<<

CHAPTER APPLICATION OF MINERALOGY AND GEOCHEMISTRY TO MINERAL 20 EXPLORATION

M.B. McClenaghan and R.C. Paulen Geological Survey of Canada, Ottawa, ON, Canada

20.1 INTRODUCTION Glacial dispersal of debris from a specific bedrock source may be the product of a broad range of transport directions produced by evolving and shifting divides and dispersal centers, and may be identified by analyzing a range of particle sizes from the clay-silt till matrix, sand-size mineral grains, pebbles in till, up to boulders and large erratic blocks. Identification of glacial dispersal trains, combined with identification of glacial landforms and ice-scoured bedrock, is a powerful exploration tool in glaciated regions where glacial sediments mask the underlying prospective bed- rock. The term ‘drift prospecting’ is defined as the use of the geochemical, mineralogical, and lithological content of glacial sediments to trace metal-rich debris ‘up-ice’ to its bedrock source (DiLabio, 1990a). This chapter will focus solely on the most commonly used type of glacial sedi- ment for drift prospecting—till—because it is a first-order sediment deposited directly by . Glaciation in the last 2 million has profoundly affected and shaped much of the northern hemisphere’s landscape (Fig. 20.1) and small isolated areas in the southern hemisphere. For this reason, drift prospecting has become a common exploration method across the glaciated regions of the northern hemisphere and is only rarely used in the southern hemisphere. The intent of this chapter is to succinctly highlight processes and features of glacial dispersal and the application of this knowledge to mineral exploration in glaciated terrain. The first half of the chapter reviews ice-flow reconstructions and glacial dispersal, and explains the specifics of dis- persal trains. A relatively newly recognized consideration is the impact that ice streams have had on the glacial landscape, and how they affected glacial dispersal patterns. A basic understanding of subglacial processes, glacial dynamics, glacial process sedimentology (discussed elsewhere in this volume) is requisite to understand glacial dispersal models. Reconstruction of ice flow history in any prospective region is extremely important because ice-flow indicators provide insights into glacial history and erosional vigor that have impacted glacial landscapes. The second half of the chapter summarizes boulder tracing, till geochemistry, and heavy mineral methods that are applied to drift prospecting and is illustrated with several examples. The drift explo- ration methods used today have evolved over time as analytical methods and our knowledge of

Past Glacial Environments. DOI: http://dx.doi.org/10.1016/B978-0-08-100524-8.00022-1 © 2018 Elsevier Ltd. All rights reserved. 689 690 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

FIGURE 20.1 Extent of the glacial maximum in the Northern Hemisphere indicated by white polygons. Modified from Ehlers, J., Gibbard, P.L., 2007. The extent and chronology of Global Glaciation. Q. Int. 164À165, 6À20. glacial transport have advanced. The chapter builds on the earlier and still relevant reviews of explo- ration methods in glaciated terrain by Shilts (1976, 1996), Kauranne (1976), Bølviken and Gleeson (1979), DiLabio and Coker (1989), Coker and Shilts (1993),andMcClenaghan et al. (1997, 2000) and summarizes principles reported in greater detail in books by Kujansuu and Saarnisto (1990), Kauranne et al. (1992), McClenaghan et al. (2001),andPaulen and McMartin (2009).

20.2 HISTORICAL PERSPECTIVE Glacial applied to the search for mineral deposits can trace its roots to historical observa- tions of boulder transport in Fennoscandia that predates the glacial theory (e.g., Tilas, 1740; Sederholm, 1911; Sauramo, 1924). These early boulder tracing studies had a profound effect on the mineral exploration in Fennoscandia, resulting in the discovery of several mineral deposits. In North America, boulder tracing was employed in the late 1800s and early 1900s to search for sul- fide deposits (Burchett, 1944; Lundberg, 1957), iron (Miller, 1901), and gold (Slaght, 1893; Rickaby, 1932). Prest (1911) applied early principles of glacial transport and tracked gold grains in till up-ice to a gold-bearing quartz vein in eastern Canada. By the 1950s, boulder tracing had become a common exploration method in glaciated terrain (e.g., Grip, 1953; Dreimanis, 1958). 20.3 ICE-FLOW RECONSTRUCTION FOR MINERAL EXPLORATION 691

Geochemical analysis of the till matrix became a common tool in the 1970s and remains so today. The tracing of specific heavy minerals in till has been carried out since the early 1900s, but only became a common exploration tool in the 1990s. Boulder tracing, till geochemistry, and till miner- alogy have facilitated the exploration for, and discovery of, many important deposits and mines, some of which are listed in Appendix A.

20.3 ICE-FLOW RECONSTRUCTION FOR MINERAL EXPLORATION Exploration in glaciated terrain relies on knowledge of ice flow directions to trace glacial debris back to its bedrock source. Reconstruction of ice-flow patterns uses information about a variety of erosional and depositional landforms and features (Table 20.1; cf., Ryder, 1995; McMartin and Paulen, 2009)

Table 20.1 Summary of Ice Flow Indicators That Are Commonly Used for Reconstruction of Former Trajectories Agent of Width Landform Classification Formation Relief Length (% of Length) Roches moutonnees´ Bedrock landform Erosion 2À100 m 10À500 m 20À100 Whalebacks Bedrock landform Erosion 1À50 m 5À500 m 10À40 Bedrock steps Bedrock landform Erosion 1À5m 5À200 m 20À100 Rock Bedrock landform Erosion 5À150 m 50À1000 m 20À50 ridges Drift landform Deposition 3À50 m 50À500 m 20À40 Flutings Bedrock & drift Erosion 1À20 m 500À.5000 m ,10 landform Fluted till Drift landform Deposition 1À20 m 500À.5000 m ,10 Megascale glacial Bedrock & drift Both 1À10 m Several km ,5 lineations landform Bedrock & drift Both 5À100 m 25À1000 m 10À40 landform Transverse ice-thrust Drift landform Both 1À200 m 100À5000 m 30À75 ridges Hill-hole pair Bedrock & drift Both 10À200 m 100À2500 m 50À100 landform Glacial grooves Bedrock landform Erosion 5À100 cm Several m ,10À40 (outcrop) Striations (outcrop) Bedrock landform Erosion ,1cm 1cmÀ20 m Few mm deep Striations (boulder Drift landform Both ,1cm 1cmÀ2 m Few mm pavement) deep Bullet-shaped boulders Sedimentary Both 20À40 cm 30À200 cm 10À50 structure Pebble fabric Sedimentary Deposition ÀÀ À structure Till fabric Sedimentary Deposition ÀÀ À (micromorphology) structure

Modified from McMartin, I., Paulen, R.C., 2009. Ice flow indicators and the importance of ice flow mapping for drift prospecting. In: Paulen, R.C., McMartin, I. (Eds.), Application of Till and Stream Sediment Heavy Mineral and Geochemical Methods to Mineral Exploration in Western and Northern Canada. Geological Association of Canada, Short Course Notes 18, pp. 15À34. 692 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

observed in the field and on airborne or satellite imagery and, together, these various forms of ice flow data are used to reconstruct past ice flow trajectories (e.g., Veillette et al., 1999; McMartin and Henderson, 2004; Paulen et al., 2013). The glacial dynamics of former continental ice sheets were complex. Glacial flow was a time-transgressive event, such that the exposed surface of a mineral deposit could undergo intense subglacial erosion and directly contribute to a dispersal train and then be buried by newly formed till deposits and be cut off from further erosion. Yet, the dispersal of min- eralized debris can have continued through comminution and entrainment by ongoing ice flow. Thus, the type, provenance, and relative age of striations, streamlined landforms, and till deposits reflect the geographic and glaciological context in the ice sheet. Ice-flow reconstruction in alpine regions is strongly influenced by topography, keeping in mind that build-up of ice masses in alpine terrain could result in switches from topographically controlled glacial flow to flow controlled by glacial dynamics (Stumpf et al., 2000). Glacial flow reconstructions based solely upon landform associations identified by image analy- sis (e.g., Boulton and Clark, 1990; Shaw et al., 2010) that do not incorporate field observations and previous research, particularly the mapping of striae, may be inaccurate and inadequate. Field observations of erosional indicators and landforms are crucial to reconstructing an accurate model to support mineral exploration (cf., McMartin and Henderson, 2004; Veillette et al., 2017). An example of an ice flow reconstruction is shown in Fig. 20.2 from the Buffalo Head Hills kimberlite field, northern Alberta, Canada, in which data from striated bedrock and boulder pavements, ori- ented landforms, and pebble fabrics were used (Paulen and McClenaghan, 2015). Ice initially flo- wed southwest during the main phase of the and then flowed to the southwest and southeast during deglaciation as the ice sheet thinned. The implication of this reconstruction to diamond exploration in this prospective region is that both phases of ice flow must be considered when tracing kimberlitic debris in till back to source. A complex relationship exists between multiple till units (i.e., depositional history) and multiple ice flow phases (i.e., erosional history). Dispersed debris from a mineral deposit may be present in one, some, or all till units in a given region. Hirvas and Nenonen (1990) published various models illustrating the influence of glacial stratigraphy and combinations of flow trajectories on the trans- port of mineralized debris in till sequences. The provenance of multiple till units, and hence prove- nance of dispersed debris from mineral deposits, can be elucidated by determining and linking the glacial erosional and depositional records. This linking can be accomplished using a variety of field and compositional evidence. For example, Bird and Coker (1987) used till geochemistry and pebble lithology to decipher gold dispersal in multiple directions within multiple down-ice of the Owl Creek gold mine in central Canada. Stea and Brown (1989) used striation orientations, landform trends, pebble lithologies, boulder pavements, and till clast fabrics to decipher and link the glacial flow history to multiple till units in eastern Canada. Rice et al. (2013) utilized striations, distinct heavy minerals, pebble lithologies, till geochemistry, clast fabrics, and till micromorphology to link the erosional history to multiple and thick till units in support of leadÀzinc exploration in the Pine Point Mining District of northern Canada. Dispersal of kimberlite indicator minerals (KIMs) within multiple till units in a glaciated terrain dominated by ice-marginal lobes and surges was documen- ted by Paulen and McClenaghan (2015) in the Buffalo Head Hills kimberlite field (Fig. 20.3) using distinct heavy minerals, till fabrics, and striations. In these, and many more examples, the recogni- tion and sampling of multiple till units is crucial for exploration success in regions of thicker and/or complex glacial sediment cover. 20.3 ICE-FLOW RECONSTRUCTION FOR MINERAL EXPLORATION 693

58° N 117° W

10 km 114°30’ W 114°30’ 56°30’ N

FIGURE 20.2 Reconstruction of the Late Wisconsin Laurentide Ice Sheet flow patterns for the Buffalo Head Hills kimberlite field (kimberlite pipes are indicated as red diamonds), on a topographic digital elevation model. Large arrows represent steady-state flow during the Late Wisconsin glacial maximum, smaller arrows are flow paths from the Peace River and Wabasca River ice lobes during various stages of deglaciation. The dashed polygon indicates a region of ice stagnation characterized by hummocky stagnant ice (MS) and the white star indicates the site shown in Fig. 20.3. Modified from Paulen, R.C., McClenaghan, M.B., 2015. Late Wisconsin ice flow history in the Buffalo Head Hills kimberlite field, north-central Alberta. Can. J. Sci. 52, 51À67.

Research over the past two decades has advanced our understanding of the complexity of conti- nental ice sheets and ice-sheet dynamics. Ice-flow trajectories, and subsequently, glacial dispersal and erosion rates can have varied significantly throughout various regions of an ice sheet. Initial ice-sheet buildup, ice-divide migration, and coalescence of ice masses all may be reflected in gla- cial dispersal trains (Nurmi, 1975; Ha¨ttestrand and Stroeven, 2002). Klassen and Thompson (1993) 694 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

FIGURE 20.3 Two distinct tills sampled for kimberlite indicator minerals and their relationship to regional ice flow patterns. The lower, sandy orange till, derived from southwestward flowing ice and down-ice of the Buffalo Head Hills kimberlite field, contains .750 KIMs of different species of minerals, as compared to the upper dark-gray silty till that is KIM-poor and derived from southeast surging ice from the Peace River Lowland. Size of pie graphs indicates the relative KIM abundances: olivine (yellow), chromite (dark gray), Mg-ilmenite (light gray), eclogitic garnet (orange), chrome diopside (green), and pyrope (red). Clast fabrics are shown as bidirectional rose diagrams. Person at top of photo for scale. Modified from Paulen, R.C., McClenaghan, M.B., 2015. Late Wisconsin ice flow history in the Buffalo Head Hills kimberlite field, north-central Alberta. Can. J. Earth Sci. 52, 51À67. documented the dispersal of iron formation clasts derived from rocks immediately under the Quebec-Labrador Ice Centre of the Laurentide Ice Sheet. The various dispersal trajectories they documented reflect the complex shifts of the former (Fig. 20.4). McMartin and Henderson (2004) documented the complex chronology and migration of the Keewatin Ice Divide of the Laurentide Ice Sheet over prospective rocks in northern Canada. Ferbey and Levson (2009) documented glacial dispersal from a CuÀMo mineral deposit within a thick till sequence, in an area underlain by an ice divide of the , western Canada. 20.3 ICE-FLOW RECONSTRUCTION FOR MINERAL EXPLORATION 695

FIGURE 20.4 An amoeboid-shaped dispersal pattern of iron formation clasts in till dispersed from the Labrador Trough (red polygons). Ice flowed radially towards the northeast, southeast, southwest, and northwest from a migrating ice divide over the region. Modified from Klassen, R.A., Thompson, F.J., 1993. Glacial history, drift composition, and mineral exploration, central Labrador. Geol. Surv. Canada, Bulletin 435, augmented with data from Rice et al. (2017).

20.3.1 ICE STREAMS Ice streams are now recognized as prominent features of continental ice sheets that played a vital role in flow dynamics and mass balance, (e.g., Marshall et al., 1996; Stokes and Clarke, 2003; Evans et al., 2008; Stokes et al., 2015). Ice streams are corridors within an ice sheet that flow more rapidly than the surrounding ice; they form the arteries of ice sheets. Palaeo-ice streams discharged large amounts of ice over large distances and were the source of well-defined tracts of far-traveled, or exotic debris (e.g., Dredge, 2000; Stokes and Clark, 2001, Bennett, 2003; Dyke, 2008; Ross et al., 2009). Few studies have focused on dispersal patterns down-ice of mineralized rocks that have been formed or modified by ice streams, in part because former tracts were not previously recognized or identified in the northern hemisphere, most notably in areas covered by the Laurentide Ice Sheet in Canada. An inventory of Laurentide Ice Sheet ice streams recently published by Margold et al. (2015a, 2015b) should provide the impetus to re-examine many unex- plained or unsourced glacial dispersal trains in northern Canada, with a better understanding of 696 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

Ice

FIGURE 20.5 Ribbon-shaped glacial dispersal train of beryllium (Be) in the ,0.063 mm fraction of till trending northeast from the Strange Lake rare earth element deposit in eastern Canada. The train was formed by a single phase of ice flow towards the northeast. Modified from Batterson, M.J., Taylor, D.M., 2009. Geochemical re-analysis of till samples from the Strange Lake area, Labrador (NTS Map Sheets 14D/5 and 24A/8). Government of Newfoundland and Labrador, Department of Natural Resources, Geological Survey, Open File LAB/1479. glacial context (cf., Stokes et al., 2016). One example of debris dispersed by an ice stream is the plume of metal-rich till down-ice of the Strange Lake rare earth element (REE) deposit in eastern Canada (Batterson and Taylor, 2009). The ice stream produced elongate streamlined landforms enriched in debris from the deposit along a 40 1 km corridor down-ice (northeast) of the deposit (Fig. 20.5).

20.4 GLACIAL DISPERSAL TRAINS Glacial debris eroded from a discrete bedrock source is deposited down-ice in what is commonly referred to as a glacial dispersal train (DiLabio, 1990a). Dispersal trains are three-dimensional (3-D) bodies of till enriched in debris from a specific source relative to the till surrounding the train. They are hundreds to thousands of times larger than their bedrock source, and therefore pro- vide ideal targets for mineral exploration (Fig. 20.6). They commonly have abrupt lateral edges; a (A) Head Tail

Ice flow

Till Background Low Mineralization Bedrock Moderate High

(B)

ice flow 2 1 till Background Low mineralization bedrock Moderate High

(C) tail Head

Ice flow

Till Background Mineralization Low Bedrock Moderate High

(D)

Ice flow 1 2

Till Background Mineralization Low Bedrock Moderate High

FIGURE 20.6 Conceptual models of glacial dispersal trains showing a buried up-ice component, a head, and a tail, and can be modified or offset by subsequent ice flow: (A) thin till cover whereby mineralized bedrock is dispersed by a single phase of ice flow, creating a dispersal train that is exposed at the surface for its entire length and trends down-ice in one direction; (B) thin till cover whereby mineralized bedrock is dispersed by two phases of ice flow, creating a palimpsest dispersal train exposed at the surface and trending down-ice in two directions; (C) thick till cover whereby mineralized bedrock is dispersed by a single phase of ice flow, creating a dispersal train that is buried proximal to source and exposed at the surface farther down ice, and trends down ice in one direction; (D) thick till cover whereby mineralized bedrock is dispersed by two phases of ice flow, creating a palimpsest dispersal train trending down-ice in two directions. The train is buried proximal to source and exposed at the surface farther down-ice. The original dispersal train created by the older ice flow (1) shown in (A) or (C) is diluted and offset by the second ice flow (2) in (B) and (D). 698 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

contrast over a very short distance between low abundances outside the dispersal train and high abundances within it. Glacial dispersal trains occur at or below the till surface, generally displaced some distance down ice of their source. Abundance of dispersed debris decreases with increasing distance down ice such that a plot of abundance versus distance down-ice commonly approximates a negative exponential curve (Fig. 20.7; Shilts, 1976; Klassen, 1997). Within an ice stream, the concentration gradient of dispersed debris decreases linearly down ice (Fig. 20.8) and not exponen- tially, as the result of rapid ice flow transporting debris far from source with little dilution

FIGURE 20.7 Plots of nickel abundance in till collected at surface in the Thetford Mines area, central Canada. Actual (top) and idealized (bottom) plots show the relationship between the source, head, and tail of a dispersal train. Modified after Shilts, W.W., 1976. Glacial till and mineral exploration. In: Legget, R.F. (Ed.), Glacial Till: An Interdisciplinary Study. Royal Society of Canada, Special Publication, vol. 12, pp. 205À223. 20.4 GLACIAL DISPERSAL TRAINS 699

FIGURE 20.8 Plots of thorium abundance in surface till down-ice of the Strange Lake REE deposit in eastern Canada. The top plot shows data from Batterson and Taylor (2009), with the best fit data curve shown as a dashed line. The lower plot is an idealized dispersal train that shows the relationship between the source, head, and tail of a dispersal train formed by a palaeo-ice stream. Lower plot is modified from Klassen, R.A., 1997. Glacial history and ice flow dynamics applied to drift prospecting and geochemical exploration. In: Gubins, A.G. (Ed.), Proceedings of Exploration 97, pp. 221À232.

(e.g., Klassen, 1997; Dredge, 2000). Dispersal trains have a ‘head’, the highest abundance of dis- persed debris at surface, which is commonly, but not universally, overlying or slightly down-ice of the mineralized source. The ‘tail’ of the train is demarcated by the lower abundance of dispersed debris that has been transported and deposited many times farther down-ice than the head, and is commonly the part of the train that is detected first in an exploration program (DiLabio, 1990a). 700 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

Glacial dispersal trains are the net result of glacial erosion of mineralized bedrock sources, transport, and subglacial deposition, all of which can be complicated by multiple cycles of till re-entrainment and deposition, as glacial dynamics change over time. The morphology of dispersal trains is influenced by a number of factors that can be grouped into four broad categories: (1) the bedrock source (composition, hardness, contrast with host lithologies, topographic position, protection from glacial erosion); (2) glacial conditions (basal ice temperature, availability of water which relates to erosion potential of the , ice velocity, ice-flow trajectories); (3) transport and depositional environment (subglacially by lodge- ment, supraglacial transport, meltout processes, etc.); and (4) postdepositional changes (e.g., metal remo- bilization by hydromorphic processes, mineral weathering, downslope movement, etc.). A single phase of ice flow typically produces a ribbon-shaped dispersal train as wide as the bed- rock source (Fig. 20.5) that is defined by sharp, straight edges (e.g., McClenaghan et al., 2002a; Batterson and Taylor, 2009). The term ‘palimpsest’ is used to refer to glacial dispersal patterns that are the net result of more than one ice flow event (Stea, 1994; Parent et al., 1996; Stea et al., 2009). Palimpsest dispersal trains have several forms. A change in ice flow direction can: (1) erode the ribbon entirely, (2) rework the ribbon into a fan-shaped train that occurs between the two ice flow vectors (Fig. 20.9) (e.g., Paulen et al., 2013), (3) create a separate new ribbon in a different orientation that results in a bilobate fan (e.g., Charbonneau and David, 1993; Stea et al., 2009) (Fig. 20.10), or (4) rework dispersed glacial debris into stellate or amoeboid-shaped trains through multiple ice flows in very different directions or flow reversals, often associated with ice divides (Klassen and Thompson 1993; Shilts, 1993; Stea and Finck, 2001)(Fig. 20.4). Some limiting factors when exploring for dispersal trains and other patterns include: (1) dispersal may not be evident or detected because mineral-rich debris was glacially transported down-ice beyond the range of sampling area, (2) the bedrock source may not have been subjected to glacial erosion (e.g., buried by glacial or preglacial sediments or covered by younger bedrock units), (3) the mineralized rocks may not contain sufficient volume of indicator minerals to produce a detectable train, and (4) some dispersal trains may have gaps, or be missing the head or some part of the tail due to differential erosion of the mineralized source (e.g., removal of a regolith or gossanous cap), erosion and entrainment of parts of the train by subsequent ice flows, or dilution and masking of a train by till derived from a nonmineralized bedrock source. Dispersal trains form from the transport of debris over distances ranging from tens of meters to tens of kilometers. The scale of dispersal is linked to regional ice flow dynamics through landformÀsediment associations. Transport distances differ spatially with respect to the ice sheet margins and ice divides, and even with landforms. Examples from Finland have shown that glacial transport distances can vary between hummocky and drumlinized terrain (Salonen, 1988). Systematic variations in glacial transport distance can be difficult to define. Shilts (1984) defined four scales of glacial dispersal which have been updated and summarized below: 1. Continental dispersal, which is measured in hundreds to more than a thousand kilometers down- ice from the source and is typically detected using a distinctive bedrock lithology or mineral. An example is the dispersal of the visually distinctive Omarolluk Formation clasts (‘omars’) from the Belcher Islands in Hudson Bay, Canada (Prest, 1990). 2. Regional dispersal, which is measured in tens to hundreds of kilometers down-ice from the source. An example is the dispersal of carbonate detritus by ice streams within the Laurentide Ice Sheet (Hicock, 1988; Dredge, 2000) or at Strange Lake (Fig. 20.5). 20.4 GLACIAL DISPERSAL TRAINS 701

FIGURE 20.9 Fan-shaped glacial dispersal train of (A) tin (Sn) and (B) tungsten (W) in whole till formed by two phases of ice flow to the southeast and to the southwest from the East Kemptville tin deposit in eastern Canada. Modified from Rogers, P.J., Chatterjee, A.K., Aucott, J.W., 1990. Metallogenic domains and their reflection in regional lake sediment surveys from the Meguma Zone, southern Nova Scotia, Canada. J. Geochem. Explor. 39, 153À174. 702 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

65°54’10” N Pyrope garnet + chrome diopside grain counts

0 – 1 111°00’00” W 111°00’00”

2 – 10

11 – 100

101 – 500

501 – 2000

Glacial flow indicators Striae Contwoyto Lake

Streamlined landforms Roche moutonnées 110°38’30” W 110°38’30” 65°45’25” N

FIGURE 20.10 Bilobate glacial dispersal train (outlined in green) of kimberlite indicator minerals in till down-ice of the Contwoyto-1 kimberlite (red star) in northern Canada formed by two phases of ice flow, first to the southwest and subsequently to the north. Modified from Stea, R.R., Johnson, M., Hanchar, D., 2009. The geometry of KIM dispersal fans in Nunavut, Canada. In: Paulen, R.C., McMartin, I. (Eds.), Application of Till and Stream Sediment Heavy Mineral and Geochemical Methods to Mineral Exploration in Western and Northern Canada. Geological Association of Canada. Short Course Notes 18, pp. 1À13.

3. Regional to local dispersal, which is measured in kilometers to ten kilometers down-ice from the source. Examples include the dispersal of visually distinct minerals from kimberlites in diamond exploration programs, such as those seen in northern Canada (Fig. 20.10). 4. Local dispersal, which is typically measured in meters to hundreds of meters and commonly is at the scale of an exploration property. 20.4 GLACIAL DISPERSAL TRAINS 703

Till thickness, which ranges from less than a meter to tens of meters thick, may also impact the surface footprint of a dispersal train, and may reduce the lower threshold at which dispersal can be detected. A dispersal pattern in till that is extremely thick may have a very diluted signal, a long transport distance, and potentially a complicated glacial history. Several case studies demonstrate the erosion and uptake of mineralized rock from a source, transportation, and deposition in an upward-climbing fashion within increasing distance down-ice (Shilts, 1976; Drake, 1983; Miller, 1984; Klassen, 1997). To better understand how a dispersal train is formed, the transport capacity and first-order pat- terns of glacial erosion and deposition for a nonsteady-state ice sheet during a glacial cycle must be considered. A measure of glacial erosion intensity is often sought as a proxy for ice-sheet configu- ration (Hildes et al., 2004; McMartin and Henderson, 2004; Kleman et al., 2008). In areas affected by warm-based glaciation, streamlined landform density (Ha¨ttestrand et al., 1999), elongation ratio (Hart and Smith, 1997), inheritance of terrestrial cosmogenic nuclide ratios (Fabel et al., 2002; Gosse et al., 2006; Margreth et al., 2016), and streamlined landform characteristics (Bradwell et al., 2008) have been used to determine erosion intensity. The intensity of glacial erosion is an impor- tant consideration for mineral exploration because regions subjected to a higher intensity of erosion can be expected to produce longer dispersal trains where the till was deposited as a thin veneer (,2 m). In regions affected by high-intensity erosion and covered by thick till (.5 m), dispersal trains will be diluted within a short distance down-ice and thus dispersal trains will be shorter. Palimpsest dispersal trains are uncommon in areas that were affected by higher erosion regimes. Regions subjected to a low intensity of erosion, such as those occurring under ice divides, are expected to have older relict landforms preserved and palimpsest dispersal trains. Till deposits, and, in particular, glacial dispersal trains, can also be a proxy for the evolution of erosion intensity. Indicator clasts within till can delimit transport distances (Dreimanis and Vagners, 1971; Dredge, 1988; Kjær et al., 2003; Plouffe et al., 2011; Trommelen et al., 2013). The magnitude of clast, mineral, or elemental dispersal signals depends on the net erosion intensity up-ice at the source, compounded with other factors such as bedrock lithology, duration of glacial transport in one or more directions relative to the source area, inheritance of debris from older deposits, subse- quent burial of the source, and dilution of mineralized debris through overprinting and entrainment (Aario and Peuraniemi; 1992; Parent et al., 1995; Stea and Finck 2001; Evans et al., 2006). Adjacent to an ice-divide, subglacial erosion intensity was probably highly variable, owing to additional factors such as initial ice buildup, ice-divide migration, and changes in mass balance with advance and retreat cycles (Kleman et al., 2008). These factors are difficult to assess, how- ever, in the interpretation of the evolution of erosional intensity, especially where till cover is wide- spread, scoured lake basins are rare, and subglacial landform flowsets are fragmented, rare, or heavily overprinted. The formation and distribution of large erosional landforms underneath former ice sheets are primarily governed by the subglacial phase state. Glacial erosion occurred when ice sheets of a given thickness and ice flow rate reached pressure melting conditions (warm-based), typically as ice flow accelerated towards the ice margins. These warm-based zones provided mate- rial for deposition for a polythermal ice sheet under steady-state conditions. There is potential for long-distance transport of subglacial debris during ice expansion and deglacial phases, but such material would have to be transported at higher velocities to remain in the migrating entrainment and transport zones of an ice sheet (Kleman et al., 2008), particularly overlying soft sediment beds (Spagnolo et al., 2016). The combination of a relatively short duration of an ice sheet contraction 704 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

phase and inward migration of erosion zones would inhibit long-distance transport during deglacia- tion, except for ice-stream corridors. Over the past 25 years, the tracing of physically robust and distinct minerals, known as ‘indica- tor minerals’ (see detailed explanation below), has provided new insights into glacial dispersal pat- terns and transport distances where the traditional methods of boulder erratics and till geochemistry have failed. The dispersal train from the Izok Lake volcanogenic massive sulfide (VMS) deposit in northern Canada is one of example of this. Here, indicator minerals were detected at least 40 km down-ice from the deposit but till geochemistry was only able to detect elements derived from the deposit up to a maximum of 6 km down-ice and boulder tracing detected clasts dispersed up to 1 km down-ice (McClenaghan et al., 2015a).

20.4.1 DISPERSAL MODELS Various models are used to estimate the length of down-ice dispersal distances and evaluate how they covary with glacial entrainment and deposition processes. Quantitative numerical models that predict glacial dispersal have been summarized by Stea and Finck (2001). These models include the exponential decay model (e.g., Shilts, 1976; Bouchard and Salonen, 1989; Parent et al., 1996; Klassen, 1999) which is commonly used to describe glacial dispersal in basal tills (Shilts, 1973; Levson, 2001), and the linear decay model (Dyke and Morris, 1988; Klassen, 1997; Batterson and Liverman, 2000; Primmer et al., 2015) which is more commonly used to describe englacial dis- persal in basal melt-out tills (Klassen, 2001), fast flowing ice (Dredge, 2000), and in regions of higher topographic relief (Paulen, 2001; Ozyer, 2011). Exponential decay models describe both enrichment of till by anomalous material and dilution of till by background material using negative exponential functions for enrichment and dilution. Most models illustrate the functional forms of these exponential equations including how the expo- nential model half-distances, also known as ‘renewal distances’, are associated with different decay constants (Bouchard and Salonen, 1990; Finck and Stea, 1995). These exponential decay model half-distances are inversely related to the ‘entrainability’ of the bedrock beneath a glacier and the basal ice velocity (Dyke, 1984; Clark, 1987; Aario and Peuraniemi, 1992). These decay constants and their associated half-distances can vary significantly with substrate lithology and hardness (Bouchard and Salonen, 1989; Puranen, 1990; Charbonneau and David, 1995). Linear decay models describe enrichment and dilution in till across a mineralized source using linear functions for enrichment and dilution (Klassen, 1997). Like the exponential decay models, the linear decay model half-distances are also inversely related to bedrock entrainability. Both of these models assume that no loss of entrained material, such as loss from ascension (shearing) into the englacial environment, took place down-ice from the source. However, if material already entrained within the glacier was lost, these increasing rates of entrainment of mineralized material would not be required with distance down-ice to achieve exponential or linear enrichment. Unfortunately, for the loss of entrained material to cause sufficient enrichment such that constant entrainment from bedrock was possible, the material lost would have to have been selectively com- posed of background material, a condition that became increasingly unlikely as the composition of the entrained material became more anomalous with increasing distance down-ice from a background-mineralized rock contact (because anomalous material was being added to the englacial load). This statement does contradict the linear and exponential decay models that appear to 20.4 GLACIAL DISPERSAL TRAINS 705

empirically fit geochemical, mineralogical, and lithological concentration data down-ice from known sources. However, Stanley (2009) stated that both models have mathematical consequences that are inconsistent with glacial processes and suggested that the exponential and linear decay models fail to adequately or reliably describe glacial dispersal trains. Hooke et al. (2013) argue that a thermal regime of basal meltwater refreezing, followed by shearing and folding of basal ice layers broaden and thicken a dispersal plume down-ice, which is later deposited. However all of these models fail to account for dispersal trains with linear enrichment and dilution patterns, especially those deposited by fast-flowing ice (Dredge, 2000). A dispersal model that appears to be a ‘best-fit’ with observed glacial dispersal trains is the aggradational-constant entrainment decay model (Stanley, 2009). This model treats units of engla- cial debris in ice and subglacial material (deformable bed) at the base of the glacier as ‘cells’. Each cell contains a specific geochemical, mineralogical, or lithological composition and is consistent with overall flow mechanisms of ice. These cells form the basal part of the glacier, including the deforming bed, as a series of layers that are stacked vertically on top of each other. Each layer is composed of cells that are adjacent to each other, and material is entrained into these cells from bedrock by regelation, glacial creep, internal thrusting, or abrasion, or from subjacent cells by inter- nal shearing processes, and then transported upwards into overlying layers by internal shearing. The amount of material entering a cell from below is equal to the amount of material transported upward to the overlying cell (Fig. 20.11). Thus, as the glacier moves, the resultant dispersal train is transported down-ice and upward through the glacier. This migration and mixing of metal-rich and

Ice flow 10 00000000236810111110 9 00000001358101111119 8 000000025810121211108 7 000000148111313121086 6 000000371114141311864 5 0 0 0 0 0 1 6 12 16 16 14 11 8 6 4 3 4 0 0 0 1 1 2 12 18 18 16 12 8 6 4 2 1 Till layers Till 3 0 0 1 3 6 9 21 22 18 13 8 5 3 2 1 0 2 0 2 7 1319263021137 4 2 1 1 0 0 1 0 173244536127125 2 1 0 0 0 0 0 0 100 100 100 100 100 0000000000 BedrockMineralization Bedrock

FIGURE 20.11 Concentration (as a percentage) of metal-rich debris in till to illustrate the aggradational-constant entrainment decay model down-ice from mineralized bedrock. Mineralized bedrock is initially eroded and smeared down-ice as till layers are accreted, the size (length and width) of the dispersal train increases down-ice, and the concentrations of metal-rich material decrease by dilution down-ice. Modified from Stanley, C.R., 2009. Geochemical, mineralogical, and lithological dispersal models in glacial till: physical process constraints and application in mineral exploration. In: Paulen, R.C., McMartin, I. (Eds.), Application of Till and Stream Sediment Heavy Mineral and Geochemical Methods to Mineral Exploration in Western and Northern Canada. Geological Association of Canada. Short Course Notes 18, pp. 35À48. 706 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

metal-poor material causes the metal-rich debris to be diluted and dispersed upwards in the till pro- file, as it is transported down-ice and upwards through the glacial sediment sequence. As a result, material eroded from mineralization is effectively dispersed by down-ice movement such that, when subsequently deposited beneath the glacier, it produces a dispersal train within the till layers. This aggradational-constant entrainment decay model is geometrically consistent with the descrip- tive 3-D model described by Miller (1984), and 3-D dispersal trains that have been described (e.g., Nurmi, 1976; Averill, 2013; Oviatt et al., 2015; Averill, 2017). Application of these models, in combination with reconstruction of ice flow history in a given prospective region, will help mineral exploration geologists better interpret dispersal trains at surface, tighten their up-ice search vector, and provide insight into transportation distances in regions of thicker till (Fig. 20.6).

20.5 MINERAL RESIDENCE SITES IN TILL Till is the end-product of glacial erosion, transport, comminution, and deposition and consists of two particle size modes: (1) rock fragments or clasts (.1mm),and(2)matrix(,2 mm) that is dominated by mineral fragments (Dreimanis and Vagners, 1971). The minerals in the till matrix are crushed and abraded to their terminal grain size mode. The characteristic terminal grain size mode for each min- eral is determined largely by the original grain size, hardness, cleavage, and the cementation or meta- morphic history of the mineral in its bedrock source (Shilts, 1993). In till, quartz, magnetite, feldspars, pyroxenes, and other minerals commonly dominate the sand-size fraction, carbonates domi- nate the silt-size fraction, and phyllosilicates and hematite dominate the clay-sized fraction. Thus, the bulk chemical compositions of different-size fractions of till reflect the dominant minerals in each size fraction. In experimental studies involving the genesis of the fine fraction of tills, Ma¨kinen (1995) observed that as the degree of sediment grinding increased the contents of some metals (cobalt, copper, iron, manganese, nickel, and zinc) decreased in comparison to the overall composi- tion of the coarser fraction. Elements that are common in coarser sand-size fractions include chromium typically in chromite, tungsten with scheelite, and tin with cassiterite (Shilts, 1993). Base metals, gold, and platinum group element contents tend to be higher in the fine (,0.06 mm) till frac- tion (Nevalainen, 1989; Salminen et al., 1989; DiLabio, 1995). Exploration programs should select sample preparation and analytical protocols based upon the target mineral deposit, the elements con- tained within or associated with it, and the likely terminal mode of ore minerals. Doing so will achieve the strongest signal:noise ratio that will better outline the head and tail of a dispersal train or other pattern.

20.6 EXPLORATION METHODS 20.6.1 BOULDER SAMPLING Mapping the distribution of mineralized boulders, cobbles, or pebbles, collectively referred to as ‘boulder tracing’, is one of the oldest exploration methods in glaciated terrain (e.g., Milthers, 1909; Dreimanis, 1958; Hyvarinen et al., 1973; Saltikoff, 1984; Sarala and Peuraniemi, 2007). Since the late 1800s, boulder tracing has provided the most obvious and direct indicators of the presence of 20.6 EXPLORATION METHODS 707

48°51’N 91°33’W

N

5 km

Ore outcrop > 2% ore pebbles in till < 2% ore pebbles in till Limits of main train Extreme limits of train Striations 92°01’N 48°32’N

FIGURE 20.12 Distribution of iron formation clasts in till down-ice of the Steep Rock Lake iron deposit in central Canada. Modified from Nichol, I., Bjo¨rklund, A., 1973. Glacial geology as a key to geochemical exploration in areas of glacial overburden with particular reference to Canada. J. Geochem. Explor. 2, 133À170, after Dreimanis, A., 1958. Tracing ore boulders as a prospecting method in Canada. Can. Inst. Min. Metallur. Trans. 61, 49À56. mineralization or a specific bedrock lithology (e.g., kimberlite, carbonatite, etc.). Boulder tracing has aided in the exploration for, and discovery of, a broad range of deposits including diamond hosted in kimberlite (Stewart et al., 1988; Carlson et al., 1999), rare metals (Ford et al., 1988; Batterson and Taylor, 2009), uranium (Steele, 1988; Kirchner and Tan, 1994; Hirvas and Ma¨kinen, 1989), VMSs (Morrison, 2004; Day et al., 1987), gold (Dreimanis, 1958; Plouffe et al., 2011), sediment- hosted leadÀzinc (Morrissey and Romer, 1973; Nawrocki and Romer, 1979), intrusion-hosted tin (Rogers et al., 1990), and iron ore (Fig. 20.12; Dreimanis, 1958). Modern boulder tracing methods involve determining the areal distribution of boulders at the surface, the rate of decrease of boulder abundance, the distance down-ice, and the geochemical, mineralogical, and isotopic composition of the boulders. Bouchard and Salonen (1990) provide an overview of the occurrence and dispersal of boulders and their utility to mineral exploration.

20.6.2 SOIL SAMPLING Postglacial weathering of till can significantly change its chemical and mineralogical composition in the zone of oxidation above the water table or above the permafrost table (Shilts, 1975; Shilts and Kettles, 1990; McMartin and McClenaghan, 2001) by destroying the labile minerals such as 708 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

sulfides. Thus, the geochemical signatures in soils formed on till are the combined result of clastic glacial dispersal and subsequent aqueous and gaseous dispersal. Therefore, metal contents may be depleted or enriched in soil in comparison to unweathered till. The 3-D geometry of glacial dis- persal trains defined by soil geochemistry may be more difficult to interpret and trace up-ice to the source if hydromorphic processes have modified the soil composition (e.g., Kaszycki et al., 1996; Paulen, 2001; Lett and Jackaman, 2002; Hall et al., 2003). B-horizon soil is the highly oxidized and chemically weathered till that forms at the surface and typically extends down to a depth of 0.5À1.0 m. It is commonly enriched in amorphous Al- and/or Fe-oxyhydroxides, organic matter, and clay (Spirito et al., 2011). C-horizon till underlies the B-horizon and is the parent material (glacial sediment) that the B-horizon soil has developed on and that has been minimally affected by soil forming processes. C-horizon till is the target sampling medium because it provides an unimpeded signal of clastic glacial dispersal that is least affected by weathering. Some of the earliest exploration programs in Canada sampled the traditional ‘B horizon soil’ (e.g., Byers, 1956; Ermengen, 1957; Govett, 1973; Miller, 1979) adapted from use in areas of residual overburden and designed to avoid/ignore the effects of glaciation (Shilts, 1984). In glaci- ated terrain, soil may be developed on a variety of sediment types in addition to till, including gla- ciofluvial, glaciolacustrine, eolian, colluvium, or weathered bedrock. Each of these soil types can have very different transport and deposition mechanisms that can complicate the interpretation of soil geochemical data. Dreimanis (1960) recognized early on that the optimal approach to mineral exploration in glaciated terrain would be to sample till—material eroded, transported, and deposited directly by glaciers—and that metal-rich debris within till is a product of mechanical dispersal. Since the early 1980s, most exploration programs, case studies, and regional geochemical surveys have sampled till and not soil.

20.6.3 TILL SAMPLING Till geochemistry and indicator mineral methods are both widely used tools for mineral exploration in glaciated terrain, and both methods can be applied to the same till sample. Reviews of drift pro- specting methods published over the last 15 years include recommendations for till sampling survey design and collection procedures and protocols (Levson, 2001; McMartin and McClenaghan, 2001; McMartin and Campbell, 2009; Paulen, 2009a; McClenaghan et al., 2013a). The size of area to be sampled, the commodity and size of deposit sought, and the expected size of a dispersed anomaly (see discussion above about scale of dispersal) dictate the sampling density (Table 20.2) and pat- tern, such as grid, random, or focused along roads. Till samples are collected at a reconnaissance to regional scale within a geological province; regional scale within a mineral district or at a regional to local scale to detect a mineralized zone or deposit (Table 20.2). Other factors that will influence the design of a till sampling survey include budget, time frame, accessibility of the area of interest, distribution and thickness of surficial materials, landforms, ice flow history, permafrost distribution, land tenure, and the mineralogical and analytical methods to be used to detect glacial dispersal (i.e., heavy minerals versus matrix geochemistry) (Coker and DiLabio, 1989; Hirvas and Nenonen, 1990; Kauranne et al., 1992; Plouffe, 1995). In areas affected by continental ice sheets, topography is generally subdued and therefore has less influence on the design of a sampling grid as compared to areas affected by alpine glaciation. A till sampling grid 20.6 EXPLORATION METHODS 709

Table 20.2 Varying Scales of Till Sampling Surveys Conducted for Mineral Exploration Sample Sample Sample Type Minimum Target Survey Scale Density Spacing and Size Analysis Comments Geologic Reconnaissance 1 sample/ 10À25 km Geochemistry Matrix Low sample density, province/ 100À500 km2 (3 kg) and geochemistry, often random sample domain heavy mineral indicator pattern (101 kg) minerals Mineralized Regional 1 sample/ 4À10 km Geochemistry Matrix Low to moderate district, 10À100 km2 (3 kg) and geochemistry, sample density, often kimberlite heavy mineral indicator sampled in offset lines field (101 kg) minerals perpendicular to regional ice flow Cluster of Regional to 1 sample/ 1À2 km Geochemistry Matrix Moderate sample deposits local 1À4km2 (3 kg) and/or geochemistry, density in either offset heavy mineral indicator grid or random sample (101 kg) minerals pattern (nearest neighbor) Individual Local to 100À1000 25À250 m Geochemistry Matrix High sample density, deposits property samples/ (3 kg) and/or geochemistry, infilling previous 1km2 heavy mineral indicator surveys, tight grid or (101 kg) minerals lines perpendicular to dominant direction of transport

Modified from McMartin, I., McClenaghan, M.B., 2001. Till geochemistry and sampling techniques in glaciated shield terrain: a review. In: McClenaghan, M.B., Bobrowsky, P.T., Hall, G.E.M., Cook, S.J. (Eds.), Drift Exploration in Glaciated Terrain. Association of Exploration Geochemistry-Geological Society of London Special Publication 185, pp. 19À43. with similar sample spacing in every direction (Fig. 20.13) is used where ice flow patterns or the character of the buried bedrock sources are not well known (McMartin and McClenaghan, 2001; McMartin and Campbell, 2009). In some areas, sample distribution will be controlled by access along roads or bodies of water (Figs. 20.12 and 20.14). In areas where ice flow patterns are known, samples may be collected along lines perpendicular or parallel to ice flow, with spacing between lines equal to or much greater than spacing between samples (Figs. 20.9 and 20.15). In mountainous terrain, ice flow can be controlled by topography, and sampling lines should therefore be oriented with respect to ice flow directions. Given the steep slopes present in these areas, till may be covered by postglacial sediments (e.g., colluvium) which must be differentiated from till and not sampled.

20.6.3.1 Sample mass A small till sample, 1À3 kg, is sufficient for bulk geochemical analysis of the till matrix and for archiving. A larger till sample, 10À20 kg if sandy till or 20À40 kg if clay-rich till, is required for satisfactory recovery of indicator minerals (e.g., McClenaghan et al., 2013a; Plouffe et al., 2013). Both large and small till samples can be collected from the same hole or overburden drill hole interval and placed in separate bags/pails to facilitate shipping to different sample preparation/ analytical facilities. If the .2 mm fraction of till is to be used for provenance studies, pebbles can be collected from the same hole at the time of field sampling or set aside during sample prep- aration at the processing laboratory. (A)

(B)

Palaeoproterozoic Neoarchean

FIGURE 20.13 Gold content in till samples collected on a regularly spaced grid in the Meliadine area of northern Canada: (A) gold content of the ,0.063 mm fraction determined by INAA; (B) number of gold grains in the ,2mm fraction normalized to a 10 kg sample mass of ,2 mm material. Ice flow arrows in top right corner indicate the trend of the three phases of ice flow that affected the area; red is the youngest ice flow. Modified from McMartin, I., 2009. Till composition along the Meliadine Trend near Rankin Inlet, Nunavut: applications to gold exploration in permafrost terrain. In: Paulen, R.C., McMartin, I. (Eds.), Application of Till and Stream Sediment Heavy Mineral and Geochemical Methods to Mineral Exploration in Western and Northern Canada. Geological Association of Canada, Short Course Notes 18, pp. 153À166. FIGURE 20.14 Zinc content in surface till samples collected on an irregular sample grid controlled by road and river access in low relief terrain in western Canada. Data are plotted on a digital elevation modal generated from the Shuttle Radar Topography Misson. Modified from Plouffe, A., Paulen, R.C., Smith, I.R., 2006. Indicator mineral content and geochemistry of glacial sediments from northwest Alberta (NTS 84L, M): new opportunities for mineral exploration. Geological Survey of Canada, Open File 5121; Alberta Energy and Utilities Board, Special Report 77 and Paulen, R.C., Paradis, S., Plouffe, A., Smith, I.R., 2011. Pb and S isotopic composition of indicator minerals in glacial sediments from NW Alberta, Canada: implications for ZnÀPb base metal exploration. Geochem. Explor. Environ. Anal. 11, 309À320. 712 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

63°05’N No. grains pyrope

91°27’W 0 ice flow N 1 2–10 11–50 51–100 101–500 501–1000

Kimberlite dyke Dyke trace

Inset 1

91°12’W

0120.5 Km 63°00’N

FIGURE 20.15 Pyrope content in till sampled along lines perpendicular to ice flow direction (southeast) with sample spacing along lines shorter than the spacing between lines on the West Churchill property in northern Canada. Modified from Strand, P., Banas, A., Baumgartner, M., Burgess, J., 2009. Tracing kimberlite indicaomineral dispersa trains: an example from the Churchill Diamond Project, Kivalliq region, Nunavut. In: Paulen, R.C., McMartin, I. (Eds.), Application of Till and Stream Sediment Heavy Mineral and Geochemical Methods to Mineral Exploration in Western and Northern Canada. Geological Association of Canada, Canada, pp. 167À175. Sampling density was increased during subsequent field seasons to follow-up on anomalous samples from previous seasons, which led to the discovery of two diamondiferous kimberlite dykes (green stars).

20.6.3.2 Till facies The facies of till sampled is another important consideration for assessing glacial dispersal. Basal (or subglacial traction) till is the optimal sample medium because it was deposited at the base of an actively flowing glacier with little or no reworking by water and is generally locally derived. It may occur at any depth between the ground surface and the top of bedrock (Levson, 2001). In local- and property-scale surveys, basal till sampling close to the bedrock surface is most effective because the composition of the till most closely resembles that of the underlying bedrock.

20.6.3.3 Sample depth In areas of simple stratigraphy and a single, thin (,2 m) till unit, glacial dispersal trains reach the surface at, or very close to, source (Fig. 20.6A and B) (e.g., Hirvas and Nenonen, 1990), and thus surface till sampling can be used to detect these trains. Surface samples should be collected at .0.5 m depth to obtain a till sample from the C-horizon. 20.6 EXPLORATION METHODS 713

In areas of complex stratigraphy with multiple till sheets, glacial dispersal trains may occur in one or more till layers (e.g., Fig. 20.6C and D). Trains may also occur deep within a single thick (.5 m) till unit. These deeper trains may not be accessible by surface till sampling. In such set- tings, profile sampling methods must be employed to collect vertically adjacent till samples below the surface, to characterize the till stratigraphy, and to determine lateral and vertical variations in the bulk geochemical composition of till (e.g., Hirvas and Nenonen, 1990; Sarala and Peuraniemi, 2007). When using deep sampling methods, the lateral sample spacing needs to be adjusted (dis- cussed below) depending on the mobility and efficiency of the sampling equipment. In permafrost areas, a till sample may be collected from a shallow (,1 m) hole dug in a mud- boil, a feature that develops in the active layer during the maximum summer thaw period (Shilts, 1978; McMartin and Campbell, 2009) when hydrostatic pressure pushes fresh till up to the surface producing a vegetation-free patch (Fig. 20.16A). Mudboils are easily recognized as round to oval patches that are barren of vegetation and surrounded by ridges of vegetation and small rock frag- ments (Fig. 20.16B). Till in mudboils is commonly well homogenized and relatively unoxidized, thus representative samples may be collected at a shallow depth (B0.3 m) from the center of the mudboil (Shilts, 1978). Where mudboils are poorly developed, old and inactive, or scarce, such as in some areas of discontinuous permafrost or coarse-grained bouldery till, samples can be collected directly below the thin soil profile on the mudboil surface. Most till sampled in glaciated terrains is a product of warm-based continental ice sheet erosion, transportation, and deposition. However, in areas affected by cold-based glacial conditions (e.g., parts of Canada’s north, or at high elevations in mountainous areas), till relict sediment or regolith may still be collected from rare mudboils (Fig. 20.16C), cryogenic sediments pushed into the center of boulder rings (sorted circles), or in between boulders in felsenmeer terrain (e.g., Tremblay and Paulen, 2012).

20.6.4 SAMPLING METHODS Surface till samples can be collected from holes dug using a hand-held shovel or mechanical exca- vator, or from vertical till sections exposed along rivers, lakes, or manmade road cuts or excava- tions. As indicated above, in thick drift areas with potentially complex stratigraphy, glacial dispersal trains may be intersected at any depth and in more than one till unit. Canadian examples of deep till sampling surveys include those conducted in prospective areas such as the Abitibi Greenstone Belt (e.g., Bird and Coker, 1987; Sauerbrei et al., 1987; McClenaghan, 1994), the Rainy River area (Averill, 2013), the central Interior Plateau of British Columbia (Ferbey and Levson, 2009; Averill, 2017), and the Athabasca Basin (Geddes, 1982; Wilson, 1985). Auger, percussion, pneumatic, and rotary drills recover till at depth, as well as bedrock samples (e.g., Paulen, 2009b; Sarala, 2015). Diamond drills can generally recover cores of compact, fine- grained till, but not of loose, sandy till. Reverse circulation drills recover a slurry of till cuttings up to 1 cm in diameter (McMartin and McClenaghan, 2001), however, most clay-sized material and approximately 30% of the silt-sized material in till is lost by this drilling method. Cross- contamination of vertically adjacent samples through the drilling fluid can occur with this drilling method. The rotasonic drilling method recovers an 8À9 cm diameter continuous core of till, through boulders and into bedrock, allowing detailed till sampling, identification of till stratigraphy, and no loss of fine-grained material (McMartin and McClenaghan, 2001; Sarala, 2015). For all drilling methods, till samples may be contaminated by the small fragments released during the (A) Sedges, Mosses, Stones Lichens Surface

Till

Permafrost

(B)

(C)

FIGURE 20.16 (A) Mudboil schematic cross-section emphasizing the mixing and push of fresh till to the surface due to hydrostatic pressure during each summer thaw period; (B) photograph of an active mudboil in till in permafrost terrain, till samples can be collected from the freshest material in the center of the mudboil; (C) photograph of a mudboil of relict sediment and/or regolith (outlined by dashed line) in a region of cold-based glaciation. (A) Modified from McMartin, I., Campbell, J.E., 2009. Near surface till sampling protocols in shield terrain, with examples from western and northern Canada. In: R.C. Paulen and I. McMartin (Eds.), Application of Till and Stream Sediment Heavy Mineral and Geochemical Methods to Mineral Exploration in Western and Northern Canada. Geological Association of Canada, Short Course Notes 18, pp. 75À95. 20.6 EXPLORATION METHODS 715

operation of drill equipment, including brass from fittings, diamonds from drill bits, tungsten- carbide fragments from drill bits (Coker and Shilts, 1993), and by drilling grease (contains molyb- denum, lead, zinc) (Averill, 1990). Deformation of till structures and stratigraphy can result from the drilling process (Smith and Rainbird, 1987).

20.6.5 TILL GEOCHEMISTRY The term ‘indicator element’ is used here to refer to an element that is an economically valuable component of the mineralization being sought and which may be used to detect a mineral deposit (Boyle, 1974; Rose et al., 1979). The term ‘pathfinder element’ is used here to refer to elements associated with the mineral deposit, but that are not an element of main economic interest (Boyle, 1974; Rose et al., 1979). Key indicator and pathfinder elements for various deposit types are sum- marized in Table 20.3 from several sources (Boyle, 1974; Rose et al., 1979; Levinson, 1980; Lehtonen et al., 2015). 20.6.5.1 Size fractions The ,0.063 mm (À250 mesh) silt 1 clay sized fraction of till is the most common size fraction geochemically analyzed for mineral exploration (e.g., Lestinen et al., 1991; McClenaghan, 1994; Tarvainen, 1995; McClenaghan et al., 2011; McClenaghan and Peter, 2016) because: (1) many minerals in the mineral deposit are readily comminuted to this size range over short distances and thus are enriched in this size fraction (Nevalainen, 1989; Coker and DiLabio, 1989; Shilts, 1995, 1996); (2) this fraction contains phyllosilicates that are common alteration or gangue minerals and can be (primary) enriched in these elements of interest; (3) phyllosilicates will scavenge cations released during till weathering (Shilts, 1993, 1996); and (4) it can be prepared for analysis rapidly and inexpensively (Lett, 1995; Levson, 2001). The ,0.002 mm (clay-sized) fraction of till may also be geochemically analyzed (e.g., McMartin et al., 1996; Plouffe et al., 2016) because of its greater capacity to retain elements released during weathering, and to avoid potential textural influences on matrix geochemistry (Shilts, 1975, 1996). Disadvantages of using the ,0.002 mm fraction include (1) the large volume of material (up to 1 kg) needed to produce a sufficient mass (1À2 g) of clay-sized material for geo- chemical analysis (Klassen, 2003), (2) the strong influence of hydromorphic processes on the geo- chemistry of this size fraction that may distort or completely mask the signature of clastic glacial dispersal, and (3) the higher cost of the centrifuging that is required to recover this size fraction. Geochemically analyzing the ,0.177 mm (À80 mesh) sand 1 silt 1 clay sized fraction of till is generally not recommended because it includes abundant sand-sized quartz and feldspar grains which will dilute and mask the geochemical signature of mineralization or specific rock lithologies. 20.6.5.2 Analytical methods

Aqua regia (3:1 HCl:HNO3) is a partial digestion that is often used to analyze till because it targets the part of the total concentration that most likely reflects mineralization or provides geochemical contrast, i.e., sulfides, native gold, platinum, palladium, arsenides, selenides, tellurides, carbonates, most sulfates, phosphates, some oxides (e.g., uranium) and their oxyhydroxides (e.g., iron, manga- nese), some silicates, and organically bound elements (Chao, 1984; Hall, 1999). Cations in some silicate minerals will be leached by aqua regia, especially in micas and clay minerals (Koljonen Table 20.3 Listing of Selected Mineral Deposit Types and Corresponding Common Indicator Minerals That Can Be Recovered From, and Pathfinder 1 Indicator Elements That Can Be Analyzed for, in Till Indicator/ Ore Pathfinder Published Reviews and Deposit Type Elements Elements Common Indicator Minerals Examples Kimberlite- C Ba, Cr, K, Cr-pyrope, Cr-diopside, McClenaghan et al. hosted LREE, Mg, Nb, eclogitic garnet, Mg-ilmenite, (2002a), McClenaghan diamonds Ni, P, Rb, Sr, chromite, diamond and Kjarsgaard (2007), Ta, Ti Nowicki et al. (2007) Volcanogenic Cu, Pb, Ag, As, Au, Ba, Chalcopyrite, sphalerite, galena, Averill (2001), massive Zn, Ag, Bi, Cd, Cu, Hg, pyrrhotite, gold, pyrite, gahnite, McClenaghan and Peter sulfide Au In, Pb, S, Sb, Tl, staurolite, cassiterite, (2016), McClenaghan Zn spessartine, sillimanite, et al. (2015a,b) andalusite, beudantite, jarosite, barite, tourmaline, hogcomite, nigerite Sediment- Ag, Cu, Ag, Cu, Pb, S, Chalcopyrite, sphalerite, galena, Tarplee and Meer (2010), hosted lead- Pb, Zn Zn pyrite, barite, spessartine, Paulen et al. (2011), zinc smithsonite, anglesite, cerussite Oviatt et al. (2015) Gold Au, Ag Ag, As, Au, B, Gold, scheelite, tourmaline, Averill (2001), Sarala Ba, Bi, Cu, Co, rutile, sulfides, tellurides, PGM, et al. (2009), Fe, Hg, Mn, Sb, barite, cinnabar McClenaghan and Se, Te, U, W Cabri (2011) Magmatic Ni, Cu, As, Au, Cr, Cu, Pentlandite, chalcopyrite, pyrite, McClenaghan and Cabri NiÀCuÀPGE PGE Mg, Ni, PGE, S millerite, platinum group (2011), Averill (2009, minerals, chromite, Cr-diopside, 2011), McClenaghan et al. enstatite, olivine, Cr-andradite (2011) Rare metals REE, Li, Be, Ce, Cl, F, Pyrochlore, columbite, Batterson and Taylor Nb, Ta, Li, Nb, U, P, Ta-minerals, allanite, zircono- (2009), Lehtonen et al. Zr REE, Ta, Th, Y, silicates, apatite, monazite, (2015), Simandl et al. Zr fluorite, rhabdophane, (2015); Mackay et al. arfvedsonite (2015) Porphyry Cu, Mo, Au, Ag, Cu, Chalocopyrite, chalcocite, Averill (2011), Kelley CuÀAuÀMo Au, Ag Mo, S pyrite, molybdenite, gold, et al. (2011), Hashmi et al. silver, epidote, tourmaline, (2015), Plouffe et al. apatite, andradite, barite, (2016) monazite, rutile, titanite, zircon, jarosite, malachite Porphyry Sn, W, As, Ag, Be, Bi, Cassiterite, scheelite, Snow and Coker SnÀW Mo Cd, Cu, F, In, wolframite, molybdenite, (1987a,b), McClenaghan Mo, Pb, S, Te, chalcopyrite, Bi sulfides, et al. (2017a,b) W, Zn sulfides, fluorite, topaz, tourmaline Uranium U As, Ba, Cu, F, Uraninite (apitchblende), Geddes (1982), Boyle La, Ni, P, Pb, thorianite, tourmaline, sulfides, (1982), Campbell (2009); Th, Ti, U, Y, Zn, monazite, allanite, zircon, Robinson et al. (2016) Zr baddelyite, niccolite, UÀTh anatase, UÀTh rutile, brannerite, magnetite

PGM, platinum group minerals; PGE, platinum group elements; REE, rare earth elements. aPitchblende—brown or black pitchy massive form of uraninite. Data compiled from Boyle (1974), Rose et al. (1979), Levinson (1980), and Lehtonen et al. (2015). 20.6 EXPLORATION METHODS 717

et al., 1992). Minerals minimally attacked by aqua regia include barite, chromite, gahnite, garnet, cassiterite, ilmenite, rutile, sphene, monazite, and zircon (Dolezalˇ et al., 1968). An aqua regia digestion is commonly used for till geochemical surveys conducted by government agencies (e.g., Koljonen et al., 1992; Lett, 1995; Sarala et al., 1998; Bajc and Hall, 2000; Parkhill and Doiron, 2003; Barnett, 2007; McClenaghan et al., 2011) and by industry (e.g., Snow and Coker, 1987a,b; Brereton et al., 1988). Another partial digestion, (8:1 HNO3: HCl) that is weaker than conventional aqua regia is commonly used to capture uranium bound in uranium oxides and loosely bound in weathered material (Campbell, 2009). A total digestion of till is carried out to determine major, minor, and trace element contents for identifying signatures related to specific bedrock lithologies, major rock-forming minerals, or resistate minerals that are not digested by aqua regia (e.g., Taipaleetal.,1986; Koljonen and Malisa, 1991; Koljonen et al., 1992; Lahtinen et al., 1993; Tarvainen, 1995; Tarvainen et al., 1996; Klassen, 2003; Plouffe et al., 2006). For example, total digestion methods are useful for determining REE contents in till when exploring for kimberlites (e.g., McClenaghan and Kjarsgaard, 2007), carbonatites, or per- alkaline granites (e.g., Lehtonen et al., 2015). Abundances of important pathfinder elements such as tungsten, tin, barium, selenium, chromium, and zirconium are best determined by ‘total’ digestion methods rather than those employing aqua regia (e.g., Chao, 1984; Chao and Sanzolone, 1992). Total/near total digestion methods include: (1) lithium meta/tetra borate fusion-nitric acid digestion which is optimal for determining the content of REE and other high field-strength elements (e.g., Cremer and Schlocker, 1976; Chao and Sanzolone, 1992); (2) four-acid digestion which is optimal for determining total copper, nickel, cobalt, lead, zinc, molybdenum, and silver, but is not a total digestion for REE, some oxides of aluminum, iron, hafnium, manganese, tin, tantalum, tungsten, and zirconium, for chromium from chrome-bearing minerals, or for barium from barium-bearing minerals (Koljonen et al., 1992; Noras, 1992). Also, volatilization during fuming may result in some loss of silicon, arsenic, antimony, and gold (Chao and Sanzolone, 1992); (3) lead fire assay on a minimum of 30 g of material which is optimal for determining total gold, platinum, and palladium content in areas with potential to host gold- or platinum group elements (PGE)-bearing mineralization (Hoffman et al., 1999; Hall and Oates, 2003). Instrumental neutron activation analysis (INAA) is a nondestruc- tive technique that can be used when the sample cannot be destroyed. Till heavy mineral concentrates (HMCs) (described below) may be analyzed for gold and other elements by INAA in order to preserve the concentrates for subsequent visual examination and determination of the abundance and character of indicator minerals present. Lead isotopic determi- nations of bulk till matrix can be useful for determining if copperÀleadÀzinc-rich till is sourced from VMS mineralization (Hussein et al., 2003 and references therein). The development of inductively coupled plasma-mass spectrometry (ICP-MS) in the late 1980s and through the 1990s was one of the key advances for applied till geochemistry as it allowed for the determination of a broad suite of trace elements including potential indicator and pathfinder elements for a wide spectrum of mineral deposit types. For example, till geochemical data may now be used to determine if glacially eroded VMS mineralization is silver-rich or gold-rich, or if it is enriched in sal- able (gold, silver, indium) or deleterious trace elements (selenium) (McClenaghan and Peter, 2016). Portable X-ray fluorescence spectrometry (pXRF) can be used to determine metal contents of till while still in the field in order to actively guide till sampling and follow-up of geochemical anomalies in the same field season. It may also be used in the laboratory to prioritize till samples for conventional lab-based geochemical or mineralogical analysis. Several recent studies report the advantages and disadvantages of applying pXRF analyses to moist versus dry till, sieved versus 718 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

unsieved till, and making determinations through plastic sample bags versus Prolene film-covered cups (Hall and McClenaghan, 2013; Kjarsgaard et al., 2014; Hall et al., 2016; Sarala, 2016). Quality assurance (QA) and quality control (QC) measures must be included at all phases of a till analytical program, including the collection of field duplicates, insertion of blanks and ‘blind’ laboratory duplicates, and insertion of certified reference standards. Recommendations for QA/QC measures for geochemical analysis of till include those suggested by Thomson (1983), Gustavsson (1992), and McClenaghan et al. (2013a).

20.6.6 INDICATOR MINERAL METHODS Similar to indicator elements described above, indicator minerals are mineral species that indicate the presence of a specific mineral deposit, alteration, or rock lithology and can be recovered from surficial sediments (Averill, 2001). Ideal indicator minerals recovered from glacial sediments are those that are found in few if any rocks other than in the mineral deposit, alteration, or lithology being sought. Their physical and chemical characteristics allow them to be readily recovered from large till samples in suf- ficient abundance for ready identification, viable grain counts, and analysis. These characteristics include: visual distinctiveness, moderate to high density (.2.9gcm23), silt to sand-size, ability to sur- vive glacial erosion and transport, and ability to survive pre- and/or postglacial weathering (Brundin and Bergstrom, 1977; Averill, 2001). Peuraniemi (1990), Averill (2001),andLehtonen et al. (2015) provide extensive lists of common heavy minerals in till that have been combined to produce Table 20.3, a list of the most common indicator minerals of the major mineral deposit types. Indicator mineral methods differ from the traditional till geochemical methods described above in that the indicator grains reflect mechanical dispersal and individual grains can be visually exam- ined, counted, and analyzed. The use of indicator mineral methods provides: (1) mineralogical evi- dence of the presence of a (unmetamorphosed or metamorphosed) mineralized target, hydrothermal alteration zone, or specific rock lithology; (2) information about the proximity to source; (3) low detection limits down to only a few grains in a single sample that is equivalent to ppb-level geo- chemical analyses; and (4) the ability to visually identify and remove, if necessary, grains from anthropogenic sources (McClenaghan, 2005). During the 1970s and 1980s, it was common for till HMCs to be pulverized and then geochemi- cally analyzed (e.g., Garrett, 1971; Peuraniemi, 1985; Bird and Coker, 1987; Peuraniemi, 1990) because there was limited capability to systematically visually examine and identify indicator minerals. For example, tin content in till was used as a proxy for the presence of cassiterite (Peuraniemi, 1987; Rogers et al., 1990), and tungsten content for scheelite (e.g., Peuraniemi, 1985; Johansson et al., 1986; Salminen and Haartikainen, 1986). Methods are now commercially available to identify suites of indicator minerals in a till HMC for a broad range of deposit types, as listed in Table 20.3. Indicator minerals can be readily identified and counted in the silt to sand-sized heavy (.3.2 g cm23) and mid-density (2.9À3.2 g cm23) fractions of till (e.g., Averill, 2001; McClenaghan, 2011), thus at present, the HMC is less commonly analyzed geochemically. Indicator minerals are sparse in till, however, the presence of one sand-sized grain of a particular indicator mineral in 10 kg of till may be significant. Till samples are processed to reduce the mass of material to be examined by removing coarse fragments (typically .2 mm), clay-sized material, and low-density minerals. Most indicator minerals have a moderate to high density, therefore processing techniques involve density separation in combination with sizing. The most common density separa- tion methods used for processing till samples include the shaking table, dense media separation, 20.6 EXPLORATION METHODS 719

jigging, centrifugal concentration, spiral concentration, and panning (e.g., Gent et al., 2011; McClenaghan, 2011). Approximately 90% of gold and platinum group mineral (PGMs) grains in source rocks and till are silt-sized (,0.063 mm) (Shelp and Nichol, 1987; Sibbick and Fletcher, 1993; Averill, 2001; McClenaghan and Cabri, 2011), thus preconcentration techniques to recover these fine-grained minerals from till must allow for the recovery of silt-sized heavy minerals. After preconcentration using one of the methods described above, the preconcentrate is further refined using a heavy liquid of a specific density to produce a sharp separation between heavy (sink) and light (float) minerals at a known density. Commonly used liquids include methylene iodide with a density of 3.3 g cm23, tetrabromoethane or the low-toxicity heavy liquid lithium heteropolytungstate, both with a density of 2.9. The density of the heavy liquids can be increased or decreased to capture specific minerals. The most common density separation for till samples is performed at 3.2 g cm23 so as to include the key indicator minerals Cr-diopside (3.25À3.55 g cm23), and olivine (3.27À3.37 g cm23)(McClenaghan and Kjarsgaard, 2007; McClenaghan, 2011). Some indicator minerals require a slightly lower density separation at 2.8 or 3.0 g cm23,suchasfluorite (3.01À3.25 g cm23), apatite (3.1À3.2 g cm23), and tourmaline (3.1À3.2 g cm23). The ferromagnetic and paramagnetic fractions are then separated from the nonmagnetic fraction to further reduce the vol- ume of material to be examined and to isolate fractions for the counting of specific mineral species. Once the HMC is prepared, the 0.25À2.0 mm nonferromagnetic fraction is examined optically to determine the indicator mineral abundance and size. In addition to mineral abundance, mineral chemistry is an important modern exploration tool that provides insights into the nature and/or grade of the bedrock source. For example, the chemistry of KIMs in till can be diagnostic of the diamond grade of a kimberlite source (McClenaghan and Kjarsgaard, 2007; Nowicki et al., 2007). Mineral grains are characterized by measuring the abundances of major oxides, and minor and trace elements by scanning electron microscopeÀenergy-dispersive X-ray spectroscopy (SEM-EDS), electron microprobe analysis, or laser -inductively coupled plasmaÀmass spectrometry (Layton Matthews et al., 2014;Lehtonen et al., 2015). Microchemical characterization techniques so well developed for placer gold (e.g., Chapman et al., 2009; Chapman and Mortensen, 2016) can now be applied to gold and PGM grains in till, to provide insights as to their bedrock source. Magnetite is a major accessory mineral in many types of mineral deposits as well as igneous, metamorphic, and sedimentary rocks, and thus is common in till. It is an ideal indicator mineral in till because it is physically robust (hardness 5 5.5À6.0), resistive to chemical alteration during gla- cial transport, and easily recovered from till because of its high density (5.1À5.2 g cm23) and ferro- magnetic character. The chemical composition of grains in till can be used to discriminate between a magmatic or hydrothermal origin of the source rocks (e.g., Nadoll et al., 2012; Boutroy et al., 2014; Dare et al., 2014; Makvandi et al., 2016). Much progress has been made in the application of ICP-MS for isotopic analysis of single min- eral grains in till. Radiogenic isotopes can provide additional information about potential minerali- zation types (e.g., VMS versus other base metal deposits) that may help refine exploration targets (e.g., Paulen et al., 2011; Oviatt et al., 2015) or determine the age of preglacial gossan formations (e.g., Kelley et al., 2011). This new lower-cost method may be faster and easier for the exploration- ist to use than traditional isotopic methods using the bulk till matrix. The rapid SEM scanning tech- niques, quantitative evaluation of minerals by SEM and mineral liberation analysis, can be used to automatically identify and characterize very small (,0.25 mm) indicator mineral grains in till HMC (Wilton and Winter, 2012; Layton-Matthews et al., 2014; Lehtonen et al., 2015). Morphology and surface textures of indicator minerals in till may provide indications of distance of 720 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

glacial transport (e.g., DiLabio, 1990b; Huhta, 1993; McClenaghan and Kjarsgaard, 2007). These features can be observed using optical microscopy and using SEM in backscatter mode. QC must be monitored at each in heavy mineral processing, indicator mineral identification, and subsequent mineral chemical analyses. Guidelines for QC procedures, as well as for the reporting of indicator mineral results for till, are reviewed in McClenaghan (2011) and Plouffe et al. (2013).

20.6.7 DIAMONDS Kimberlite, the primary host rock for diamond deposits in glaciated terrain, is a mineralogically and chemically distinct point source. A kimberlite’s variable size (50À200 m in diameter), shape (ellipti- cal pipe or narrow, elongated dyke), competence, and degree of postglacial weathering influence the size and areal extent of dispersed kimberlitic debris (McClenaghan and Kjarsgaard, 2001, 2007). In most cases, kimberlite is softer than its surrounding host rock and may be capped by soft pregla- cially weathered kimberlite. As a result of both factors, dispersal trains from kimberlite pipes are typically kilometers in length. Trains can be ribbon-shaped (Fig. 20.17), fan-shaped (e.g., Lehtonen et al., 2005; McClenaghan et al., 2012), or bilobate (Fig. 20.10), depending on ice flow patterns.

65º19' N 111º26' W <------Ice flow

Ranch Lake kimberlite

Cr-pyrope/20 kg till 0 1 – 10 11 – 50 50 – 150 151 – 300 Geological Survey of Canada till sample 301 – 445 2 km Canamera Geological Ltd. till sample

111º59' W 111º59' 65º11' N

FIGURE 20.17 Ribbon-shaped glacial dispersal train trending westward from the Ranch Lake kimberlite, Lac de Gras as defined by Cr-pyrope content in the 0.25À0.5 mm HMC fraction of till. The train was formed by a single phase of ice flow towards the west. Modified from McClenaghan, M.B., Ward, B.C., Kjarsgaard, I.M., Kjarsgaard, B.A., Kerr, D.E., Dredge, L.A., 2002a. Indicator mineral and till geochemical dispersal patterns associated with the Ranch Lake kimberlite, Lac de Gras region, NW Territories, Canada. Geochem. Explor. Environ. Anal. 2, 299À320. 20.6 EXPLORATION METHODS 721

In contrast, dispersal trains from kimberlite dykes that trend perpendicular to ice flow are typically wide and short (,1 km) (e.g., Kirkley et al. 2003; Strand et al., 2009). Glacial dispersal from kimberlite is typically traced using the abundance of the visually and chemically distinct indicator minerals Cr-pyrope, eclogitic garnet, Mg-ilmenite, chromite, Cr-diopside, forsteritic olivine, and diamond (Fig. 20.18 and Table 20.3) in the 0.25À2.0 mm nonferromagnetic fraction of till. Kimberlite boulders are also used to identify dispersal. Mineral chemistry is used to differentiate KIMs in till from visually similar minerals from other bedrock sources and to indicate the potential presence of diamond in the kimberlite source (e.g., McClenaghan and Kjarsgaard, 2007; Nowicki et al., 2007, and references in both). Surface textures on mineral grains can provide indications that their source is kimberlitic, or of glacial transport dis- tance (Mosig, 1980; Afanase’ev et al., 1984). For example, resorbed octahedral crystal faces on chromite grains (Fig. 20.18F) are typical of a kimberlitic genesis versus magmatic or hydrothermal origin (McClenaghan and Kjarsgaard, 2001, 2007). The presence of soft kelyphite reaction rims on Cr-pyrope (Fig. 20.18B) indicates that its source is kimberlitic and that glacial transport has been relatively short (Garvie and Robinson, 1984; McCandless, 1990). Kimberlites have distinctive major and trace element compositions compared to their host rocks, thus till geochemistry may be useful for detecting the presence of kimberlitic debris down-ice (McClenaghan and Kjarsgaard, 2007). Its usefulness will depend on the geochemical contrast between the kimberlite and the surrounding bedrock such that elements in the kimberlite with con- centrations that are 10 times greater than the host rock(s) they are emplaced into will be the most useful pathfinder elements (e.g., McClenaghan et al., 2002b,2004; McClenaghan and Kjarsgaard, 2007). Elements such as magnesium, nickel, chromium, niobium, tantalum, titanium, barium, potas- sium, rubidium, strontium, phosphorous, and light REE (Table 20.3) should be the most viable kim- berlite pathfinder elements. Element abundances in till should be determined using a total digestion method such as lithium meta/tetra borate fusion (discussed above) in order to fully digest the oxide and silicate minerals and to determine the total concentrations of elements.

20.6.8 VOLCANOGENIC MASSIVE SULFIDE DEPOSITS VMS deposits are important sources of copper, lead, zinc, and silver. Glacial dispersal from VMS deposits has typically been traced using mineralized boulders and till geochemistry (McClenaghan and Peter, 2013, 2016, and examples therein). Indicator and pathfinder elements for VMS deposits that have proven effective for till geochemistry are listed in Table 20.3, and include the metals of economic interest in these deposits: copper, lead, zinc, and silver. Over the past 15 years, a suite of indicator minerals has been developed for VMS deposits (Table 20.3) that includes minerals present within mineralization and associated hydrothermal alter- ation zones (Averill, 2001; McClenaghan et al., 2015b). Detection of glacial dispersal of sulfide minerals down-ice is complicated by their chemical instability in the surface weathering environ- ment. Thus, oxide and silicate indicator minerals can be more useful for this deposit type (Table 20.3). Gahnite is a zincian spinel that occurs around and within metamorphosed VMS deposits (Spry and Scott, 1986; McClenaghan et al., 2015a). It is a useful indicator mineral of VMS deposits because it is physically robust (hardness 5 8), chemically stable in oxidizing environments, and it can be easily recovered from till because of its high density (4À4.6 g cm23) and distinctive bluish green color (Fig. 20.19)(Lalonde et al., 1994; Paulen et al., 2013). An example of gahnite glacial dispersal down-ice from the Izok Lake VMS deposit in northern Canada is shown in Fig. 20.20. 722 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

FIGURE 20.18 Color photographs of KIMs: (A) purple to pink Cr-pyrope; (B) Cr-pyrope with dark green-gray kelyphite rims (k); (C) Cr-diopside; (D) eclogitic garnet; (E) Mg-ilmenite; (F) chromite showing resorbed crystal faces; (G) forsteritic olivine; and, (H) diamond. Mineral photography by Michael J. Bainbridge. Eclogitic garnet grains provided by Mineral Services. FIGURE 20.19 Color photographs of VMS indicator mineral gahnite (Zn-spinel) recovered from till down-ice of the Izok Lake VMS deposit in northern Canada: (A) gahnite with adhering minerals, an indication of proximity to source mineralization; and, (B) gahnite without adhering minerals and subrounded shape, an indication of longer glacial transport distance. Mineral photography by Michael J. Bainbridge.

N 65°40’N No. gahnite grains/10 kg (0.25–0.5 mm)

113°00’W Ice flow Range No. of samples 4 51–1739 (9) ? Iznogoudh Lake 21–50 (19) ? 6–20 (22) ? 2–5 (32) 1 (18) 0 (3)

Phase 3 dispersal (NW flow) Izok ? Lake Izok Deposit ?

?

Phase 1 112°43’W dispersal 1 km (SW flow) 65°36’N

FIGURE 20.20 Fan-shaped glacial dispersal train of gahnite abundance in the 0.25À0.5 mm HMC fraction of till normalized to a 10-kg sample mass down-ice of the Izok Lake volcanogenic massive sulfide deposit in arctic Canada. The fan- shaped train was formed by two phases of ice flow towards the southwest (blue polygon) and subsequent reworking by a younger ice flow towards the northwest (yellow polygon). Green polygon indicates the part of the train modified by both phases of ice flow. Modified from McClenaghan, M.B., Paulen, R.C., Layton-Matthews, D., Hicken, A.K., Averill, S.A., 2015a. Glacial dispersal of gahnite from the Izok Lake ZnÀCuÀPbÀAg VMS deposit, northern Canada. Geochem. Explor. Environ. Anal. 15, 333À349. 724 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

The morphology of gahnite shape may provide clues to relative distance of glacial transport. At source, grains in till are commonly interlocked with quartz and mica (Fig. 20.19A). As transport distance down-ice increases, glacial crushing of grains liberates the interlocking minerals to leave only gahnite (Fig. 20.19B) (McClenaghan et al., 2014a). The trace element chemistry of indicator minerals may be useful for establishing that the bedrock source of the mineral is VMS mineraliza- tion (e.g., gahnite: Heimann et al., 2005; O’Brien et al., 2015; magnetite: Makvandi et al., 2016). Secondary minerals such as beudantite, jarosite, and goethite form during the oxidation of pri- mary sulfide mineralization. Where preglacial gossans have formed on VMS mineralization, these gossan minerals can survive glacial erosion and transportation and be useful indicator minerals in till down-ice (e.g., McClenaghan et al., 2015b).

20.6.9 SEDIMENT-HOSTED LEADÀZINC DEPOSITS Sediment-hosted leadÀzinc deposits include sedimentary exhalative (SEDEX) and Mississippi -type (MVT) deposits, as well as Irish-type deposits that can be considered to be transitional between the first two (Wilkinson, 2014). SEDEX and MVT deposits are major sources of global zinc resources (Goodfellow and Lydon, 2007; Wilkinson, 2014). Glacial dispersal from SEDEX deposits has been traced using mineralized boulders and till geochemistry (e.g., Morrissey and Romer, 1973; Nawrocki and Romer, 1979; Tarplee and Meer, 2010). These studies have identified indicator and pathfinder elements, including the main metals of economic interest (zinc, lead, sil- ver). Indicator minerals for this deposit type include sulfide minerals and may also include spessar- tine, cassiterite, and barite (Table 20.3). MVT and Irish-type deposits are carbonate-hosted leadÀzinc deposits (Paradis et al., 2007; Wilkinson, 2014), thus, till derived from these deposits has a carbonate-rich matrix which buffers the till as it weathers and allows the less-stable sulfide minerals to survive. As a result, sphalerite and galena (Fig. 20.21) can be detected down-ice from these deposits using both till geochemistry (Fig. 20.14) and indicator mineral methods. Secondary minerals such as smithsonite formed from the oxidation of sphalerite, and anglesite as well as cerussite formed from the oxidation of galena may also be useful indicator minerals in till (Table 20.3)(Oviatt et al., 2013). Lead isotopic compo- sitions of individual galena and sulfur isotopic compositions of discrete sphalerite grains in till may

FIGURE 20.21 Color photographs of (A) sphalerite and (B) galena grains recovered from till down-ice of the O-28 lead-zinc deposit in the Pine Point Mining District of northern Canada. Mineral photography by Michael J. Bainbridge. 20.6 EXPLORATION METHODS 725

be useful for characterizing the bedrock source (e.g., Paulen et al., 2011; Oviatt et al., 2015). A few case studies and regional geochemical surveys have been carried out around, or in search of, carbonate-hosted leadÀzinc deposits (e.g., Allan, 1974; Hannon and Scott, 1975; Nawrocki and Romer, 1979; Paulen et al., 2011; Oviatt et al., 2013). Glacial dispersal from these deposits is char- acterized mainly by the ore-forming elements lead and zinc (Table 20.3).

20.6.10 LODE GOLD DEPOSITS Glacial dispersal from lode gold deposits was initially traced using mineralized boulders (e.g., Dreimanis, 1958). Today, gold grains recovered from till are the best indicator of the presence of gold mineralization (e.g., McMartin, 2009; Averill, 2017). Sulfides, arsenides, tellurides, PGM, and other oxide and silicate minerals listed in Table 20.3 are also useful indicators (McClenaghan and Cabri, 2011 and references therein). In rare cases, secondary minerals formed during pre- or post- glacial oxidation may be useful indicator minerals (e.g., pyromorphite, malachite, goethite) (Averill and Zimmerman, 1986; Peuraniemi, 1990). Gold grain methods for till include documenting gold grain abundance, size, shape, and compo- sition (e.g., Chapman et al., 1990; Huhta, 1993; McClenaghan, 2001; Sarala et al., 2009). Because gold is malleable, the degree of rounding and bending of the gold grains recovered from till pro- vides useful information about glacial transport distance (Averill, 2001). DiLabio (1990b) presented a graphically descriptive classification scheme that is widely used today in Canada for describing gold grain condition (shape, crystal face, surface texture) that relates to glacial transport distance (pristine, modified, or reshaped; Fig. 20.22).

(A) (B) (C)

10 µm 10 µm 10 µm

FIGURE 20.22 Secondary electron images of gold grains showing examples of the three ‘conditions’ used to classify gold grains recovered from till: (A) pristine gold grain with equant molds suggestive of former quartzÀfeldsparÀcarbonateÀsulfide gangue; (B) modified gold grain with vestiges of equant gangue molds and edges that are slightly curled; and, (C) reshaped gold grain showing pitted surfaces and well curled edges. Modified from McClenaghan, M.B., 2001. Regional and local-scale gold grain and till geochemical signatures of lode Au deposits in the western Abitibi Greenstone Belt, central Canada. In: McClenaghan, M.B., Bobrowsky, P.T., Hall, G.E.M., Cook, S.J. (Eds.), Drift Exploration in Glaciated Terrain. Geological Society of London, Special Publication 185, pp. 201À223. 726 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

Noteworthy examples of the successful application of indicator mineral methods to the discov- ery of gold deposits in Canada include the Casa Berardi deposits (Sauerbrei et al., 1987), the Rainy River deposit (Averill, 2013), and most recently, the Blackwater Lake deposit (Averill, 2017). Electrum, native copper, chalcocite, bornite, molybdenite, galena, and pyromorphite grains in till were used to identify a glacial dispersal train which led to the discovery of the Partridge Zone at Waddy Lake in central Canada (Averill and Zimmerman, 1986). Fig. 20.13B is an example of gold grain distribution in a regional till survey conducted over a prospective greenstone belt in northern Canada, showing a significant gold anomaly in till formed by south and southeast ice flow (McMartin 2009). Consistent gold grain recovery, counting, and shape classification methods have been used for more than 30 years in Canada. Use of these consistent methods allows for the com- parison of industry- and government-led surveys results within and between projects over time, and the establishment of background thresholds for gold grain content in some regions such as the Abitibi Greenstone Belt (e.g., Averill, 1988). Till geochemistry is also an important gold exploration tool (e.g., Bird and Coker, 1987; Sarala and Peuraniemi, 2007; Sarala et al., 2009). Fig. 20.13A is an example of gold contents in the ,0.063 mm fraction of regional till samples around gold occurrences along the Meliadine Trend in northern Canada. Indicator and pathfinder elements for gold deposits are listed in Table 20.3, and include gold and silver. The optimal analytical methods for determining the total gold content of a till sample include fire assay-ICP-MS or INAA on a 30 g aliquot to reduce the nugget effect. Several studies have demonstrated only poor to moderate correlations between gold concentration in the ,0.063 mm frac- tion and number of gold grains in the heavy mineral fraction (e.g., MacEachern and Stea, 1985; McClenaghan, 1992, 1994). These studies indicate that both heavy mineral and geochemical methods should be used in concert for gold exploration programs.

20.6.11 MAGMATIC NICKELÀCOPPER DEPOSITS Magmatic nickelÀcopper deposits are significant sources of nickel, copper, and PGE. Recent Canadian examples of till geochemistry studies around magmatic nickelÀcopper deposits include those conducted in the Thompson Nickel Belt, Sudbury (Bajc and Hall, 2000; McClenaghan et al., 2011, 2014b), and Lac des Isles regions (Barnett, 2007) in central Canada. Table 20.3 lists indicator and pathfinder elements summarized from these studies. A suite of indicator minerals for magmatic nickelÀcopper deposits has been developed that includes nickel, copper, and PGE-bearing sulfides, arsenides, antimonides, and gold, together with one or more of Cr- and Mg-bearing oxides and silicates (Table 20.3)(Averill, 2009, 2011; McClenaghan and Cabri, 2011; McClenaghan et al., 2013b). Of these minerals, chromite, Cr- diopside, forsterite, and enstatite have distinguishing color, size, surface textures, and mineral inclu- sions that allow them to be visually distinguished from those found in kimberlite (Averill, 2009). Sperrylite is a common PGM that has been recovered from till down-ice from several magmatic nickelÀcopper deposits (McClenaghan and Cabri, 2011). It is a particularly useful indicator mineral because it is relatively stable in the surface weathering environment, is physically robust (hardness 5 6À7), visually distinct (tin white), and a direct indicator of the presence of PGE. The composition of some indicator minerals such as chromite, olivine, and magnetite, can be useful for identifying the presence of nearby nickelÀcopperÀPGE mineralization (e.g., McClenaghan et al., 2013b; Dare et al., 2014; Sappin et al., 2014). One of the better examples of glacial dispersal from 20.6 EXPLORATION METHODS 727

49°13’N 0 89°32’W

0

N 0 Lac des Iles 0 0 North Lac des 0 28 10 4 Iles Intrusion 0 2200 34 5858 10 633 3232 0 8 5757 1414 4400 50 8 0 0 1 0 76 45 4 1 30 1 0 7 0 235 0 2 0 20 4 Mine block 00 0 3434 1 intrusionIntrusion 0 222 3 9 0 Ice flow 0 0 1 0 7 0 2222 0 3737 53535 07 3535 6 0 2424 20 1 3838 0 3636 0 330 0 1 0 2 0

Zone of Pd mineralization Study area 36 Sample site and number of chromite grains normalized to 1 km 10 kg sample mass

30 Chromite abundance contour line 89°40’W 49°09’N

FIGURE 20.23 Pink ribbon-shaped glacial dispersal train showing the number of chromite grains in the 0.25À0.5 HMC mm fraction of till, normalized to 10-kg sample mass, at the Lac des Isles palladium deposits in central Canada. The train was formed by a single phase of ice flow towards the southwest. Modified from Barnett, P.J., Averill, S., 2010. Heavy mineral dispersal trains in till in the area of the Lac des Iles PGE deposit, northwestern Ontario, Canada. Geochem. Explor. Environ. Anal. 10, 391À399.

a PGE deposit is shown for the Lac des Iles area of central Canada in Fig. 20.23 (Barnett and Averill, 2010). A ribbon-shaped glacial dispersal train defined by Cr-andradite and chromite con- tent in till extends at least 5 km down-ice (southwest) from the palladium mineralization (Averill, 2009; Barnett and Averill, 2010).

20.6.12 RARE METAL DEPOSITS Rare metal deposits are significant sources of REE and other high field strength elements such as tantalum, niobium, and zirconium, and are hosted in peralkaline intrusions, carbonatites, pegma- tites, and peraluminous granites (Simandl et al., 2012, 2015). Indicator and pathfinder elements for these deposits include various combinations of the elements listed in Table 20.3. One geochemical example of glacial dispersal from an REE deposit is shown in Fig. 20.5 for the Strange Lake REE 728 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

deposit in eastern Canada. A ribbon-shaped glacial dispersal train defined by Be content in till extends at least 30 km down-ice (northeast) from the intrusion (Batterson and Taylor, 2009). Other useful examples of glacial dispersal from REE-rich sources include Ford et al. (1988), and Sarapa¨a¨ and Sarala (2013). Airborne or ground gamma-ray spectrometry may be used in combination with till geochemistry to detect glacial dispersal trains of uranium, potassium, or thorium-rich debris from these unusual host rocks (e.g., Ford et al., 1988). Indicator minerals of rare metal deposits contain high concentrations of the REE, niobium, tan- talum, yttrium, and zirconium and are listed in Table 20.3. They include carbonates, phosphates, Na-pyroxenes, sodium and potassium amphiboles, titanium-bearing minerals, and zirconosilicates (e.g., Birkett et al., 1992, 1996; Makin et al., 2014).

20.6.13 PORPHYRY DEPOSITS Porphyry deposits are important sources of copper and molybdenum, as well as gold, silver, and tin (Sinclair, 2007). Till geochemical and mineralogical case studies and surveys carried out around porphyry deposits in glaciated terrain are described here in two sections: (1) the well-known copperÀgoldÀmolybdenum porphyry deposits of the North American Cordillera and (2) tinÀtungsten deposits. 20.6.13.1 CopperÀgoldÀmolybdenum porphyry deposits Copper porphyry deposits are significant sources of copper, molybdenum, gold, and silver. Till geo- chemistry is an important exploration tool for these deposits as demonstrated in several studies and regional till surveys (e.g., Nurmi and Isohanni, 1984; Sibbick and Kerr, 1995; Sibbick et al., 1997; Levson, 2001). More recent Canadian case studies include examples from the Highland Valley, Gibraltar, and Mount Polley 1 Huckleberry porphyry deposits, and the Woodjam prospect (Ferbey and Levson, 2009; Hashmi et al., 2015; Plouffe et al., 2016). From these studies a list of the most common indicator and pathfinder elements has been summarized in Table 20.3. Indicator minerals of porphyry deposits include the minerals chalcopyrite, chalcocite, molybde- nite, and gold, as well as high-density oxide and silicate minerals such as andradite garnet, epidote, rutile, titanite, and zircon (Table 20.3) (e.g., Averill, 2011; Kelley et al., 2011; Hashmi et al., 2015; Plouffe et al., 2016). In contrast to other deposit types described here, the indicator mineral suite for porphyry deposits also includes mid-density (2.9À3.2 g cm23) minerals such as tourmaline, apa- tite, and fluorite (Averill, 2011), as well as jarosite, that formed prior to glaciation from the oxida- tion and supergene enrichment of iron sulfides (Kelley et al., 2011). The composition of some indicator minerals such as epidote and magnetite in till can provide further indications of the pres- ence of porphyry mineralization nearby (e.g., Cooke et al., 2014; Nadoll et al., 2015; Canil et al., 2016). 20.6.13.2 TinÀtungstenÀmolybdenum porphyry deposits Intrusion-hosted tinÀtungstenÀmolybdenum deposits are an important source of strategic metals. The use of till geochemistry to explore for these deposits in glaciated terrain is well documented from case studies and exploration programs carried out in the 1970À1990s (e.g., Szabo et al., 1975; Snow and Coker, 1987a,b). One notable example is the use of till geochemistry combined with boulder tracing to outline a tin-rich dispersal fan (Fig. 20.5) and discover what would become the 20.6 EXPLORATION METHODS 729

East Kemptvlille tin deposit in eastern Canada (Rogers et al., 1990). Recent studies by McClenaghan et al. (2014c, 2017a,b) demonstrate the multielement geochemical signatures of two deposits (Sisson, Mount Pleasant) in eastern Canada. In general, indicator and pathfinder elements include tin, tungsten, molybdenum, and in some cases, indium (Table 20.3). In contrast to geochemical methods, the use of indicator mineral methods to explore for tungsten deposits in general is a more recent development. Older reports of the recovery of tin- or tungsten- bearing minerals from a till sample are mostly from follow-up investigations to determine the miner- alogical composition of a single metal-rich till sample (e.g., Brundin and Bergstrom, 1977; Lindmark, 1977; Toverud, 1984; Johansson et al., 1986). Indicator minerals include the main ore minerals, scheelite (Fig. 20.24A and B), wolframite (Fig. 20.24B), cassiterite (Fig. 20.24C), and molybdenite, as well as accessory sulfides, Bi-bearing minerals, fluorite, and topaz (Table 20.3)(Averill, 2001). The visual identification of the scheelite in till HMC is greatly aided by the mineral’s bluish-white fluorescence under short-wave ultraviolet light (Fig. 20.24B). Recent studies by McClenaghan et al. (2017a,b) have documented the indicator mineral signatures of tin and tungsten-bearing porphyry deposits through case studies around the Sisson and the Mount Pleasant deposits.

FIGURE 20.24 Color photographs of: (A) scheelite grains recovered from till overlying the Sisson tungstenÀmolybdenite deposit in eastern Canada, under visible light; (B) scheelite grains from (A) under short-wave ultraviolet light; (C) wolframite, and (D) cassiterite grains recovered from till-down ice of the Mount Pleasant tinÀtungsten deposit in eastern Canada. Mineral photography by Michael J. Bainbridge. 730 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

20.6.14 URANIUM DEPOSITS The highest-grade uranium deposits in the world occur at, or in close proximity to, the unconfor- mity between quartz-rich sandstones in basins and underlying Aphebian and basement rocks, including those in the Athabasca Basin in central Canada (Jefferson et al., 2007; Potter and Wright, 2015). These are termed unconformity uranium deposits. Other uranium deposit types include pegmatites and other granitic deposits, uraniferous iron oxide copperÀgold deposits, ancient fluviatile placers, and uraninite veins and lodes (Boyle, 1982; Tauchid and Underhill, 1997). The tracing of radioactive boulders is a relatively simple and widely used exploration method for uranium deposits in glaciated terrain (e.g., Lundberg, 1973; Gustafsson and Minell, 1977; Hirvas and Ma¨kinen, 1989). The radioactive boulder trains of the Athabasca Basin are among the most numerous and well documented in the world (e.g., Kirwan, 1978; Ramaeker et al., 1982; Wilson, 1985; Kirchner and Tan, 1994). The geochemistry of glacial erratics (composite sampling) in the Athabasca Basin has also proven to be useful (Earle, 2001). Till geochemistry is also an important uranium exploration tool (e.g., Geddes, 1982; Simpson and Sopuck, 1983; Wilson, 1985; Campbell, 2009) and employs indicator and pathfinder elements that are summarized in Table 20.3. Uranium is highly mobile in the oxidizing surface environment, thus the interpretation of uranium geochemical anomalies in till must take into account the potential for both clastic dispersal and hydromorphic dispersion (e.g., Bjo¨rklund, 1976; Boyle, 1982; Peuraniemi and Aario, 1991). Because the ,0.002 mm fraction of till has a greater cation exchange capacity than the ,0.063 mm fraction (Shilts, 1975), it has greater potential to scavenge uranium from hydromorphic solutions, and thus it is a commonly analyzed size fraction. Airborne or ground gamma-ray spectrometry, in combination with surficial geology and geochemistry, can help delin- eate glacial dispersal from radioactive sources and prioritize areas of interest for uranium explora- tion (Fortin et al., 2015). Boyle (1982) suggested a list of potential indicator minerals that is summarized in Table 20.3 and vary with uranium deposit type. Geddes (1982) reported one of the only descriptions of the use of indicator minerals to outline a uranium-rich dispersal train using niccolite and pitchblende in till. This till sampling led to the discovery of a uranium deposit in the eastern Athabasca Basin.

20.7 CONCLUSIONS In glaciated terrain, glaciers have eroded metal-rich debris from mineralization and transported and deposited the debris down-ice in dispersal trains, fans, and other patterns, all of which form larger exploration targets than their bedrock sources. Till geochemical and indicator mineral methods, combined with bedrock mapping and geophysical methods, are used to detect and trace the dis- persed debris up-ice to mineralization. Glacial dispersal may be the result of one or more phases of ice flow and vary in distance down-ice from a few tens of meters to in excess of 100 km. Knowledge of the complexity of continental ice sheets and ice-sheet dynamics during their growth, evolution, and deglaciation, provides additional clues to elucidating the nature of dispersal in complex glacial stratigraphy and/or glacial history such as corridors of fast-flowing ice. The dis- persal pattern often detected at the surface is the net result of all cycles of erosion, transportation, and deposition by one or more glacial events. REFERENCES 731

Boulder tracing and till geochemistry have been exploration tools in glaciated terrain for more than 60 years. The development and advancements of ICP-MS analytical instruments since the 1980s now allow for the determination of an extensive suite of indicator and pathfinder elements for a broad range of commodity and deposit types. Portable XRF is increasingly being used in the field to characterize till geochemistry and guide sampling. Since the early 1990s, the use of indicator mineral methods in glaciated terrain has expanded rapidly such that they are now routinely used to explore for a broad range of commodities and deposit types. Indicator minerals include the more traditional gold and sulfide minerals, as well as more physically robust and chemically stable oxide and silicate minerals. In some cases secondary sulfate and phosphate minerals, formed during preglacial weathering of mineralization, survive gla- cial erosion and transport and can be useful indicator minerals. Microanalytical techniques have evolved such that indicator mineral species may now be rapidly identified in the very fine size frac- tion in addition to the traditional .0.25 mm size range of HMCs. Isotopic compositions may now be determined for single mineral grains and compared to the known characteristics of various min- eral deposit types. Utilizing indicator mineral methods in tandem with till geochemical methods allows for the simultaneous exploration of a broad range of commodities. Although not discussed in this chapter, till compositional data generated during an exploration program can also provide useful data for environmental assessments and planning.

ACKNOWLEDGEMENTS This chapter reflects the cumulative experiences of the authors in drift prospecting and till sampling surveys, but also draws on the published work of fellow researchers at the Geological Survey of Canada (GSC) and the Geological Survey of Finland. The GSC’s Geomapping for Energy and Minerals (GEM) and Targeted Geoscience Initiative (TGI) programs at Natural Resources Canada have supported the recent research in drift prospecting and glacial studies reported here. The authors thank the following individuals and companies for contributing information and/or figures: Pertti Sarala (Geological Survey of Finland); Jan Peter, Isabelle McMartin, Janet Campbell (GSC); and Stu Averill (Overburden Drilling Management Limited). Tom Nowicki (Mineral Services) provided the eclogitic garnets that were photographed in Fig. 20.18. Pertti Sarala is also thanked for his contributions to Appendix A. Michael Bainbridge produced the color photographs of the min- eral grains. Natural Resources Canada Earth Sciences Library, Ottawa staff are thanked for their help in retrieving some of the older reference material used in this review. We thank Tracy Barry and Paul Champagne (GSC) for their production of some of the digital figures included in this chapter. Alain Plouffe, Jan Peter, and Wendy Spirito (GSC) are thanked for their thorough and thoughtful reviews of the manuscript. Natural Resources Canada Earth Sciences Sector (NRCan ESS) Contribution No. 20160420.

REFERENCES Aario, R., Peuraniemi, V., 1992. Glacial dispersal of till constituents in morainic landforms of different types. Geomorphology 6, 9À25. Afanase’ev, V.P., Varlamov, V.A., Garanin, V.K., 1984. The abrasion of minerals in kimberlites in relation to the conditions and distance of their transportation. Geologiya i Geofizika 25/10, 119À125. 732 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

Alcock, F.J., 1941. Jacquet River and Tetagouche River map area, New Brunswick. Geological Survey of Canada, Memoir 227, 34 pp. Allan, R.J., 1974. Trace metal dispersion in an arctic desert landscape: a PbÀZn deposit on Little Cornwallis Island, District of Franklin. Geological Survey of Canada, Paper 74-1, Part B, pp. 51À56. Averill, S.A., 1988. Regional variations in the gold content of till in Canada. In: MacDonald, D.R., Mills, K.A. (Eds.), Prospecting in Areas of Glaciated Terrain-1988. Canadian Institute of Mining and Metallurgy, pp. 271À284. Averill, S.A., 1990. Drilling and sample process methods for deep till geochemistry surveys: making the right choices. Application de la geologie´ du Quaternaire a` l’exploration minerale,´ Association Professionnelle des geologues´ et des geophysiciens´ du Quebec,´ pp. 139À173. Averill, S.A., 2001. The application of heavy indicator minerals in mineral exploration with emphasis on base metal indicators in glaciated metamorphic and plutonic terrain. In: McClenaghan, M.B., Bobrowsky, P.T., Hall, G.E.M., Cook, S. (Eds.), Drift Exploration in Glaciated Terrain. Geological Society of London, London, pp. 69À82, Special volume 185. Averill, S.A., 2009. Useful NiÀCuÀPGE versus kimberlite indicators in surficial sediments: similarities and differences. In: Paulen, R.C., McMartin, I.M. (Eds.), Application of Till and Stream Sediment Heavy Mineral and Geochemical Methods to Mineral Exploration in Western and Northern Canada. Geological Association of Canada, Canada, pp. 125À139, Short Course Notes 18. Averill, S.A., 2011. Viable indicators in surficial sediments for two major base metal deposit types: NiÀCuÀPGE and porphyry Cu. Geochem. Explor. Environ. Anal. 11, 279À291. Averill, S.A., 2013. Discovery and delination of the Rainy River gold deposit using glacially dispersed gold grains sampled by deep overburden drilling: a 20 odyssey. In: Paulen, R.C., McClenaghan, M.B. (Eds.), New Frontiers for Exploration in Glaciated Terrain. Geological Survey of Canada, Canada, pp. 37À46, Open File 7374. Averill, S.A., Zimmerman, J.R., 1986. The riddle resolved: the discovery of the Partridge gold zone using sonic drilling in glacial overburden at Waddy Lake, Saskatchewan. Can. Geol. J. Can. Inst. Min. Metallur. 1, 14À20. Averill, S.A., 2017. Blackwater gold-spessartine-pyrolusite dispersal trains, British Columbia, Canada; influ- ence of sampling depth on indicator mineralogy and geochemistry. Geochem. Explor. Environ. Anal. 17, 43À60. Bajc, A.F., Hall, G.E.M., 2000. Geochemical responses of surficial media, north and East Ranges, Sudbury Basin. Ontario Geological Survey, Open File Report 6033. Barnett, P.J., 2007. Overburden geochemical signature of the Lac des Iles platinum group element deposit, northwestern Ontario, Canada. Can. J. Earth Sci. 44, 1151À1168. Barnett, P.J., Averill, S., 2010. Heavy mineral dispersal trains in till in the area of the Lac des Iles PGE deposit, northwestern Ontario, Canada. Geochem. Explor. Environ., Anal. 10, 391À399. Barr, D.A., 1978. Chapelle gold-silver deposit, British Columbia. Can. Inst. Min. Metallur. Bull. 71, 66À79. Batterson, M.J., Liverman, D., 2000. The contrasting styles of glacial dispersal in Newfoundland and Labrador: methods and case studies. Current Research (2000), Newfoundland Department of Mines and Energy, Geological Survey, Report 2000-1, pp. 1À31. Batterson, M.J., Taylor, D.M., 2009. Geochemical re-analysis of till samples from the Strange Lake area, Labrador (NTS Map Sheets 14D/5 and 24A/8). Government of Newfoundland and Labrador, Department of Natural Resources, Geological Survey, Open File LAB/1479. Bennett, M.R., 2003. Ice streams as the arteries of an ice sheet: their mechanics, stability and significance. Earth Sci. Rev. 61, 309À339. Bird, D.J., Coker, W.B., 1987. stratigraphy and geochemistry at the Owl Creek Gold Mine, Timmins, Ontario, Canada. In: Garrett, R. G. (Ed.), Geochemical Exploration 1985. J. Geochem. Explor. 28: 267À284. REFERENCES 733

Birkett, T.C., McCandless, T.E., Hood, C.T., 2004. Petrology of the Renard igneous bodies: host rocks for dia- mond in the northern Otish Mountains region, Quebec. Lithos 76 (1À4), 475À490. Birkett, T.C., Miller, R.R., Roberts, A.C., Mariano, A.N., 1992. Zirconium-bearing minerals of the Strange Lake intrusive complex, Quebec-Labrador. Can. Mineral. 30, 191À205. Birkett, T.C., Trzcienski Jr., W.E., Stirling, J.A.R., 1996. Occurrence and compositions of some Ti-bearing minerals in the Strange Lake intrusive complex, Quebec-Labrador boundary. Can. Mineral. 34, 779À801. Bjo¨rklund, A., 1976. Use of till in geochemical uranium exploration. Exploration for Uranium Ore Deposits, Proceedings of a Symposium. International Atomic Energy Agency, Vienna, pp. 297À309, March 29ÀApril 2, 1976. Bølviken, B., Gleeson, C.F., 1979. Focus on the use of soils for geochemical exploration in glaciated terrain. In: Hood, P.J. (Ed.), Geophysics and Geochemistry in the Search for Metallic Ores, Proceedings of Exploration 77. Geological Survey of Canada, Economic Geology Report 31, Ottawa, pp. 295À326. Bouchard, M.A., Salonen, V.-P., 1989. Glacial dispersal of boulders in the James Bay lowlands of Quebec,´ Canada. Boreas 18, 189À199. Bouchard, M.A., Salonen, V.-P., 1990. Boulder transport in shield areas. In: Kujansuu, R., Saarnisto, M. (Eds.), Glacial Indicator Tracing. A.A. Balkema, Rotterdam, pp. 87À107. Boulton, G.S., Clark, C.D., 1990. A highly mobile Laurentide ice sheet revealed by satellite images of glacial lineations. Nature 346, 813À817. Boutroy, E., Dare, S.A.S., Beaudoin, G., Barnes, S.J., Lightfoot, P.C., 2014. Minor and trace element composi- tion of magnetite from NiÀCuÀPGE deposits worldwide and its application to mineral exploration. J. Geochem. Explor. 145, 64À81. Boyle, R.W., 1974. Elemental Associations in Mineral Deposits and Indicator Elements of Interest in Geochemical Prospecting (Revised). Geological Survey of Canada, Canada, Paper 74À145. Boyle, R.W., 1982. Geochemical Prospecting for Thorium and Uranium Deposits, Developments in Economic Geology 16. Elsevier Science Ltd, New York. Bradwell, T., Stoker, M., Krabbendam, M., 2008. Megagrooves and streamlined bedrock in NW Scotland: the role of ice streams in landscape evolution. Geomorphology 97, 135À156. Brereton, W.E., Briggs, D.N., Rollinson, J.P., 1988. Till prospecting in the area of the Farley Lake gold depos- its, northwestern Manitoba, Canada. In: MacDonald, D.R., Mills, K.A. (Eds.), Prospecting in Areas of Glaciated Terrain-1988. Canadian Institute of Mining and Metallurgy, Canada, pp. 225À239. Brown, E.L., 1947. Prospecting in the Granville Lake mineral area, Manitoba. 20 (2), 4. Brummer, J.J., Gleeson, C.F., Hansuld, J.A., 1987. A historical perspective of exploration geochemistry in CanadaÀthe first 30 years. J. Explor. Geochem. 28, 1À39. Brummer, J.J., MacFadyen, D.A., Pegg, C.C., 1992a. Discovery of kimberlites in the Kirkland Lake area, north- ern Ontario, Canada; Part 1, Early surveys and the surficial geology. Explor. Min. Geol. 1 (4), 339À350. Brummer, J.J., MacFadyen, D.A., Pegg, C.C., 1992b. Discovery of kimberlites in the Kirkland Lake area, northern Ontario, Canada; Part II, Kimberlite discoveries, sampling, diamond content, ages and emplace- ment. Explor. Min. Geol. 1 (4), 351À370. Brundin, N.H., Bergstrom, J., 1977. Regional prospecting for ores based on heavy minerals in glacial till. J. Geochem. Explor. 7, 1À19. Burchett, W., 1944. The discovery of the Sullivan Mine. Western Miner. 17 (8), 38À39. Byers, A.R., 1956. Geochemical investigations in the Flin Flon area, northern Saskatchewan. Can. Min. J. 77 (4), 83À86. Campbell, J., 2009. Drift prospecting for uranium in the Athabasca Basin, Saskatchewan. In: Paulen, R.C., McMartin, I. (Eds.), Application of Till and Stream Sediment Heavy Mineral and Geochemical Methods to Mineral Exploration in Western and Northern Canada. Geological Association of Canada, Canada, pp. 207À214, Short Course Notes 18. Canil, D., Grondahl, C., Lacourse, T., Pisiak, L.K., 2016. Trace elements in magnetite from porphyry CuÀMoÀAu deposits in British Columbia, Canada. Ore Geol. Rev. 72 (1), 1116À1128. 734 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

Carlson, J.A., Kirkley, M.B., Thomas, E.M., Hillier, W.D., 1999. Recent Canadian kimberlite discoveries. In: Gurney, J.J., Gurney, J.L., Pascoe, M.D., Richardson, S.H. (Eds.), The J.B. Dawson Volume, Proceedings of the VIIth International Kimberlite, Cape Town. pp. 81À89. Chao, T.T., 1984. Use of partial dissolution techniques in geochemical exploration. J. Geochem. Explor. 20, 101À135. Chao, T.T., Sanzolone, R.F., 1992. Decomposition techniques. J. Geochem. Explor. 44, 65À106. Chapman, R., Curry, G., Sopuck, V., 1990. The Bakos deposit discoveryÀa case history. In: Beck, L.S., Harper, C.T. (Eds.), Modern Exploration Techniques. Saskatchewan Geological Society, pp. 195À212, Special Publication 10. Chapman, R.J., Mortensen, J.K., 2016. Characterization of gold mineralization in the Northern Cariboo Gold District, British Columbia, Canada, through integration of compositional studies of lode and detrital gold with historical placer production: a template for evaluation of orogenic gold districts. Econ. Geol. 111, 1321À1345. Chapman, R.J., Leake, R.C., Bond, D.P.G., Stedra, V., Fairgrieve, B., 2009. Chemical and mineralogical signa- tures of gold formed in oxidizing chloride hydrothermal systems and their significance within populations of placer gold grains collected during reconnaissance. Econ. Geol. 104, 563À585. Charbonneau, R., David, P.P., 1993. Glacial dispersal of rock debris in central Gaspesie,´ Quebec, Canada. Can. J. Earth Sci. 30, 1697À1707. Charbonneau, R., David, P.P., 1995. A shear-diffusion model of till genesis based on the dispersal pattern of indicator rocks in the Grand-Volume till of central Gaspesie,´ Quebec, Canada. Boreas 24, 281À292. Clark, P.U., 1987. Subglacial sediment dispersal and till composition. J. Geol. 95, 527À541. Coker, W.B., DiLabio, R.N.W., 1989. Geochemical exploration in glaciated terrain: geochemical responses. In: Garland, G.D. (Ed.), Proceedings of Exploration ’87. Ontario Geological Survey, Canada, pp. 336À383, Special Volume 3. Coker, W.B., Shilts, W.W., 1993. Geochemical exploration for gold in glaciated terrain. In: Foster, R.P. (Ed.), Gold Metallogeny and Exploration. Chapman and Hall, pp. 336À359. Cooke, D.R., Baker, M., Hollings, P., Sweet, G., Chang, Z., Danyuskevsky, L., et al., 2014. New advances in detecting the distal geochemical footprints of porphyry systemsÀepidote mineral chemistry as a tool for vectoring and fertility assessments. In: Kelley, K.D., Golden, H.C. (Eds.), Building Exploration Capability for the 21st Century. Economic Geology, pp. 127À152, Special Publication 18. Cremer, M., Schlocker, J., 1976. Lithium borate decomposition of rocks, minerals and ores. Am. Mineral. 61, 318À321. Dare, S.A.S., Barnes, S.-J., Beaudoin, G., Meric,´ J., Boutroy, E., Potvin-Doucet, C., 2014. Trace elements in magnetite as petrogenetic indicators. Mineral. Deposit. 49, 785À796. Day, S.J., Broster, B.E., Sinclair, A.J., 1987. Sulphide erratics applied to subglacial exploration: St Elias Mountains, British Columbia. Can. J. Earth Sci. 24, 723À730. DiLabio, R.N.W., 1989. Terrain geochemistry in Canada. In: Fulton, R.J. (Ed.), Quaternary Geology of Canada and Greenland. Geological Survey of Canada, Canada, pp. 645À663, Geology of Canada Series, No. 1. DiLabio, R.N.W., 1990a. Glacial dispersal trains. In: Kujansuu, R., Saarnisto, M. (Eds.), Glacial Indicator Tracing. A.A. Balkema, Rotterdam, pp. 109À122. DiLabio, R.N.W., 1990b. Classification and interpretation of the shapes and surface textures of gold grains from till on the . Current Research, Part C. Geological Survey of Canada, Canada, pp. 323À329, Paper 90-1C. DiLabio, R.N.W., 1995. Residence sites of trace elements in oxidized tills. In: Bobrowsky, P.T., Sibbick, S.J., Newell, J.H., Matysek, P.F. (Eds.), Drift Exploration in the Canadian Cordillera. British Columbia Ministry of Energy, Mines and Petroleum Resources, Canada, pp. 139À148, Paper 1995-2. DiLabio, R.N.W., Coker, W.B. (Eds.), 1989. Drift Prospecting. Geological Survey of Canada, Canada. REFERENCES 735

Dolezal,ˇ J., Povondra, P., Sulcek,ˇ Z., 1968. Decomposition Techniques in Inorganic Analysis. American Elsevier Publishing Company Inc., New York. Donovan, P.R., James, C.H., 1967. Geochemical dispersion in glacial overburden over the Tynagh (Northgate) base metal deposit, west-central Eire. In: Cameron, E.M. (Ed.), Proceedings, Symposium on Geochemical Prospecting, Paper 66-54. Geological Survey of Canada, Ottawa, pp. 89À110, April 1966. Drake, L.D., 1983. Ore plumes in till. J. Geol. 91, 707À713. Dredge, L.A., 1988. Drift carbonate on the Canadian Shield. II: carbonate dispersal and ice flow patterns in northern Manitoba. Can. J. Earth Sci. 25, 783À787. Dredge, L.A., 2000. Carbonate dispersal trains, secondary till plumes, and ice streams in the west Foxe Sector, Laurentide Ice Sheet. Boreas 29, 144À156. Dreimanis, A., 1956. Steep Rock iron ore boulder train. Geological Association of Canada, Proceedings, 8, 32À68. Dreimanis, A., 1958. Tracing ore boulders as a prospecting method in Canada. Can. Inst. Min. Metallur. Trans. 61, 49À56. Dreimanis, A., 1960. Geochemical prospecting for Cu, Pb and Zn in glaciated areas, eastern Canada. International Geological Congress XXI, Norden, Part II, Geological Results of Applied Geochemistry and Geophysics. pp. 7À19. Dreimanis, A., Vagners, U.J., 1971. Bimodal distribution of rock and mineral fragments in basal tills. In: Goldthwait, R.P. (Ed.), TillÀA Symposium. Ohio State University Press, USA, pp. 237À250. Dyke, A.S., 1984. Quaternary geology of Boothia peninsula and northern District of Mackenzie, central Canadian Arctic. Geological Survey of Canada, Canada, Bulletin 407. Dyke, A.S., Morris, T.F., 1988. Drumlin fields, dispersal trains and ice streams in arctic Canada. Can. Geogr. 32, 86À90. Dyke, A.S., 2008. The Steensby Inlet ice stream in the context of the deglacation of Northern Baffin Island, Eastern Arctic Canada. Earth Surf. Process. Landform. 33, 573À592. Earle, S., 2001. Application of composite glacial boulder geochemistry to exploration for unconformity-type ura- nium deposits in the Athabasca Basin, Saskatchewan, Canada. In: McClenaghan, M.B., Bobrowsky, P.T., Hall, G.E.M., Cook, S. (Eds.), Drift Exploration in Glaciated Terrain. Geological Society of London, pp. 225À235, Special Volume 185. Eccles, D.R. 2011. Northern Alberta kimberlite province: the first 20 years. Alberta Geological Survey, Bulletin 65. Ehlers, J., Gibbard, P.L., 2007. The extent and chronology of Cenozoic Global Glaciation. Quater. Int. 164À165, 6À20. Ermengen, S.V., 1957. Geochemical prospecting in Chibougamau. Can. Min. J. 78 (4), 99À104. Evans, D.J.A., Clark, C.D., Rea, B.R., 2008. Landform and sediment imprints of fast glacial flow in the south- west Laurentide Ice Sheet. J. Quater. Sci. 23, 249À272. Evans, D.J.A., Phillips, E.R., Hiemstra, J.F., Auton, C.A., 2006. Subglacial till: formation, sedimentary charac- teristics and classification. Earth Sci. Rev. 78, 115À176. Fabel, D., Stroeven, A., Harbor, J., Kleman, J., Elmore, D., Fink, D., 2002. Landscape preservation under Fennoscandian ice sheets determined from in situ produced 10Be and 26Al. Earth Planet. Sci. Lett. 201, 397À406. Ferbey, T., Levson, V.M., 2009. The influence of ice flow reversals on the vertical and horizontal distribution of trace element values in tills, Huckleberry Mine area, west-central British Columbia. In: Paulen, R.C., McMartin, I. (Eds.), Application of Till and Stream Sediment Heavy Mineral and Geochemical Methods to Mineral Exploration in Western and Northern Canada. Geological Association of Canada, Canada, pp. 177À183, Short Course Notes 18. Finck, P.W., Stea, R.R., 1995. The Compositional Development of Tills Overlying the South Mountain Batholith. Nova Scotia Department of Natural Resources, Mines and Minerals Branch, Nova Scotia, Paper 95-1. 736 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

Fipke, C.E., Gurney, J.J., Moore, R.O., 1995. Diamond exploration techniques emphasising indicator mineral geochemistry and Canadian examples. Geological Survey of Canada, Canada, Bulletin 423. Ford, K.L., DiLabio, R.N.W., Rencz, A.N., 1988. Geological, geophysical and geochemical studies of the Allan Lake carbonatite, Algonquin Park, Ontario. J. Geochem. Explor. 30, 99À121. Fortin, R., Campbell, J.E., Harvey, B.J.A., McCurdy, M.W., Sinclair, L.E., Hanson, M.A., et al., 2015. Ground-truthing of the ‘Eastern Athabasca Basin’ regional airborne gamma-ray survey: context for explo- ration of deeply buried unconformity-related uranium deposits in the Athabasca Basin of northern Saskatchewan. In: Potter, E.G., Wright, D.M. (Eds.), Targeted Geoscience Initiative 4: Unconformity- Related Uranium Systems. Geological Survey of Canada, Canada, pp. 43À51, Open File 7791. Fox, P., Cameron, R., Hoffman, S., 1987. Geology and soil geochemistry of the Quesnel River gold deposit, British Columbia. In: Elliott, I.L., Smee, B.W. (Eds.), Geoexpo ‘86. The Association of Exploration Geochemists, Amsterdam, pp. 61À67. Garrett, R.G., 1971. The dispersion of copper and zinc in glacial overburden at the Louvem deposit, Val d’Or, Quebec. Geochemical Exploration. Canadian Institute of Mining and Metallurgy, Canada, pp. 157À158, Special Volume 11. Garvie, O.G., Robinson, D.N., 1984. The formation of kelyphite and associated subkelyphitic and sculpted sur- faces on pyrope from kimberlite. In: Kornprobst, J. (Ed.), Proceedings of the Third International Kimberlite Conference. Volume 1, Kimberlites and Related Rocks. Elsevier, Amsterdam, pp. 371À382. Geddes, R.S., 1982. The Vixen Lake indicator train, northern Saskatchewan. In: Davenport, P.H. (Ed.), Prospecting in Glaciated Terrain. Canadian Institute Mining Metallurgy, Canada, pp. 264À283. Gent, M., Menendez, M., Toran˜o, J., Torno, S., 2011. A review of indicator minerals and sample processing methods for geochemical exploration. J. Geochem. Explor. 110, 47À60. Goodfellow, W.D., Lydon, J.W., 2007. Sedimentary exhalative (SEDEX) deposits. In: Goodfellow, W.D. (Ed.), Mineral Deposits of Canada: A Synthesis of Major Deposit Types, District Metallogeny, the Evolution of Geological Provinces, and Exploration Methods. Geological Association of Canada, Mineral Deposits Division, Canada, pp. 163À183, Special Publication No. 5. Gosse, J., Bell, T., Gray, J., Klein, J., Yang, G., Finkel, R., 2006. Using cosmogneic isotopes to interpret the landscape record of glaciation: nunataks in Newfoundland? In: Knight, P. (Ed.), Glacier Science and Environmental Change. Blackwell Publishing, Malden, MA/Oxford, pp. 442À446. Govett, G.J.S., 1973. Geochemical exploration studies in glaciated terrain, New Brunswick, Canada. In: Jones, M.J. (Ed.), Prospecting in Areas of Glacial Terrain. Institution of Mining and Metallurgy, Canada, pp. 11À24. Grip, E., 1953. Tracing of glacial boulders as an aid to ore prospecting in Sweden. Econ. Geol. 48, 715À725. Gustafsson, B., Minell, H., 1977. Case history of discovery and exploration of Pleutajokk uranium deposit, northern Sweden. Prospecting in Areas of Glaciated Terrain 1977. Institution of Mining and Metallurgy, London, pp. 72À79. Gustavsson, N., 1992. Recognition of anomaly patterns in regional geochemical investigations. In: Kauranne, K., Salminen, R., Eriksson, K. (Eds.), Regolith Exploration Geochemistry in Arctic Temperate Terrains. Handbook of Exploration Geochemistry, Volume 5. Elsevier, Amsterdam, pp. 217À262. Halet, R.A., 1948. Malartic gold fields mine. Structural Geology of Canadian Ore Deposits. Canadian Institute of Mining and Metallurgy, Jubilee Volume, Mercury Press Limited, Montreal, pp. 868À895. Hall, G.E.M., McClenaghan, M.B., 2013. Field portable XRF in exploration and mining. In: Paulen, R.C., McClenaghan, M.B. (Eds.), New Frontiers for Exploration in Glaciated Terrain. Geological Survey of Canada, Canada, pp. 75À81, Open File 7374. Hall, G.E.M., 1999. Near-total acid digestions. EXPLORE Newsletter, Association of Applied Geochemists, No. 104, pp. 15À19. Hall, G.E.M., Oates, C.J., 2003. Performance of commercial laboratories in analysis of geochemical samples for gold and the platinum group elements. Geochem. Explor. Environ. Anal. 3, 107À120. Hall, G.E.M., McClenaghan, M.B., Page,´ L., 2016. Application of portable XRF to the direct analysis of till samples from various deposit types in Canada. Geochem. Explor. Environ. Anal. 16, 62À84. REFERENCES 737

Hall, G.E.M., Parkhill, M.A., Bonham-Carter, G.F., 2003. Conventional and selective leach geochemical exploration methods applied to humus and B horizon soil overlying the Restigouche VMS deposit, Bathurst Mining Camp, New Brunswick. In: Goodfellow, W.D., McCutcheon, S.R., Peter, J.M. (Eds.), Massive Sulfide Deposits of the Bathurst Mining Camp, New Brunswick and northern Maine. Society of Economic Geologists, Monograph 11, pp. 763À782. Hannon, P., Scott, F., 1975. Lead-zinc exploration in the Gays River District of Nova Scotia. Soc. Min. Eng. Trans. 258, 209À212. Harron, G.A., Middleton, R.S., Durham, R.B., Philipp, A., 1987. Geochemical and geophysical gold explora- tion in the Timmins area, Ontario: a case history. Can. Inst. Min. Metallur. Bull. 80, 52À57. Hart, J.K., Smith, B., 1997. Subglacial deformation associated with fast ice flow, from the Columbia Glacier, Alaska. Sediment. Geol. 111, 177À197. Hashmi, S., Ward, B.C., Plouffe, A., Leybourne, M.I., Ferbey, T., 2015. Geochemical and mineralogical dis- persal in till from the Mount Polley CuÀAu porphyry deposit, central British Columbia, Canada. Geochem. Explor. Environ. Anal. 15, 234À249. Ha¨ttestrand, C., Stroeven, A.P., 2002. A relict landscape in the centre of Fennoscandian glaciation: geomor- phological evidence of minimal Quaternary glacial erosion. Geomorphology 44, 127À143. Ha¨ttestrand, C., Goodwillie, D., Kleman, J., 1999. Size distribution of two cross-cutting drumlin systems in northern Sweden: a measure of selective erosion and formation time length. Annal. Glaciol. 28, 146À152. Heimann, A., Spry, P.G., Teale, G., 2005. Zincian spinel associated with metamorphosed Proterozoic base metal sulphide occurrences, Colorado: A re-evaluation of gahnite composition as a guide in exploration. Can. Mineral. 43, 601À622. Hicock, S.R., 1988. Calcareous till facies north of Lake Superior, Ontario: implications for Laurentide ice streaming. Geograph.´ Phys. Quatern. 42, 120À135. Hildes, D.H.D., Clarke, G.K.C., Flowers, G.E., Marshall, S.J., 2004. Subglacial erosion and englacial sediment transport modelled for North American ice sheets. Quatern. Sci. Rev. 23, 409À430. Hirvas, H., Ma¨kinen, K., 1989. Tracing of uranium-rich boulders at Pahtavuoma, Northern Finland. In: DiLabio, R.N.W., Coker, W.B. (Eds.), Drift Prospecting. Geological Survey of Canada, Canada, pp. 7À12, Paper 89-10. Hirvas, H., Nenonen, K., 1990. Field methods for glacial indicator tracing. In: Kujansuu, R., Saarnisto, M. (Eds.), Glacial Indicator Tracing. A.A. Balkema, Rotterdam, pp. 217À248. Hoffman, E.L., Clark, J.R., Yeager, J.R., 1999. Gold analysis-fire assaying and alternative methods. Explor. Min. Geol. 7, 155À160. Hooke, R.L., Cummings, D.I., Lesemann, J.-E., Sharpe, D.R., 2013. Genesis of dispersal plumes in till. Can. J. Earth Sci. 50, 847À855. Huhta, P., 1993. The Use of Heavy Mineral Concentrates from Till in Gold Exploration in the Late Archean Hattu Schist belt, Ilomantsi, Eastern Finland. Geological Survey of Finland, Finland, pp. 363À372, Special Paper 17. Hussein, A.A., Lochner, C., Bell, K., 2003. Application of Pb isotopes to mineral exploration in the Halfmile Lake area, Bathurst, New Brunswick. Econ. Geol. Monograph 11, 670À688. Hyvarinen, L., Kauranne, K., Yletyinen, V., 1973. Modern boulder tracing in prospecting. Prospecting in Areas of Glacial Terrain. Institue of Mining and Metallurgy, London, pp. 87À95. Isohanni, M., Ohenoja, V., Papunen, H., 1985. Geology and NickelÀCopper Ores of the Nivala Area. Geologic Survey of Finland, Finland, pp. 211À228, Bulletin 333. Jefferson, C.W., Thomas, D.J., Gandhi, S.S., Ramaekers, P., Delaney, G., Brisban, D., et al., 2007. Unconformity-associated uranium deposits of the Athabasca Basin, Saskatchewan and Alberta. In: Goodfellow, W.D. (Ed.), Mineral Deposits of Canada: A Synthesis of Major Deposit-Types, District Metallogeny, the Evolution of Geological Provinces, and Exploration Methods. Geological Association of Canada, Mineral Deposits Division, Canada, pp. 273À305, Special Publication No. 5. 738 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

Johansson, P., Keinanen, V., Lehmuspelto, P., 1986. Geochemical exploration of tungsten in glaciogenic deposits in Soretiapulju, western Finnish Lapland. Prospecting in Areas of Glaciated Terrain-1986. Institution of Mining and Metallurgy, London, pp. 61À67. Kaszycki, C.A., Nielsen, E., Gobert, G., 1996. Surficial geochemistry and response to volcanic-hosted massive sulphide mineralization in the Snow Lake region. In: Bonham-Carter, G.F., Galley, A.G., Hall, G.E.M. (Eds.), EXTECH I: A Multidisciplinary Approach to Massive Sulphide Research in the Rusty Lake-Snow Lake Greenstone Belts, Manitoba. Geological Survey of Canada, Canada, pp. 139À154, Bulletin 426. Kauranne, K. (Ed.), 1976. Conceptual models in exploration geochemistry-Norden 1975. J. Geochem. Explor., 5. pp. 173À420. Kauranne, K., Salminen, R., Eriksson, K., 1992. Regolith exploration geochemistry in arctic temperate terrains, Handbook of Exploration Geochemistry, vol. 5. Elsevier, Amsterdam. Kelley, K.D., Eppinger, R.G., Lang, J., Smith, S.M., Fey, D.L., 2011. Porphyry Cu indicator minerals in till as an exploration tool: example from the giant Pebble porphyry CuÀAuÀMo deposit, Alaska, USA. Geochem. Explor. Environ. Anal. 11, 321À334. Kirchner, G., Tan, B., 1994. Key Lake: the trail of its discovery. Can. Inst. Min. Metallur. Bull. 87, 57À61. Kirkley, M., Mogg, T., McBean, D., 2003. Snap Lake field trip guide. VIIIth International Kimberlite Conference, Slave Province and Northern Alberta Field Trip Guide Book, pp. 67À78. Kirwan, L.D., 1978. The discovery of the Midwest Lake uranium deposit. In: Parslow, G.R. (Ed.), Uranium Exploration Techniques. Saskatchewan Geological Society, Canada, pp. 59À79, Special Publication No. 4. Kjær, K.H., Houmark-Nielsen, M., Richardt, N. 2003. Ice flow patterns and dispersal of erratics at the south- western margin of the late Scandinavian Ice Sheet: signature of palaeo-ice streams. Boreas, 32, 130À148. Kjarsgaard, B.A., Knight, R.D., Plourde, A.P., Reynen, A.M.G., 2014. Portable XRF Spectrometry of Surficial Sediment Samples in the Region of East Arm, Great Slave Lake, Northwest Territories, Canada. Geological Survey of Canada, Canada, Open File 7607. Klassen, R.A., 1997. Glacial history and ice flow dynamics applied to drift prospecting and geochemical exploration. In: Gubins, A.G. (Ed.), Proceedings of Exploration 97, pp. 221À232. Klassen, R.A., 1999. The application of glacial dispersal models to the interpretation of till geochemistry in Labrador, Canada. J. Geochem. Explor. 67, 245À269. Klassen, R.A., 2001. A Quaternary geological perspective on geochemical exploration in glaciated terrain. In: McClenaghan, M.B., Bobrowsky, P.T., Hall, G.E.M., Cook, S.J. (Eds.), Drift Exploration in Glaciated Terrain. Association of Exploration Geochemistry-Geological Society of London, Canada, pp. 1À17, Special Publication 185. Klassen, R.A., 2003. The geochemical and physical properties of till, Bathurst Mining Camp, New Brunswick, Canada. In: Goodfellow, W.D., McCutcheon, S.R., Peter, J.M. (Eds.), Massive Sulfide Deposits of the Bathurst Mining Camp, New Brunswick, and Northern Maine. Society of Economic Geologists, Monograph 11, pp. 763À782. Klassen, R.A., Thompson, F.J., 1993. Glacial History, Drift Composition, and Mineral Exploration, Central Labrador. Geological Survey of Canada, Bulletin 435. Kleman, J., Stroeven, A.P., Lundqvist, J., 2008. Patterns of Quaternary ice sheet erosion and deposition in Fennoscandia and a theoretical framework for explanation. Geomorphology 97, 73À90. Koljonen, T., Malisa, E., 1991. Solubility in aqua regia of selected chemical elements occurring in the fine fraction of till. In: Pulkkinen, E. (Ed.), Environmental Geochemistry in Northern Europe. Geological Survey of Finland, Finland, pp. 49À52, Special Paper 9. Koljonen, T., Gustavsson, N., Noras, P., Tanskanen, H., 1992. Sampling, analyses and data processing. In: Koljonen, T. (Ed.), The Geochemical Atlas of Finland Part 2: Till. Geological Survey of Finland, Espoo, Finland, pp. 14À27. Kontoniemi, O., 1998. An Overview of the Geology in the Osikonma¨ki Area, Rantasalmi, South-Eastern Finland: Especially as a Promising Environment for Epigenetic Gold Mineralisation in Geological Setting REFERENCES 739

and Characteristics of the Tonalite-Hosted Gold Deposit at Osikonmaa¨ki, Rantasalmi, South-Eastern Finland. Geological Survey of Finland, Finland, pp. 7À18, Special Paper 25. Kontoniemi, O., Mursu, J., 2006. Hirsikangas gold prospect in Himanka, western Finland, claim areas Hirsi 1 and 2 (7847/1 and 8036/1). Geological Survey of Finland, Report M19/2413/2006/1/10, 37 p. Korkiakoski, E.A., Karvinen, A., Pulkkinen, E., 1989. Geochemistry and hydrothermal alteration of the komatiite-hosted Pahtavaara gold mineralization, Finnish Lapland. Geological Survey of Finland, Current Research 1988, Special Paper 10, pp. 83À89. Kujansuu, R. and Saarnisto, M. (eds), 1990. Glacial Indicator Tracing, A.A. Balkema, Rotterdam, 252 p. Lahtinen, R., Lestinen, P., Savolainen, H., 1993. The use of total and partial dissolution till geochemical data in delineating favorable areas for Ni prospects: an example from the Tampere-Ha¨meenlinna area, southern Finland; Geological Survey of Finland, Special Paper 18, pp. 107À111. LaLonde, J-P., Bernier, L., Choinie`re, J., Hebert,´ C., 1994. Dispersion de gahnite a` partir des gıˆte polymetalli-´ ques de Montauban et du lac Dussault, Comte´ de Portneuf. Quebec Ministe`re des Ressources naturelles, MB 94-42. Layton-Matthews, D., Hamilton, C., McClenaghan, M.B., 2014. Mineral chemistry: modern techniques and applications to exploration. In: McClenaghan, M.B., Plouffe, A., Layton-Matthews, D. (Eds.), Application of Indicator Mineral Methods to Mineral Exploration. Geological Survey of Canada, Canada, pp. 9À18, Open File 7553. Lehtonen, M.L., Lahaye, Y., O’Brien, H., Lukkari, S. Marmo, J.S., and Sarala, P. 2015. Novel technologies for indicator mineral-based exploration. Geological Survey of Finland, Special Paper 57, pp. 23À62. Lehtonen, M.L., Marmo, J.S., Nissinen, A.J., Johanson, B.S., Pakkanen, L.K., 2005. Glacial dispersal studies using indicator minerals and till geochemistry around two eastern Finland kimberlites. J. Geochem. Explor. 87, 19À43. Lestinen, P., Kontas, E., Niskavaara, H., Virtassalo, J., 1991. Till geochemistry of gold, arsenic and antimony in the Seina¨joki district, western Finland. J. Geocehm. Explor. 39, 343À361. Lett, R.E., 1995. Analytical methods for drift. In: Bobrowsky, P.T., Sibbick, S.J., Newell, J.H., Matysek, P.F. (Eds.), Drift Exploration in the Canadian Cordillera. British Columbia Ministry of Energy, Mines and Petroleum Resources, Canada, pp. 215À228, Paper 1995-2. Lett, R.E., Jackaman, W., 2002. Geochemical exploration models, Volume 1. VMS deposits in south-central British Columbia (NTS 82L/13, L/14, 82M/4. M/5, M/6 and 92P/1E). British Columbia Ministry of Energy, Mines and Petroleum Resources, Open File 2000-31. Levinson, A.A., 1980. Introduction to Exploration Geochemistry. Applied Publishing, Wilmette, II. Levson, V.M., 2001. Regional till geochemical surveys in the Canadian Cordillera: sample media, methods and anomaly evaluation. In: McClenaghan, M.B., Bobrowsky, P.T., Hall, G.E.M., Cook, S.J. (Eds.), Drift Exploration in Glaciated Terrain. Association of Exploration Geochemistry-Geological Society of London Special Publication 185, pp. 45À68. Lundberg, B., 1973. Exploration for uranium through glacial drift in the Arjeplog district, northern Sweden. In: Jones, M.J. (Ed.), Prospecting in Areas of Glaciated Terrain 1973, pp. 31À43. Lindmark, B., 1977. Till-sampling methods used in exploration for scheelite in Kaustinen, Finland. Prospecting in Areas of Glaciated Terrain-1977. Institution of Mining and Metallurgy, London, pp. 46À48. Lundberg, H., 1957. The discovery of large lead-zinc deposits at Buchans, Newfoundland. Methods and Case Histories in Mining Geophysics, 6th Commonwealth Mining and Metallurgical Congress 141À154. MacEachern, I.J., Stea, R.R., 1985. The Dispersal of Gold and Related Elements in Tills and Soils at the Forest Hill Gold District, Guysborough County, Nova Scotia. Geological Survey of Canada, Canada, pp. 85À118. MacKenzie, G.S., 1958. History of mining exploration in the Bathurst-Newcastle District, New Brunswick. Can. Inst. Min. Metallur. Bull. 61, 156À161. 740 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

Mackay, D.A.R., Simandl, G.J., Ma, W., Redfearn, M., Gravel, J., 2015. Indicator mineral-based exploration for carbonatites and related specialty metals deposits—a QEMSCANs orientation survey, British Columbia, Canada. J. Geochem. Explor. 165, 159À173. Maier, W., Lahtinen, R., O’Brien, H. (Eds.), 2015. Mineral Deposits of Finland. Elsevier, Amsterdam. Makin, S.A., Simandl, G.J., Marshall, D., 2014. Fluorite and its potential as an indicator mineral for carbonatite-hosted rare earth element deposits. Geological Fieldwork 2013. British Columbia Ministry of Energy and Mines, British Columbia Geological Survey, Paper 2014-1, pp. 207À212. Ma¨kinen, J., 1995. Effects of grinding and chemical factors on the generation and composition of the till fine fraction: an experimental study. J. Geochem. Explor. 54, 49À62. Makvandi, S., Ghasemzadeh-Barvarz, M., Beaudoin, G., Grunsky, E.C., McClenaghan, M.B., Duchesne, C., 2016. Principal component analysis of magnetite composition from volcanogenic massive sulfide deposits: case studies from the Izok Lake (Nunavut, Canada) and Halfmile Lake (New Brunswick, Canada) deposits. Ore Geol. Rev. 72, 60À85. Margold, M., Stokes, C.R., Clark, C.D., 2015a. Ice streams in the Laurentide Ice Sheet: identification, charac- teristics and comparison to modern ice sheets. Earth-Sci. Rev. 143, 117À146. Margold, M., Stokes, C.R., Clark, C.D., Kleman, J., 2015b. Ice streams in the Laurentide Ice Sheet: a new mapping inventory. J. Maps 11, 380À395. Margreth, A., Gosse, J.C., Dyke, A.S., 2016. Quantification of subaerial and episodic subglacial erosion rates on high latitude upland plateaus: Cumberland Peninsula, Baffin Island, Arctic Canada. Quatern. Sci. Rev. 133, 108À129. Marshall, S.J., Clarke, G.K., Dyke, A.S., Fisher, D.A., 1996. Geologic and topographic controls on fast flow in the Laurentide and Cordilleran Ice Sheets. J. Geophys. Res. 101, 17,827À17,839. McCandless, T.E., 1990. Kimberlite xenocryst wear in high-energy fluvial systems: experimental studies. J. Geochem. Explor. 37, 323À331. McClenaghan, M.B., 1992. Surface till geochemistry and implications for exploration, Black RiverÀMatheson area, northeastern Ontario. Explor. Min. Geol. 1, 327À337. McClenaghan, M.B., 1994. Till geochemistry in areas of thick drift and its application to gold exploration, Matheson area, northeastern Ontario. Explor. Min. Geol. 3, 17À30. McClenaghan, M.B., 2001. Regional and local-scale gold grain and till geochemical signatures of lode Au deposits in the western Abitibi Greenstone Belt, central Canada. In: McClenaghan, M.B., Bobrowsky, P.T., Hall, G.E.M., Cook, S.J. (Eds.), Drift Exploration in Glaciated Terrain. Geological Society of London, London, pp. 201À223, Special Publication 185. McClenaghan, M.B., 2005. Indicator mineral methods in mineral exploration. Geochem. Explor. Environ. Anal. 5, 233À245. McClenaghan, M.B., 2011. Overview of common processing methods for recovery of indicator minerals from sediment and bedrock in mineral exploration. Geochem. Explor. Environ. Anal. 11, 265À278. McClenaghan, M.B., Cabri, L.J., 2011. Gold and Platinum Group Element indicator minerals in surficial sedi- ments. Geochem. Explor. Environ. Anal. 11, 251À263. McClenaghan, M.B., Kjarsgaard, B.A., 2001. Indicator mineral and geochemical methods for diamond explora- tion in glaciated terrain in Canada. In: McClenaghan, M.B., Bobrowsky, P.T., Hall, G.E.M., Cook, S.J. (Eds.), Drift Exploration in Glaciated Terrain. Geological Society of London, London, pp. 83À123, Special Publication 185. McClenaghan, M.B., Kjarsgaard, B.A., 2007. Indicator mineral and surficial geochemical exploration methods for kimberlite in glaciated terrain: examples from Canada. In: Goodfellow, W.D. (Ed.), Mineral Deposits of Canada: A Synthesis of Major Deposit-Types, District Metallogeny, the Evolution of Geological Provinces, and Exploration Methods. Geological Association of Canada, Mineral Deposits Division, Canada, pp. 983À1006, Special Publication No. 5. McClenaghan, M.B., Peter, J.M., 2016. Geochemical signatures of volcanogenic massive sulfide deposits: an overview of Canadian examples. Geochem. Explor. Environ. Anal. 16, 27À47. REFERENCES 741

McClenaghan, M.B., Thorleifson, L.H., DiLabio, R.N.W., 1997. Till geochemical and indicator mineral meth- ods in mineral exploration, In: Gubins, A.G., (Ed.), Proceedings of Exploration 97: Fourth Decennial Conference on Mineral Exploration, pp. 233À248. McClenaghan, M.B., Thorleifson, L.H., DiLabio, R.N.W., 2000. Till geochemical and indicator mineral meth- ods in mineral exploration. Ore Geol. Rev. 16, 145À166. McClenaghan, M.B., Bobrowsky, P.T., Hall, G.E.M., Cook, S. (Eds.), 2001. Drift Exploration in Glaciated Terrain. Geological Society of London, London, Special Volume 185. McClenaghan, M.B., Ward, B.C., Kjarsgaard, I.M., Kjarsgaard, B.A., Kerr, D.E., Dredge, L.A., 2002a. Indicator mineral and till geochemical dispersal patterns associated with the Ranch Lake kimberlite, Lac de Gras region, NW Territories, Canada. Geochem. Explor. Environ. Anal. 2, 299À320. McClenaghan, M.B., Kjarsgaard, B.A., Kjarsgaard, I.M., 2002b. Kimberlite indicator mineral content and geo- chemistry of till around the Peddie kimberlite, Lake Timiskaming, Ontario. Geological Survey of Canada, Open File 4262. McClenaghan, M.B., Kjarsgaard, I.M., Kjarsgaard, B.A., 2004. Kimberlite indicator mineral chemistry and till geochemistry around the Seed and Triple B kimberlites, Lake Timiskaming, Ontario, Geological Survey of Canada, Open File 4822. McClenaghan, M.B., Layton-Matthews, D., Matile, G., 2011. Till geochemical signatures of magmatic NiÀCu deposits, Thompson Nickel Belt, Manitoba, Canada. Geochem. Explor. Environ. Anal. 11, 145À159. McClenaghan, M.B., Kjarsgaard, I.M., Kjarsgaard, B.A., 2012. Kimberlite indicator mineral chemistry of the Bucke and Gravel kimberlites and associated indicator minerals in till, Lake Timiskaming, Ontario, Geological Survey of Canada, Open File 5815. McClenaghan, M.B., Plouffe, A., McMartin, I., Campbell, J.E., Spirito, W.A., Paulen, R.C., et al., 2013a. Till sampling and geochemical analytical protocols used by the Geological Survey of Canada. Geochem. Explor. Environ. Anal. 13, 285À301. McClenaghan, M.B., Kjarsgaard, I.M., Averill, S.A., Layton-Matthews, D., Crabtree, D., Matile, G., et al., 2013b. Indicator mineral signatures of magmatic Ni-Cu deposits, Thompson Nickel Belt, Manitoba part 2- till data, Geological Survey of Canada, Open File 7200. McClenaghan, M.B., Holmes, D.R.S., Averill, S.A., Paulen, R.C., Layton-Matthews, D., 2014a. Physical fea- tures indicating the glacial transport distance of gahnite from the Izok Lake CuÀZnÀPbÀAg VMS deposit, Nunavut. Geological Survey of Canada, Open File 7603. McClenaghan, M.B., Ames, D.E. Buckle, J.L., Bajc, A.F., 2014b. Till geochemical signatures of the Broken Hammer Cu-(Ni)-PGE occurrence, North range, Sudbury Structure, Ontario. Geological Survey of Canada, Open File 7602. McClenaghan, M.B., Seaman, A.A., Parkhill, M.A., Pronk, A.G., 2014c. Till geochemical signatures associated with the Sisson W-Mo deposit, New Brunswick, Canada. Geol. 50, 116À137. McClenaghan, M.B., Paulen, R.C., Layton-Matthews, D., Hicken, A.K., Averill, S.A., 2015a. Glacial dispersal of gahnite from the Izok Lake ZnÀCuÀPbÀAg VMS deposit, northern Canada. Geochem. Explor. Environ. Anal. 15, 333À349. McClenaghan, M.B., Peter, J.M., Layton-Matthews, D., 2015b. Overview of VMS exploration in glaciated ter- rain using indicator minerals, till geochemistry, and boulder tracing: a Canadian perspective. In: Peter, J. M., Mercier-Langevin, P. (Eds.), Targeted Geoscience Initiative 4 Contributions to the Understanding of Volcanogenic Massive Sulphide Genesis and Exploration Methods Development. Geological Survey of Canada, Canada, pp. 81À100, Open File 7853. McClenaghan, M.B., Parkhill, M.A., Pronk, A.G., Seaman, A.A., McCurdy, M., Leybourne, M.I., in 2017 a. Indicator mineral and geochemical signatures associated with the Sisson W-Mo deposit, New Brunswick, Canada. Geochem. Explor. Environ. Anal. Online doi: 10.1144/geochem2015-396. McClenaghan, M.B., Parkhill, M.A., Pronk, A.G., Sinclair, W.D., in 2017b. Indicator mineral signatures of SnÀW deposits: examples from the Mount Pleasant SnÀWÀMoÀBiÀIn deposit, New Brunswick, Canada. J. Geochem. Explor. 172, 151À166. 742 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

McMartin, I., 2009. Till composition along the Meliadine Trend near Rankin Inlet, Nunavut: Applications to gold exploration in permafrost terrain. In: Paulen, R.C., McMartin, I. (Eds.), Application of Till and Stream Sediment Heavy Mineral and Geochemical Methods to Mineral Exploration in Western and Northern Canada. Geological Association of Canada, Canada, pp. 153À166, Short Course Notes 18. McMartin, I., Campbell, J.E., 2009. Near surface till sampling protocols in shield terrain, with examples from western and northern Canada. In: Paulen, R.C., McMartin, I. (Eds.), Application of Till and Stream Sediment Heavy Mineral and Geochemical Methods to Mineral Exploration in Western and Northern Canada. Geological Association of Canada, Canada, pp. 75À95, Short Course Notes 18. McMartin, I., Henderson, P.J., 2004. Evidence from Keewatin (central Nunavut) for paleo-ice divide migra- tion. Geograp.´ Phys. Quatern. 58, 163À186. McMartin, I., McClenaghan, M.B., 2001. Till geochemistry and sampling techniques in glaciated shield ter- rain: a review. In: McClenaghan, M.B., Bobrowsky, P.T., Hall, G.E.M., Cook, S.J. (Eds.), Drift Exploration in Glaciated Terrain. Association of Exploration Geochemistry-Geological Society of London, London, pp. 19À43, Special Publication 185. McMartin, I., Paulen, R.C., 2009. Ice flow indicators and the importance of ice flow mapping for drift pro- specting. In: Paulen, R.C., McMartin, I. (Eds.), Application of Till and Stream Sediment Heavy Mineral and Geochemical Methods to Mineral Exploration in Western and Northern Canada. Geological Association of Canada, Canada, pp. 15À34, Short Course Notes 18. McMartin, I., Henderson, P.J., Nielsen, E., and Campbell, J.E., 1996. Surficial geology, till and humus compo- sition across the Shield margin, north-central Manitoba and Saskatchewan: geospatial analysis of a glaci- ated environment; Geological Survey of Canada, Open File 3277. Miller, J.K., 1979. Geochemical dispersion over massive sulphides within the continuous permafrost zone, Bathurst Noresemines, Canada. Prospecting in Areas of Glaciated Terrain 1979. Institution of Mining and Metallurgy, London, pp. 101À110. Miller, J.K., 1984. Model for clastic indicator trains in till. Prospecting in Areas of Glaciated Terrain 1984. Institute of Mining and Metallurgy, London, pp. 69À77. Miller, W.G., 1901. Iron ores of Nipissing District. Ontario Bureau of Mines, vol. 10, 165 pp. Milthers, V., 1909. Scandinavian indicator boulders in the Quaternary deposits; extension and distribution. Danmarks Geologiske Undersøgelse, Series 2 (23), 154. Money, P.L., Heslop, J.B., 1976. Geology of the Izok Lake massive sulphide deposit. Can. Min. J. 97, 24À28. Moore, E.S., Armstrong, H.S., 1948. Iron deposits in the District of Algoma. Ontario Department of Mines, 55th Annual Report, Part IV, pp. 49. Morrison, I.R., 2004. Geology of the Izok Lake massive sulphide deposit, Nunavut Territory, Canada. Explor. Min. Geol. 13, 25À36. Morrissey, C.J., Romer, D.M., 1973. Mineral exploration in glaciated regions of Ireland. In: Jones, M.J. (Ed.), Prospecting in Areas of Glacial Terrain 1973. Institute of Mining and Metallurgy, London, pp. 45À53. Mosig, R.W., 1980. Morphology of indicator minerals, a guide to proximity of source. In: Glover, J.E., Grovers, D.E. (Eds.), Kimberlites and Diamonds. Geology Department and Extension Service. University of Western Australia Publication No. 5, Australia, pp. 81À87. Nadoll, P., Mauk, J.L., Hayes, T., Koenig, A.E., Box, S.E., 2012. Geochemistry of magnetite from hydrother- mal ore deposits and host rocks of the Belt Supergroup, United States. Econ. Geol. 107, 1275À1292. Nadoll, P.N., Mauk, J.L., Leveille, R.A., Koenig, A.E., 2015. Chemistry of magnetite from porphyry and skarn deposits in the southwestern United States. Mineral. Deposit. 50, 493À515. Nawrocki, P.E., Romer, D.M., 1979. A buried anomaly associated with a float train in central Ireland. Prospecting in Areas of Glaciated Terrain 1979. Institution of Mining and Metallurgy, London, pp. 40À44. Neary, G.N., 1981. Mining history of the Buchans area. In: Swanson, E.A., Strong, D.F., Thurlow, J.G. (Eds.), The Buchans Orebodies: Fifty Years of Geology and Mining. Geological Association of Canada, Canada, pp. 1À64, Special Paper 22. REFERENCES 743

Nevalainen, R., 1989. Lithology of fine till fractions in the Kuhmo greenstone belt area, eastern Finland. In: Perttunen, M. (Ed.), Transport of Glacial Drift in Finland. Geological Survey of Finland, Finland, pp. 59À65, Special Paper 7. Ney, C.S., Anderson, J.M., Panteleyev, A., 1972. Discovery, geologic setting and style of mineralization, Sam Goosly deposit, British Columbia. Can. Inst. Min. Metallur. Bull. 65, 53À64. Nichol, I., Bjo¨rklund, A., 1973. Glacial geology as a key to geochemical exploration in areas of glacial over- burden with particular reference to Canada. J. Geochem. Explor. 2, 133À170. Noras, P., 1992. Analytical aspects. In: Kauranne, K., Salminen, R., Eriksson, K. (Eds.), Regolith Exploration Geochemistry in Arctic and Temperate Terrains. Handbook of Exploration Geochemistry, vol. 5. Elsevier, Amsterdam, pp. 185À216. Nowicki, T.E., Moore, R.O., Gurney, J.J., Baumgartner, M.C., 2007. Diamonds and associated heavy minerals in kimberlite: a review of key concepts and applications. Dev. Sedimentol. 58, 1235À1267. Nurmi, A., 1975. Geochemical exploration in a glacial ice-divide region: Riikonkoski copper ore deposit, Kittila¨, Finnish Lapland. In: Jones, M.J. (Ed.), Prospecting in Areas of Glaciated Terrain, 1975. Institution of Mining and Metallurgy, London, pp. 54À59. Nurmi, A. 1976. Geochemistry of the till blanket at the Talluskanava Ni-Cu ore deposit, Tervo, Central Finland. Geological Survey of Finland, Report of Investigation No. 15, 84 p. Nurmi, P.A., Isohanni, M., 1984. Rock, till and stream-sediment geochemistry in the search for porphyry-type Mo-Cu-Au deposits in the Proterozoic Rautio Batholith, western Finland. J. Geochem. Explor. 20, 209À228. O’Brien, J.J., Spry, P.G., Teale, G.S., Jackson, S.E., Koenig, A.E., 2015. Gahnite composition as a means to fingerprint metamorphosed massive sulfide and non-sulfide zinc deposits. J. Geochem. Explor. 159, 46À61. Oviatt, N.M., McClenaghan, M.B., Paulen, R.C., and Gleeson, S.A. 2013. Till geochemical signatures of the Pine Point Pb-Zn Mississippi Valley-type District, Northwest Territories; Geological Survey of Canada, Open File 7320. Oviatt, N.M., Gleeson, S.A., Paulen, R.C., McClenaghan, M.B., Paradis, S., 2015. Characterization and dis- persal of indicator minerals associated with the Pine Point Mississippi Valley-type (MVT) district, Northwest Territories, Canada. Can. J. Earth Sci. 52, 776À794. Ozyer, C.A., 2011. Ice-movement history and kimberlite indicator mineral dispersal study, Pelly Bay, lower Boothia Peninsula, and Wager Plateau areas, Nunavut, Canada. PhD Thesis, Western University, Electronic Thesis and Dissertation Repository, Paper 226. Pankka, H., Keina¨nen, V., Valkama, J., 2006. Kaarestunturin kultakonglomeraateista Suurikuusikon kultaesiin- tyma¨a¨n À Ilkka Ha¨rko¨sen panos kultageologina (with English summary). Geologi 58, 4À8. Papunen, H. 1986. One hundred years of ore exploration in Finland. Finland Geological Survey, Bulletin 336, pp. 165À203. Paradis, S., Hannigan, P., Dewing, K., 2007. Mississippi Valley-Type lead-zinc deposits. In: Goodfellow, W. D. (Ed.), Mineral Deposits of Canada: A Synthesis of Major Deposit-Types, District Metallogeny, the Evolution of Geological Provinces, and Exploration Methods. Geological Association of Canada, Mineral Deposits Division, Canada, pp. 185À203, Special Publication No. 5. Parent, M., Paradis, S., Doiron, A., 1996. Palimpsest glacial dispersion trains and their significance to drift pro- specting. J. Geochem. Explor. 56, 123À140. Parent, M., Paradis, S.J., Boisvert, E., 1995. Ice flow patterns and glacial transport in the eastern Hudson Bay region: implications for the late Quaternary dynamics of the Laurentide Ice Sheet. Can. J. Earth Sci. 32, 2057À2070. Parkhill, M.A., Doiron, A., 2003. Quaternary geology of the Bathurst Mining Camp and implications for base metal exploration using drift prospecting. In: Goodfellow, W.D., McCutcheon, S.R., Peter, J.M. (Eds.), Massive Sulfide Deposits of the Bathurst Mining Camp, New Brunswick, and Northern Maine. Society of Economic Geologists, Economic Geology Monograph 11, pp. 631À660. 744 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

Paulen, R.C., 2001. Glacial transport and secondary hydromorphic metal mobilization: examples from the southern interior of British Columbia, Canada. In: McClenaghan, M.B., Bobrowsky, P.T., Hall, G.E.M., Cook, S.J. (Eds.), Drift Exploration in Glaciated Terrain. Association of Exploration Geochemistry- Geological Society of London, London, pp. 323À337, Special Publication 185. Paulen, R.C., 2009a. Drift prospecting in northern Alberta- a unique terrain for exploration. In: Paulen, R.C., McMartin, I. (Eds.), Application of Till and Stream Sediment Heavy Mineral and Geochemical Methods to Mineral Exploration in Western and Northern Canada. Geological Association of Canada, Canada, pp. 185À205, Short Course Notes 18. Paulen, R.C., 2009b. Sampling techniques in the Western Canada Sedimentary Basin and the Cordillera. In: Paulen, R.C., McMartin, I. (Eds.), Application of Till and Stream Sediment Heavy Mineral and Geochemical Methods to Mineral Exploration in Western and Northern Canada. Geological Association of Canada, Canada, pp. 49À74, Short Course Notes 18. Paulen, R.C., McClenaghan, M.B., 2015. Late Wisconsin ice flow history in the Buffalo Head Hills kimberlite field, north-central Alberta. Can. J. Earth Sci. 52, 51À67. Paulen, R.C., McMartin, I. (Eds.), 2009. Application of Till and Stream Sediment Heavy Mineral and Geochemical Methods to Mineral Exploration in Western and Northern Canada. Geological Association of Canada, Canada, Short Course Notes 18. Paulen, R.C., Paradis, S., Plouffe, A., Smith, I.R., 2011. Pb and S isotopic composition of indicator minerals in glacial sediments from NW Alberta, Canada: implications for Zn-Pb base metal exploration. Geochem. Explor. Environ. Anal. 11, 309À320. Paulen, R.C., McClenaghan, M.B., Hicken, A.K., 2013. Regional and local ice flow history in the vicinity of the Izok Lake ZnÀCuÀPbÀAg deposit, Nunavut. Can. J. Earth Sci. 50, 1209À1222. Pell, J., Gru¨tter, H., Neilson, S., Lockhart, G., Dempsey, S., Grenon, H., 2013. Exploration and discovery of the Chidliak Kimberlite Province, Baffin Island, Nunavut, Canada’s Newest Diamond District. In: Pearson, D. G., Gru¨tter, H.S., Harris, J.W., Kjarsgaard, B.A., O’Brien, H., Rao, N.V.C., Sparks, S. (Eds.), Proceeding of the 10th International Kimberlite Conference, Bangalore, India, vol. 2. Springer, India, pp. 209À227. Peter, J.M., Layton-Matthews, D., Piercey, S., Bradshaw, G., Paradis, S., Boulton, A., 2007. Volcanic-hosted massive sulphide deposits of the Finlayson Lake District, Yukon. In: Goodfellow, W.D. (Ed.), Mineral Deposits of Canada: A Synthesis of Major Deposit Types, District Metallogeny, the Evolution of Geological Provinces, and Exploration Methods. Geological Association of Canada, Mineral Deposits Division, Canada, pp. 471À508, Special Publication 5. Peuraniemi, V., 1985. Interpretation of heavy mineral geochemical results from till. Geological Survey of Finland, Special Paper 3, pp. 169À179. Peuraneimi, V., 1987. Interpretation of heavy mineral geochemical results from till. Geological Survey of Finland, Special Paper 3, pp. 169À179. Peuraniemi, V., 1990. Heavy minerals in glacial material. In: Kujansuu, R., Saarnisto, M. (Eds.), Glacial Indicator Tracing. A.A. Balkema, Rotterdam, pp. 165À185. Peuraniemi, V., Aario, R., 1991. Hydromorphic dispersion of uranium in a surficial environment in northern Finland. In: Rose, A.W., Taufen, P.M. (Eds.), Geochemical Exploration 1989, J. Geochem. Explor., 41. pp. 197À212. Plouffe, A., 1995. Drift-prospecting sampling methods. In: Bobrowsky, P.T., Sibbick, S.J., Newell, J.M., Matysek, P.F. (Eds.), Drift Exploration in the Canadian Cordillera, British Columbia. British Columbia Ministry of Energy, Mines and Petroleum Resources, Canada, pp. 43À52, Paper 1995-2. Plouffe, A., Paulen, R.C., Smith, I.R., 2006. Indicator mineral content and geochemistry of glacial sediments from northwest Alberta (NTS 84L, M): new opportunities for mineral exploration. Geological Survey of Canada, Open File 5121; Alberta Energy and Utilities Board, Special Report 77. Plouffe, A., Anderson, R.G., Gruenwald, W., Davis, W.J., Bednarski, J.M., Paulen, R.C., 2011. Integrating ice flow history, , geology, and geophysics to trace mineralized glacial erratics to their bedrock source: an example from south-central British Columbia. Can. J. Earth Sci. 48, 1113À1130. REFERENCES 745

Plouffe, A., McClenaghan, M.B., Paulen, R.C., McMartin, I., Campbell, J.E., Spirito, W.A., 2013. Processing of unconsolidated glacial sediments for the recovery of indicator minerals protocols used at the Geological Survey of Canada. Geochem. Explor. Environ. Analy. 13, 303À316. Plouffe, A., Ferbey, T., Hashmi, S., Ward, B.C., 2016. Till geochemistry and mineralogy: vectoring towards Cu porphyry deposits in British Columbia, Canada. Geochem. Explor. Environ. Anal. 16, 213À232. Available from: http://dx.doi.org/10.1144/geochem2015-398. Potter, E.G., Wright, D.M., 2015. Unconformity-related uranium deposits synthesis: tools to aid deep exploration and refine the genetic model. In: Potter, E.G., Wright, D.M. (Eds.), Targeted Geoscience Initiative 4: Unconformity-related Uranium Systems. Geological Survey of Canada, Canada, pp. 1À13, Open File 7791. Prest, V.K., 1990. Laurentide ice flow patterns: a historical review, and implications of the dispersal of Belcher Island erratics. Geogr.´ Phys. Quatern. 44, 113À136. Prest, W.H., 1911. Prospecting in Nova Scotia, Nova Scotia Min. Soc. J., 16. pp. 73À91. Primmer, S., Bell, T., Batterson, M., 2015. The role of ice dynamics on drift dispersal in the Newfoundland : Preliminary investigations. Current Research (2015), Newfoundland and Labrador Department of Natural Resources, Geological Survey Report 15-1, pp. 277À285. Puranen, R., 1990. Modelling the glacial transport of tills. In: Kujansuu, R., Saarnisto, M. (Eds.), Glacial Indicator Tracing. Balkema, pp. 15À34. Ramaeker, P., Mackie, B.W., Roy, C.E., 1982. The study of glacial erratics in prospecting for unconformity type uranium deposits benath thick sandstone cover, Athabasca Basin, Saskatchewan. In: Davenport, P.H. (Ed.), Prospecting in Glaciated Terrain. Canadian Institute Mining Metallurgy, Canada, pp. 249À263. Rice, J.M., Paulen, R.C., Menzies, J.M., McClenaghan, M.B., Oviatt, N.M., 2013. Glacial stratigraphy of the Pine Point Pb-Zn mine site, Northwest Territories. Geological Survey of Canada, Current Research, 2013-5. Rice, J.M., McClenaghan, M.B., Paulen, R.C., Ross, M.A., Pyne, M.D., and Lion, A.J., 2017. Indicator mineral abundance data for till samples from the south Core Zone, Quebec and Labrador (NTS 23P and 23I): sam- ples collected in 2015. Geological Survey of Canada, Open File 8187. Rickaby, H.C., 1932. Bannockburn gold area. Ontario Department of Mines, 41st Annual Report, Part II, pp. 19. Robinson, S.V.J., Jefferson, C.W., Paulen, R.C., Layton-Matthews, D., Joy, B., and Quirt, D. 2016. Till and bedrock heavy mineral signatures of the Kiggavik uranium deposits, Nunavut. Geological Survey of Canada, Open File 7771. Rogers, P.J., Chatterjee, A.K., Aucott, J.W., 1990. Metallogenic domains and their reflection in regional lake sedi- ment surveys from the Meguma Zone, southern Nova Scotia, Canada. J. Geochem. Explor. 39, 153À174. Rose, A.W., Hawkes, H.E., Webb, J.S., 1979. Geochemistry in Mineral Exploration. Academic Press, London, p. 657. Ross, M., Campbell, J.E., Parent, M., Adams, R.S., 2009. Palaeo-ice streams and the subglacial landscape mosaic of the North American mid-continental prairies. Boreas 38, 421À439. Ryder, J.M., 1995. Recognition and interpretation of flow direction indicators for former glaciers and meltwa- ter streams. In: Bobrowsky, P.T., Sibbick, S.J., Newell, J.H., Matysek, P.F. (Eds.), Drift Exploration in the Canadian Cordillera. British Columbia Ministry of Energy, Mines and Petroleum Resources, Canada, pp. 1À22, Paper 1995-2. Salminen, R., Hartikainen, A., 1986. Tracing gold, molybdenum and tungsten mineralization by use of a step by step geochemical study in llomantsi, eastern Finland. Prospecting In Areas of Glaciated Terrain-1986. Institution of Mining and Metallurgy, London, pp. 201À209. Salminen, R., Hartikainen, A., Lestinen, P., 1989. Enrichment of As and S in the finest fractions of rock sam- ples—an experimental study simulating the formation of till. J. Geochem. Explor. 32, 287À290. Salonen, V.-P., 1988. Application of glacial dynamics, genetic differentiation of glacigenic deposits and their landforms to indicator tracing in the search for ore deposits. In: Goldthwait, R.P., Matsch, C.L. (Eds.), Genetic Classification of Glacigenic Deposits. A. A. Balkema, Rotterdam, pp. 183À190. Saltikoff, B., 1984. Boulder tracing and mineral identification data bank in Finland. Prospecting in Areas of Glaciated Terrain 1984. Institution of Mining and Metallurgy, London, pp. 179À191. 746 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

Sappin, A.-A., Dupuis, C., Beaudoin, G., McMartin, I., McClenaghan, M.B., 2014. Optimal ferromagnetic frac- tion in till samples along ice flow paths: case studies from the Sue-Dianne and Thompson deposits, Canada. Geochem. Explor. Environ. Anal. 14, 315À329. Sarala, P., 2015. Comparison of different techniques for basal till sampling in mineral exploration, in: Geological Survey of Finland, Special Paper 57, pp. 11À22. Sarala, P., 2016. Comparison of different portable XRF methods for determining till geochemistry. Geochem. Explor. Environ. Anal. 16, 181À192, http://dx.doi.org/10.1144/geochem2012-162. Sarala, P., Peuraneimi, V., 2007. Exploration using till geochemistry and heavy minerals in the ribbed moraine area of southern Finland. Geochem. Explor. Environ. Anal. 7, 195À205. Sarala, P., Peuraniemi, V., Aario, R., 1998. Glacial geology and till geochemistry in ore exploration in the Tervola area, southern Finnish Lapland. Bull. Geol. Soc. Finland 70, 19À41. Sarala, P., Pulkkinen, E., Ojala, J.V., Peltoniemi-Taivalkoski, A., 2009. Gold exploration using till at Peta¨ja¨lehto, northern Finland. Geochem. Explor. Environ. Anal. 9, 247À255. Sarapa¨a¨, O., Sarala, P., 2013. Rare earth element and gold exploration in glaciated terrain: example from the Ma¨ka¨ra¨ area, northern Finland. Geochem. Explor. Environ. Anal. 13, 131À143. Sarapa¨a¨, O., Ani, T.A., Lahti, S.I., Lauri, L.S., Sarala, P., Torppa, A., et al., 2013. Rare earth exploration potential in Finland. J. Geochem. Explor. 133, 25À41. Sauerbrei, J.A., Pattison, E.F., Averill, S.A., 1987. Till sampling in the Casa-Berardi gold area, Quebec: a case history in orientation and discovery. J. Geochem. Explor. 28, 297À314. Sauramo, M., 1924. Tracing of glacial boulders and its application in prospecting. Bull. Commission Geol.´ Finlande 67, 37. Sederholm, J.J. 1911. Extension du glacier continental dans l’Europe septentrionale et transport de blocs erra- tiques fennoscandiens (Atlas de Finlande 1910, Carte No. 5). Fennia, 30, pp. 56À61. Shaw, J., Sharpe, D., Harris, J., 2010. A flowline map of glaciated Canada based on remote sensing data. Can. J. Earth Sci. 47, 89À101. Shelp, G.S., Nichol, I., 1987. Distribution and dispersion of gold in glacial till associated with gold mineraliza- tion in the Canadian Shield, J. Geochem. Explor., 27. pp. 315À336. Shilts, W.W., 1973. Glacial dispersal of rocks, minerals and trace elements in Wisconsinan till, southeastern Quebec, Canada. Geological Society of America, Memoir 136, pp. 189À219. Shilts, W.W., 1975. Principles of geochemical exploration for sulphide deposits using shallow samples of gla- cial drift, Can. Inst. Min. Metallur. Bull., 68. pp. 73À80. Shilts, W.W., 1976. Glacial till and mineral exploration. In: Legget, R.F. (Ed.), Glacial Till: An Interdisciplinary Study. Royal Society of Canada, Canada, pp. 205À223, Special Publication, 12. Shilts, W.W., 1978. Nature and genesis of mudboils, central Keewatin, Canada. Can. J. Earth Sci. 15, 1053À1068. Shilts, W.W., 1984. Till geochemistry in Finland and Canada. J. Geochem. Explor. 21, 95À117. Shilts, W.W., 1993. Geological Survey of Canada’s contributions to understanding the composition of glacial sediments. Can. J. Earth Sci. 30, 333À353. Shilts, W.W., 1996. Glacial drift exploration. In: Menzies, J. (Ed.), Past Glacial Environments: Sediment Forms and Techniques. Butterworth-Heineman, UK, pp. 411À438. Shilts, W.W., Kettles, I.M., 1990. Geochemical-mineralogical profiles through fresh and weathered till. In: Kujansuu, R., Saarnisto, M. (Eds.), Glacial Indicator Tracing. A.A. Balkema, Rotterdam, pp. 187À216. Sibbick, S.J., Fletcher, W.K., 1993. Distribution and behaviour of gold in soils and till at the Nickel Plate Mine, southern British Columbia, Canada. J. Geochem. Explor. 47, 183À200. Sibbick, S.J., Kerr, D.E., 1995. Till geochemistry of the Mount Milligan area, north-central British Columbia; recommendations for drift exploration for Porphyry CuÀAu mineralization. In: Bobrowsky, P.T., Sibbick, S.J., Newell, J.H., Matysek, P.F. (Eds.), Drift Exploration in the Canadian Cordillera. British Columbia Ministry of Energy, Mines and Petroleum Resources, Canada, pp. 167À180, Paper 1995À2. Sibbick, S.J, Balma, R.G., Dunn, C.E., 1997. Till Geochemistry of the Mount Milligan area (Parts of 93N/1 and 93O/4). Geological Survey of Canada, Open File 3291. REFERENCES 747

Simandl, G.J., Prussin, E.A., Brown, N., 2012. Specialty metals in Canada. British Columbia Geological Survey, Open File 2012-7. Simandl, G.J., Mackay, D.A.R., Ma, X., Luck, P., Gravel, J., Grcic, B., et al., 2015. Direct and indirect indica- tor minerals in exploration for carbonatite and related ore deposits—an orientation survey, British Columbia, Canada. In: Application of Indicator Mineral Methods to Mineral Exploration, 27th International Applied Geochemistry Symposium, April 20À24, 2015, Tucson, Arizona. Association of Applied Geochemists, Short course SC02, 7 pp. Simpson, M.A., Sopuck, V.J., 1983. Till geochemistry near the Midwest uranium deposit. In: Cameron, E.M. (Ed.), Uranium Exploration in Athabasca Basin, Saskatchewan, Canada. Geological Survey of Canada, Canada, pp. 207À214, Paper 82-11. Sinclair, W.D., 2007. Porphyry deposits. In: Goodfellow, W.D. (Ed.), Mineral Deposits of Canada: A Synthesis of Major Deposit-Types, District Metallogeny, the Evolution of Geological Provinces, and Exploration Methods. Geological Association of Canada, Mineral Deposits Division, Canada, pp. 223À243, Special Publication No. 5. Slaght, A., 1893. Second report of the Bureau of Mines (Ontario), Legislative Assembly of Ontario. Ontario Bureau of Mines, 2nd Report, 257 pp. Smith, S.L. and Rainbird, R.H., 1987. Soft-sediment deformation structures in overburden drill core, Quebec. Curr. Res. A: Geol. Surv. Canada, 87-1A, p. 53À50. Snow, R.J., Coker, W.B., 1987. Overburden geochemistry related to W-Cu-Mo mineralization at Sisson Brook, New Brunswick, Canada: an example of short- and long-distance glacial dispersal. In: Garrett, R.G. (Ed.), Geochemical Exploration 1985. J. Geochem. Explor., 28. pp. 353À368. Spagnolo, M., Phillips, E., Piotrowski, J.A., Rea, B.R., Clark, C.D., Stokes, C.R., et al., 2016. Ice stream motion facilitated by a shallow-deforming and accreting bed. Nat. Commun. 6. 7, 10723, http://dx.doi.org/ 10.1038/ncomms10723. Spirito, W.A., McClenaghan, M.B., Plouffe, A., McMartin, I. Campbell, J., Paulen, R.C., et al., 2011. Till sampling and analytical protocols for GEM Projects: from field to archive. Geological Survey of Canada, Open File 6850. Spry, P.G., Scott, S.D., 1986. The stability of zincian spinels in sulfide systems and their potential as explora- tion guides for metamorphosed massive sulfide deposits. Econ. Geol. 81, 1446À1463. Stanley, C.R., 2009. Geochemical, mineralogical, and lithological dispersal models in glacial till: physical pro- cess constraints and application in mineral exploration. In: Paulen, R.C., McMartin, I. (Eds.), Application of Till and Stream Sediment Heavy Mineral and Geochemical Methods to Mineral Exploration in Western and Northern Canada. Geological Association of Canada, Canada, pp. 35À48, Short Course Notes 18. Stea, R.R., 1994. Relict and palimpsest glacial landforms in Nova Scotia, Canada. In: Warren, W.P., Croot, D.G. (Eds.), Formation and Deformation of Glacial Deposits. A.A. Balkema, Rotterdam, pp. 141À158. Stea, R.R., Brown, Y., 1989. Variation in drumlin orientation, form and stratigraphy relating to successive ice flows in southern and central Nova Scotia. Sedimentary Geol. 62, 223À240. Stea, R.R., Finck, P.W., 2001. An evolutionary model of glacial dispersal and till genesis in Maritime Canada. In: McClenaghan, M.B., Bobrowsky, P.T., Hall, G.E.M., Cook, S.J. (Eds.), Drift Exploration in Glaciated Terrain. Geological Society, London, pp. 237À265, Special Publication no. 185. Stea, R.R., Johnson, M., Hanchar, D., 2009. The geometry of KIM dispersal fans in Nunavut, Canada. In: Paulen, R.C., McMartin, I. (Eds.), Application of Till and Stream Sediment Heavy Mineral and Geochemical Methods to Mineral Exploration in Western and Northern Canada. Geological Association of Canada, Canada, pp. 1À13, Short Course Notes 18. Steele, K.G., 1988. Utilizing glacial geology in uranium exploration; Dismal Lake, Northwest Territories, Canada. Boreas 17, 183À194. Stewart, R.A., Mayberry, S.W., Pickerill, M.J., 1988. Composition of till in the vicinity of the Lake Ellen kim- berlite and implications for the source of diamonds in glacial sediments of eastern Wisconsin. In: MacDonald, D.R., Mills, K.A. (Eds.), Prospecting in Areas of Glaciated Terrain-1988. Canadian Institute of Mining and Metallurgy, Canada, pp. 103À120. 748 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

Stokes, C.R., Clark, C.D., 2001. Palaeo-ice streams. Quatern. Sci. Rev. 20, 1437À1457. Stokes, C.R., Clark, C.D., 2003. The Dubawnt Lake palaeo-ice stream: evidence for dynamic ice sheet behaviour on the Canadian Shield and insights regarding controls on ice-stream location and vigour. Boreas 32, 263À279. Stokes, C.R., Tarasov, L., Blomdin, R., Cronin, T.M., Fisher, T.G., Gyllencreutz, R., et al., 2015. On the reconstruction of palaeo-ice sheets: Recent advances and future challenges. Quatern. Sci. Rev. 125, 15À49. Stokes, C.R., Margold, M., Clark, C.D., Tarasov, L., 2016. Ice stream activity scaled to ice sheet volume dur- ing Laurentide Ice Sheet deglaciation. Nature 530, 322À326. Available from: http://dx.doi.org/10.1038/ nature16947. Strand, P., Banas, A., Baumgartner, M., Burgess, J., 2009. Tracing kimberlite indicator mineral dispersal trains: an example from the Churchill Diamond Project, Kivalliq region, Nunavut. In: Paulen, R.C., McMartin, I. (Eds.), Application of Till and Stream Sediment Heavy Mineral and Geochemical Methods to Mineral Exploration in Western and Northern Canada. Geological Association of Canada, Canada, pp. 167À175, Short Course Notes 18. Stumpf, A.J., Broster, B.E., Levson, V.M., 2000. Multiphase flow of the late Wisconsinan Cordilleran ice sheet in western Canada. Geol. Soc. Am. Bull. 112, 1850À1863. Szabo, N.L., Govett, G.J.S., Lajtai, E.Z., 1975. Dispersion trends of elements and indicator pebbles in glacial till around Mt. Pleasant, New Brunswick, Canada. Can. J. Earth Sci. 12, 1534À1556. Taipale, K., Nevalainen, R., Saarnisto, M., 1986. Silicate analyses and normative compositions of the fine frac- tion of till: examples from eastern Finland. J. Sediment. Petrol. 56, 370À378. Tan, B., 1977. Geochemical history case in the Key Lake area. In: Dunn, C.E. (Ed.), Uranium in Saskatchewan. Saskatchewan Geological Society, pp. 323À330, Special Publication no. 3. Tarplee, M.F.V., Meer, J.J.M. van der, 2010. Irish Ice sheet dynamics as indicated by the Tynagh mineral deposit erratic assemblages. Proceedings of the Geologists’ Association 121, 32À42. Tarvainen, T., 1995. The geochemical correlation between coarse and fine fractions of till in southern Finland. J. Geochem. Explor., 54, 187À198. Tarvainen, T., Aatos, S., Ra¨isa¨nen, M.-L., 1996. A method for determining the normative mineralogy of tills. Appl. Geochem. 11, 117À120. Tauchid, M., Underhill, D.H., 1997. Uranium deposits of the world. In: A.G. Gubins (Ed.), Proceedings of Exploration 97: Fourth Decennial Conference on Mineral Exploration, pp. 177À180. Thomson, J.E. 1936. Gold deposits on Shoal Lake (West). Ontario Department of Mines, Annual Report, vol. XLV, Part III, pp. 44À53. Thomson, M., 1983. Control procedures in geochemical analysis. In: Howarth, R.J. (Ed.), Statistics and Data Analysis, Volume 2, Handbook of Exploration Geochemistry. Elsevier Scientific Publishing Company, Amsterdam, pp. 39À58. Tilas, D., 1740. Tanckar om Malmletande, i anledning af lo¨se gra˚stenar. Kongl. Svenska Vetenskaps. Acad. Handlingar 1739À1740, I 190À193. Toverud, O¨ ., 1984. Dispersal of tungsten in glacial drift and humus in Bergslagen, south central Sweden. J. Geochem. Explor. 21, 261À272. Tremblay, L.P., 1950. Fiedmont map-area, Abitibi County, Quebec. Geological Survey of Canada, Canada, p. 99, Memoir 253. Tremblay, T., Paulen, R., 2012. Glacial geomorphology and till geochemistry of central Melville Peninsula, Nunavut. Geological Survey of Canda, Open File 7115. Trommelen, M.S., Ross, M., Campbell, J.E., 2013. Inherited clast dispersal patterns: implications for palaeo- of the SE Keewatin Sector of the Laurentide Ice Sheet. Boreas 42, 693À713. Van Alstine, R.E., 1948. Geology and Mineral Deposits of the St. Lawrence Area, Burin Peninsula, Newfoundland. Newfoundland Geological Survey, p. 35, Bulletin 23. Veillette, J.J., Dyke, A.S., Roy, M., 1999. Ice flow evolution of the Labrador Sector of the Laurentide Ice Sheet: a review, with new evidence from northern Quebec. Quatern. Sci. Rev. 18, 993À1019. APPENDIX A 749

Veillette, J.J., Roy, M., Paulen, R.C., Menard,´ M., St-Jacques, G. 2017. Uncovering the hidden part of a large ice stream of the Laurentide Ice Sheet, northern Ontario, Canada. Quatern. Sci.Rev. 155, 136À158. Wennervirta, H., 1967. Geochemical methods in uranium prospecting in Finland. In: Kvalheim, A. (Ed.), Geochemical Prospecting in Fennoscandia. Interscience, New York, NY, pp. 155À169. Wilkinson, J.J., 2014. Sediment-hosted zinc-lead mineralization: processes and perspectives. Geochemistry of mineral deposits. Treat. Geochem. 13, 219À249. Wilson, H.S., 1948. Lamaque mine. In: Structural Geology of Canadian Ore Deposits. Canadian Institute of Mining and Metallurgy, Jubilee Volume, Mercury Press Limited, Montreal, pp. 882À891. Wilson, J.S., 1985. Exploration in the carswell structure. In: Laine,´ R., Alonso, D., Svab, M. (Eds.), The Carswell Structure Uranium Deposits, Saskatchewan. Geological Association of Canada, Canada, pp. 175À188, Special Paper 29. Wilton, D.H.C., Winter, L.S., 2012. SEM-MLA (Scanning electron microscope-mineral liberation analyzer) research on indicator minerals in till and stream sedimentsÀan example from the exploration for awaruite in Newfoundland and Labrador. Mineralogical Association of Canada, Short Course 42, pp. 265À284.

APPENDIX A: MINERAL DEPOSITS DISCOVERED BY TILL SAMPLING AND BOULDER TRACING A significant number of mineral deposits (e.g., precious metals, base metals, uranium, iron, diamonds) have been discovered in glaciated terrain using various drift prospecting techniques. DiLabio (1989) compiled an extensive list of discoveries that were made in Canada and Fennascandia prior to the diamond rush in the early 1990s and the associated expansion in the use of indicator mineral methods. This appendix is an updated list of selected examples of the discov- ery of major mineral deposits and mines in which boulder tracing, till geochemistry, and indicator mineral methods have contributed. The deposits are organized by commodity. Note that the map- ping of mineralized boulders, cobbles, or pebbles is collectively referred to here as ‘boulder tracing.’

Deposit Locale Commodity Method Reference Taivalja¨rvi Eastern Finland Ag Boulder tracing Papunen (1986) Storbodsund Sweden Ag Boulder tracing Grip (1953) Equity Ore British AgÀCu Soil (till) Ney et al. Columbia, geochemistry (1972) Canada Chappell British AgÀAu Soil (till) Barr (1978) Columbia, geochemistry Canada Hirsikangas Western Au Boulder tracing Kontoniemi and Finland Mursu (2006) Osikonma¨ki Southeast Au Boulder tracing Kontoniemi Finland (1998) Kittila¨ Central Au Till geochemistry Pankka et al. Lapland, (2006) Finland (Continued) 750 CHAPTER 20 APPLICATION OF TILL TO MINERAL EXPLORATION

Continued

Deposit Locale Commodity Method Reference Pahtavaara Northern Au Till geochemistry Korkiakoski Finland et al. (1989) McGill Mines Ontario, Au Boulder tracing Rickaby (1932) Canada Duport Ontario, Au Boulder tracing Thomson Canada (1936) Lamaque Mines Quebec, Au Boulder tracing Wilson (1948) Canada Fisher-Quebec Quebec, Au Boulder tracing Tremblay Canada (1950) Malarctic Gold Quebec, Au Boulder tracing Halet (1948) Fields Canada Rainy River Ontario, Au Indicator minerals Averill (2013) Canada Quesnel River British Au Soil (till) Fox et al. Columbia, geochemistry (1987) Canada Casi Beradi Area Quebec, Au Indicator minerals Sauerbrei et al. Canada (till) (1987) Hoyle Township Ontario, Au Indicator minerals Harron et al. Canada (till) (1987) Waddy Lake EP Saskatchewan, Au Indicator minerals Averill and Zone Canada (till) Zimmerman (1986) Kemi Mine Northern Cr Boulder tracing Maier et al. Finland (2015) Otanma¨ki Central Finland FeÀTiÀV Boulder tracing Sarapa¨a¨ et al. (2013) Steep Rock Ontario, Fe Boulder tracing Dreimanis Canada (1956) Josephine Ontario, Fe Boulder tracing Moore and Canada Armstrong (1948) Burin Peninsula Newfoundland, Fluorite Boulder tracing Van Alstine, Canada (1948) Hitura Mine Central Finland Ni Boulder tracing Isohanni et al. (1985) Lynn Lake Manitoba, NiÀCu Boulder tracing Brown (1947) Canada East Kemptville Nova Scotia, Sn Till geochemistry Rogers et al. Canada and boulder tracing (1990) Rabbit Lake Saskatchewan, U Boulder tracing Tan (1977) Canada Key Lake Saskatchewan, U Boulder tracing Tan (1977) Canada APPENDIX A 751

Continued

Deposit Locale Commodity Method Reference Cluff Lake Saskatchewan, U Boulder tracing Wilson (1985) (D deposit) Canada Midwest Lake Saskatchewan, U Boulder trains Kirwan (1978) Canada Paukkajanaar a Eastern Finland U Boulder tracing Wennervirta Mine (1967) Sisson New WÀMo Till geochemistry Snow and Brunswick, Coker (1987a,b) Canada Nigadoo Mine New ZnÀPbÀCu Boulder tracing MacKenzie Brunswick, (1958) Canada Photo Lake Manitoba, ZnÀCuÀAu Till geochemistry Kaszycki et al. Canada (1996) Mogador Quebec, ZnÀPb Boulder tracing Dreimanis (Vendome) Canada (1958) Izok Lake Nunavut, ZnÀPbÀCu Boulder tracing Money and Canada Heslop (1976) Buchans Newfoundland, ZnÀPbÀCuÀAg Boulder tracing Neary (1981) Canada Fyre Lake Yukon, Canada ZnÀPbÀCuÀAgÀAu Boulder tracing Peter et al. (2007) Newfoundland Zinc Newfoundland, Zn Soil (till Brummer et al. Mines Canada geochemistry) (1987) New Calumet New CuÀPbÀZn Boulder tracing Alcock (1941) Brunswick, Canada Sullivan Mine British PbÀZn Boulder tracing Burchett (1944) Columbia, Canada Tynagh Ireland CuÀPbÀZn Boulder tracing Donovan and James (1967) Chidliak kimberlite Nunavut, Diamonds Indicator minerals Pell et al. province Canada (2013) Buffalo Head Hills Alberta, Diamonds Indicator minerals Eccles (2011) kimberlite province Canada ( and till) Lac de Gras Northwest Diamonds Indicator minerals Fipke et al. kimberlite province Territories, (eskers and till) (1995) Canada Otish Mountains Quebec, Diamonds Indicator minerals Birkett et al. kimberlite field Canada (till) (2004) Kirkland Lake Ontario, Diamonds Indicator minerals Brummer et al. kimberlite field Canada (eskers and till) (1992a,b)