, Ag, Sb, Bi, As and Cu for Pb were added one by one to the regression model and their accumulative effects on Pb and Zn predictions are calculated. Correlated elements will make positive contributions to the regression model and they will be retained in the model. The assessment can be done using a desired "Adjusted R Square". The aim is to select a model with a higher Adjusted R Square and a lesser number of predictor elements. Some of the statistical terms of the LMR are briefly described below: 1. The selected predictor elements may show correlation with Pb or Zn but they need not necessarily have a strong correlation with the other predictor elements. Otherwise, a strong correlation between some of the predictor elements may
ADE
Chapter 5- Multi-Element Relationships 167 obscure the relative contributions of each predictor element to the success of the model. This can be checked using the collinearity diagnostics section of SPSS. The collinearity diagnostics measure the tolerance and Variance Inflation Factor (VIF) value. The results are presented in the prediction models of Pb and Zn. A value less than 0.01 for tolerance and more than 10 for VIF implies a strong correlation between predictor elements, meaning they have the collinearity problem, and 2. The beta value represents the intensity of influence of each predictor element on the variation of Pb and Zn. A large beta value indicates a higher impact of the predictor element on variations of Pb and Zn. The beta value is defined in units of the standard deviation. For example, a negative beta value of -0.274 of Fe for Pb (Table 5.11) indicates that a positive change of one standard deviation in Fe will result in a negative change of 0.274 standard deviations in Pb. 5.5.1 Results of LMR for Pb A significant model with three predictor elements emerged in the third model (Table 5.11) that is characterised by F , = 572.974 and p < 0.0005, where 713.787 438.092 713.787 and 438.092 in F (Fisher test) are sum of the squares for regression and residual respectively. The Adjusted R Square is 0.619 and collinearity statistics are: Tolerance (0.3 < X < 1) and VIF < 3. Table 5.11: Significant predictor elements of Pb in the third model of LMR analysis. Elements Beta p-value Fe -0.275 p < 0.0005 Ag 0.410 p = 0.0005 Bi -0.245 p < 0.0005 Elements of Fe, Ag and Bi are part of the third prediction model for Pb with the Adjusted R Square of 61.9 % indicating that the variability of Fe, Ag and Bi taken together as an exploration model for Pb can predict 61.9 % of the variability (variance) of the Pb concentration in the Western Mineralisation.
ADE
168 Chapter 5- Multi-Element Relationships 5.5.2 Interpretation of LMR results for Pb The Adjusted R Square improved when more elements were added successfully to the model. The final LMR model for Pb contains the following predictor elements in decreasing order of their significant contributions: Fe, Ag, Bi, Zn, As, Cu, Cd, Ag and Sb. Sulphur made little contribution to the prediction of Pb and therefore it was not included in the final model. In comparison with the bivariate analysis (Table 5.9), the following conclusions can be drawn based on LMR analysis of the Pb: 1. Iron and Bi concentrations have a negative impact (negative beta) on the variability of the Pb concentration. The same conclusion was also reached based on Table 5.9, and 2. Antimony concentrations did not appear in the third model as predictor elements for Pb though it was in the last model but with little contribution (small beta). On the contrary, Sb shows a strong negative correlation coefficient with Pb from the bivariate analysis (Table 5.9). From Table 5.1, the average atomic percentage of Fe, As, Bi and Sb in galena samples of the Western Mineralisation is higher than the other Broken Hill orebodies. The result of LMR analysis suggests Fe, Ag and Bi are appropriate predictor elements for the prediction of Pb variation in the Western Mineralisation. 5.5.3 Results of LMR for Zn For Zn concentration, the third model (Table 5.12) is characterised by F , = 572.974 and p < 0.0005, adjusted R square of 0.816 and collinearity 1317.403 296.121 statistics: Tolerance (0.5 < X< 1) and VIF < 2. Table 5.12: Significant predictor elements of Zn in the third model of LMR analysis. Elements Beta p-value Cd 0.487 p < 0.0005 Sb -0.578 p = 0.0005 Ag -0.134 p < 0.0005
ADE
Chapter 5- Multi-Element Relationships 169 Elements that contribute to the quality of the prediction model for Zn are (in decreasing order of importance): Cd, Sb, Ag, Fe, As, Pb, Cu and Bi. 5.5.4 Interpretation of LMR results for Zn A linear combination of the three elements Cd, Sb and Ag can provide a good predictor model (81.6 %) for Zn concentration in the Western Mineralisation and similar mineralisation. 1. The LMR model for Zn shows that Ag has a negative correlation (negative beta) with Zn in contrast to the correlation coefficient in Table 5.9 where it shows a weak positive correlation of Ag with Zn, 2. Bismuth has only a minor contribution to the final prediction model for Zn and consequently this model suggests that Bi is not an appropriate predictor element for tracking the variability of Zn concentration, while in Table 5.9, the Bi and Zn show a strong negative correlation, and 3. Sulphur content is not recognised as a significant predictor in the final model (9) for Zn and was thus removed from the model. Based on these assessments, it can be concluded that LMR is more suitable for identifying the major predictor elements for tracking ore minerals and Pearson correlation coefficient should be used with care in geological interpretation. 5.6 Cluster analysis Cluster analysis (Davis 1973; Hartigan 1975; Templ, Filzmoser & Reimann 2008) is an investigative multivariate data analysis that can be used to classify geochemical variables or samples. Cluster analysis can support the development of geochemical models by identifying multi-element relationships or by clustering element concentrations based on the amount of their percentage of similarities. Furthermore, it reduces the number of variables that need to be considered in subsequent analysis and provides a visual description of combined variables that is easier to understand than results obtained through more traditional analysis such as PCA.
ADE
170 Chapter 5- Multi-Element Relationships Cluster analysis needs to specify a final partitioning by selecting the number of sub-clusters (user defined) and the problem is to select an optimum number of sub-clusters for partitioning (Fraley & Raftery 1988). It is possible to group the elements into 1 to n7 clusters, a procedure that is called "hierarchical" clustering. Hierarchical clustering provides n cluster solutions and the user should make a decision which model is the most appropriate. The principal aim of cluster analysis in the Western Mineralisation is to divide element concentrations into a number of groups to better understand the multivariate behaviour and the structures of multi-element interactions. A good cluster analysis attempts to classify the elements into groups with high levels of similarity while at the same time the differences between the individual groups are kept as large as possible. 5.6.1 Distance measures for cluster analysis Cluster analysis uses distance measures to quantify the similarity among the element concentrations in a multivariate space based on the entered assayed samples. The "distance" in cluster analysis is not the same as a geographical distance among samples and the important issue is how best they can measure distance between the elements. Several distance measures exist in this method such as Euclidean, Manhattan, Ward, Gower (Gower 1966), Canberra (Lance & Williams 1966), the random forest proximity (Breiman 2001) and the correlation coefficient. The Euclidian is a straight-line distance in geometry that is calculated by the root of the sum of the squares and the Manhattan is the sum of linear distances. More detail about distance measures and examples can be found in Arabie, Hubert and De Soete (1996), Gordon (1999), Mark and Roger (1984) and Swan and Sandilands (1995). However, there are no theoretical reasons to prefer one of the distances over the other. In this thesis, the correlation coefficient of the elements was used as the distance measure for quantifying the amount of similarity of 10 element concentrations. 5.6.2 Cluster algorithm After selection of a distance measure (e.g. correlation distance) a cluster algorithm should be selected in order to determine cluster membership for each element based on their distance measures. The cluster algorithms can be "divisive" or "agglomerative". A divisive method starts with all elements in one cluster and the cluster gradually splits the 7 Maximum n is equal to the number of variables (i.e. elements)
ADE
Chapter 5- Multi-Element Relationships 171 smaller groups step by step. In contrast, in the agglomerative technique, at first each element forms its own cluster and this process continues for "n" single element clusters and then the number of clusters is reduced by joining the two closest elements and continues to find another similar element for gradually joining the first two or two other elements. This procedure will proceed until one large cluster is formed. There are several algorithms for linking two clusters, such as average linkage, complete linkage, single linkage, median linkage, centroid linkage and so on. In this study, the agglomerative average linkage distance was selected for the calculation of the cluster algorithm. The average linkage algorithm calculates the mean of all distances of two elements between the elements of two sub-clusters and then combines two sub-clusters with the minimum average distance into one new cluster. 5.6.3 Result of cluster analysis for the 1,059 clr-transformed data Cluster analysis was calculated for the 1,059 clr-transformed data using the average linkage algorithm and correlation coefficient distance (Figure 5.12). Figure (5.12) shows the following three main groups for 10 elements: 1. Group 1 (red colour): Zn, Pb, S, Ag and Cd, 2. Group 2 (green colour): Cu, and 3. Group 3 (blue colour): Fe, As, Sb and Bi. It is possible to define two main groups instead of three for the cluster algorithm in Figure 5.12 and the percentage of similarity among the 10 elements does not change but Cu will be considered as part of the group 3. The horizontal line in the dendrogram (Figure 5.12) represents the similarity level between two or more elements while the vertical line indicates the distance level or differences. For example, the elements of Sb and Bi in group 3 shows the highest similarity level with 97.7 %, whereas all elements of the groups 1 and 2 reveal the lowest similarity (26.30 %) with all elements of the group 3. Minor abundance of chalcopyrite, pyrrhotite, pyrite, gudmundite, tetrahedrite8 and arsenopyrite in the Western Mineralisation may relate to the amount of dissimilarities of Fe, Cu, As and S in Figure 5.12. 8 (Cu, Fe, Ag, Zn) Sb S 12 4 13
ADE
172 Chapter 5- Multi-Element Relationships Figure 5.12: Hierarchical horizontal cluster algorithm with three main groups for 1,059 clr-transformed data of each element. According to Figure 5.12, there are a high percentage of similarities among the following elements: 1. Zn-Cd with 89.4 %, 2. Pb-Ag with 89 %, 3. Fe-Sb-Bi with 83.4 %, and 4. S-Zn-Cd with 77 %. Some of those similarities between the elements are seen in the sulphide minerals of the Western Mineralisation e.g. sphalerite containing Cd and galena containing Ag. 5.7 Principal Component Analysis (PCA) PCA is a non-spatial multivariate procedure that is most commonly applied with the intention of recognising a small number of interesting sub-dimensional elements, which may then be examined by spatial approaches, exploring for spatial patterns and correlations. The aim of PCA is to reduce the complexity and dimensions of variables to a smaller number of uncorrelated principal components (Anderson 1984; Johnson & Wichern 1992; Rencher 1995). PCA helps to describe the greatest degree of variance for the lowest amount of the uncorrelated variables. In geochemical study, the goal of PCA is to facilitate the interpretation and explanation of the underlying data structure from large
ADE
Chapter 5- Multi-Element Relationships 173 assay data. Moreover, PCA investigates the degree of continuity or clustering of samples and identifies element concentrations whose significance is that they can be broken down into some distinct groups. 5.7.1 Result of PCA The 1,059 clr-transformed data of each element of the Western Mineralisation were examined by a subprogram of the PCA and correlation matrix in SPSS 17. The correlation matrix is preferred over the covariance method if the element concentrations are measured by different scales (e.g. percentage and ppm) and need to be standardised. For solving this problem, in the subprogram of PCA, the standardisation part of the program was selected. In comparison to factor analysis, PCA deals with the total variation in the correlation matrix rather than part of it and does not require showing if the PCA is appropriate for structural detection of the data. In Table 5.13, PC explains 51.6 % of the variance of ten element concentrations 1 and it is thus a relatively important PC, while the other PCs have less important roles in describing the variation of ten element concentrations. Table 5.13: Summary of PCs in the Western Mineralisation. Percentage of Percentage of PCs9 Eigenvalues proportional variance cumulative variance (PPV) (PCV) PC 5.1594 51.6 51.6 1 PC 1.5696 15.7 67.3 2 PC 1.2503 12.5 79.8 3 PC 0.9561 9.6 89.4 4 A combination of PC , PC , PC and PC provides sufficient detail about major 1 2 3 4 geochemical variation (89.4 %) in the Western Mineralisation, rather than studying ten geochemical elements and their behaviours individually. It should be noted that additional PCs, such as PC and PC , if used, should increase overall variance. The increment in the 5 6 variance will be case dependent but in general smaller and smaller contributions will be expected, see Figure 5.13. 9 Principal Components
ADE
174 Chapter 5- Multi-Element Relationships Figure 5.13: The scree plot of eigenvalues versus number of PCs. Figure 5.13 represents the relative size of each eigenvalue in descending order versus the number of the respective PC. It shows visually which of the 10 PCs accounted for most of the variability in the Western Mineralisation. The first component shows maximum variance and the successive components make an increasingly smaller contribution to the total variance. According to the Figure, the first four eigenvalues explain the major geochemical variations of the Western Mineralisation. It should be noted that the remaining accumulated 11.6 % variances cannot be explained by the results of the first four PCs and the remaining six PCs fulfil only a small portion of the accumulated 11.6 % variances. Hence, each of them contributes only minimally to the variation of the element concentrations and cannot be interpreted meaningfully. Those PCs constitute random variations or an unsystematic chemical process within the sulphide orebody of the Western Mineralisation. Spry, Plimer and Teal (2008) suggested that the Western Mineralisation is a stratabound deposit formed by the replacement of sediments at the top of an upward- coarsening sequence beneath an aquifer cap (Unit 4.6). If this were the case, sulphide bearing ore must have pulsated numerous times between the deposition of C Lode and A Lode and they can be interpreted as geochemical variations evident from the PCs of this study (Table 5.13). The Western Mineralisation has also a variation in grade and zonation and Plimer (2006b) argued that the intersection of F3 and F4 structures created dilation
ADE
Chapter 5- Multi-Element Relationships 175 during the Olarian Orogeny. Plimer (2006b) also suggested that more ductile minerals such as galena and pyrrhotite might have flowed into these dilation zones. The geochemical variations evident from the PCs of this study can also be attributed to the dilation zones. 5.7.2 The amount of PC loadings In Table 5.14, the PC loadings denote the degree of correlation between individual element concentrations and the PCs (i.e. PC , PC , PC and PC ). Thus, elements with 1 2 3 4 higher PC loadings in each PC play more roles in geochemical variation of the PC and have more potential for mineralisation. For example, in Table 5.14, Zn has maximum positive PC loading in PC (0.857) in comparison with the weights of other elements in 1 PC and even the weights of Zn in PC , PC and PC . 1 2 3 4 Table 5.14: PC loadings10 of 10 elements for each PC (Table 5.13) of the Western Mineralisation. PCs loading PC loading PC loading PC loading PC loading 1 2 3 4 Elements Zn 0.857 0.346 - 0.224 - 0.063 Pb 0.697 - 0.584 0.056 0.071 Cu 0.029 0.221 0.929 0.057 Fe - 0.842 0.271 0.073 0.228 S 0.681 0.479 0.302 - 0.051 Ag 0.511 - 0.724 0.083 0.266 Cd 0.677 0.460 - 0.449 0.180 As - 0.549 - 0.147 - 0.055 - 0.810 Sb - 0.932 0.008 - 0.122 0.269 Bi - 0.936 0.010 - 0.106 0.238 This indicates a strong correlation of Zn with geochemical variation of PC . Therefore, 1 major sulphide mineralisation containing Zn (sphalerite) should occur during geochemical variation evident from PC in the Western Mineralisation. 1 In regards to the sulphide mineralogy of the Western Mineralisation, the elements with absolute PC loading values ≥ 0.4 in Table 5.14 are considered to have potential for mineralisation or are considered to have a significant influence on the geochemical variation of the PCs. According to the absolute PC loading values ≥ 0.4, the following elements have major chemical contributions to each PC: 10 Weights of elements in each PC
ADE
176 Chapter 5- Multi-Element Relationships 1. PC : Bi, Sb, Zn, Fe, Pb, S, Cd, As and Ag, 1 2. PC : Ag, Pb, S and Cd, 2 3. PC : Cu and Cd, and 3 4. PC : As. 4 5.7.3 The sign of PC loadings of the elements Interpretation of the PC loadings of the elements in Table 5.14 based on their signs (i.e. negative or positive) depends on dominant sulphide minerals in the Western Mineralisation. For example, the positive PC loadings of Zn and Pb in PC (with 51.6 % 1 1 contribution in the sulphide mineralisation; Table 5.13) should be interpreted as sulphide mineralisation that is consistent with the dominant sulphide mineralogy (galena and sphalerite) of the Western Mineralisation.  Elements with the positive PC loadings 1 In Table 5.14, the positive loadings of Zn, Pb, Cu, Ag, Cd and S indicate the following possibilities: 1. These elements initially may tend to incorporated as a major constituent of the mineral compositions of the Western Mineralisation (e.g. Pb in galena or Zn in sphalerite), 2. These elements may be as mineral inclusions containing the elements inside other major minerals (e.g. chalcopyrite in galena), or 3. These elements are incorporated into the atomic lattice of other minerals (solid solution) as trace elements. This is most likely when the amount of an element concentration is naturally very low to form an independent mineral. For example, the average concentration of Cd is very low against the average concentrations of Zn and Pb in the Western Mineralisation (Table 3.2). The negative PC loadings of those elements are interpreted as dilution effects or depletion or removal of those elements from sulphide bearing fluids during geochemical variation evident from the PC.
ADE
Chapter 5- Multi-Element Relationships 177  Elements with the negative PC loadings 1 In PC of Table 5.14, the semi-metallic elements of As, Bi and Sb and metallic 1 element of Fe have negative PC loadings indicating that these elements tend to be incorporated in the atomic structure of other minerals (solid solution). In contrast, positive weight values of the elements indicate that those elements trend to form independent sulphide minerals (e.g. pyrite) and sulfosalt minerals (e.g. tetrahedrite) during geochemical variations evident from the PCs. 5.7.4 Interpretation of PCs  PC 1 In Table 5.14, PC separates the elements of Zn, Pb, S, Ag and Cd with positive 1 PC loadings from the elements of Fe, As, Sb and Bi with negative PC loadings. 1 1 According to Table 5.13, major sphalerite and galena of the Western Mineralisation occurred during the processes that can be inferred from the PC loadings of the elements. 1  PC 2 In Table 5.14, PC separates the elements of Zn, Cu, Fe, S and Cd with positive 2 PC loadings from the elements of Pb, Ag and As with negative PC loadings. PC has 2 2 2 week positive correlations with Fe and Cu and moderate positive correlation with S, Pb and Zn.  PC 3 PC is characterised by highly distinct contrasts between Cu and Cd. Cu has the 3 significant positive loading (0.929 in Table 5.14) in PC that means a strong correlation 3 with the geochemical variation evident from PC . 3  PC 4 In Table 5.14, PC was affected dominantly by As content because As is the only 4 element with a highly negative PC loading that has a strong correlation with PC . In 4 4 contrast, the PC loading of S is very low which suggests S content was not effective in the 4 sulphide mineral formation during the geochemical variation evident from PC . This 4 suggests that the lack of S content should have been replaced by excess As.
ADE
178 Chapter 5- Multi-Element Relationships 5.7.5 Interpretation of elements within PCs  Zn Maximum positive PC loading is associated with Zn and this element also has a 1 positive PC loading in contrast with a negative PC loading of Pb (Table 5.14). This is 2 2 consistent with a higher abundance of sphalerite relative to galena in the Western Mineralisation.  Pb Major Pb-rich sulphide minerals (such as galena) had to be produced during the geochemical variation evident from PC . Pb shows smaller PC loading (0.697 in Table 1 1 5.14) relative to that of Zn (0.857 in Table 5.14) that means smaller potential for mineralisation relative to Zn. PC loading of Pb is a negative value (-0.584 in Table 5.14) 2 that means the removal of Pb from sulphide mineralisation. According to Table 5.14, PC 3 and PC loadings of Pb are so small that this indicates no linear relationship with 4 geochemical variations evident from PC and PC . Therefore, major galena should be 3 4 produced during the processes that can be inferred from the loading of PC , but the amount 1 of galena is smaller than sphalerite in the Western Mineralisation.  S Table 5.14 shows that the PC loading of S is reduced when comparing PC to PC . 1 3 This indicates the reduction of potential for sulphide mineralisation from PC to PC and 1 3 there is no linear correlation between S content and the geochemical variation of other elements in PC . 4  Cu Copper has a minimum PC loading (0.029 in Table 5.14) that suggests Cu 1 concentration is not correlated with PC . Accordingly, there is a lack of incorporation of 1 Cu during the formation of major galena and sphalerite minerals in the Western Mineralisation. This result is consistent with the mineralogy of the Western Mineralisation and Cu occurred there in the following form of minerals: 1. Chalcopyrite as major Cu-bearing sulphide mineral, 2. Incorporation of Cu in the atomic structure (solid solution) of sulphide minerals (e.g. gudmundite, galena, sphalerite and pyrrhotite) and sulfosalt minerals (e.g. tetrahedrite, tennantite and bournonite), and
ADE
Chapter 5- Multi-Element Relationships 179 3. Appearance in form of inclusion in other sulphide minerals. Spry, Plimer & Teale (2008, p.234) explained the presence of lamellar twins and inclusions of chalcopyrite along the cleavage and twinplanes of galena samples within the Broken Hill orebodies and they claimed that the inclusions of chalcopyrite have been generated from post- peak modification to mineral assemblage involving galena and sphalerite.  Fe Iron has a high negative PC loading (-0.842 in Table 5.14) which indicates a 1 strong negative correlation of Fe with the geochemical variation of PC . Iron in PC has 1 1 strongly concentrated in the atomic structure of major and minor sulphide minerals. This result of PC is very consistent with the result of a high average atomic percentage of Fe 1 within the atomic structures of major sulphide minerals of the Western Mineralisation (see the results of Fe concentration in Tables 5.1 and 3.2).  Cd and Ag In Table 5.14, the PC loadings of Cd and Ag, being 0.677 and 0.511 respectively, 1 show a good positive correlation with PC that suggests the occurrence of minerals 1 containing Cd and Ag resulting from geochemical variation of PC . Cadmium also shows a 1 moderate positive correlation with geochemical variation of elements in PC . In contrast, 2 Ag reveals a strong negative correlation with PC . This indicates a removal of Ag 2 concentration during the geochemical variation evident from PC . In the Western 2 Mineralisation, Cd and Ag contributed mostly to the atomic structure of sphalerite and galena respectively rather than forming major independent minerals. This may be because of the low amount of Cd and Ag (Table 3.2).  Bi and Sb In Table 5.14, Bi and Sb show strong negative correlations with PC and they are 1 incorporated significantly in the atomic lattice of galena (Table 5.1) and other sulphide minerals of the Western Mineralisation.  As Arsenic also has a moderate negative PC loading and high negative PC loading 1 4 (Table 5.14). This indicates that during the processes associated with PC and PC ,arsenic 1 4 may be highly concentrated in the atomic structure of sulphide minerals (e.g. galena; Table 5.1). Plimer (2006b) stated that it is possible that As, Mo, W and Au were added to sulphide rocks during the Olarian Orogeny.
ADE
180 Chapter 5- Multi-Element Relationships 5.7.6 Maps of PC loadings A map of PC loadings visualises distances between elements (points) in ratios to the two dimensional space of two specific PCs. The maps of PC loadings help us to understand the following issues: 1. Identification of the relative real distances of elements in different dimensional spaces, 2. Evaluation of the probability of occurrence of mineral paragenesis, and 3. Identification of the potential mineralisation during the geochemical interaction of the two PCs.  Similarity of the map of PC loadings to correspondence maps They both aim to reduce dimensional spaces and chemical complexity. This helps to better interpret the multi-influences of elements.  Differences between the map of PC loadings and correspondence maps The map of PC loadings is used to show the relationship of one type of variable (e.g. elements) within two PC spaces at one time; however, a correspondence map is used to show the relationships between two different types of variable (e.g. the Broken Hill orebodies with the elements of a mineral) simultaneously within two dimensional spaces. Important factors for interpretation of the maps of PC loadings are: 1. The total variance of each map that indicates the degree of total mineralisation during the effect of geochemical interaction between two PCs, 2. The magnitude value of PC loading for each element that indicates the degree of potential for mineral formation of the element within the map in comparison with the other elements, and 3. Distances of elements (points) from each pair of PCs in each map that reveal the possible mineral paragenesis or occurrence of different minerals during the geochemical interactions of the two PCs.
ADE
Chapter 5- Multi-Element Relationships 181 5.7.6.1 Map of PC loadingand PC loading (Figure 5.14) 1 2  Chemical variation PC loading and PC loading jointly account for 67.3 %11of the total geochemical 1 2 variation in the Western Mineralisation.  Position of important elements relative to PC loading scale 1 By dividing this map into two parts from the zero value of the PC loading scale, 1 the group of elements Fe, Bi, Sb and As lies at the extreme end of the negative scale of the PC loading and the group of elements Pb, Zn, Cd, S and Ag lies at the extreme end of the 1 positive scale of the PC loading In this map, there is an intense polarization or divergence 1 . between the two groups of elements on the negative and positive sides of the PC loading 1 scale. This is because of the relatively high distances between the two groups of elements.  PC loading as a chemical discriminator 1 In Figure 5.14, it is obvious that PC loading acts as a good geochemical 1 discriminator between the elements Zn-Pb-Ag-Cd-S and Bi-Sb-Fe-As in the Western Mineralisation. Figure 5.14: Map of PC loading, PC loading and the legend of 10 elements. 1 2 11 Calculated by PPV (PC ) + PPV (PC ) from Table 5.13 1 2
ADE
182 Chapter 5- Multi-Element Relationships  Position of important elements in relation to PC loadingscale 2 If Figure 5.14 is divided into two parts from the zero value of the PC loading scale, 2 it shows distinct polarization between the group elements of Pb-Ag at the extreme end of the negative scale of the PC loading and the group elements of S-Zn-Cd at the extreme 2 end of the positive scale of the PC loading. This result is very consistent with the chemical 2 composition of the sphalerite and galena samples collected from the Western Mineralisation because the sphalerite samples (Table 5.2) contain a higher average atomic percentage of Cd and S relative to the galena samples (Table 5.1). In contrast, the galena samples (Table 5.1) contain more Ag relative to the sphalerite samples (Table 5.2). The occurrence of arsenopyrite and löllingite12 can be associated with the geochemical interaction between PC loading and PC loading. However, the abundance of 2 1 these minerals is very low in the Western Mineralisation because PC has only 15.7 % 2 contribution to sulphide mineralisation of the orebody. 5.7.6.2 Map of PC loading and PC loading (Figure 5.15) 1 3  Chemical variation PC loadingand PC loading account for 64.1 %13 of the total geochemical 1 3 variation.  Position of important elements in relation to PC loadingscale 3 PC loading is characterised by significant loading of Cu, which is positioned at the 3 upper end of the positive scale of the PC loading and to a lesser extent Cd at the lower end 3 of negative scale of the PC loading. 3 Figure 5.15 also shows that the geochemical interaction between PC loading and 3 PC loading can produce arsenopyrite and löllingite. However, PC has only 12.5 % 1 3 contribution to sulphide mineralisation of the orebody. 12 FeAs 2 13 Calculated by PPV (PC ) + PPV (PC ) from Table 5.13 1 3
ADE
184 Chapter 5- Multi-Element Relationships 5.7.6.4 Map of PC loading and PC loading (Figure 5.17) 2 3  Chemical variation PC loading and PC loading only account for 28.2 %15 of the total geochemical 2 3 variation of the Western Mineralisation. Figure 5.17: Map of PC loading and PC loading. 2 3  PC loading as a chemical discriminator 2 Based on this map, PC loading again separates the elements of Pb and Ag from the 2 elements of Zn, S and Cd. PC loading appears to act as a discriminator of Cu from other 3 elements. 5.7.6.5 Map of PC loading and PC loading (Figure 5.18) 2 4  Position of important elements in relation to PC scale 2 As shown in Figure 5.14 and 5.17, PC loading in Figure 5.18 also shows the 2 positive loading of Zn as opposed to the negative loading of Pb in relation to the PC 2 loading scale. This model highlights more enrichment of sphalerite relative to galena that is very consistent with the Western Mineralisation. 15 Calculated by PPV (PC ) + PPV (PC ) from Table 5.13 2 3
ADE
Chapter 5- Multi-Element Relationships 185  Paragenesis In Figure 5.18, the distance between Cu and the group elements of Fe, S, Sb, Zn, Cd and Bi is small. This indicates that, during the geochemical interaction of PC loading 2 with PC loading, there was potential for occurrence of chalcopyrite, tetrahedrite and 4 gudmundite. However, they are minor minerals in the Western Mineralisation and this conclusion is consistent with the present interpretation of this map. This is because the map shows the elements have contributed only 25.3 % of the total chemical variation of the Western Mineralisation and also that the elements show very weak loadings in this map (i.e. low potential for occurrence of minerals with those elements). Figure 5.18: Map of PC loading and PC loading. 2 4 5.7.6.6 Map of PC loading and PC loading (Figure 5.19) 3 4  Chemical variation PC loading and PC loading account for 22.1%16 of the total geochemical variation 3 4 of the Western Mineralisation.  PC loading and PC loading as chemical discriminators 4 3 In this map, PC loadingseparates As from other elements and PC loading appears 4 3 to be a discriminator of Cu from other elements. 16 Calculated by PPV (PC ) + PPV (PC ) from Table 5.13 3 4
ADE
186 Chapter 5- Multi-Element Relationships Figure 5.19: Map of PC loading and PC loading. 3 4 5.7.7 Map of PC scores A map of PC scores visualises the distribution of samples (observations) in ratios of two dimensional spaces of specific PCs. The projections of samples in different maps that resulted from two different PCs, constitute the possibility for evaluating the probability distribution of samples based on the geochemical interaction of the two PCs (Figure 5.20). The kth PC score of a sample is calculated by Equation (5.1): 10 PC score (PC loading  Element concentration ) (5.1) i k j k j j1 In Equation (5.1), the PC score is calculated for the ith sample (i =1 to 1,059 in this data set) and the kth PC (k =1 to 4), and j (j = 1 to 10) is the PC loading and element concentration respectively listed in Table 5.14. For example, the PC score for PC is 1 calculated as: PC score = 0.857 × Zn + 0.697 × Pb + 0.029 × Cu - 0.842 × Fe + 0.681 × S + 0.511 × i 1 Ag + 0.677 × Cd - 0.549 × As - 0.932 × Sb - 0.936×Bi
ADE
188 Chapter 5- Multi-Element Relationships positive scale of PC score) from Pb- and Ag- rich sulphide samples (at the negative scale 2 of PC score). The distribution of samples in Figure 5.20d does not show a linear 2 relationship relative to PC score and PC score. This indicates that almost all of the 2 3 samples are independent from the two PCs. The data sets of PC scores have been provided in supplementary file to this thesis. 5.7.8 Biplots of PC loadings and PC scores for 1,059 samples of the Western Mineralisation One of the most popular statistical approaches for the visualisation of compositional data sets in geochemistry is seen in ternary diagrams. It can be used effectively to evaluate the variability of three element concentrations or three oxide elements. However, the technique has difficulty in handling more than three components (Aitchison 1986), such as is the case in this study. For this reason, biplot diagrams are used for this study instead of ternary diagrams. Biplots (Aitchison & Greenare 2002; Gabriel 1971; Greenacre 2010) can display samples (points), element concentrations (vectors) and their degrees of correlation in a diagram. It is a very efficient tool to visualise multiple element concentrations (more than 3 elements) and samples in a low dimensional space. In a biplot, the lengths and directions of vectors are important for its interpretation. The cosine of the angle between two vectors approximates the amount of correlation between the two element concentrations. The length of a vector represents the standard deviation of the element concentration. A short length of a vector for an element concentration in a biplot indicates that the element is relatively constant in the data set, whereas a long one indicates a greater relative variation of the element (Labus 2005). The 1,059 clr-transformed values of this chapter were used for the construction of biplots of 10 element concentrations and 1,059 samples in relation to for PC and PC loadings using 1 2 Minitab software (Figures. 5.21 and 5.22).
ADE
190 Chapter 5- Multi-Element Relationships The distance between two vertices is approximately proportional to the variance of their log-ratios, ln (X / X). The vertices of Fe, Sb, Bi and As are close to each other which i j indicates the geochemical variation of these elements are closely correlated. Figures 5.21 and 5.22 demonstrate the following four groups of correlated elements in association with their PCs loadings: 1. Group 1: Zn, Cd and S have positive PC and PC loadings, 1 2 2. Group 2: Pb and Ag have positive PC loadings and negative PC loadings, 1 2 3. Group 3: Fe, As, Bi and Sb display clear negative PC loadings, and 1 4. Group 4: Cu shows a clear positive PC loading and very low correlation with other 2 elements. 5.8 Three-dimensional biplots for chemical composition of galena and sphalerite samples of the Western Mineralisation 3D biplots can also be generated in relation to element concentrations and sample distribution using the software packages of Matlab and CoDaPack3D for Excel. In these software packages, it is possible to rotate orthogonally the whole biplot in space and demonstrate the distribution of samples and elements within different angles. The spatial rotation helps to separate different distributed samples within elements into suitable angles. This helps to identify possible mineral generation, mineral alteration and remobilisation of one or more elements. For this investigation, 92 galena and 103 sphalerite samples from Kitchen (2001) and Patchett (2003) were used. Kitchen’s samples were collected from drill cores 4001 and 4002 and Patchett’s samples were collected from drill core 4003. The galena and sphalerite samples were analysed by measuring the atomic percentage of 11 elements using EMPA. The measured elements in galena and sphalerite have been shown in Figures 5.23 and 5.25 respectively. The data sets of Kitchen and Patchett have been provided in a supplementary file to this thesis. 5.8.1 Procedure for preparation of data for the CoDaPack3D In the CodaPack3D, when introducing the geochemical data, the summation of the atomic percentage of the elements should be 100. However, the summation does not always reach 100 percent and calculation is required to find the remaining percentage
ADE
Chapter 5- Multi-Element Relationships 191 value. That remaining value is calculated by the subtraction of the summation of other percentage element concentrations from 100. The title row of each column in a standard spreadsheet of CoDaPack3D was labelled according to the element concentrations and the remaining percentage value. The atomic percentage of 11 element concentrations and the remaining percentage value of each sample were inputted in columns of a standard spreadsheet of CoDaPack3D. This program calculates the clr-transformation of the constrained geochemical data automatically and generates a 3D biplot. Current CoDaPack3D does not show visually the three axes of PC , PC and PC in 1 2 3 a 3D biplot (Figures 5.23a, 5.24a, 5.25a and 5.26a) but by programming in Matlab software can show the PCs scales (Figures 5.23b, 5.24b, 5.25b and 5.26b). It should be noted that a 3D biplot is not limited only to three PCs in a tetragonal model and a 3D biplot can be defined and interpreted for more than three PCs, for example in a pentagonal or hexagonal model. 5.8.2 Galena Figures 5.23 and 5.24 show the same 3D biplot for galena but viewed from different angles to give a sense of distance between vectors and show their real lengths. Samples inside the red line of Figures 5.23a or 5.24a comprise a small number of galena samples identified by the influence of chemical variation of Fe and Zn. Figures 5.23 and 5.24 show two types of galena generation containing Zn- and Fe-rich galena (samples inside the red line) and Zn- and Fe-poor galena. The following observations can be drawn from Figures 5.23 and 5.24 which are consistent with the results from previous analyses: 1. Strong positive correlation among Pb, Bi and S, 2. Strong positive correlation between Zn and Fe, 3. Strong negative correlation of Fe and Zn other eight elements, 4. Moderate positive correlation of Ag, Sb and As, 5. Moderate positive correlation of Co and Cd, 6. Strong negative correlation of Cd with Ag, Sb and As, 7. Strong negative correlation of Co with Ag, Sb and As, and 8. High atomic variation of Zn, Fe and Cd in galena samples.
ADE
Chapter 5- Multi-Element Relationships 195 5.8.3 Sphalerite The biplot of the sphalerite samples is given in Figures 5.25 and 5.26. They show clearly two generations of sphalerite samples in the Western Mineralisation. One generation of sphalerite samples was strongly affected by the chemical variation of Bi (samples inside the red region) and they were enriched in Bi. The following observations can be drawn from Figures 5.25 and 5.26: 1. Strong positive correlation among Zn, Cd, Fe and S, 2. Strong positive correlation among Pb, Co and Cu, 3. Strong positive correlation among Ag and Sb, 4. Strong negative correlation of Bi with the other 9 elements, 5. Strong negative correlation of Ag with Co, Pb and Cu, 6. Strong negative correlation of Sb with Co, Pb and Cu, and 7. High atomic variation of Bi in sphalerite samples. Again, these observations agree well with the results from previous analyses. Table 5.16: Summary of decomposition sphalerite samples into seven PCs based on variation of eleven elements. PCs PPV PCV PC 29 29 1 PC 16 45 2 PC 15 60 3 PC 12 72 4 PC 11 83 5 PC 9 92 6 PC 7 99 7 According to Table 5.16, a combination of PC , PC and PC accounts for 60 % of 1 2 3 geochemical variation of sphalerite in the Western Mineralisation.
ADE
198 Chapter 5- Multi-Element Relationships 5.9 Summary In this chapter, simple correspondence analysis was used to provide a picture of the internal relationships between the chemical composition of the galena samples and the Broken Hill orebodies and between the chemical composition of the sphalerite samples and the Broken Hill orebodies. Bivariate and multivariate analyses were calculated for 1,059 assayed samples of each element concentration. In order to generate unconstrained data, the compositional data of elements were transformed to a clr log-ratio. The clr transformed data of the elements also showed normal distributions. The following conclusions can be drawn from the analyses conducted in this chapter: 1. The results of 3D correspondence maps showed that galena samples of the Western Mineralisation contain higher Bi, Fe, As and Sb, and sphalerite samples of the Western Mineralisation contain higher Pb in comparison with the other Broken Hill orebodies, 2. According to bivariate analysis, Zn showed a strong positive correlation with Cd and high negative correlations with Fe, Sb and Bi. Lead showed a strong positive correlation with Ag and high negative correlations with Fe, Sb and Bi, 3. Based on multiple linear regressions: 3.1 The three key predictor elements for Pb are Fe, Ag and Bi concentrations with an Adjusted R Square value of 0.619, 3.2 The three key predictor elements for Zn are Cd, Sb and Ag concentrations with an Adjusted R Square value of 0.816, and 3.3 This means that chemical variation of Pb and Zn in the Western Mineralisation can be estimated by chemical variation of Fe, Ag, Bi, Sb and Cd. 4. The cluster algorithm grouped 10 elements based on their similarity. Three groups were identified: group 1 shows high similarity among the elements of Zn, Pb, S, Ag and Cd; group 2 contains Cu only; and group 3 shows high similarity among the elements of Fe, As, Sb and Bi,
ADE
Chapter 5- Multi-Element Relationships 199 5. PCA was used to reduce the chemical complexity of 10 elements (10 dimensional spaces or 10 components) into four dimensional spaces or four PCs. The following four PCs were derived: 5.1: PC : Bi, Sb, Zn, Fe, Pb, S, Cd, As and Ag, 1 5.2: PC : Ag, Pb, S and Cd, 2 5.3: PC : Cu and Cd, and 3 5.4: PC : As. 4 6. The four PCs were considered as major geochemical discriminators for the sulphide mass of the Western Mineralisation: PC separated Zn and Pb from the other 1 elements, PC separated Zn from Pb, PC discriminated Cu from the other nine 2 3 elements and finally PC act as a good discriminator for As, 4 7. The biplot of PC and PC classified ten elements into the following groups 1 2 (Section 5.7.6): 7.1: Group-1: Zn, Cd and S, 7.2: Group-2: Pb and Ag, 7.3: Group-3: Fe, As, Bi and Sb, and 7.4: Group-4: Cu. 8. The 3D biplot of chemical composition of galena samples of the Western Mineralisation showed two generations of Zn- and Fe- rich galena and Zn- and Fe- poor galena samples, and 9. The 3D biplot of chemical composition of sphalerite samples of the Western Mineralisation revealed two generations of Bi- rich and poor sphalerite samples.
ADE
CHAPTER 6 Variogram Analysis for the Western Mineralisation 6.1 Introduction Statistical analysis for variables which are spatially correlated uses geostatistics rather than classical statistics. The technique uses a model (variogram) that describes the spatial geological or geophysical continuity of the variables of mineralised zone. The variogram model can be constructed provided there are sufficient samples available, generally the case for a mining operation. The geostatistical approach enables the quantification of structural and random variations of the variables in space, and more importantly the approach retains the effects of significant anomalous samples and therefore helps to identify the anomalous locations. Combining variogram models with techniques, such as kriging or simulation, geostatistics can then be used to generate the 3D block model for the orebody. Compared with classical statistics, such as multivariate regression, the block model is a more effective tool to visualise the continuity and distribution of grade variables in space. In general, mining engineers’ main interests are to model only economic elements in the mineralisation (e.g. Pb, Zn and Ag or combination of them) in order to estimate tonnage and grade of future production. For example, at Broken Hill, uneconomic elements such as Bi, As, Sb, Fe are usually not considered for any geostatistical analysis unless they are treated as contaminant elements in the final product (for which the smelter may charge a penalty). In this chapter, 136 directional variogram and down-hole variogram models were calculated for 43 variables (assays, minerals, rock types, magnetic susceptibility, specific gravity and sulphide textures) measured in the Western Mineralisation. The software used for these calculations is Geostatistics for Windows (Dowd & Xu 2006), a geostatistical package developed at The University of Adelaide. The variogram models derived are then used for the estimation of 43 block models using ordinary kriging. The specific objective of this chapter is to determine variogram parameters of the 43 measured variables, including the degree of spatial variability and grade continuity, the directional anisotropy and the proportion of between the structural and random variations of the variables.
ADE
202 Chapter 6-Variogram Analysis 6.2.1 Variogram model-spherical scheme In minerals applications, the variogram of grade values near the origin (i.e. zero distance) in general appears to be a linear. It increases with distance and eventually reaches a stable value (or fluctuating around a stable value) which is termed the sill value of the variogram. The sill value, in stationary case, is also equal to the total variance. The distance when the variogram reach the sill value is termed the range of the variogram. The initial variogram value when h0 [Equation (6.1)] is termed the nugget variance. The proportion of the nugget variance with the sill value is termed "nugget effect" (Dowd 2006a, p.81), which together with the slope of the variogram are the two most significant factors affecting the outcome of the kriging estimation. A variogram model fitting this type of variogram behaviour is called the spherical model (see the red dashed curve in Figure 6.1). Spherical model is the most popular variogram model used in minerals application. In Figure 6.1, three structural components are fitted to achieve the best-of-fit. Figure 6.1: An experimental variogram ("+") and a fitted spherical model (red dashed curve) for Zn concentration of the Western Mineralisation and its variogram parameters. Formulae used for calculation of the three structural components and the ranges of influence for a variogram model is given in Table 6.1.
ADE
Chapter 6-Variogram Analysis 203 Table 6.1: Formulae for calculation of variogram models with three structural components. γ(h)C C F(h,α )C F(h,α )C F(h,α ) for 0hα 0 1 1 2 2 3 3 1 γ(h)C C F(h,α )C F(h,α )C F(h,α ) for 0hα 0 1 1 2 2 3 3 1 γ(h)C C C F(h,α ) for α hα 0 1 2 2 1 2 γ(h)C C C C F(h,α ) for α hα 0 1 2 3 3 2 3 γ(h)C C C C for hα 0 1 2 3 3  3 3 h 1  h   F(h,a i)  2 α  2  α     i i  where C and a are the structural variance and the corresponding range for the i i ith structure and i=1, 2, 3 , C = The nugget variance , C , C , C = The 0 1 2 3 structural variance, C +C +C +C =Total variance or sill value of the 0 1 2 3 variogram, α ,α The range of influence, α The full range of influence 1 2 3 Nugget variance describes random variation of uncertain phenomena at small scale. Sampling errors and geological discontinuity will also contribute to the value of nugget variance. A high nugget effect is the case when the nugget variance is greater than 50 % of the total variance1 (Dominy, Stephenson & Annels 2001). In this case, the application of geostatistics will not have significant improvement in estimation over classical statistics. In this application, the nugget effect for all 43 variables (element concentrations, minerals, rocks and geophysical measurement) are significantly smaller than 50 %. 6.2.2 The range of influence The range of influence is an indication of the extent of spatial correlations. The maximum range is identified in a variogram model at the maximum distance when the variogram value reaches the sill (e.g. a in Figure 6.1). The spatial correlation of the 3 concentration at a distance greater than the maximum range of the variogram model is considered to be non-existence (Walter, Christensen & Simmelsgaard 2002). 1 Sill value
ADE
204 Chapter 6-Variogram Analysis 6.2.3 Advantages of application of geostatistics versus classic statistics Compared with geostatistics, classical statistics assume spatial independence of grade values i.e., any spatial correlation is disregarded. The variogram in this case is a horizontal line (Figure 6.2a) which equals to the total variance of the variable. This type of variogram model is termed the "pure nugget effect" (Carrasco 2010; Dominy, Stephenson & Annels 2001) with zero range of correlation. In this case, the search neighbourhood imposed in geostatistics for block estimation is irrelevant as no correlation is considered at any scale (Vann, Jackson & Bertoli 2003). In geostatistics, the general appearance of the variogram is shown in Figure 6.2b. The value of the variogram starts at a lower value and it will increase gradually with distance until it reaches the sill value, which is the total variance of the variable. The value of variogram smaller than the total variance indicates that there is spatial correlation for the distance as variogram is indirectly proportional to correlation. Figure 6.2: The change of variance (the red line) versus distance in (a) classic statistics and in (b) geostatistics. 6.2.4 Variogram calculations A variogram is directional. In order to calculate variogram at different directions, the following parameters should be defined: 1. The direction of variogram (azimuth and dip angles, or trend and plunge of the directional line). The directions of variograms are defined by an azimuth angle that is measured horizontally clockwise from north (0 to 360 degrees) and a dip angle that
ADE
Chapter 6-Variogram Analysis 205 is measured from the horizontal plane (-90 to 90 degrees) as positive (if the dipping direction is in the positive elevation direction). 2. Conical search angle (0-90 degrees) and maximum search distance. In practice, it is unlikely pairs of samples will be aligned exactly in the variogram direction. It is then necessary to determine an approximate tolerance angle  for capturing pairs of samples in a particular direction. This tolerance angle  help classify any pairs of θ samples within ± either side of a specified direction as being in that direction 2 (Figure 6.3). This angle is termed the angle of regularisation. Maximum search distance will effectively turn the conical search into a truncated conical search (Figure 6.3). Figure 6.3: Conical search volume and effect of direction approximation in a cone. 3. Lag distance for variogram. In Figure 6.3, all sample locations falling within a slice of width "dh" on the top of the cone are considered to be at a lag of "h" from sample "X".
ADE
206 Chapter 6-Variogram Analysis 6.2.4.1 Down-hole variogram Down-hole variograms refer to the variograms calculated along each drill hole. As samples are aligned closely with each drill hole, the variograms calculated in general will be the best candidates for identifying possible variogram structures. The down-hole variogram also help derive a more reliable estimate for the nugget variance for the related three-dimensional variogram model, which is in general difficult based on erratic three- dimensional experimental variogram (Guibal 2001). In order to work out variogram in a particular direction closely aligned with the drill hole direction, drill holes are classified (projected) on cross-sections and directions on the cross-sectional plane. The average of individual down-hole variograms is calculated for each specified direction within each specified cross-section and then those directional variograms are averaged over all cross-sections to determine the global average directional variograms of the entire mineralised zone or some specified part of it (Dowd 2006b, p 32). The drilling section of the Western Mineralisation is perpendicular to the strike of the mineralisation. The alignment of drill cores along drilling section makes easier to calculate directional variograms with minimum distance and angle approximations. The down-hole variograms in this case are calculated based on cross-sections perpendicular to the strike direction of the mineralisation (NNE-SSW). 6.2.4.2 Three-dimensional (3D) variograms in different directions The ultimate purpose of variogram modelling is to construct the three dimensional variogram structures that quantify the 3D spatial correlation. The full 3D variogram model can be constructed based on directional variograms calculated along different directions. The common representation for a 3D variogram model is using an ellipsoid with different lengths of axes, representing different variogram ranges in the three main axes: major, intermediate and minor. Major axis is the direction of longest range (maximum continuity) in the direction and minor axis is the direction of the shortest range (minimum continuity). In minerals applications, major axis normally coincides with the strike-plunge direction of the orebody, intermediate axis with the down-dip direction and minor axis with the cross-dip direction. This is also the case for the Western Mineralisation based on its 3D variogram calculations.
ADE
210 Chapter 6-Variogram Analysis Table 6.2: The approximate variogram model of Pb in raw data scale for the strike-plunge direction resulted from the variogram model of the logarithms of the data. σ2 C C C 22.910 213010 210010 2252.910 2 Logarithm , Pb 0 1 2 σ2  6.59 (%) 2 Raw data , Pb C 0, Pb σ2  22.910 2 6.59 (%) 2 0.60 (%) 2 σ2 Raw data , Pb 252.910 2 Logarithm , Pb C 13010 2 1, Pb σ2  6.59 (%) 2 3.40 (%) 2 σ2 Raw data , Pb 252.910 2 Logarithm , Pb C 10010 2 2, Pb σ2  6.59 (%) 2 2.62 (%) 2 σ2 Raw data , Pb 252.910 2 Logarithm , Pb 6.3 The strike, plunge and dip of the orebody in the Western Mineralisation The strike, plunge and dip of the orebody of the Western Mineralisation were estimated for 3D visualisation of the sulphide sample locations (Figure 2.2), observations of structural geology and variogram analysis. The primary estimated ranges were: 1. Strike direction of the orebody: between 5º and 25º from north, 2. Plunge of the orebody: between 15º and 35º from horizontal plane, and 3. Dip of the orebody: between 30° and 50° from horizontal plane. Figure 6.7 shows an average intersection of the mineralisation zone in a couple of dashed red planes in the Western Mineralisation which was simplified to depict the possible strike, plunge and dip of the orebody. Therefore, the spatial outline of the orebody is not a symmetrical plane. In Figure 6.7, plunge and dip are used as a term of structural geology. In structural geology, plunge is used for lineation (e.g. fold hinges, mineral lineations, cleavage and bedding intersections) and dip is used for planar features. Plunge and dip in structural geology always have positive values. More accurate orientation angles for the orebody of the Western Mineralisation (Figure 6.8) were determined based on derived directional variogram models for the three perpendicular planes of strike-plunge, down-dip and cross-dip (Figure 6.9).
ADE
212 Chapter 6-Variogram Analysis On the strike-plunge plane, the variogram ranges construct the ellipse as shown in Figure 6.9a. The major axis of the ellipse is the strike-direction of the orebody (azimuth 15°) and the minor axis is the dip-direction (285°). In the down-dip plane (Figure 6.9b), the variogram model will have an ellipse with the major axis representing the strike-direction of the 3D variogram model and the minor axis of the ellipse is in the intermediate axis. The intermediate axis of the 3D variogram model in the down-dip direction is worked out to be 40o. On the cross-dip plane (Figure 6.9c), the major axis of the ellipse is in the intermediate axis direction of the ellipse and the minor axis represents the cross-dip variogram range. Figure 6.9: (a) shows a horizontal plane ellipse that is used for displaying the strike-direction of the strike-plunge variogram and dip-direction of the down-dip variogram. (b) shows the vertical plane ellipse for the dip angles of the down-dip (-40°) variogram and strike-direction of the strike-plunge variogram and (c) displays the vertical plane ellipse for the dip angles of the cross-dip (50°) and the down-dip (-40°) variograms. It should be noted that a directional variogram model with an azimuth angle of X° is identical to the variogram model with the azimuth angle of X°+180°, provided the sign of the dip angle for the direction is also reversed. 6.4 The use of variogram ranges for the design of the optimal sampling grid A simple application of the derived variogram ranges is in the design of the optimal sampling grid which is cost effective. The optimal sampling grid derived based on the variogram models from the Western Mineralisation can help the design of an optimal
ADE
Chapter 6-Variogram Analysis 213 sampling grid for similar Zn-Pb-Ag type of mineralisation within the Broken Hill district or in other areas. Once the full 3D variogram model is constructed, it is possible to 2 calculate the optimal distance between sampling drill holes. As a rule of thumb, of the 3 variogram range is usually considered an appropriate distance to capture the spatial grade continuity of mineralisation (Flatman & Yfantis 1984, p.346). In practice, it is unlikely variogram anisotropies will coincide exactly with the coordinate system (easting, northing and elevation) used and the calculated sampling grid will have to be adjusted accordingly. The optimal surface geochemical sample spacing for detecting anomalous concentration and geochemical zonation in the Western Mineralisation are calculated and shown in Tables 6.3 and 6.4. Table 6.3: An appropriate surface geochemical sampling grid for Pb, S, Bi, Fe and Zn when the real dip of orebody is not clear for the Western Mineralisation. Elements The full range of influence (metre) Pb S Bi Fe Zn The full range of influence at the strike direction=15° 304.36 227.49 190.97 179.26 149.84 The full range of influence at the strike direction=105° 151.27 39.31 130.19 84.30 104.09 The full range of influence at the 2 strike direction=15° × 3 202.91 151.66 127.31 119.51 99.89 The full range of influence at the 2 100.85 26.21 86.79 56.20 69.39 strike direction=105° × 3 Anisotropy ratios 2.01 5.79 1.47 2.13 1.44 As discussed above, the longer variogram range on the plane is in the strike-plunge direction and the shorter range is in the cross-dip direction.
ADE
214 Chapter 6-Variogram Analysis Table 6.4: An appropriate surface geochemical sampling grid for Cd, Cu, Sb, As and Ag when the real dip of orebody is not clear for the Western Mineralisation. Elements The full range of influence (metre) Cd Cu Sb As Ag The full range of influence at the strike direction=15° 115.92 100.85 79.00 51.98 33.58 The full range of influence at the strike direction=105° 84.13 66.08 19.61 15.17 15.63 The full range of influence at the 2 strike direction=15° × 3 77.28 67.23 52.67 34.65 22.39 The full range of influence at the 2 56.09 44.05 13.07 10.11 10.42 strike direction=105° × 3 Anisotropy ratios 1.38 1.53 4.03 3.43 2.15 For example, in order to detect the extension of geochemical halo zoning for Zn in the Western Mineralisation, the optimal surface sampling grid should be about 100 m (99.89 m in Table 6.3) along the azimuth of 15° and 69 m (69.39 m in Table 6.3) along the azimuth of 105° (Figure 6.10). Tables 6.3 show that S has the greatest variogram anisotropy ratio. This indicates that spatial continuity of S concentration in the strike direction of 105o is 5.75 times that of the strike direction of 15o in the Western Mineralisation Figure 6.10: A schematic geochemical sampling grid for detection of Zn concentration.
ADE
Chapter 6-Variogram Analysis 215 It should be noted that different variogram anisotropies are detected for different elements in the Western Mineralisation. An overall optimal sampling grid in this case will be a compromise considering variograms for all elements, if the spatial variations for all elements are to be appropriately quantified. If all ten elements are considered in this example, the optimal sampling grid will be 10 m in the strike direction of 105° (the minimum range for As in Table 6.4) and 22 m in the strike direction of 15° (the minimum range for Ag in Table 6.4), 6.5 Comparison of variogram parameters of the Western Mineralisation with other Pb and Zn deposits One way to characterise the spatial structures and variations of mineralisation is application of variogram parameters (range of influence, nugget effect and sill values) derived for the elements concerned. Deposits with more irregular grade variations such as gold or vein type deposits may have a large nugget effect and a short range of influence. Relatively uniform deposits such as stratabound sedimentary Pb-Zn mineralisation are characterised by very low nugget variance and a large variogram range. Figure 6.12 compares published variogram parameters from other types of lead and zinc sulphide deposits such as Irish, Mississippi Valley-type deposits (MVT), sedimentary exhalative deposits (Sedex) and vein-type Pb-Zn deposits with variogram parameters of the Western Mineralisation.
ADE
216 Chapter 6-Variogram Analysis C Figure 6.12: The nugget effect ( 0 ) versus the full range of variogram for Zn and Pb Sill concentrations (modified from Wellmer 1998). Figure 6.12 shows that the Western Mineralisation is a more continuous mineralisation and has higher degree of spatial correlation relative to other lead and zinc sulphide ore deposits. Therefore, the use of geostatistical analysis for the Western Mineralisation is more appropriate technique in comparison with classic statistical methods. In Figure 6.12, the Western Mineralisation has a long range of influence and the nugget effect is only higher than the Mount Isa deposit in Australia. It should be noted that relative nugget effect is also related to the composite length (size) of core samples. Smaller nugget effect will be obtained for larger sample size (Guibal 2001). 6.6 Variogram anisotropy of different elements in the Western Mineralisation Variogram anisotropy reveals different spatial variability of the elements in different directions. The anisotropy can be represented by an ellipsoid where the major axis is in the direction of the longest range and the minor axis is in the direction of the shortest variogram range. In general, the longest variogram range coincides with the strike-plunge direction of the orebody, the intermediate range with the down-dip direction and the shortest variogram range with the cross-dip direction. However, this is not necessarily to be always the case. In the following discussion, the variogram ellipsoids for different elements are plotted using the following conventions (Figures 6.13 and 6.14):
ADE
Chapter 6-Variogram Analysis 219 3. b > a > c: the longest range "b" is in the down-dip direction of the orebody and the shortest range is in the cross-dip direction. Only the element of Sb is under this category. Anisotropy ratios can also be calculated for all the elements by Equation (6.3) in Table 6.5. These ratios can help to analyse the spatial continuity of the elements. For this exercise, the ratios are calculated for each ellipse of Figure 6.14 by dividing the major radius of ellipse by its minor radius. For example, the anisotropy ratios of Zn (Figure 6.14) are calculated in Table 6.5. Table 6.5: Formula used for calculation of the anisotropy ratios and its results for Zn concentration of the Western Mineralisation. Max. radius of the ellipse Anisotropy ratio of each ellipse (6.3) Min. radius of the ellipse a 149.8 At SD ellipse  1.4 b 109.7 a 149.8 At SC ellipse  5.5 c 27.2 b 109.7 At DC ellipse  4 c 27.2 The anisotropy ratios thus measure the ratios of the spatial correlation along the major radius to the spatial correlation along the minor radius on the corresponding plane. For instance, in the case of the SD ellipse, the anisotropy ratio of S is 4.6 indicating that the extent of spatial correlation of S along the major radius with direction of 015° from north and plunge of 25° is 4.6 times its minor radius with direction of 285° from north and plunge of 40°. For the SC ellipse, the anisotropy ratios of Pb, Cu, Cd, S, Fe and Zn are generally high. The highest value of 23 is observed for Pb. For the DC ellipse, the ratios show high values for Pb, Cu, Cd and Zn. The highest value is 11.53 for Pb in this ellipse. Bismuth, Sb, As and Ag have lower anisotropy ratios in comparison with those of the other elements.
ADE
224 Chapter 6-Variogram Analysis Table 6.6: Calculation of the anisotropic ratios 1 and 2 for Pb concentration. a 15.9 1, Down-dip   0.52 a 30.5 1, Strike -plunge a 150.7 2, Down-dip   0.49 The anisotropic ratio 1: a 304.3 2, Strike -plunge a  The first range of influence 1 a  The second range of influence 2 a 3.3 1, Cross-dip   0.11 a 30.5 1, Strike -plunge a 13 2, Cross-dip   0.04 The anisotropic ratio 2: a 304.3 2, Strike -plunge a  The first range of influence 1 a  The second range of influence 2 6.10 Block Model 6.10.1 Orebody outline The first step in the construction of block model is to define an orebody wireframe (skin) which is then filled with blocks that divide up the orebody. The orebody skin are constructed from orebody outlines defined on cross-sections based on drill hole projection on the cross-sections. During the 3D wireframe construction process, the Dijkstra algorithm method was selected in the Geostatistics for Windows as the algorithm ensure the surface reconstructed is optimal in the sense that the surface area is minimised (Xu & Dowd 2001). 6.10.2 Kriging parameters Data search strategy and search parameters are also important for the successful implementation of a kriging estimation regime. In this research, the largest variogram range, together with the anisotropic parameters described above, are used to define the data search neighbourhood for the kriging estimation. For example, 304 m is used to define the search radius for the estimation of Pb as that is the largest range for the variogram for Pb. For discrete block representation, 4×4×4 points are used.
ADE
Chapter 6-Variogram Analysis 225 6.10.3 Block size determination The use of appropriate block size is important for deriving a suitable block model to be used. If the block size is too large, the resolution of orebody representation will be too low, and to the contrary, if the block size is too small, the variation of estimated block values will be greatly reduced (over-smoothed). The determination of a suitable block size, however, is not an easy issue. From the geostatistical point of view, a suitable block size is ultimately determined by the sample spacing. From the mining operation point of view, the block size ideally should be identical to the actual selective mining unit to be used in the mining operation. The actual spatial continuity of the grad values also plays a part in the final decision. In the following section, it will try to find a suitable block size based only on geostatistics. Kriging variance in this case can be used as an effective tool to derive a suitable block size. For smaller block size, correlations between samples and block will be lower and the kriging variance is expected to be higher. A suitable block size can be found by examining the variation of kriging variance against different block sizes. As a rule of thumb, a suitable block size will be one that has the smallest possible size under the condition that the kriging variance is reasonable compared with larger block sizes. The results of kriging variances against different block size for three elements are given in Table 6.7 and Figure 6.17 below. Based on these investigations, the block size of 20 × 20 × 10 m (or the volume of 4000 m3) is considered to be the most suitable. Table 6.7: The selected discretisation grid, the corresponding volume and the mean value of kriging variance for three elements of Pb, Zn and Bi. Discretisation Volume The mean value of kriging variance grid (m) (m3) Pb (%2) Zn(%2) Bi (ppm2) 5 ×5×10 250 3.49 8.11 951.11 10 ×10×10 1000 2.97 7.16 895.35 15 ×15×10 2250 2.72 6.36 847.79 20 ×20×10 4000 2.36 5.78 810.85 25 ×25×10 6250 2.31 5.32 780.7 30 ×30×10 9000 2.19 5 752.06 40 ×40×10 16000 2.1 4.52 714.48 50 ×50×10 25000 2.05 4.21 679.04
ADE
226 Chapter 6-Variogram Analysis Figure 6.17: Visualisation of the mean value of kriging variance estimation versus different block volumes for Zn, Pb and Bi. It is interesting to note that the average drilling spacing in the Western Mineralisation is about 50 m and it is unusual for a block model with block size smaller than half of the drilling spacing. This is only suitable for cases where the variogram shows a low nugget effect and large correlation range i.e. mineralisation with high grade continuity, such as the case for the Western Mineralisation. The block size of 20×20×10 m is also suitable in this case for the identification of geochemical halo pattern, spatial variation of minerals, rocks and geophysical features which will be discussed in Chapters 7 and 8. The 43 block model as defined will produce in total of 424 horizontal and vertical cross-sections cutting through the orebody at different elevation and directions. Those cross-sections will be examined in more detail in Chapters 7 and 8 for spatial variability of all the 43 variables concerned. 6.10.4 Optimal number of samples for kriging estimation In general, fewer number of samples used for kriging will produce higher kriging (estimation) variance than greater number of samples, provided samples are all within the correlation range to the point to be estimated. However large number of samples will increase the computation cost and in some cases the improvement is negligible due to the screen effect of kriging. In this context, the number of samples to be used for kriging can be optimised. This can also be done by cross-validation. For this study, the cross- validations for the elements are run using 10, 20, 30, 40, 50, 60 and 70 numbers of samples. The average estimation errors for different cases are plotted against the number of
ADE
228 Chapter 6-Variogram Analysis interpretation of geological, geochemical and geophysical data and will be used in the next chapters for kriging estimations. Ten elements were selected to evaluate their degree of similarity based on their variogram parameters and their spatial anisotropic characteristics. The result showed that the 10 elements can be classified in three similar groups, including: 1. Group 1: Zn, Pb, S, Fe, Cu, Cd and Bi, 2. Group 2: Ag and As, and 3. Group 3: Sb. The variogram parameters for Pb and Zn in the Western Mineralisation were compared with those of other similar significant Pb-Zn deposits and the result showed that Pb and Zn concentrations in the orebody of the Western Mineralisation have a higher degree of spatial correlation or greater degree of continuity. This indicates that application of a spatial correlation tool, such as geostatistics rather than the classical statistics in this case is a more appropriate choice for the modelling of the variables. The variogram ranges of ten elements in the Western Mineralisation were also used to derive the optimal sampling grid for this mineralisation, which is found to be 22 m in the strike-plunge direction and 10 m in the cross-dip direction of the orebody. The analysis as described has not been performed before at Broken Hill or similar types of deposits. The study is the first to have comprehensive analyses of combined geological, geophysical and geochemical characteristics for an unmined orebody. Suitable variogram models for all 43 variables were calculated, modelled and cross- validated to produce suitable models to be used in the kriging estimation. For a few minerals, such as pyrite, arsenopyrite and red garnet, appropriate variogram models could not be produced due to insufficient number of samples. The kriging parameters for the linear estimation method chosen were also optimised. The optimal numbers of samples to be used for kriging were found by examining the kriging variance against the number of samples used. The optimal block size to be used to construct the block model for the deposit was found to be 20×20×10 m, which was calculated by examining the variation of kriging variance versus different block sizes.
ADE
CHAPTER 7 Spatial Geochemical Models for the Western Mineralisation 7.1 Introduction The study of spatial models of mineral deposits based on archived data of drill core from mine sites enables geologists to recognise and evaluate the scale of spatial correlation of each type of mineralisation using geological, geochemical and geophysical information. Thus, exploration guidelines for similar deposits can be developed on the base of existing data and their quantitative statistical interpretation. In spatial geochemical models, element concentrations are treated as spatial variables i.e. their variations are location dependent. A spatial geochemical model of an orebody can present significant support to exploration geochemistry when it is used for identification of the following issues: 1. Geometrical properties of geochemical haloes (e.g. distribution, size, orientation, shape and dimension), 2. The spatial variability of geochemical haloes with depth in different cross-sections, 3. Separation of threshold1 level from background and anomalous concentrations, and 4. Identification of zonation sequence in different directions. In this study, Western Mineralisation drill cores provide a valuable opportunity to evaluate dispersion and zonation of the primary geochemical haloes and their geological and geophysical associations for this type of Zn and Pb mineralisation. Since the spatial geochemical model of the Western Mineralisation is derived directly from its mineralised samples at different depths, it provides useful information for future geochemical survey. This approach is widely used in mining operations for the estimation of in situ mineral resource/reserve in relation to grade-tonnage of the orebody. However, it is still a rare practice to use the spatial models for the appraisal of zonation patterns of the orebody. More details concerning the geochemical halo zoning can be found in Beus and Grigorian (1977), Chen, Huang and Liang (2008), Chen and Zhao (1998), Grigorian (1974), Gundobin (1984), Huang and Zhang (1989), Kashirtseva (1967), Lawrie and Hinman (1998), Liu and Xu (1995) and Walters (1998). 1 Minimum anomalous value
ADE
230 Chapter 7-Spatial Geochemical Models The specific geochemical characteristics of the Western Mineralisation may be controlled by its structural environment (dislocations, faults and fractures) and formation of strike equivalent Broken Hill deposit. However, the different orebodies at the Broken Hill deposit may be generated and controlled by different geological and geochemical parameters with different scaling properties. For example, based on this study it is difficult to define a universal zonation halo system valid for the entire Broken Hill deposit. The main purposes of spatial geochemical modelling in this chapter are: 1. Construction of cross-sections for evaluating geochemical halo patterns of Pb, Zn, As, Cu, Fe, S, Sb, Bi, Ag and Cd in different directions, 2. Separation of the concentration range of threshold values from its background and anomalous levels in the orebody of the Western Mineralisation, 3. Quantitative comparison of the dimensional distribution patterns, orientation and anisotropies of the 10 geochemical haloes in order to determine zonation sequence of the orebody, and 4. Identification of pathfinder (indicator) elements associated with Pb and Zn ore of the Western Mineralisation. 7.2 Construction of cross-sections for evaluating geochemical halo patterns A zonation of a geochemical halo has a spatial nature and vectorial context that can be defined by the three following parameters: 1. Dimension (space), 2. Direction, and 3. Element concentration. The halo zonation can be a distinct spatial representation of the effects of the ore- bearing solution (Beus & Grigorian 1977) plus any effects of secondary redistribution and remobilisation (e.g. metamorphism and deformation). In order to study of the geochemical haloes and the zonation patterns of the steeply dipping mineralised zone of the Western Mineralisation, the following types of sections (Beus & Grigorian 1977) were constructed inside the 3D block models of the 10 elements: 1. Transverse sections to show the variation of halo patterns in the horizontal sections (Figure 7.1),
ADE
Chapter 7-Spatial Geochemical Models 231 2. Longitudinal sections to show the variation of halo patterns in north-south vertical sections along the strike of the mineralised zone (Figure 7.2), and 3. Axial sections2 to show the variation of halo patterns along east-west vertical sections (Figure 7.3). For identification of transverse zonation, the spatial models of the 10 element concentrations of the Western Mineralisation were intersected by transverse (horizontal) directions at an elevation of 10218 m close to the surface and an elevation of 9848 m next to the bottom of the 3D mineralised sample locations and two cross-sections between them at 10078 m and at 9958 m (Figure 7.1). Figure 7.1: The position of four transverse (horizontal) sections (dashed red rectangles) and locations of the mineralised drill core intersections. For demonstrating of the longitudinal zonation, 3D block models of the 10 element concentrations were cut vertically along north-south directions by two cross-sections at east = 9357 m and east = 9467 m. In general, two longitudinal sections were mapped for each of the spatial models (Figure 7.2). 2 Vertical zonation
ADE
234 Chapter 7-Spatial Geochemical Models Ziaii 1997; Miesch 1981; Sinclair 1974, 1976; Stanley 1988; Stanley & Sinclair 1989), fractal concentration-area method (Cheng 1999; Cheng, Agterberg & Ballantyne 1994; Cheng, Agterberg & Bonham-Carter 1996; Cheng, Xu & Grunsky 2000), the multifractal inverse distance weighted (Lima et al. 2003), the element concentration-distance method (Li, Ma & Shi 2003) and finally, spatial statistical methods such as kriging, moving average procedures and spatial factor analysis (Grunsky & Agterberg 1988). The spatial methods (geostatistical framework) such as moving averages and kriging have largely overtaken non-spatial methods because of consideration of sample size, sample locations, the scale of structural relationship of intrinsic variables and the degree of spatial continuity of mineralisation as well as their anisotropism. For determination of local or regional threshold, the following situations can be considered, depending on the number and location of samples within the mineralisation or non-mineralisation area and their amount of concentrations: 1. The first situation arises in regional geochemical prospecting for detecting secondary haloes (e.g. in soil and regolith). In this situation, most often the number of samples representing the regional background concentration is greater than the anomalous samples and recognition of a reliable regional anomalous grade is a very difficult practice. The regional background levels are usually spread broadly over an area and reflect regional geological processes with a wide-range of correlations. In this case, it would be necessary to separate the regional threshold concentration from a large number of background grades and to define this as maximum deviation from the regional background contents, and 2. Unlike the above situation, when a large number of samples are extracted from a mineralisation zone at different depths, the local threshold level should be separated from a large number of local anomalous samples. In this case, the local threshold level is deduced to be the minimum local anomalous content appearing in a mineralisation zone and the local anomalous grades are confined to the mineralised samples with a narrow-range of correlations. For this situation, the application of the spatial statistical methods is more efficient. For determination of local threshold in the Western Mineralisation, a similar concept of concentration-area (Cheng, Agterberg & Ballantyne 1994) was integrated with a
ADE
Chapter 7-Spatial Geochemical Models 235 3D kriged block model for each element. The combination of concentration-area and 3D kriged block model provide a very powerful and robust technique for geochemical anomaly separation and for minimising misclassification of threshold concentrations and background levels. 7.3.1 Procedure of separating threshold from background in the Geostatistics for Windows software Figures 7.4 to 7.7 show 10 geochemical haloes of the Western Mineralisation in different sections with a colour index distinguishing background level and threshold grade. The method used for detection of threshold level relies on adjusting the colour index of the Geostatistics for Windows software between minimum concentration (blue area) and a relative maximum concentration (red area) for each of the 10 elements. In this method, the relative maximum concentration of each element is changed manually in the colour index until the red area becomes the largest area (Figures 7.4 to 7.7). In this situation, a slightly increase or decrease of the relative maximum concentration causes an extreme reduction of the red area. The largest red area for each of the 10 elements was obtained from implementation of the above procedure at eight different cross-sections to identify the commonest threshold level among them. In colour indices of Figures 7.4 to 7.7, the minimum concentration of elements are not zero except for Pb, Zn and Cu in the Western Mineralisation and this suggests the elements will have larger dispersal haloes if geochemical sampling or the drilling network spread is broader in the vicinity of the Western Mineralisation. Although the local threshold level can be specified to the value of red colour of each element in Figures 7.4 to 7.7; however, because the regional threshold concentration is always lower than the local threshold concentration, so a range of concentrations between green and red (in colour index of Figures 7.4 to 7.7) were considered arbitrarily as the threshold concentrations range for each element, rather than one specific concentration.
ADE
240 Chapter 7-Spatial Geochemical Models The geochemical haloes in Figure 7.6 are characterised by an approximately 40° dip toward the west. Figure 7.7 show that the geochemical haloes have a southward-dip of about 25°. Although all the 10 geochemical haloes of the Western Mineralisation are related to common geological features, geochemical environment, structure and lithostratigraphy, they delineated various distinct intensifications and expansion dispersal halo patterns with changes of depth and distance (Figures 7.4 to 7.7). Antimony and Bi delineate maximum and most pronounced geochemical haloes in all cross-sections. In addition, the orientation and distribution shape of Sb and Bi haloes display a good conformity with the shape of Pb and Zn haloes. Therefore, they can be considered as possible pathfinder elements for exploration of Pb and Zn in the Western Mineralisation and similar types of mineralisation. Although there is a loose correlation between lode horizon rocks (quartz-gahnite, quartz-garnet, plumbian orthoclase, tourmalinite etc), such a correlation could not be determined in this statistical study. These lode horizon rocks are commonly interpreted as proximal to sulphides and their presence demonstrates a near miss. However, this may be flawed and the use of predictor elements such as Sb and Bi may be more fruitful. 7.4 Quantitative comparison of the geometrical characteristics of the geochemical halo patterns at different cross-sections of the Western Mineralisation The geometrical dimensions of the 10 analysed elements within the Western Mineralisation are distinctive and can be compared quantitatively with each other. For this purpose, the maximum width and maximum length of each geochemical halo were measured and their anisotropy ratios were calculated for each cross-section (Figures 7.4 to 7.7). The anisotropy ratio for each geochemical halo was calculated from the ratio of the maximum length of the geochemical halo to its maximum width. The maximum length and maximum width were measured in that part of the threshold ranges of each element that was marked by colours in the range from green to red. In Figures 7.4 to 7.7, a few geochemical haloes such as As and Ag display discontinuous dispersal haloes in their images. In these cases, the maximum length of their haloes was measured in exactly the same way as continuous geochemical haloes regardless of the gaps in their haloes. This is mostly because the need to understand the maximum
ADE
242 Chapter 7-Spatial Geochemical Models Table 7.1: Maximum lengths, maximum widths and anisotropy ratios on the transverse sections at different elevations (Figures 7.4 and 7.5). Elevations Transverse sections Length (m) Width (m) Anisotropy ratio Sb 900 300 3.00 Bi 855 240 3.56 Pb 780 165 4.73 Zn 600 165 3.64 10218 m Cu 585 120 4.88 Cd 510 105 4.86 Fe 495 120 4.13 S 495 135 3.67 Ag 450 67.5 6.67 As 435 75 5.80 Sb 1050 375 2.80 Pb 990 180 5.50 Bi 960 300 3.20 S 840 180 4.67 10078 m Zn 840 180 4.67 Fe 840 135 6.22 Cu 750 135 5.56 As 660 135 4.89 Ag 615 135 4.56 Cd 585 150 3.90 Sb 990 360 2.75 Bi 975 285 3.42 Pb 945 165 5.73 Zn 855 180 4.75 9958 m Fe 780 165 4.73 Cd 765 142.5 5.37 Cu 765 135 5.67 S 750 165 4.55 As 630 120 5.25 Ag 630 105 6.00 Sb 840 300 2.80 Bi 795 255 3.12 Pb 735 195 3.77 Zn 360 157.5 2.29 9848 m Fe 315 142.5 2.21 Cu 315 120 2.63 S 300 135 2.22 Ag 255 78 3.27 As 240 45 5.33 Cd No halo No halo No halo
ADE
Chapter 7-Spatial Geochemical Models 243 Table 7.2: The sequence of geochemical haloes based on their maximum lengths, maximum widths and anisotropy ratios resulted from Table 7.1. Elevation Maximum lengths 10218 m Sb > Bi > Pb > Zn > Cu > Cd > Fe = S > Ag > As 10078 m Sb > Pb > Bi > S = Zn = Fe > Cu > As > Ag > Cd 9958 m Sb > Bi > Pb > Zn > Fe > Cd = Cu > S > As = Ag 9848 m Sb > Bi > Pb > Zn > Fe = Cu > S > Ag > As Elevation Maximum widths 10218 m Sb >Bi > Pb = Zn > S > Cu = Fe > Cd > As > Ag 10078 m Sb > Bi > Pb = Zn = S > Cd > Fe = Cu = As = Ag 9958 m Sb > Bi > Zn > Pb = Fe = S > Cd > Cu > As > Ag 9848 m Sb > Bi > Pb > Zn > Fe > S > Cu > Ag > As Elevation Anisotropy ratios 10218 m Ag > As > Cu > Cd > Pb > Fe > S > Zn > Bi > Sb 10078 m Fe > Cu > Pb > As > S = Zn > Ag > Cd > Bi > Sb 9958 m Ag > Pb > Cu > Cd > As > Zn > Fe > S > Bi > Sb 9848 m As > Pb > Ag > Bi > Sb > Cu > Zn > S > Fe Figure 7.8 shows the results of Tables 7.1 to 7.2.  Maximum length According to Figures 7.8a, b, c, d, group (i) includes elements of Sb, Bi and Pb with greater lengths than groups (ii) and (iii). Group (ii) consists of Zn, Cu, S and Fe with greater lengths than group (iii) which contains As and Ag. Apart from Figure 7.8b, Cd can be attributed to group (ii).  Maximum width According to Figures 7.8e, f, g, h, elements of group (i) have greater widths and wider distribution relative to the elements of group (ii). snoitces esrevsnart eht gnola ecneuqes noitanoZ
ADE
244 Chapter 7-Spatial Geochemical Models  The anisotropy ratios The anisotropy ratios in Figures 7.8i, j, k, l compare the amount of dispersion of each element in two directions. In Figures 7.8i, j, k, l, Sb and Bi have relatively constant anisotropy ratios at four elevations. The anisotropy ratio of Sb varies between 2.75 and 3 and for Bi between 3.12 and 3.56. Accordingly, Sb and Bi are showing very low anisotropy ratios at all elevations and this highlights their greater dispersion within and around the Western Mineralisation in all directions and depths relative to other elements. At elevation 9848 m, Pb, Zn, Cu, S, Fe and Ag show a much lower anisotropy ratio relative to other elevations. This is because the maximum lengths of geochemical haloes are highly reduced at this depth while their maximum widths only vary a little.  Pathfinder elements In Figures 7.8a, c, d, e, f, g, h, at all elevations, Sb and Bi display a greater length and width relative to Pb and Zn. At elevation 10078 m (Figures 7.8b and 7.8f), Sb shows higher length and width relative to Pb and Zn and Bi larger dispersion than Zn; however, in Figure 7.8b, Bi shows shorter length relative to Pb. According to Figures 7.8a to 7.8h, Sb and Bi can be considered as geochemical pathfinders for the Western Mineralisation and similar Pb and Zn ores. By contrast, the amount of length and width of other elements is between moderate and small and they show shorter dispersion in comparison with haloes of Pb and Zn. This situation reduces appreciably their effectiveness and the reliability of their applications as geochemical pathfinders of the respective type of mineralisation. The threshold concentrations of Bi within the mineralisation zone in the Western Mineralisation are estimated to be between 3 and 5 ppm and for Sb, a range between 5.23 and 10 ppm is suggested. Figures 7.8a to 7.8h suggest that in lithogeochemical surveys of Pb and Zn mineralisation targets similar to the Western Mineralisation, the priority should be given to detecting threshold content of Sb and Bi or additive of Sb and Bi (Sb+ Bi) or composite (multiplicative) haloes (Sb × Bi) in surface sampling.
ADE
246 Chapter 7-Spatial Geochemical Models The composite and additive haloes intensify the extent of geochemical haloes and present closer correlations with the structure of the mineralisation relative to monoelement haloes. The application of multi-element haloes is more efficient in the mapping of weak geochemical anomalies particularly, in superficial geochemical fingerprints (Beus & Grigorian 1977). The multi-element haloes are more robust against random and analytical sampling errors and improve the contrast of halo zoning reliability and the geochemical interpretation (Beus & Grigorian 1977). Another important exploration guide for Western Mineralisation type ore is Sb and Bi in the sequence of transverse zonation. 7.4.2 Geometrical characteristics of the axial zoning haloes Table 7.3: Maximum lengths, maximum widths and anisotropy ratios on the E-W axial sections at N = 2109 m and N = 1639 m (Figure 7.6). Anisotropy North E-W axial sections Length (m) Width (m) ratio Sb 630 330 1.91 Bi 630 240 2.63 Pb 615 135 4.56 Zn 525 135 3.89 Cu 480 105 4.57 2109 m S 465 135 3.44 Fe 465 120 3.88 Cd 435 105 4.14 Ag 345 105 3.29 As 330 90 3.67 Bi 600 240 2.50 Sb 570 300 1.90 Pb 555 180 3.08 Zn 420 150 2.80 S 360 150 2.40 1639 m Cu 315 105 3.00 Fe 315 135 2.33 As 195 60 3.25 Cd 180 60 3.00 Ag 180 90 2.00
ADE
Chapter 7-Spatial Geochemical Models 247 Table 7.4: The sequence of geochemical haloes based on maximum lengths, maximum widths and anisotropy ratios resulted from Table 7.3. North Maximum lengths 2109 m Sb = Bi > Pb > Zn > Cu > S = Fe > Cd > Ag > As 1639 m Bi > Sb > Pb > Zn > S > Cu = Fe > As > Cd > Ag North Maximum widths 2109 m Sb > Bi > Pb = Zn = S > Fe > Cu = Cd =Ag > As 1639 m Sb > Bi > Pb > Zn = S > Fe > Cu > Ag > As = Cd North Anisotropy ratios 2109 m Cu > Pb > Cd > Zn > Fe > As > S > Ag > Bi > Sb 1639 m As > Pb > Cu > Cd > Zn > Bi > S > Fe > Ag > Sb Figure 7.9 show the results of Tables 7.3 and 7.4.  Maximum lengths In Figures 7.9a and 7.9b, group (i) comprises Sb, Bi and Pb with greater lengths than groups (ii) and (iii). Group (ii) contains Zn, Cu, S and Fe with greater lengths than group (iii) which consists of As and Ag. Cadmium can be attributed to either group (ii) in Figure 7.9a or group (iii) in Figure 7.9b.  Maximum widths In Figures 7.9c and 7.9d, group (i) includes Sb and Bi with greater widths than group (ii).  Anisotropy ratios Figures 7.9e and 7.9f show that the anisotropy ratios of Sb and Bi are lower than other elements. Hence, the anisotropic ratio of Sb is 1.90 and for Bi, the anisotropy ratios vary between 2.50 and 2.63. This means distribution of Sb and Bi occurs more easily and is larger than the other eight elements inside and around the mineralisation zone in the Western Mineralisation. snoitces laixa eht gnola ecneuqes noitanoZ
ADE
Chapter 7-Spatial Geochemical Models 249 7.4.3 Geometrical characteristics of the longitudinal zoning haloes Table 7.5: Maximum lengths, maximum widths and anisotropy ratios on the N-S longitudinal sections at E = 9467 m and E = 9357 m (Figure 7.7). East N-S longitudinal sections Length (m) Width (m) Anisotropy ratio Sb 1100 440 2.50 Bi 1100 360 3.06 Pb 1060 200 5.30 Zn 1000 200 5.00 9467 m S 840 220 3.82 Cu 840 140 6.00 Fe 840 200 4.20 Cd 680 160 4.25 As 660 100 6.60 Ag 660 100 6.60 Sb 1120 480 2.33 Bi 1080 340 3.18 Pb 1080 260 4.15 Zn 1000 200 5.00 9357 m Cu 960 140 6.86 Fe 940 220 4.27 S 940 280 3.36 Cd 880 120 7.33 As 640 100 6.40 Ag 600 100 6.00 Table 7.6: The sequence of geochemical haloes based on maximum lengths, maximum widths and anisotropy ratios resulted from Table 7.5. East Maximum lengths 9467 m Sb = Bi > Pb > Zn > S = Cu = Fe > Cd > As = Ag 9357 m Sb > Bi = Pb > Zn > Cu > Fe = S > Cd > As > Ag East Maximum widths 9467 m Sb > Bi > S > Pb = Zn = Fe > Cd > Cu > As = Ag 9357 m Sb > Bi > S > Pb > Fe > Zn > Cu > Cd > As = Ag East Anisotropy ratios 9467 m As = Ag > Cu > Pb > Zn > Cd > Fe > S > Bi > Sb 9357 m Cd > Cu > As > Ag > Zn > Fe > Pb > S > Bi > Sb eht gnola ecneuqes noitanoZ snoitces lanidutignol
ADE
250 Chapter 7-Spatial Geochemical Models Figure 7.10 shows the results of Tables 7.5 and 7.6.  Maximum lengths In Figures 7.10a and 7.10b, group (i) includes Sb, Bi, Pb and Zn with greater lengths relative to groups (ii) and (iii). Group (ii) contains Cu, S and Fe with greater lengths than group (iii) which consists of As and Ag. Cadmium can be attributed to either group (ii) in Figure 7.10a or group (iii) in Figure 7.10b.  Maximum widths In Figures 7.10c and 7.10d, group (i) includes Sb and Bi with greater widths relative to groups (ii) and (iii). Group (ii) encompasses Pb, Zn, Cu, S and Fe with greater widths than group (iii) which consists of As and Ag. Cadmium can be attributed to either group (ii) in Figure 7.10c or the group (iii) in Figure 7.10d.  Anisotropy ratios The dispersal elements (Sb and Bi) show very low anisotropy ratios in comparison with other dispersal elements (Figures 7.10e and 7.10f) thus, indicating that the elements’ distribution is widespread within the mineralisation zone. The anisotropy ratios of Sb range between 2.33 and 2.5 and, for Bi, vary between 3.06 and 3.18.
ADE
252 Chapter 7-Spatial Geochemical Models 7.5 Evaluation of similarity levels among the geometrical patterns of haloes Although the geometrical similarities of the geochemical haloes can be identified visually in Figures 7.4 to 7.7, one of the useful quantitative methods for identification of the similarity level is cluster analysis. The maximum lengths and widths of all geochemical haloes within all corresponding sections were used as entry data for the cluster analysis (Tables 7.1, 7.3 and 7.5). The cluster algorithm in Figure 7.11 classified the geochemical haloes into a number of groups to ensure that the similarity of haloes based on maximum length and maximum width inside each group is as greater as possible, while at the same time within the groups, the differences are as large as possible. The method of average linkage and correlation coefficient distance was used for amalgamation steps of the cluster algorithm using Minitab software. Figure 7.11 shows the amount of similarities among the spatial distribution of geochemical haloes in the Western Mineralisation. For example, S and Fe show very similar dispersal shape, orientation and dimension in 8 different cross-sections (Figures 7.4 to 7.7) and they show maximum similarity level in Figure 7.11 with 99.89 % as well. Figure 7.11: The percent of similarity among 10 elements based on the maximum lengths and the maximum widths of their geochemical haloes.