The Nature and Origin of Pebble Dikes An

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 30

THE NATURE AND ORIGIN OF PEBBLE DIKES AND ASSOCIATED

ALTERATION: TINTIC MINING DISTRICT (AG-PB-ZN-AU), UTAH


Douglas M. Johnson1 and Eric H. Christiansen2

ABSTRACT
In many ore deposits throughout the world, brecciation often accompanies or occurs in association with minerali-
zation. Such is the case in the Tintic mining district (Ag-Pb-Zn-Au) of north-central Utah, where unique breccia
features called pebble dikes occur alongside significant mineralization. Pebble dikes are tabular bodies of breccia,
which consist of angular to rounded clasts of quartzite, shale, carbonate, and minor igneous rock cemented in a fine-
grained clastic matrix. All clasts now lie above or adjacent to corresponding source rocks. Dikes are thin, typically
less than 0.3 m wide to as much as 1 m, and can exceed 100 m in length. The average of the largest clast size is
less than 3 cm but correlates positively with pebble dike width. Contacts are sharp and envelopes of fine breccia sur-
round roughly half of the dikes; parallel sub-vertical fractures are also common. Pebble dikes are mostly hosted in an
Eocene rhyolite lava flow, which displays argillic to silicic alteration where in contact with pebble dikes, but the dikes
are also hosted in an assortment of deformed Paleozoic sedimentary rocks. The pebble dikes have a strong northeast
trend, following a regional fabric of northeast-trending strike-slip and oblique-slip faults.

The formation of pebble dikes has been historically attributed to the intrusion of the Silver City stock, the Tintic
district’s main productive intrusion. However, pebble dikes in the eastern part of the Tintic district are spatially asso-
ciated with a previously unrecognized andesitic unit, informally named the porphyry of North Lily, which is texturally,
mineralogically, and chemically distinct from the Silver City stock, and like pebble dikes, was emplaced in northeast-
trending plugs and dikes. In addition to the strong spatial association, there are clasts of the porphyry of North Lily in
the pebble dikes, and quartzite clasts like those in the pebble dikes are found as xenoliths in the porphyry of North
Lily. Some of the igneous dikes are comingled with the pebble dikes. These similarities and interactions suggest
simultaneous formation.

Geochemical alteration patterns revealed by isocon analysis show no enrichments in Cu, Pb, or Zn related to the peb-
ble dikes. This is consistent with their formation before major mineralization. Low-grade alteration associated with
pebble dikes indicates that they formed at elevated temperatures (100-170°C). Stable isotope (O and H) composi-
tions of rhyolite altered by the pebble dikes show they formed in the presence of heated groundwater, with very low
water/rock ratios, and little to no magmatic water association. The overall physical, spatial, and chemical character-
istics of pebble dikes of the Tintic mining district suggest that they formed by the mobilization of breccia during the
explosive expansion of groundwater that had been heated by the intrusion of the porphyry of North Lily. This escape
occurred along pre-existing, northeast-trending faults and fractures and created new fractures with similar orienta-
tions. These phreatic explosions released early hydrothermal fluids and enhanced permeability, serving as “ground
preparation” by forming pathways for later mineralizing fluids.

INTRODUCTION productive mining districts throughout the world. These


include, but are not limited to, Sudbury, Ontario, the por-
Breccias are common accompaniments to hydrothermal phyry copper systems of the Basin and Range, Chile, and
ore deposits worldwide, especially those formed during Butte, Montana, the uranium and gold deposits of the
the emplacement of subduction-related magmas (Silli- Colorado Plateau, and, the lead-zinc deposits of Yugo-
toe, 1985). The formation of breccias is often a critical slavia (Bryner, 1961). Although breccias form through a
step in the creation of necessary permeability for the variety of processes, this study focuses on defining the
transmission of ore fluids (Byrne and Tosdal, 2014). origin of a specific set of intrusive breccias – breccias
Without brecciation, the concentration of metals in eco- formed “by the fragmentation of a rock and its [subse-
nomically-viable quantities could be severely hindered. quent] mobilization by magma or [fluids]” (Wright and
Breccia features, including diatremes, breccia pipes, Bowes, 1963, p. 84) – found in the Tintic mining district
and breccia dikes, have been recognized in multiple (Ag-Pb-Zn-Au) of north-central Utah.

1
Department of Geology 2
Department of Geological Johnson, D.M. and Christiansen, E.H., 2016, The Nature and
Brigham Young University- Sciences Origin of Pebble Dikes and Associated Alteration: Tintic
Idaho, Romney 150 Brigham Young University Mining District (Ag-Pb-Zn-Au), Utah, in Comer, J.B.,
Rexburg, Idaho, 83460 S389 ESC Inkenbrandt, P.C., Krahulec, K.A., and Pinnell, M.L., edi-
doug.m.johnson@gmail.com Provo, Utah 84602 tors, Resources and Geology of Utah’s West Desert: Utah
Geological Association Publication 45, p. 13-42.
14 The Nature and Origin of Pebble Dikes and Associated Alteration: Tintic Mining District (Ag-Pb-Zn-Au), Utah

The Tintic mining district (including the historical Main, tons of silver, lead, zinc, gold, and copper ore total-
North, East, and Southwest subdistricts (Krahulec and ing $3.5 billion at 2006 prices (Krahulec and Briggs,
Briggs, 2006)), lies 50 km southwest (figure 1) of Provo, 2006). Tintic ores are dominated by carbonate-hosted
Utah. It is presently recognized as the second most Ag-Pb-Zn-Au replacement bodies, but minor quartzite-
productive mining district in the state of Utah. Total pro- hosted Cu-Au veins have also been exploited (Bush and
duction, to the present, has resulted in nearly 20 million others, 1960).

Quaternary alluvium
40°15’N

73
Fairfield
Undifferentiated Tertiary
igneous rocks
Laguna Springs Volcanic Group
33.29 ± 0.09 Ma

Laguna Springs intrusive latite


40°10’N

Porphyry of North Lily


~33.29 ± 0.09 Ma
Latite of Rock Canyon
33.80 ± 0.10 Ma
Silver City Stock
40°05’N

~33.80 ± 0.10 Ma
Tintic Mountain Volcanic Group
North 34.64 ± 0.17 Ma
Sunrise Peak intrusion and flows
34.70 ± 0.30 Ma
Packard Quartz Latite
40°00’N

35.27 ± 0.03 Ma
Undifferentiated Paleozoic
East sedimentary rocks
Eureka
6 Fault
36
39°55’N

Main

Rocky
Figure 3 and 9
Mountains
Basin
and
Range
South
39°50’N

Provo
West

0 2 4 8 km

Colorado
39°45’N

Plateau

112°15’W 112°10’W 112°05’W 112°00’W

Figure 1. Simplified geologic map of the East Tintic Mountains, modified from Morris and Lovering (1961), Proctor (1985),
and McKean (2011); unit ages as discussed in text. The red box shows the boundaries of the maps in figures 3 and 9. The
black box on the shaded relief map shows the location of the geologic map relative to Provo, Utah. The boundaries of the
historic Main, North, East, and Southwest Tintic Mining Districts are also shown (Krahulec and Briggs, 2006).

UGA Publication 45 (2016)—Resources and Geology of Utah’s West Desert


Johnson, D.M. and Christiansen, E.H. 15

Breccia features in the Tintic mining district include After deposition, these sedimentary rocks were deformed
tabular alloclastic dikes (non-volcanic rocks disrupted during the Cretaceous Sevier orogeny into large north-
by volcanic processes) composed of rounded clasts of south-trending anticlines and synclines. Sevier orogenesis
multiple rock types set in a fine-grained clastic matrix also created north-south-trending thrust faults, as well
(figure 2). Due to the rounded nature of the individu- as northeast-trending strike- and oblique-slip faults
al clasts, they have been called pebble dikes, as clast (Krahulec and Briggs, 2006), which appear to have
roundness approximates the appearance of streambed accommodated differential movement in the eastward-
pebbles (Farmin, 1934; Kildale, 1938). Similar peb- verging Sevier thrust sheet (McKean and others, 2011).
ble dike features have been found and described in ore All these faults later became important in ore deposition,
deposits at Butte, Montana; Urad, Mt. Emmons, Central as most major replacement ore bodies in the district are
City, and Leadville, Colorado; Bisbee, Arizona; Cuajone now found within carbonate lithologies near the intersec-
and Toquepala, Peru; El Salvador, Chile; and Mt. Morgan, tions of strike-slip faults and thrust faults (Morris and
Australia. Most pebble dike features are associated with Lovering, 1979). Following deformation, a lengthy period
porphyry Cu-Mo deposits, with a smaller amount asso- of geologic quiescence resulted in the erosion of Paleo-
ciated with Ag-Pb-Zn-Au replacement deposits similar zoic strata.
to the Tintic district (Sillitoe, 1985). This study seeks
to define the origin of the pebble dikes specifically in This was followed by intense late Eocene to Oligocene
the East Tintic subdistrict by addressing these primary subduction-related magmatism, which intruded into and
questions: What is the nature of pebble dikes in the East covered Paleozoic strata in the area (figure 3). Given the
Tintic subdistrict? Are these pebble dikes related to a late Tertiary eastward rotation of the Tintic range, vol-
specific intrusive body? How are host rocks altered near canic cover is thickest to the east and thins to the west to
pebble dikes? What is the character and origin of the expose the Paleozoic sedimentary rocks at higher eleva-
fluid involved in the emplacement of pebble dikes? Is tions (Morris and Mogensen, 1978; Morris and Lovering,
there any connection between pebble dike formation and 1979; Krahulec and Briggs, 2006).
ore-generating processes in the Tintic mining district?
The oldest volcanic unit produced during Eocene to
Oligocene magmatism is the Packard Quartz Latite, a
METHODS rhyolite lava flow. Within the boundaries of the East Tin-
tic subdistrict, the Packard covers approximately 60% of
To answer these questions, a comprehensive field and the surface area east of the crest of the East Tintic Moun-
laboratory analysis of pebble dikes and associated intru- tains and is the unit most commonly cut and altered by
sive and extrusive igneous rocks was conducted in the pebble dikes (Morris and Lovering, 1979). A 40Ar/39Ar age
eastern portion of the Tintic mining district, mostly with- on sanidine shows that the Packard Quartz Latite was
in the boundaries of the historical East Tintic subdistrict emplaced 35.27 ± 0.03 Ma (Allen, 2012). The Sunrise
(figures 1 and 3). Field analysis included the loca- Peak intrusion and flows followed the Packard Quartz
tion and description of both pebble dikes and igneous Latite at 34.7 ± 0.3 Ma, bearing no association to Tin-
dikes followed by mapping and analysis in Geographical tic mineralization (Keith and others, 1997). The Tintic
Information System (GIS) software. Laboratory analysis Mountain Volcanic Group, consisting of three distinct
included petrographic microscopy, X-ray diffractometry ash-flow tuffs, came shortly after. A 40Ar/39Ar age on bio-
(XRD), X-ray fluorescence spectrometry (XRF), and sta- tite from one of the three units is 34.64 ± 0.17 Ma (UGS
ble-isotope mass spectrometry. For a full treatment of and NMGRL, 2007).
methods see Johnson (2014).
The monzonite to quartz monzonite of the Silver City
stock was emplaced next, and due to its close associa-
GEOLOGIC SETTING tion with the district’s significant Ag-Pb-Zn-Au resources,
is thought to be the source of both the Tintic district’s
The Tintic mining district is underlain by a basement metals and mineralizing fluids (Morris and Lovering,
sequence of phyllitic slate and quartzite belonging to 1979; Hildreth and Hannah, 1996; Kim, 1997). The
the Neoproterozoic Big Cottonwood Formation that is age of the Silver City stock is not well known, but it has
presumably underlain by Mesoproterozoic schist, gneiss,
been chemically correlated (Kim, 1997) to the latite of
and granite as exposed in the nearby Wasatch Range. A
Rock Canyon (figure 3, previously called the monzonite
thick section of Paleozoic sedimentary rocks lies uncon-
formably on top of this Proterozoic basement (Hintze and porphyry of Gough Sill by Morris and Lovering, 1979).
Kowallis, 2009). These sedimentary rocks are mostly The latite of Rock Canyon is a dark-colored complex of
Cambrian to Mississippian carbonates, but also include lava flows and shallow intrusions (Kim, 1992). This vol-
the massive Cambrian Tintic Quartzite, a buff-colored, canic equivalent to the Silver City stock has a 40Ar/39Ar
medium-grained quartzite, and the Cambrian Ophir age on biotite of 33.80 ± 0.10 Ma, which suggests that
Formation, an interbedded assortment of limestones, the stock was emplaced at or before this age (Moore and
sandstones, and micaceous shales (Morris and Lovering, others, 2007).
1979). The sedimentary rocks of the Tintic mining dis-
trict appear to have accumulated on a rifted continental Pebble dikes in the East Tintic subdistrict, the focus of
margin that subsided throughout much of the Paleozoic this study, were thought to have formed at some point
(Hintze and Kowallis, 2009). during the intrusion of the Silver City stock by Farmin

Comer, J.B., Inkenbrandt, P.C., Krahulec, K.A., and Pinnell, M.L.


16 The Nature and Origin of Pebble Dikes and Associated Alteration: Tintic Mining District (Ag-Pb-Zn-Au), Utah

A B

C D

Figure 2. Photographs of pebble dikes of the East Tintic subdistrict. A – Pebble dike in a roadcut on Highway 6. This
dike is hosted in Packard Quartz Latite and demonstrates the characteristic limonite stain and subvertical dip common
to most pebble dikes. B – Close-up of the pebble dike in A, showing that dikes consist of rounded quartzite fragments in
a fine-grained matrix. This pebble dike is zoned in clast size, with finer fragments in the margin and coarser fragments in
the interior. C – Outcrop of silicified pebble dike, created when the erosional resistance of the pebble dike exceeds that
of its host rock. D – Pebble dike hosted in Cambrian dolomite. A significant breccia envelope extends up and away from
this vertical dike. Breccia on the sides of the dike grades outward into sub-parallel fractures.

UGA Publication 45 (2016)—Resources and Geology of Utah’s West Desert


Johnson, D.M. and Christiansen, E.H. 17

Tls
Qa
Qa
Quaternary
Pu
» North Standard alluvium

Tss
Water Lillie
Tls
39°59'N

» Silver Shield
Quartz Latite
Tls Figure 5
Homansville » Central Standard Tls
» Tap Laguna Springs
Tss Volcanic Group
Tp Qa
» Copper Leaf »
Silver Shield Tnl
39°58'N

6 Porphyry of
Tnl North Lily
Tp 6
Qa North Lily
Trc
» Latite of
Rock Canyon
Eureka Big Hill » Qa
39°57'N

Tsc
» Burgin
Silver City Stock
Pu
Tsc Ttm
Pu Tp Tintic Mountain
Volcanic Group
39°56'N

Tsp
Trixie
Ttm » Sunrise Peak
intrusion and flows
Pu
Tp
Nevada Ttm Packard Quartz
» Tsp Latite
39°55'N

Tsp Trc Tap


Qa
Apex
Tsc Conglomerate
Trc
112°7'W 112°6'W 112°5'W 112°4'W 112°3'W 112°2'W Pu
Undifferentiatied
0 0.5 1 2 Km Faults » Mineshaft Paleozoic
Normal sedimentary
1:58,000 Strike-slip Pebble Dikes rocks

Figure 3. Simplified geologic map of the Tintic mining district modified from Lindgren and Louglin (1919), Morris
(1964), Morris and Lovering (1979), and Keith and others (2009). This geologic map shows the relative locations of
important igneous units and mines, with the town of Eureka included on the western edge of the map for reference.
All Paleozoic sedimentary units are combined into one unit. Pebble dikes (black lines, size exaggerated) mostly occur
within the boundaries of the red box (pebble dikes not shown within the red box, see figure 5), but also occur to the
south between the Trixie and Nevada mineshafts. The intrusive units within the red box (red) were originally mapped
as Silver City stock.

Comer, J.B., Inkenbrandt, P.C., Krahulec, K.A., and Pinnell, M.L.


18 The Nature and Origin of Pebble Dikes and Associated Alteration: Tintic Mining District (Ag-Pb-Zn-Au), Utah

(1934), Kildale (1938), Lovering (1949), Cook (1957), but large glassy masses are locally preserved. Moderate
Morris and Mogensen (1978), and Morris and Lovering alteration consists of mild iron staining and argillization
(1979). This idea was based on the very close spatial of plagioclase. More intense alteration includes the near
association of pebble dikes to a swarm of porphyritic complete destruction of original texture through intense
plugs and dikes that were mapped as Silver City stock argillization or silicification. These areas are easily recog-
by Morris (1979). However, no conclusive chemical or nized due to significant bleaching accompanied by iron
mineralogical correlation was made between these por- staining. Areas of the Packard Quartz Latite that display
phyritic dikes and the main Silver City stock which lies more intense alteration often host swarms of pebble
5 km to the southwest. Below, we show that these dikes dikes.
in the East Tintic subdistrict are related to a separate
intrusion. Silver City Stock
The Laguna Springs Volcanic Group, a group of latite lava The Silver City stock, identified by most as the produc-
flows and tuffs, followed the intrusion of the Silver City tive intrusion of the Tintic mining district, crops out
stock. A 40Ar/39Ar age on hornblende indicates that the primarily in the Main and Southwest subdistricts over a
Laguna Springs Volcanic Group was emplaced 33.29 ± 5 km2 area (figures 1 and 3). It is mostly phaneritic, with
0.09 Ma (Moore and others, 2007). a generally equigranular texture; but it was emplaced at
a shallow level—it intruded and altered volcanic rocks
Magmatism continued after the emplacement of this related to the Sunrise Peak intrusion (Keith and others,
group; however, the composition of subsequent Mio- 2009). Pervasive propylitic alteration applies a greenish
cene volcanic units is transitional to extension-related hue to some of the stock. The stock includes 1 to 2 mm
types, signaling the end of subduction-related mag- crystals of (in order of decreasing abundance) plagio-
matism in the Tintic mining district (Christiansen and clase, K-feldspar, augite, biotite, and quartz. Accessory
others, 2007). Bimodal magmatism and deformation phases include magnetite, apatite, zircon, and titanite.
related to Basin and Range extension followed, creating Secondary uralitic hornblende is common. It is a dis-
north-south-trending, high-angle normal faults that cut tinctly potassic monzonite with high concentrations of
all pre-Miocene rocks and features (McKean and others, Rb, Th, and other incompatible elements compared to
2011; Allen, 2012). other units with similar silica concentrations (figure 4).
Numerous irregular apophyses of the stock extend away
from the main body along bedding planes and fractures
RESULTS in sedimentary wall rock. This possibly indicates that the
stock was emplaced by the stoping of wall rock into the
Characteristics of Igneous Rocks Related intrusion (Morris and Lovering, 1979; Kim, 1992).
to Pebble Dikes
A series of porphyritic plugs and dikes extend away from
Packard Quartz Latite the main body of the Silver City stock to the northeast
(figure 3). There are a few short dikes near the intrusion,
The Packard Quartz Latite, a rhyolite by the current IUGS but other more abundant dikes lie 5 km to the northeast,
chemical classification (Le Bas and Streckeisen, 1991), in a zone of more abundant pebble dikes (figure 3). All
covers the majority of the of the East Tintic subdistrict. of these dikes were mapped as Silver City stock by Morris
From the crest of the East Tintic Mountains in the west, (1964) and Morris and Lovering (1979). With increas-
the Packard Quartz Latite extends eastward in a thicken- ing distance from the main stock, these plugs and dikes
ing wedge between the Trixie and North Standard shafts become increasingly porphyritic and hornblende-rich.
until it is buried by young sediments in the Goshen Valley Despite differences in mineralogy and texture, the plugs
(figure 3). The thickness of the formation is highly vari- and dikes were correlated with the Silver City stock (Mor-
able, varying from under 10 m in the west to greater than ris and Lovering, 1979; Kim, 1992). Due to their close
1000 m in the east based on drilling results summarized spatial association to pebble dikes, these plugs and dikes
by Morris and Lovering (1979). The Packard Quartz Latite are discussed in more detail below.
includes local basal flow breccias and a main porphyritic
unit that is strongly flow foliated and locally brecciated Latite of Rock Canyon
and vitrophyric. Textures and flow structures show that
it was emplaced as a lava flow. Phenocrysts (0.5 to 1 The latite of Rock Canyon is exposed in the southern
mm) comprise roughly 20 to 30% of the rock and con- portions of the Main and East Tintic subdistricts, directly
sist of (in order of decreasing abundance) andesine, east of the Silver City stock (figure 3). It is comprised
biotite, quartz, and sanidine. Accessory phases include of latitic lava flows, lahars, and a vent-filling complex
magnetite, apatite, zircon, and titanite. The groundmass of lava flows and intrusions that vary in thickness from
consists of a microcrystalline aggregate of feldspar and 100 to 300 m. The latite of Rock Canyon is dominantly
quartz or glass (Morris and Lovering, 1979). porphyritic. Phenocrysts comprise approximately 25 to
30% of the rock and consist of 2 to 3 mm crystals of (in
Significant portions of the Packard Quartz Latite display order of decreasing abundance) plagioclase, augite, bio-
moderate to intense alteration (Lovering, 1949). In fact, tite, hornblende, and orthopyroxene. Accessory phases
it is difficult to find truly fresh Packard Quartz Latite, include magnetite and titanite. The latite of Rock Canyon

UGA Publication 45 (2016)—Resources and Geology of Utah’s West Desert


Johnson, D.M. and Christiansen, E.H. 19

12 8
11 A B
7
Trachydacite
10 Quartz Monzonite
6
Latite Rhyolite
9 Monzonite Shoshonitic
Na2O + K2O (wt%)

K2O (wt%)
8 nl
nl
nl nl nlnlnl nl
4 nl nlnl High K
7 Monzodiorite nl' Silic nl nl nl
nl nl
nl
nl nl nl nlnlnl ifica
tion nl
nl
nlnl nlnl nlnl
nl nl nl nl nl' nl
nl nl

Argillization
nl 3
nl nlnl
nl nl nlnl nl
6 nl
nl nl
nl nl nlnlnl Medium K
2
5 nl Dacite
Andesite 1
4 B. Andesite
Low K
3 0
50 55 60 65 70 75 80 50 55 60 65 70 75 80
SiO2 (wt %) SiO2 (wt %)

300 40
C D
250 35

30
200
Rb (ppm)

Th (ppm)
25
150 nl
nl 20
nl nl nl
nl nl
100 nlnl nl
nlnl
nl nl nl nl
nl nl nl nl 15 nl
nl nl nl nl
nl nl nl
nlnl nl
50 10

0 5
50 55 60 65 70 75 80 150 170 190 210 230 250 270 290
SiO2 (wt %) Zr (ppm)

nl

Packard Quartz Latite Silver City Stock Latite of Rock Canyon Porphyry of Laguna Springs
35.3 Ma ~33.8 Ma 33.8 Ma North Lily Volcanic Group
~33.3 Ma 33.3 Ma
Figure 4. These variation diagrams demonstrate the chemical differences between the Silver City stock and the porphyry
of North Lily. A – The Silver City stock is a monzonite to quartz monzonite while the porphyry of North Lily is dominantly
andesite (Le Bas and Streckeisen, 1991). This total alkali-silica diagram also shows alteration trends seen in the Packard
Quartz Latite due to interaction with pebble dikes. B, C – The porphyry of North Lily has lower concentrations of alkali
elements than the Silver City stock, as seen in these K2O (Le Bas and Streckeisen, 1991) and Rb diagrams. D – The
porphyry of North Lily also has lower concentrations of Th and Zr than the Silver City stock. These diagrams identify
volcanic equivalents for the Silver City stock and the porphyry of North Lily, which are the latite of Rock Canyon and
Laguna Springs Volcanic Group, respectively. This equivalency is based on chemical similarity. Actual porphyry of North
Lily data points are marked by ‘nl’. Data sources are referenced in the text.

is generally free of alteration; however, outcrops near the tify the latite of Rock Canyon as the volcanic equivalent
Silver City stock display mild argillic to silicic alteration of the Silver City stock. The accepted age (33.80 ± 0.10
(Kim, 1992). Based on published whole-rock and miner- Ma) for the latite of Rock Canyon is used to estimate the
al compositions (Morris and Lovering, 1979; Kim, 1992; age of the Silver City stock.
Kim, 1997; Moore and others, 2007) and new whole-
rock analyses (table 1 and table 2), the latite of Rock Porphyry of North Lily (Monzonite Porphyry Plugs and
Canyon is very similar to the Silver City stock. In contrast Dikes)
to other igneous rocks considered, both the latite of Rock
Canyon and the Silver City stock are ferroan and sho- Porphyritic plugs and dikes, formerly referred to as “mon-
shonitic. Both have similar enrichments in incompatible zonite porphyry plugs and dikes” (Morris and Lovering,
elements. Moreover, compositions of rocks from each 1979; Kim, 1992; Kim, 1997), occupy a ~4 km-long
unit plot in overlapping fields on many variation diagrams northeast-trending zone in the East subdistrict, 5 km
(figure 4). In agreement with mineral chemistry evidence northeast of the main body of the Silver City stock (figure
from Kim (1997), whole-rock chemistry appears to iden- 3). Only the most northeastern plugs and dikes, bounded

Comer, J.B., Inkenbrandt, P.C., Krahulec, K.A., and Pinnell, M.L.


20 The Nature and Origin of Pebble Dikes and Associated Alteration: Tintic Mining District (Ag-Pb-Zn-Au), Utah

Table 1. Major element compositions (wt %) normalized to 100% on a volatile-free basis of rocks from the TIntic mining district.

Sample Rock SiO2 TiO2 Al2O3 Fe2O3 MnO MgO CaO Na2O K2 O P2 O5 Total
EUREKA-DGK-13 Tnl* 65.39 0.77 15.29 5.36 0.08 1.97 3.62 3.53 3.65 0.35 100.00
EUREKA-WJJ-13 Tp 72.33 0.46 14.98 2.07 0.03 0.43 2.03 2.64 4.90 0.13 100.00
EUREKA-RAF-13 Tp 75.03 0.44 14.81 2.65 0.02 0.41 0.71 1.74 4.06 0.12 100.00
EUREKA-KAG-13A Tp 76.54 0.46 15.84 0.90 0.00 0.18 0.20 0.88 4.93 0.07 100.00
EUREKA-BB-13 Tsc 60.62 1.03 15.66 7.28 0.12 2.70 4.70 3.36 4.11 0.42 100.00
EUREKA-TAK-13 Tsc 63.08 0.89 15.46 6.13 0.10 2.33 3.81 3.47 4.40 0.35 100.00
EUREKA-TDG-13A Tsc 64.07 0.98 16.21 6.78 0.10 2.77 1.02 3.14 4.56 0.37 100.00
EUREKA-DJ-01 Tnl 57.97 1.21 15.30 8.66 0.13 4.15 6.43 3.22 2.48 0.44 100.00
EUREKA-DJ-02 Tnl 61.74 1.04 16.10 7.62 0.10 2.29 3.92 3.24 3.55 0.39 100.00
EUREKA-DJ-03 Tnl 57.10 1.24 15.64 8.67 0.16 3.92 7.20 2.85 2.77 0.46 100.00
EUREKA-DJ-04 Tnl 60.64 1.13 15.48 6.69 0.12 2.48 6.72 2.86 3.34 0.55 100.00
EUREKA-DJ-05 Tnl 59.47 1.23 15.74 7.06 0.10 2.32 6.96 3.25 3.25 0.61 100.00
EUREKA-DJ-06 Tnl 59.71 1.22 15.54 7.17 0.07 3.09 6.04 3.26 3.30 0.61 100.00
EUREKA-DJ-07 Tnl 57.71 1.07 13.75 7.88 0.13 6.92 6.45 2.63 3.01 0.44 100.00
EUREKA-DJ-08 Tnl 55.29 1.16 13.30 8.87 0.22 8.26 7.44 2.48 2.52 0.47 100.00
EUREKA-DJ-09 Tnl 60.71 1.00 15.08 7.31 0.12 3.75 5.33 3.09 3.24 0.37 100.00
EUREKA-DJ-10 Tnl 62.66 0.97 15.59 7.00 0.10 3.08 3.75 3.17 3.33 0.35 100.00
EUREKA-DJ-11 Tnl 65.86 0.80 15.55 4.02 0.09 1.37 5.94 2.95 3.12 0.31 100.00
EUREKA-DJ-12 Tp 98.77 0.47 0.40 0.28 0.00 0.00 0.00 0.00 0.01 0.06 100.00
EUREKA-DJ-PD2a Tp 75.31 0.38 14.26 1.78 0.01 0.65 1.23 1.70 4.55 0.12 100.00
EUREKA-DJ-PD2b Tp 75.60 0.38 14.10 1.45 0.03 0.69 1.60 1.68 4.35 0.12 100.00
EUREKA-DJ-PD2c Tp 71.97 0.48 15.31 1.31 0.07 0.70 2.91 2.20 4.93 0.12 100.00
EUREKA-DJ-PD12a Tp 69.78 0.49 16.51 2.72 0.04 0.79 2.71 1.65 5.19 0.13 100.00
EUREKA-DJ-PD12b Tp 72.16 0.48 15.07 2.83 0.02 0.57 1.45 2.20 5.11 0.11 100.00
EUREKA-DJ-PD12c Tp 72.16 0.43 15.10 2.68 0.02 0.46 1.68 2.59 4.78 0.10 100.00
EUREKA-DJ-PD5 PD 88.92 0.27 5.63 1.99 0.04 0.26 1.78 0.00 1.03 0.07 100.00
EUREKA-DJ-PD12 PD 88.02 0.27 5.15 1.56 0.04 0.26 3.47 0.06 1.08 0.07 100.00
EUREKA-DJ-PD18 PD 88.67 0.31 6.60 1.70 0.01 0.30 0.42 0.34 1.57 0.08 100.00
EUREKA-DJ-PD29 PD 89.35 0.30 6.33 2.41 0.03 0.33 0.33 0.05 0.77 0.11 100.00
EUREKA-DJ-PD30 PD 40.92 0.33 7.19 2.63 0.30 19.31 27.98 0.01 1.20 0.13 100.00
*Tnl = Porphyry of North Lily, Tp = Packard Quartz Latite, Tsc = Silver City stock, PD = Pebble Dike

by the North Lily shaft in the southwest and Copper Leaf Groundmass texture is also highly variable. Although it
shaft in the northeast, are considered here (figure 5). usually comprises 70 to 80% of the rock, the groundmass
Plugs and dikes in this area, all hosted in the Packard ranges from a microgranular aggregate of plagioclase and
Quartz Latite, were studied due to their close spatial quartz to dark black glass. Phenocryst size drops dramat-
association to pebble dikes. ically in smaller plugs and dikes, and mild but pervasive
iron-staining and argillization is present in the majority
Exposures are commonly tabular dikes, ranging in width of thinner dikes. Most dikes strike northeast and appear
from a few centimeters to 2 m. Plugs are less common, to have brecciated their Packard Quartz Latite host, sug-
covering areas of a few hundred square meters. The min- gesting that the emplacement of the porphyry of North
eralogy of the plugs and dikes can vary greatly across Lily was somewhat violent (Morris and Lovering, 1979;
individual exposures, but all have 3 to 5 mm crystals Kim, 1992; Krahulec and Briggs, 2006).
of plagioclase, biotite, and quartz (figure 6). Varying
amounts of hornblende and/or pyroxene are also often New and published whole-rock and mineral data (Mor-
present, but are entirely absent in some samples. Where ris and Lovering, 1979; Kim, 1992; Kim, 1997; Moore
present, hornblende is much more abundant than pyrox- and others, 2007) show there are significant differences
ene. Accessory phases include magnetite and titanite. between the monzonitic Silver City stock and the por-

UGA Publication 45 (2016)—Resources and Geology of Utah’s West Desert


Johnson, D.M. and Christiansen, E.H. 21

Table 2. Trace element concentrations in rocks from the Tintic mining district (ppm).

Sample Unit Ba Ce Cr Cu Ga La Nb Nd Ni Pb Rb Sc Sm Sr Th U V Y Zn Zr

EUREKA-DGK-13 Tnl* 998 95 30 14 19 49 13 37 13 28 119 12 8 589 19 3 88 18 65 170

EUREKA-WJJ-13 Tp 2437 158 1 4 16 92 13 55 2 29 120 5 10 323 21 4 29 21 49 209

EUREKA-RAF-13 Tp 2316 134 1 4 15 82 12 48 0 21 102 7 10 260 19 4 26 22 39 206

EUREKA-KAG-13A Tp 1973 141 1 2 16 92 13 49 2 26 114 3 10 234 18 5 30 13 12 198

EUREKA-BB-13 Tsc 1020 100 13 37 21 50 12 43 11 21 147 18 8 632 20 4 142 25 97 220

EUREKA-TAK-13 Tsc 946 103 27 7 21 51 13 43 13 17 155 16 9 564 24 4 113 24 71 199

EUREKA-TDG-13A Tsc 1044 86 15 18 19 41 12 45 13 13 128 15 9 261 16 5 149 30 107 200

EUREKA-DJ-01 Tnl 885 90 88 27 19 46 13 42 23 17 78 20 8 662 12 2 173 26 88 187

EUREKA-DJ-02 Tnl 954 91 18 17 19 47 13 42 10 20 108 16 9 514 16 4 132 26 88 188

EUREKA-DJ-03 Tnl 1050 96 83 28 20 50 13 43 22 17 72 22 9 657 13 3 171 27 76 189

EUREKA-DJ-04 Tnl 1060 127 55 21 20 65 16 52 26 20 84 17 9 683 14 3 126 23 79 227

EUREKA-DJ-05 Tnl 1622 143 67 26 21 73 17 58 30 20 82 18 11 936 13 3 142 24 90 254

EUREKA-DJ-06 Tnl 1529 144 71 25 21 69 17 59 32 19 87 17 10 933 13 2 147 24 86 263

EUREKA-DJ-07 Tnl 1280 121 536 48 18 68 15 56 166 15 84 21 10 660 15 3 166 23 97 195

EUREKA-DJ-08 Tnl 990 115 560 51 17 60 15 54 203 15 65 21 10 670 12 2 194 24 74 193

EUREKA-DJ-09 Tnl 930 91 81 30 19 48 12 42 22 28 97 18 8 585 16 4 145 23 77 169

EUREKA-DJ-10 Tnl 1053 90 44 22 19 49 13 43 14 20 81 17 8 504 15 3 130 24 84 178

EUREKA-DJ-11 Tnl 979 99 37 74 18 49 13 38 12 70 83 14 7 387 17 4 79 27 55 183

EUREKA-DJ-12 Tp 1238 64 2 0 3 44 15 25 5 6 2 0 7 198 12 3 5 7 2 250

EUREKA-DJ-PD2a Tp 2322 151 1 4 14 89 11 53 2 23 108 4 10 243 19 4 26 18 30 192

EUREKA-DJ-PD2b Tp 2104 139 1 3 14 88 11 50 2 23 97 5 10 254 19 16 27 20 36 190

EUREKA-DJ-PD2c Tp 3077 133 1 3 15 87 14 49 2 21 118 6 9 317 19 3 28 21 35 197


EUREKA-DJ-
Tp 2298 135 1 3 16 80 13 46 2 25 118 5 9 245 22 4 29 17 55 205
PD12a
EUREKA-DJ-
Tp 2413 117 2 3 15 69 13 40 2 25 124 5 9 289 21 4 30 14 44 200
PD12b
EUREKA-DJ-
Tp 2356 118 2 2 15 70 12 43 3 25 117 5 9 339 21 4 27 14 42 210
PD12c
EUREKA-DJ-PD5 PD 134 63 19 71 8 39 6 26 7 46 38 5 6 74 9 5 32 13 36 157

EUREKA-DJ-PD12 PD 874 52 15 11 7 30 6 25 5 40 39 6 5 96 7 2 30 17 12 132

EUREKA-DJ-PD18 PD 778 83 13 8 9 54 7 37 7 11 52 3 8 110 10 3 32 15 16 157

EUREKA-DJ-PD29 PD 68 52 17 15 9 26 6 29 17 25 23 2 6 63 9 2 39 12 78 159

EUREKA-DJ-PD30 PD 74 38 19 13 7 19 8 0 11 13 29 0 0 42 6 2 39 14 85 97
*Tnl = Porphyry of North Lily, Tp = Packard Quartz Latite, Tsc = Silver City stock, PD = Pebble Dike

phyritic plugs and dikes. Unlike the Silver City stock, Laguna Springs Volcanic Group
the porphyritic plugs and dikes are magnesian and high-
K, but not shoshonitic (figure 4). The plugs and dikes The Laguna Springs Volcanic Group is exposed primarily
have significantly lower concentrations of alkali elements in the north to northeastern portions of the East Tintic
(particularly K and Rb) than the Silver City stock, as well subdistrict, and continues on into the North subdistrict
as generally higher concentrations of MgO, Ni, and Cr. (figure 3). Locally, the group is subdivided into the North
The Silver City stock is best described as a monzonite Standard Latite, the Pinyon Queen Latite, and the Tintic
to quartz monzonite, whereas the porphyritic plugs and Delmar Latite. Small intrusive plugs of porphyritic latite
dikes are better described as andesites. Given these are also included (Morris and Lovering, 1979). The three
significant chemical differences, along with previously- volcanic members consist of basal lithic-rich tuffs with
described textural and mineralogical differences, the overlying lava flows, each ranging in thickness to as much
most northeastern porphyritic plugs and dikes are here as 300 m (Morris and Lovering, 1979). Textures are
separated into a new lithologic unit, distinct from the dominantly porphyritic, and the mineral assemblage is
Silver City stock, termed the porphyry of North Lily due similar in all flows. Phenocrysts account for 20 to 25% of
to its proximity to the North Lily shaft. the rock, and include 2 to 5 mm crystals of plagioclase,

Comer, J.B., Inkenbrandt, P.C., Krahulec, K.A., and Pinnell, M.L.


22 The Nature and Origin of Pebble Dikes and Associated Alteration: Tintic Mining District (Ag-Pb-Zn-Au), Utah

Qa
Qa
Pu Quaternary
Qa
alluvium
» Central Standard
Tls
Tp
» Homansville Tls
Laguna Springs
39°58'30"N

DJ-PD30 DJ-PD29 Volcanic Group


6
Pu
Tnl

KAG-13A Pebble
DJ-12 Tap dikes
DJ-PD18 DJ-11

DJ-PD12 RAF-13
Tnl
DJ-06 » Porphyry of
Copper Leaf
DJ-PD5 North Lily
Tp DJ-PD2abc
WJJ-13 6
Tp
DJ-01
DJ-08 Packard Quartz
39°58'0"N

DJ-07 Latite
DJ-04 DJ-05

Tap
DJ-02 Apex
DJ-03
Conglomerate
Tnl
DGK-13
Tp Pu
DJ-09 DJ-10 Undifferentiated
Paleozoic
sedimentary
Tnl Ct rocks
Qa
» North Lily
39°57'30"N

Ct
Tintic
Tp
Quartzite
Big Hill Pu
(off map)
» Mineshaft
Tintic Standard
»
112°4'30"W » 112°4'0"W 112°3'30"W
0 0.25 0.5 1 Km
Sample
1:15,000 location

Figure 5. Simplified geologic map of the main pebble dike swarm. This map shows the strong spatial association
between dikes and plugs of the porphyry of North Lily (red) and pebble dikes (black lines). Pebble dikes appear to occur
in greatest number near outcrops of the porphyry of North Lily. Modified from Morris and Lovering (1979).

UGA Publication 45 (2016)—Resources and Geology of Utah’s West Desert


Johnson, D.M. and Christiansen, E.H. 23

Silver City Stock Porphyry of North Lily

Figure 6. Photographs and photomicrographs (20x magnification, cross-polarized light) of the Silver City stock
(EUREKA-TAK-13) and the porphyry of North Lily (EUREKA-DGK-13). These images show the textural and mineralogical
differences between these two units. The Silver City stock is phaneritic and equigranular with plagioclase, K-feldspar,
biotite, pyroxene, and very minor hornblende. The porphyry of North Lily is strongly porphyritic with phenocrysts of
plagioclase, biotite, quartz, and abundant hornblende.

Comer, J.B., Inkenbrandt, P.C., Krahulec, K.A., and Pinnell, M.L.


24 The Nature and Origin of Pebble Dikes and Associated Alteration: Tintic Mining District (Ag-Pb-Zn-Au), Utah

biotite, hornblende, ±augite, ±quartz, and ±titanite (Mor- well exposed in road cuts (figure 2D). Some dikes are
ris and Lovering, 1979; Kim, 1992). Even though the isolated lenses within fracture systems that extend away
formal unit name calls them latites, all members of the from the lenses (Lovering, 1949). Commonly, unfilled
Laguna Springs Volcanic Group are andesites using the but iron-stained fractures and minor breccia extend for
current IUGS chemical classification (figure 4). several meters above and below some dikes segments. A
few dikes have lateral off shoots or buds consistent with
Examination of whole-rock data from the literature (Mor- propagation to the northeast. Zones of intense fractur-
ris and Lovering, 1979; Kim, 1992; Moore and others, ing, but with no pebble-dike fill, parallel some of the
2007), in conjunction with new whole-rock analyses, dikes or connect dike segments and are most obvious in
reveals distinct similarities between the Laguna Springs carbonate wall rocks. The fracture zones are comparable
Volcanic Group and the porphyry of North Lily. Both are in width and structure to those enclosing pebble dikes
magnesian and high-K, and have similar concentrations with microbreccia cores grading outward to mesobreccia
of alkali elements. The compositions of the two units plot and ultimately to massive wallrock with a few paralleling
in overlapping fields on many variation diagrams (figure fractures. Some fracture zones are slightly altered, show-
4). Based on similar compositions and mineral assem- ing coarser carbonate grains and iron stains formed by
blages, the Laguna Springs Volcanic Group appears to be oxidation of minor sulfide.
the volcanic equivalent of the porphyry of North Lily. Kim
(1997) also correlated the Laguna Springs Volcanic Group As mentioned, most pebble dike clasts are rounded.
and the porphyritic plugs and dikes (now recognized as Roundness covers a spectrum from subangular to rounded
the porphyry of North Lily) through an examination of with more resistant clast types having the highest degree
mineral chemistry. Accordingly, the age (33.29 ± 0.09 of rounding (Farmin, 1934). Clasts include quartzite,
Ma, Moore and others, 2007) for the Laguna Springs Vol- limestone/dolomite, shale, and igneous rock (including
canic Group is used to estimate the age of the porphyry fragments of Packard Quartz Latite and porphyry of North
of North Lily. Lily). As confirmed by XRD and XRF, quartzite makes up
Pebble Dikes the most abundant clast type, representing well over
90% of clasts. Carbonate and shale clasts make up ~4%
General Characteristics each, with the remaining ~2% being igneous clasts. The
percentage of one particular clast type increases where
Pebble dikes in the East Tintic subdistrict are tabular the dike is hosted in the same rock type, e.g., dolomite
bodies of breccia dominated by fragments of multiple clast abundance is higher where a pebble dike is hosted
rock types enclosed in a fine-grained clastic matrix. in dolomite. The average longest dimension of the ten
They are mainly found in a swarm of northeast-trending largest clasts measured in a 1 m long section of pebble
dikes that occurs over a 6 km2 area (figure 5) between dike is 2 cm, but ranges from <0.5 cm to at least 15 cm
the Big Hill and Water Lillie shafts (Morris and Lover- across. Clast size correlates positively with pebble dike
ing, 1979). Pebble dikes are on average 25 cm wide, width; wider pebble dikes carry larger clasts (figure 7).
but can range from 5 to 90 cm (see Johnson (2014) All clasts appear to have originated from common sedi-
for complete pebble dike data). Lengths of many exceed mentary rocks in the area, including the Tintic Quartzite,
100 m; the longest continuous pebble dike measured the Ophir Shale, and a series of Cambrian to Mississip-
exceeds 400 m in length, but some may be even longer pian limestones and dolomites. All clasts now lie well
as outcrops are often difficult to trace. Dips are generally above or adjacent to their point of origin (Farmin, 1934;
subvertical and cut across flow foliation in the Packard Kildale, 1938; Morris and Lovering, 1979).
lava flows. A pervasive tan to brown limonite stain makes
the dikes stand out in outcrop. In roadcuts, pebble dikes Quartzite clasts commonly have an onionskin structure,
are readily recognizable; however, most surface expres- where clast exteriors spall off in fragments that approxi-
sions consist solely of linear trains of quartzite pebbles mate the overall rounded shape of the original clast. This
in the soil (Morris and Lovering, 1979). Pebble dike has been interpreted as a thermal spallation effect, a
thickness and strike are generally consistent throughout product of the rapid heating of the clasts (Farmin, 1934;
outcrop exposures, but locally some pebble dikes display Kildale, 1938; Morris and Lovering, 1979; McCallum,
large deviations in thickness and strike due to fracture 1985). Given quartzite’s poor thermal conductivity, any
or bedding control. Pebble dikes mostly cut the Packard increase in ambient temperature will result in the pref-
Quartz Latite but are also hosted in various Paleozoic erential heating of the exterior of the clast. If heating is
sedimentary rocks (figure 3). Contacts with host rocks extreme enough, thermal expansion of the clast’s exterior
are sharp and pebble dikes commonly have envelopes will result in shallow fractures that parallel the exterior
of closely spaced fractures in the surrounding host rock of the clast. This spallation effect has been experimen-
(figure 2D). Zones of finely brecciated and iron-stained tally shown to occur at temperature differentials of only
rock lie directly above some dikes and grade upward 100°C in sandstone and granite (Smith and Pells, 2008;
into zones of closely spaced (1-5 cm) subvertical frac- Walsh and Lomov, 2013). Some rounding may also be
tures. Microbreccias are immediately adjacent to the the result of abrasion during transport and emplacement.
dikes that grade out to mesobreccia with centimeter- and
decimeter-sized fragments over distances of less than a Pebble dike clasts are hosted in a matrix of fine-grained
few meters. The three-dimensional shapes of the dikes clastic material (figure 8). Matrix types include two varie-
are blade-like; the upper and lower limits of several are ties: a loose, friable medium- to fine-grained matrix of

UGA Publication 45 (2016)—Resources and Geology of Utah’s West Desert


Johnson, D.M. and Christiansen, E.H. 25

the same material as the clasts themselves, or a dark- 8


colored jasperoid of clastic material that was likely A y = 0.077x
cemented by hydrothermal fluids (Farmin, 1934). Pebble 7 r2 = 0.69

Average largest clast size (cm)


dikes with friable matrices are easily weathered and form n = 40
recessive channels in outcrop. Dikes with silicified matri- 6
ces are much more resistant to erosion, to the point that 5
they exceed the resistance of their host rock and form
easily-recognizable linear protrusions (figure 2C). Pebble 4
dikes with friable matrices are much more abundant. Of
special note, some pebble dikes are zoned in clast and 3
matrix sizes (figure 2B). These dikes have fine-grained
2
margins and coarser-grained interiors that are separated
by sharp contacts. This suggests multiple generations of 1
pebble dike formation, with younger pebble dikes form-
ing in the same fractures occupied by older pebble dikes. 0
0 10 20 30 40 50 60 70 80 90 100
Pebble dike width (cm)
Most pebble dikes have pervasive argillic to silicic altera-
tion that permeates the dikes and extends into rhyolite 20
host rock (figure 2A). Alteration haloes typically extend 18 B Mean: 2.4 ± 1.6 cm 1σ
only centimeters from the pebble dike/host rock contact,
16 Median: 2.3 cm
but in some places where pebble dikes are densely spaced
over a distance of a few meters, alteration effects overlap Mode: 2.1 cm
14 n = 40
producing wider swaths of argillized or silicificified rock.
Little to no alteration occurs in carbonate hosts, although 12
Frequency
most carbonates near pebble dikes are lightly silicified or 10
dolomitized, which is probably related to fluids flowing
into carbonate wall rock after pebble dike emplacement. 8
Additionally, most pebble dikes host small amounts of 6
disseminated sulfides, which are mostly pyrite. Second-
ary oxidation of sulfides in pebble dikes provides the 4
characteristic limonite stain. Although sulfides are often 2
present, whole-rock chemical analysis shows that the
shallow pebble dikes studied here do not host any signifi- 0
0 1 2 3 4 5 6
cant Pb or Zn mineralization characteristic of the Tintic Average largest clast size (cm)
mining district (table 2). Concentrations of Pb range from
11 to 46 ppm and Zn from 12 to 85 ppm; such concen- 18
trations are very similar to or only slightly higher than 16
C
Mean: 22 ± 18 cm 1σ
the host Packard Quartz Latite, indicating no remarkable
Median: 17 cm
enrichment. 14
Mode: 15 cm
12 n = 40
Relationship of Pebble Dikes and the
Frequency

Porphyry of North Lily 10

8
As noted above, the area of densest pebble dike occur-
rence overlaps with outcrops of the porphyry of North 6
Lily (figure 3). Within the bounds of the map in figure
5, wherever pebble dikes are found, exposures of the 4
porphyry of North Lily occur nearby. Pebble dikes have 2
been mapped extending into or emanating directly from
North Lily dikes (Morris, 1964). A few composite dikes 0
0 10 20 30 40 50 60 70 80 90
crop out and linear trains of quartzite pebbles can almost Pebble dike width (cm)
always be found within a few meters of a North Lily dike
outcrop. Likewise, Lovering (1949) found that some peb- Figure 7. Characteristics of pebble dikes. A – The average
ble dikes “rest on the top” of igneous dikes, some pebble largest clast size increases with pebble dike width,
dikes have intrusive cores, many formed alongside igne- although the correlation is somewhat weak. The slope and
ous dikes, and some form selvages on sills. Some pebble r2 values were calculated using Excel’s LINEST function,
dikes do crop out to the southwest around the northeast- bounded to zero. B – The average largest clast size in
ern margins of the Silver City stock, with no apparent pebble dikes is 2.4 ±1.6 cm 1σ, with a significant skew
spatial relationship to the porphyry of North Lily, but their towards smaller clasts sizes. C – The average pebble dike
abundance is dramatically less than porphyry of North width is 22 ± 18 cm 1σ, once again with a skew towards
smaller widths.
Lily-associated pebble dikes (figure 3 and 9). These
southernmost pebble dikes may represent an earlier gen-

Comer, J.B., Inkenbrandt, P.C., Krahulec, K.A., and Pinnell, M.L.


26 The Nature and Origin of Pebble Dikes and Associated Alteration: Tintic Mining District (Ag-Pb-Zn-Au), Utah

the clasts in pebble dikes (figure 11). Rounded clasts


A of porphyritic andesite removed from pebble dikes show
sufficient chemical and petrographic similarity to the
porphyry of North Lily to conclude they are North Lily
fragments (table 1, EUREKA-DJ-11). Additionally, angu-
lar xenoliths of Tintic Quartzite, identical to the quartzite
in pebble dikes, are found entrained in andesite dikes of
the porphyry of North Lily (figure 11).

Alteration and Pebble Dikes


Although the Tintic mining district is host to multiple
alteration types, specific emphasis is placed herein on
describing the low-grade alteration haloes in rhyolite host
rocks (the Packard Quartz Latite) that are directly related
to pebble dikes. Alteration varies from argillic to silicic,
with argillic being the dominant type.

B Within argillic alteration envelopes, variation in the


intensity of alteration is common. Rhyolite outcrops that
host individual pebble dikes have weak argillic alteration
in the few centimeters nearest to the margin of the dike.
Original volcanic textures are almost entirely preserved;
however, most plagioclase phenocrysts are soft and milky
white. Plagioclase is partially pseudomorphed by fibrous,
blocky masses of smectitic clay. Biotite is partially con-
verted to clay as well (figure 12).

Rhyolite outcrops that host swarms of pebble dikes


display stronger argillic alteration, resulting in the
destruction and replacement of all phenocrysts except
quartz and K-feldspar. Plagioclase and mafic silicates are
completely replaced by smectitic clay (figure 12). XRD
Figure 8. Photomicrographs (20x magnification, cross-
analysis reveals that strong argillic alteration is accom-
polarized light) of a pebble dike (EUREKA-DJ-PD18).
A – This thin section shows two rounded quartzite clasts panied by kaolinite as an additional alteration phase.
on the left and an unidentified highly-birefringent clast Although K-feldspar persists into the strong argillic stage,
on the right (shale?). Clasts are separated by an aggregate it displays minor evidence of alteration to clay. Strong
of angular quartz grains and other clastic debris. B – This argillic alteration also results in the partial destruction of
thin section has one quartzite clast in the lower left and original volcanic texture; nonetheless, the persistence of
a dark clast of porphyritic igneous rock (probably Packard quartz and K-feldspar phenocrysts still reveal a rhyolite
Quartz Latite) on the right. Again, clasts are separated by parentage. Rhyolite that displays strong argillic altera-
an altered matrix of clastic debris. tion crumbles very easily and is often white to yellow in
outcrop. Minor amounts of disseminated sulfides are also
eration of dike formation more related to the Silver City found in strong argillic assemblages.
stock, but are not directly addressed here.
In rare cases, swarms of pebble dikes produce jasper-
In addition to a strong spatial association, the orienta- oid from rhyolite as a result of pervasive silicification.
Similar to strong argillic assemblages, rhyolite jasperoids
tions of pebble dikes and dikes of the porphyry of North display very little of their original volcanic texture. Of the
Lily are very similar. Like North Lily dikes, pebble dikes original phenocrysts only quartz remains; however, the
commonly strike northeast (figure 10). Both dike types persistence of magmatic quartz is only readily appar-
approximate the orientation of Cretaceous-age strike- ent in thin section (figure 12). Rhyolite jasperoids are
and oblique-slip faults hosted in Paleozoic rock (Morris, exceptionally hard, and are composed entirely of a mosa-
1964). A subordinate north-south trend is also seen in ic of small interlocking quartz grains, which permeate
the original groundmass and fill fractures and cavities.
both dike types. The dikes displaying more northerly Pebble dikes with silicified matrices probably developed
trends may be part of a radial set of dikes around the from the same fluids that produced the silicification of
main North Lily stock that is obscured by the dominant rhyolite into jasperoid.
NE-trend. Isocon analysis (Grant, 2005) of fresh and altered
Packard Quartz Latite is consistent with the alteration
As an additional indication of the relationship between mineralogy and textures (figure 13). The isocons are
pebble dikes and the porphyry of North Lily, fragments defined by a wide variety of elements that were essen-
of the porphyry of North Lily make up a small portion of tially immobile, including Al, Y, Zr, Nb, Th, and LREE.

UGA Publication 45 (2016)—Resources and Geology of Utah’s West Desert


Johnson, D.M. and Christiansen, E.H. 27

»North Standard Pebble


dike

Selma
an
n Pe ak

xm
Pinyo

Pa
»Water Lillie
39°59'N

nsville Tnl
Homa
Central Standard Porphyry of
Homansville » North Lily
» Copper Leaf
Canyon e
or

tic
lti
m » Tsc

Tin
Ba tic
39°58'N

in
East T

st
Silver City

Ea
ote
ial
Coy
nn

Stock
nte
Ce

Tint
ic
le

Ba
il » North Lily Tsp
Leadv

Stan

llpa
di ate Big Hill
rme Sunrise Peak

rk
Inte

ard
»
Eureka Lilly
h
Addie

ut intrusion and flows


So
39°57'N

South
»Burgin
20t
l

Faults:
tra

hC

ing
en

nK le
ent

idd
dC

Iro
ay

yon M
ury
an

Can
Yan Normal

z
yD

Ine
Gr

kee rd
Pine a
Ma

nd Strike-slip
rd ta
oth

n da e xS
Sta Ap Thrust
mm
39°56'N

re ka ex
Ma

Eu Ap
th
Sioux-Ajax S ou
» Mineshaft
»
Trixie

Nevada
»
39°55'N

d
on
am
Di
39°54'N

112°7'W 112°6'W 112°5'W 112°4'W 112°3'W 112°2'W

0 0.75 1.5 3 Km

1:60,000
Figure 9. Fault and dike map of the Tintic mining district modified from Lindgren and Louglin (1919), Morris (1964),
Morris and Lovering (1979), and Keith and others (2009). This map shows the fault network for the region, complete
with abundant northeast-trending strike- and oblique-slip faults (blue), north-south to east-west-trending normal faults
(orange), and eastward-verging thrust faults (purple). This map also shows the prevalence of pebble dikes (black lines,
size exaggerated) near the porphyry of North Lily and the relative lack of pebble dikes near the Silver City stock.

Weak argillic alteration, marked by the partial destruc- shows systematic increases in CaO and Na2O concentra-
tion of plagioclase feldspar, results in the loss of CaO and tions away from the dike (figure 14), showing that pebble
Na2O from the rock. In fact, when sampled at intervals dikes were flow paths for fluids that caused alteration. Up
from the margin of a pebble dike outward, the rhyolite to 60% of the original CaO and 40% of the original Na2O

Comer, J.B., Inkenbrandt, P.C., Krahulec, K.A., and Pinnell, M.L.


28 The Nature and Origin of Pebble Dikes and Associated Alteration: Tintic Mining District (Ag-Pb-Zn-Au), Utah

Pebble Dikes Porphyry of North Lily Dikes


n = 400 n = 83
Mean = 37.0° ± 3.2° 1σ Mean = 29.1° ± 6.2° 1σ

Faults
n = 26
Mean = 45.4° ± 6.5° 1σ
Figure 10. Rose diagrams showing orientations of pebble
dikes, porphyry of North Lily dikes, and strike-slip and normal
faults in the area shown in figure 9. These rose diagrams
were created in RockWorks using spatial data collected from
ArcMap. Each feature was approximated by a two-point
straight line, and endpoints of the line were used to calculate
the strike and length of the feature. Features were then
separated into 10° bins based on strike. The length of each
10° bin represents the cumulative length of all features in the
bin. The red lines show the mean and one standard deviation
for each diagram. On the faults diagram, only the mean and
standard deviation of strike-slip faults are shown. These rose
diagrams show that pebble dikes, igneous dikes, and strike-
slip faults share a common northeast orientation. Pebble dikes
and igneous dikes were perhaps guided by and emplaced
along pre-existing northeast-trending strike-slip faults.

Strike-slip faults Normal faults

A B

Figure 11. Clasts in dikes near the North Lily mine. A – These rounded clasts (EUREKA-DJ-11) were removed from a
pebble dike hosted in Packard Quartz Latite. Their texture, mineralogy, and chemistry indicate that they represent the
porphyry of North Lily. B – This sample of porphyry of North Lily contains a xenolith of Tintic Quartzite (marked by the
black arrow), the same material that makes up the bulk of pebble dike clasts.

UGA Publication 45 (2016)—Resources and Geology of Utah’s West Desert


Johnson, D.M. and Christiansen, E.H. 29

Weak argillization
A B

Strong argillization
C D

Silicification (jasperoid)

E F

Figure 12. Photomicrographs (20x magnification, cross-polarized light) of pebble dike-related alteration in Packard
Quartz Latite host rocks. A – Weak argillization (EUREKA-RAF-13) involves the partial decomposition of plagioclase
phenocrysts. This is seen in the brown, fibrous stripes in the large plagioclase phenocryst. B – Mafic silicates, like this
biotite, are also partially decomposed. C – Strong argillization (EUREKA-KAG-13A) is characterized by the complete
decomposition of plagioclase and mafic silicates. Plagioclase shapes remain, but are entirely replaced by clay. D –
K-feldspar phenocrysts persist into strong argillic alteration assemblages. E – Silicification (EUREKA-DJ-12) converts
all phenocrysts and groundmass to quartz. Secondary quartz often approximates the shape of the phenocryst it replaced
(here probably a plagioclase grain). F – Magmatic quartz phenocrysts persist in silicified assemblages (black line is a
fracture in the thin section).

Comer, J.B., Inkenbrandt, P.C., Krahulec, K.A., and Pinnell, M.L.


30 The Nature and Origin of Pebble Dikes and Associated Alteration: Tintic Mining District (Ag-Pb-Zn-Au), Utah

70
Weak argillization TiO2
SiO2
Fe2O3
60 % 1:1
50

Altered Packard Quartz Latite


+
CO
N Al2O3
50 ISO
Added P2O5
Cr MgO
40 Cu K2O

Ce Ba
Nb Na2O
30 Ni Ga -50%
Sc MnO
Sm Nd La
20 CaO
Th Rb Pb
U
10 Y V Sr Subtracted
Zn
Zr
0
0 10 20 30 40 50 60 70
Unaltered Packard Quartz Latite

70
TiO2
Strong argillization
Al2O3
60 0% SiO2
Altered Packard Quartz Latite

+5 N 1:1
CO
ISO
50
Added Cr
K2O
40
Ce
Ni Nb P2O5
30 Ga -50%
Pb La
Rb Ba Fe2O3
Sm Nd Cu
20 U MgO
Na2O
Th
V Sc
10 Sr Subtracted
Zn Y CaO
Zr MnO
0
0 10 20 30 40 50 60 70
Unaltered Packard Quartz Latite

90
SiO2
Silicification (jasperoid)
80
Altered Packard Quartz Latite

0%
70 +5 1:1
Added TiO2
60
N
CO
ISO
50 Cr

40 Nb
Ni
30 -50%
P2O5
20 Ce Ba
Nd La Subtracted
U Sm Na2O Fe2O3
10 Zn Y Sr Pb Ga MgO
Zr V K2O CaO Al2O3
Th Sc Rb Cu MnO
0
0 10 20 30 40 50 60 70
Unaltered Packard Quartz Latite

Figure 13. Isocon diagrams (Grant, 2005) of different types of alteration related to pebble dikes in Packard Quartz Latite
host. Immobile elements and oxides form the isocon (blue line). Elements and oxides that plot below the isocon have
been removed during alteration, while those that plot above have been added. Weak argillization removed Na2O and CaO.
Strong argillization removed Na2O, CaO, MgO, and Fe2O3. Silicification resulted in the loss of most elements and oxides
with the significant addition of SiO2. 50% gained and lost lines are included to reference approximately how much of
each component was gained or lost during alteration. The lack of enrichment of Cu, Pb, or Zn is evidence that these
dikes are not directly related to mineralization. All element and oxide concentrations have been arbitrarily scaled using
different but consistent scaling factors in order to fit on the same diagram.
UGA Publication 45 (2016)—Resources and Geology of Utah’s West Desert
Johnson, D.M. and Christiansen, E.H. 31

are removed from the rhyolite due to weak argillic altera- Pebble Dike Packard Quartz Latite
tion. Sr, Rb, Pb, Ba, K, Mg, and Mn were also leached -130
from the rhyolite to varying degrees. Strong argillic alter-
ation, characterized by the destruction of all phenocrysts
-140 δDVSMOW (‰)
except for quartz and K-feldspar, results in the more
extreme loss of CaO (-90%), Na2O (-70%), MgO (-65%),
and Fe2O3 (-60%) as plagioclase and biotite break down -150
further. Minor losses of Ba and Sr are also observed as a
result of feldspar destruction. K2O concentration remains 8
more or less constant in strong argillic assemblages due
to the persistence of K-feldspar. Mn, P, Sc, and Cu also 7 δ18OVSMOW (‰)
show notable depletions. No elements were significantly
enriched during this low temperature alteration. Notably,
elements enriched in Tintic ores, e.g., Pb, Zn, and Cu, 6
are not enriched and instead are quite similar between
fresh and altered rhyolite. In rhyolite jasperoids, all major 3
elements (except for Si and Ti) and most trace elements
(except Zr, Nb, Ni, and Cr) were removed from the rock, 2 Na2O (wt %)
not just diluted by the addition of SiO2. This is indicated
by the position of most elements below the isocon. Pb,
Zn, and Cu are strongly depleted in the jasperoids. On 1
the other hand, SiO2 was introduced by the hydrothermal 3
fluids; it is not enriched simply by leaching. Concentra-
tions increase substantially in jasperoids, upwards of
35%, due to the addition of quartz (figure 12). 2 CaO (wt %)

Stable Isotopes and Pebble Dikes


1
Given that pebble dike emplacement resulted in signifi- 0 50 100
Distance from pebble dike
cant changes in rhyolite host rock chemistry, changes in
into host rhyolite (cm)
stable isotope composition may reveal more about the
causes and conditions of alteration. Fresh and altered Figure 14. Chemical and stable isotope trends near
host rocks (collected at or near the margins of pebble pebble dikes. In the Packard Quartz Latite CaO, Na2O,
dikes), pebble dikes, quartz phenocrysts, hydrothermal and δ18O all increase away from the pebble dike. Lower
quartz crystals, and dolomite clasts from pebble dikes values in each case represent more altered rock, with the
were analyzed for oxygen, hydrogen, and carbon iso- most altered samples occurring right on the margin of the
tope compositions. All stable isotope data, referenced to pebble dike. This observation identifies pebble dikes as a
VSMOW and VPDB, are summarized in table 3. possible alteration source.

Quartz phenocrysts separated from fresh (albeit still a δ18O value of 9.1‰ and probably formed at low tem-
displaying minor alteration) and altered Packard Quartz perature from isotopically exchanged fluids (see below).
Latite range from 8.8 to 8.9‰ δ18O. There is no dis- Pebble dike whole-rocks range from 8.6 to 10.9‰ δ18O.
cernible difference between quartz in fresh and altered The higher δ18O values seen in pebble dikes (which are
rhyolite samples. Because of the resistance of quartz to composed of over 90% quartzite) and quartz relative to
alteration and weathering, the δ18O of analyzed quartz rhyolite whole-rock are explained by the natural enrich-
phenocrysts approximates the original δ18O value of ment of18O in quartz compared to other minerals (Sharp,
quartz in the Packard Quartz Latite at ~8.9‰. This 2007; Taylor, 1974). Pebble dike values cover a larg-
value does not accurately estimate the δ18O value of the er range of δ18 O than quartz separates due to varying
original magma, as quartz is always richer in 18O than a amounts of carbonate contamination.
coexisting silicate melt (Taylor, 1968). The δ18O value of
the original magma should have been about 0.8‰ lower
than the quartz phenocrysts, at ~8.1‰ (Zhao and Zheng, Overall, rhyolite whole-rocks have δD values ranging from
2003). This δ18O value is similar to that of many rhyo- -135 to -147‰. δD for altered rhyolites overlaps with,
lites, which typically range from 6‰ to 10‰. but extends to somewhat lower values than for fresh
rhyolite. δD in the freshest rhyolite ranges from -138 to
δ18O values in rhyolite whole-rocks that fall below -139‰ with more altered rhyolite from -135 to -147‰
this magmatic value could be the result of isotopic (figure 15). These δD values are much lower than typical
reequilibration with pebble dike-related fluids. Rhyolite fresh igneous rocks which range in δD from -50 to -85‰.
whole-rocks have δ18O values ranging from 5.8 to 8.1‰. Since these rhyolites fall significantly outside this range,
δ18O in fresher rhyolite ranges from 7.4 to 8.1‰ and is
fairly consistent with the value estimated from quartz. they must have experienced isotopic reequilibration—
More altered rhyolite occupies a lower range of δ18O from probably with meteoric fluids at low water/rock ratios
5.7 to 7.7‰ (figure 15). Hydrothermal quartz formed which causes strong decreases in δD with little change in
in crystal-lined cavities in a silicified pebble dike had δ18O (figure 15).

Comer, J.B., Inkenbrandt, P.C., Krahulec, K.A., and Pinnell, M.L.


32 The Nature and Origin of Pebble Dikes and Associated Alteration: Tintic Mining District (Ag-Pb-Zn-Au), Utah

Table 3. Stable isotope compositions via mass spectrometry.

δ18O δ13C δ18O Wt. Calculated Calculated


δ18O Whole- δD Whole-
Sample Rock Type Dolomite Dolomite Quartz % Water δ18O Water δD Lab
Rock (‰) Rock (‰)
(‰) (‰) (‰) H2O (‰) (‰)

PD-Quartz Quartz 9.08 -8.7 Oregon

EUREKA-WJJ-13 Packard 8.91 8.09 -137.7 1.22 -8.9*/-9.2 -114.0 Oregon

EUREKA-RAF-13 Packard 8.84 6.87 -144.5 1.90 -9.0*/-8.6 -121.0 Oregon

EUREKA-KAG-13 Packard 8.84 7.65 -143.8 3.03 -9.0*/-8.8 -121.7 Oregon

EUREKA-DJ-PD2a Packard 5.95 -139.8 1.73 -9.5 -116.1 Oregon

EUREKA-DJ-PD2b Packard 5.77 -135.0 1.86 -9.6 -111.2 Oregon

EUREKA-DJ-PD2c Packard 5.66 -135.5 1.52 -9.8 -111.7 Oregon

EUREKA-DJ-PD12a Packard 6.37 -142.7 2.30 -9.1 -119.1 Oregon

EUREKA-DJ-PD12b Packard 6.83 -147.1 1.42 -8.6 -123.6 Oregon

EUREKA-DJ-PD12c Packard 7.40 -138.8 1.13 -9.9 -115.4 Oregon

Pebble
EUREKA-DJ-PD5 8.64 -128.2 1.41 -9.2 -104.2 Oregon
dike
Pebble
EUREKA-DJ-PD12 10.20 -125.9 1.12 -7.6 -101.8 Oregon
dike
Pebble
EUREKA-DJ-PD30 10.90 -135.0 1.68 -7.0 -111.2 Oregon
dike
PD6a Dolomite 9.9 -2.5 -9.5** BYU
PD26 Dolomite 27.7 -2.5 BYU
PD26 rind Dolomite 16.3 -7.0 -3.3** BYU
PD28 in Dolomite 23.2 -1.1 BYU

PD28 in rind Dolomite 14.5 -5.1 -5.0** BYU

PD28 out Dolomite 24.2 -1.0 BYU


PD29 in Dolomite 24.3 0.0 BYU
PD29 out Dolomite 23.2 -0.2 BYU
* Calculated Water δ18O values calculated from δ18O Quartz
**δ18O only calculated for clasts that experienced isotopic reequilibration (evidenced by low δ18O values)

Rounded clasts of Cambrian dolomite removed from peb- rocks taken from the same outcrops that hosted carbon-
ble dikes range widely in δ13C from 0.0 to -7.0‰ and ate-bearing pebble dikes were isotopically similar to the
in δ18O from 9.9 to 27.7‰. Two groups are apparent: elevated δ18O group. The elevated δ18O group compares
an elevated δ18O group with δ18O values that range from well to the isotopic composition of unaltered Cambrian
23.2 to 27.7‰ and δ13C values that range from 0.0 to carbonate of the Basin and Range at 16 to 23‰ δ18O
-2.5‰, and a group with lower δ18O ranging from 9.9 to
and 0 to -1‰ δ13C (Brasier, 1993; Saltzman and others,
16.3‰ and δ13C ranging from -2.5 to -7.0‰ (figure 16).
1998), suggesting that this group has retained its origi-
While most clasts appeared to be unaltered gray marine
carbonate, some had tan to brown alteration rinds which nal Cambrian isotopic signature. The departure of the
varied in thickness from 1 to 2 mm. When analyzed lower δ18O and δ13C group away from the elevated group
independently, these alteration rinds fall into the group indicates that some dolomite clasts experienced isotopic
with lower δ18O and δ13C, but the cores of these same reequilibration with a hydrothermal fluid, even if only on
clasts are in the elevated δ18O group. Carbonate wall their exteriors.

UGA Publication 45 (2016)—Resources and Geology of Utah’s West Desert


Johnson, D.M. and Christiansen, E.H. 33

0 0
SMOW
A B
-20 -20

e
in

Lin
L
er

er
at

at
-40 -40

W
Initial rhyolite

ic

ric
r
composition

eo
Magmatic

eo
et
δDVSMOW (‰)

et
-60 -60

δDVSMOW (‰)
Water

M
ite
in
-80

ol
-80

Ka
ite
ne
in
ge

ol
po
-100

Ka
-100

ght
0.003

Hy
ene

wei
rg
pe
Initial

ite
-120 -120

Su

lin
y 0.01
Water-Rock meteoric

ao

b
s
atio

gK
Exchange H2O Wa t e r / R o c k r

in
-140 -140

er
0.1

th
0.9 0.6 0.4 0.3 0.2

ea
0.03

W
-160 -160
-20 -15 -10 -5 0 5 10 15 -20 -15 -10 -5 0 5 10 15
δ18OVSMOW (‰) δ18OVSMOW (‰)

Modern Tintic Freshest rhyolite Argillically altered rhyolite Pebble dike Quartz δ18O Dolomite clast
groundwater Calculated water Calculated water Calculated water Calculated δ18O in water Calculated δ18O in water

Figure 15. δD versus δ18O of pebble dikes and rhyolites altered by pebble dikes. A – Solid symbols represent measured
compositions, while hollow symbols represent calculated compositions of equilibrium fluids. Tie lines connect measured
values to the corresponding fluid. Calculations were performed at 100°C. Altered rhyolite from the walls of the pebble
dikes plots near the line intended to separate supergene kaolinite on the right from hypogene kaolinite on the left
(Sheppard and others, 1969). Kaolinite formed by weathering (Savin and Epstein, 1970) is more enriched in δ18O than
the kaolinite of the altered rhyolites. Calculated water compositions fall to the right of the meteoric water line (Craig,
1961), far from the field for magmatic water, and approach the composition of modern Tintic groundwater (Norman and
others, 1991; Hamaker and Harris, 2007). These attributes suggest the unlikely participation of magmatic water in the
generation of pebble dikes and associated alteration. The general lowering of both δ18O and δD in more altered rhyolite
samples when compared to fresher rhyolite samples demonstrates an important role for groundwater (produced through
meteoric recharge) in alteration. Although they have no corresponding δD values, the δ18O of water calculated to be in
equilibrium with hydrothermal quartz and dolomite clasts are shown on the x-axis and are similar to that calculated for
the other materials. Measured values for the hydrothermal quartz are shown, but the measured δ18O of dolomite clasts
falls off the graph to the right. B – Calculated water/rock ratios involved in alteration assuming the initial values shown
and fractionation factors of 8.5 for oxygen and 30 for hydrogen using the equations of Taylor (1977).
0
Dolomite clast Cambrian Carbonates
of the Basin and Range
Dolomite clast rind
-1
Dolomite wall rock

Hydrothermal calcite
-2

-3
δ13CVPDB (‰)

Hydrothermal Calcite
-4

-5

-6

-7

-8
0 5 10 15 20 25 30
18
δ OVSMOW (‰)
Figure 16. δ13C versus δ18O of dolomite clasts in pebble dikes and dolomite wall rock shows two isotopically distinct
groups. Those that likely did not experience isotopic reequilibration with pebble-dike fluids plot near to the field for
Cambrian dolomite of the Basin and Range (Brasier, 1993; Saltzman and others, 1998). One clast and the rinds of two
other clasts that experienced isotopic reequilibration plot near to the field for hydrothermal calcite from the Tintic district
(Norman and others, 1991). Hypothetical trends for equilibration are shown with an arrow.

Comer, J.B., Inkenbrandt, P.C., Krahulec, K.A., and Pinnell, M.L.


34 The Nature and Origin of Pebble Dikes and Associated Alteration: Tintic Mining District (Ag-Pb-Zn-Au), Utah

DISCUSSION Nevertheless, maximum horizontal stress in the western


United States maintained a northeast orientation until
Pebble dikes are hypothesized to form when superheated 30 Ma (Kowallis and others, 1995), thereby allowing the
water explosively flashes to steam, most likely due to a continued emplacement of dikes with a northeast strike,
sudden release of confining pressure (as would occur even in post-Sevier rocks. Thus, we conclude that old
when a fracture opens to the surface, releasing lithostatic fractures were re-opened where they aligned with the
load), and escapes to an area of lower pressure while Eocene stress regime, influencing new fractures with
carrying rock debris that was generated by the explo- similar orientations to form.
sions (Farmin, 1934; Kildale, 1938; Cook, 1957; Morris
and Mogensen, 1978; Morris and Lovering, 1979). Our The final evidence of porphyry of North Lily and peb-
observations further refine this hypothesis. ble dike interaction is seen in the presence of rounded
North Lily clasts in pebble dikes and quartzite xenoliths
The Role of the Porphyry of North Lily in in North Lily dikes (figure 11). The genetic relationship
Pebble Dike Formation suggested by the spatial correlation of pebble dikes to
the porphyry of North Lily is even more strongly suggest-
Historically, the formation of pebble dikes in the Tintic ed by this direct incorporation of each rock type in the
mining district has been attributed to the intrusion of other. Pebble dikes and the porphyry of North Lily would
the Silver City stock, whether named specifically or sim- likely have to be forming at the same time in order to
ply referred to as the ‘monzonite’ or ‘quartz monzonite’ interact in this way.
of the Tintic district. Early workers proposed that peb-
ble dikes formed as water interacted with the intruding Pebble Dike Characteristics
Silver City stock (Farmin, 1934; Kildale, 1938; Cook,
1957; Morris and Mogensen, 1978; Morris and Lover- The physical characteristics of the pebble dikes provide
ing, 1979). Although the heating of water by an intrusive further refinements of the formation model. Breccia
body appears to be a requirement in pebble dike forma- envelopes that surround some pebble dikes (figure 2D)
tion, the porphyry of North Lily, rather than the Silver City indicate that pebble dike emplacement was violent, as
stock, appears to have been responsible for those in the host rocks fractured to make space for pebble dike mate-
East Tintic subdistrict. rial. Pebble dikes only contain clasts that originated
below or adjacent to their present-day structural posi-
The identification of the porphyry of North Lily as the heat
tion, suggesting that the clasts were created by explosive
source was driven by the predominance of pebble dikes
in an area 5 km northeast of the main outcrops of the Sil- events that occurred at depth, and that pebble dikes
ver City stock, near a swarm of igneous plugs and dikes were formed by the general upward movement of mate-
now identified as the porphyry of North Lily due to their rial. The rounding of the clasts likely occurred by thermal
unique texture, hydrous mineralogy, and andesitic com- spallation and abrasion as they were transported to their
position. The strong spatial association between pebble current positions. The abundance of quartzite clasts rela-
dikes and the porphyry of North Lily is easily seen in map tive to any other clast type could be explained by the
view, where pebble dikes mostly overlap with the extent resistance of quartzite to the violence of emplacement,
of the porphyry of North Lily (figure 3, 5, and 9).This but also likely indicates that the main explosive events
spatial association strongly suggests that the porphyry of occurred within the Tintic Quartzite, which is estimated
North Lily was fundamental in pebble dike formation; to have been at a depth of 150 to 500 m (Morris, 1964).
however, the Silver City stock may have its own aureole
Fluid involvement in the formation of pebble dikes is
of minor pebble and igneous dikes.
strongly suggested by the nearly universal presence of
Additional lines of evidence indicate that the porphyry of hydrothermal alteration within pebble dikes and along
North Lily was directly involved in pebble dike formation. their margins in rhyolite host rocks (Kildale, 1938) and
The first evidence is the similarity in orientation between by our new chemical and isotopic data for the altered
North Lily dikes and pebble dikes (figure 10). Orientation rocks.
similarity, possibly brought about by the influence of a
series of subparallel, northeast-trending, Sevier-related These observations suggest that the pebble dikes devel-
strike- and oblique-slip faults hosted in Paleozoic strata oped as subvertical NE-trending fractures opened (some
(figure 9), suggests that both dike types were emplaced pre-existing and some newly developed), followed by
when regional stresses were similar. This might indi- explosive fragmentation of the walls of the fracture. Once
cate that both dike types were actively forming at the
the carapace breccia formed, a fluidized stream of peb-
same time. Although dike emplacement likely utilized
bles and clastic debris filled the fracture. Escape of the
northeast-trending zones of weakness in Paleozoic stra-
ta, similar northeast orientations continue in dikes that boiling fluid from the slurry caused further disruption of
crosscut Cenozoic volcanic and plutonic rocks. Cenozoic the wall rocks. Finally, sheets of andesite magma fol-
volcanic rocks were emplaced long after Sevier deforma- lowed up the same conduits. The heated water moving in
tion ended, and should therefore not have inherited the and from the dikes caused minor alteration in surround-
structural weaknesses of the deformed Paleozoic strata. ing host rocks.

UGA Publication 45 (2016)—Resources and Geology of Utah’s West Desert


Johnson, D.M. and Christiansen, E.H. 35

The Character of the Fluid Involved in The small change of 1 to 2‰ in δ18O from fresher to more
Pebble Dike Formation altered rhyolite samples indicates that fluid interaction
occurred at low water-rock ratios. Using the equations of
Alteration characteristics and stable isotope composi- Taylor (1977) for open system exchange of oxygen and
tions can be used to determine the character of the fluid hydrogen isotopes, the calculated water-rock ratios are
involved in the formation of the pebble dikes. Alteration about ~0.03 by weight. Of particular note, a sequence of
effects in rhyolite host rocks, including the removal of altered rhyolite samples taken at ~50 cm intervals from
CaO, Na2O, MgO, MnO, Fe2O3, Ba, Sr, and Cu, and the the margin of a pebble dike outward showed a consist-
production of smectite, kaolinite, and jasperoid indicate ent increase in δ18O away from the dike (figure 14). This
the participation of a hydrothermal fluid in pebble dike observation, of more altered, lower δ18O rhyolite occur-
formation. Assuming that this alteration mineral assem- ring closer to pebble dike margins, strongly suggests that
blage reached equilibrium with the fluid that caused pebble dikes contained and were flow paths for fluids
alteration, the fluid temperature for a system dominated capable of argillic alteration, silicification, and lower
by smectite, kaolinite, and jasperoid is estimated to have δ18O.
been at or below 150°C (Reed and Spycher, 1984). As a
further constraint on fluid temperature, pebble dike-relat- δD also appears to be lowered by alteration. All measured
ed alteration in the Packard Quartz Latite is very similar δD values are exceptionally low when compared to hydro-
to hydrothermal alteration observed in drill cores of the gen isotope compositions of fresh igneous rocks around
Biscuit Basin rhyolite flow of Yellowstone National Park the world. Altered rocks with δD values below the range
(Sturchio and others, 1986). At Yellowstone, alteration for typical igneous rocks (-50 to -85‰ δD) have usually
results in the loss of the same suite of elements (CaO, experienced post-crystallization exchange with heated
Na2O, MgO, MnO, K2O, Ba, and Sr) from the rhyolite, meteoric water, during which secondary OH-bearing min-
with the accompanying alteration of rhyolite to smectitic erals are produced. The presence of OH-bearing minerals
clay, zeolites, celadonite, and secondary quartz. Altera- results in the lowering of the overall δD value (Sheppard
tion temperatures at Yellowstone are estimated directly and Taylor, 1974; Taylor, 1974). The exceptionally low
by water temperatures measured during drilling, which δD values of these rhyolite and pebble dike samples can
ranged from 130° to 170°C. This temperature range be partially attributed to the presence of smectite and
overlaps with the ≤150°C estimated by the pebble dike- kaolinite, as shown by XRD. Even the δD value of the
related alteration mineral assemblage. Fluid temperature freshest rhyolite analyzed (EUREKA-WJJ-13), at -135‰,
can be further constrained by the requirement for boiling falls significantly outside the range of unaltered igneous
in pebble dike formation, which occurs in water under rock, and thereby suggests that it has interacted signifi-
atmospheric pressure at 100°C. The boiling point of cantly with meteoric water. Alternatively, δD depletion
water increases to 190°C at 150 m of hydrostatic pres- could be attributed to eruptive degassing (figure 15),
sure, the estimated depth of the Tintic Quartzite below which lowers δD over time during an eruptive sequence if
this part of the Eocene-Oligocene volcanic field. This a discrete vapor phase escapes (Taylor and others, 1983).
temperature exceeds that estimated by alteration min- We consider this less likely because the lower δ18O values
eralogy, which indicates that boiling did not occur under found in the lower δD rhyolites are not predicted by the
a full hydrostatic load, or that the altered rocks we sam- degassing hypothesis.
pled represent more distal, and thus lower temperature,
alteration. The presence of onionskin structure in quartz- Some deuterium depletion might also be explained by
ite clasts, created at a minimum temperature differential the location of the Tintic mining district. The δD value
of 100°C, may also constrain fluid temperature. of modern groundwater of meteoric origin in the region
is significantly depleted in deuterium (average of -112‰
Stable isotope compositions can be used to provide fur- per Norman and others (1991) and Hamaker and Har-
ther insight into the character of the fluid. The effect ris (2007)). Prolonged interaction of igneous rocks with
of pebble dike-related alteration on δ18O is most eas- depleted groundwater of this sort would certainly lower δD
ily seen in the comparison of the freshest to the most values below their typical range. The δD values of middle
altered rhyolite samples, as it appears that quartzite in Cenozoic-age igneous rocks (and associated calculated
the pebble dikes and quartz phenocrysts in the rhyolite waters) in the western United States have been shown
experienced little to no reequilibration. The freshest rhy- to be geographically correlated with modern groundwater
olite analyzed (EUREKA-WJJ-13) has a higher δ18O value composition (Sheppard and others, 1969). Although the
than any other rhyolite at 8.1‰. The δ18O values in more Packard Quartz Latite was emplaced during the Eocene,
altered samples were consistently lower (5.7 to 7.7‰, groundwater during the time of emplacement was appar-
figure 15), which is in agreement with observations from ently isotopically similar to present-day groundwater. This
other areas, like Yellowstone, where near surface volcanic idea is supported by the analysis of Tintic district fluid
rocks experience hydrothermal alteration (Sturchio and inclusions by Norman and others (1991), which provides
others, 1986; Sharp, 2007). This trend continues in an estimate of -105 to -120‰ δD for mineralization-
pebble dike dolomite clasts, where the lowest δ18O val- age fluids. Middle Cenozoic groundwater was probably
ues (10 to 16‰) are measured in samples displaying the this depleted because of the high elevation of the pro-
most alteration. In most cases, rocks that exhibit a trend posed Great Basin altiplano (Best and others, 2009) and
of decreasing δ18O with alteration achieved such by inter- the negative correlation between altitude and the δD of
action with meteoric water (Taylor, 1968; Taylor, 1974). meteoric water.

Comer, J.B., Inkenbrandt, P.C., Krahulec, K.A., and Pinnell, M.L.


36 The Nature and Origin of Pebble Dikes and Associated Alteration: Tintic Mining District (Ag-Pb-Zn-Au), Utah

The overall isotopic trend of lower δ18O and δD values heated by the intruding porphyry of North
in Tintic district rocks altered by pebble dikes suggests Lily until it approached the boiling point
the participation of meteoric water in pebble dike forma- at hydrostatic pressure (190°C at 150 m).
tion (Taylor, 1968; Taylor, 1974), or more specifically, of During heating, the water experienced isotopic
groundwater produced though meteoric water recharge. exchange with surrounding rocks, which pulled
Given a fluid temperature estimate for this groundwater the isotopic composition of the water off of the
of 100° to 170°C, the oxygen and hydrogen isotope com- meteoric water line.
positions of water in equilibrium with rhyolite and pebble
dikes were calculated from analytical results at 100°C 2. New and pre-existing northeast-trending
(figure 15). Calculations and fractionation factors used fractures and faults that were aligned
are summarized in Johnson (2014). Calculated water for favorably with the Eocene stress field opened
the fresher rhyolite samples ranges from -9.2 to -9.9‰ to accommodate the rising andesitic magma.
δ18O and -114 to -115‰ δD while more altered rhyo- When any one of these fractures opened to the
lite samples range from -8.6 to -9.8‰ δ18O and -111 surface (or to any area of overall lower pressure),
to -124‰ δD. Calculated water for pebble dike samples confining pressure on the superheated water
ranges from –7.0 to -9.2‰ δ18O and -102 to -111‰ δD. dropped, allowing it to flash into steam. Multiple
On the whole, calculated waters plot near but to the right explosions likely occurred in an area of about
of present-day regional groundwater, which ranges from 6 km2, producing several generations of pebble
-13 to -16‰ δ18O and -96 to -121‰ δD (Norman and dikes.
others, 1991; Hamaker and Harris, 2007), indicating
that the fluid that participated in pebble dike formation 3. The large volume expansion related to this phase
likely originated as meteoric water. The rightward shift change created more fractures and brecciated
away from the meteoric water line is probably the result the surrounding rock (Morgan and others,
of the meteoric water experiencing isotopic exchange 2009). Brecciated rock material was fluidized
with rocks at elevated temperatures and low water-rock and carried upward along fractures and faults
ratios (Taylor, 1974) prior to participation in pebble in the escaping stream of decompressing and
dike formation. Such shifts are common in geothermal expanding fluid. Quartzite clasts were rounded
fluids and hot springs. For example, the water at Steam- into pebbles during transport, probably in large
boat Springs, Colorado, has δD and δ18O values (Sharp, part by thermal spalling. The fluidized mixture
2007) very similar to those calculated for the waters in traversed fractures in Paleozoic sedimentary
the pebble dikes and is shifted to the right compared rocks and continued along newly formed
to local precipitation. Calculated water values plot far fractures in the overlying Packard Quartz Latite.
from the magmatic water field, indicating the unlikely
4. As the fluid left the system, transported material
participation of magmatic water in the formation of the
became cemented in place, filling fractures
pebble dikes in the East Tintic subdistrict. Increasing the
temperature of calculation to 150 or 170°C moves calcu- far above its point of origin. Wall rocks directly
lated values further to the right of the meteoric water line adjacent to pebble dikes were altered by the low
in the direction of the actual measured values. Although temperature fluids emanating from the pebble
isotopic evidence indicates the participation of warm dikes to clay-rich argillic assemblages and minor
groundwater in pebble dike formation, the limit of host jasperoid. Soluble elements were leached out of
rock alteration to the nearest few centimeters from the the rhyolite and silica was locally introduced.
pebble dike/host rock contact and the presence of only 5. If the explosive events that created pebble dikes
partially reequilibrated dolomite clasts in the dikes sug- reached the surface, fluidized breccia may have
gests that fluid presence in pebble dikes was relatively erupted to create hydrothermal explosion craters
short-lived. or trench-like features bordered by rings or
mounds of accumulated loose breccia (Morgan
Model for the Formation Pebble Dikes and others, 2009).
Any model created to describe the formation of peb- 6. Magma followed up through the breccia in some
ble dikes in the East Tintic subdistrict should take into places creating pebble carapaces on dikes and
account the new observations and interpretations pre- sills. Eruption of lava followed and formed the
sented in this study. Some of the more important include: Laguna Springs Volcanic Group which partially
(1) the participation of the porphyry of North Lily as a buried the older Packard Quartz Latite.
heat source, (2) the involvement of heated groundwater
in the mobilization of pebble dike material and hydro- In summary, pebble dikes were formed by the mobiliza-
thermal alteration, and (3) the influence of the regional tion of breccia in the explosive escape of groundwater
fault system in pebble dike orientation. With these con- that had been heated by the porphyry of North Lily. Using
straints, pebble dike formation probably followed these the correlation of the porphyry of North Lily to the Lagu-
steps (figure 17): na Springs Volcanic Group, pebble dike formation can be
1. Following ~2.5 m.y. of preceding rhyolite, estimated to have occurred at 33.29 ± 0.09 Ma, about
monzonite, and latite magmatism, groundwater 500,000 y after the intrusion of the monzonitic Silver
hosted in fractures in the Tintic Quartzite was City stock.

UGA Publication 45 (2016)—Resources and Geology of Utah’s West Desert


Johnson, D.M. and Christiansen, E.H. 37

A 0 500 m CONCLUSION
In order to better understand the origin of the pebble dikes
W in the East Tintic subdistrict, this study focused on these pri-
mary questions: (1) What is the nature of pebble dikes in the
Packard Quartz Latite East Tintic subdistrict? (2) Are these pebble dikes related to
E a specific intrusive body? (3) How are host rocks altered near
Undifferentiated Paleozoic
pebble dikes? (4) What is the character and origin of the fluid
sedimentary rocks involved in the emplacement of pebble dikes? and (5) Is there
any connection between pebble dike formation and ore-gener-
STAN
DAR
D ating processes in the Tintic mining district?
TINTIC HRUST
T
ation In response to question (1), pebble dikes are tabular bodies
Form
Ophir Groundwater of intrusive breccia (Wright and Bowes, 1963) dominated by
Tintic Quartzite rounded fragments of quartzite, carbonate, shale, and igne-
ous rock which are enclosed in a fine-grained clastic matrix.
Also, pebble dikes are mostly hosted in the Packard Quartz
Porphyry of North Lily
Latite, are bounded by sharp contacts with host rocks, and are
typically bordered by significant brecciation in host rocks. To
answer question (2), pebble dikes share many relationships
with the porphyry of North Lily, a new intrusive unit identified
B and named in this study. These relationships include a strong
spatial association, a similar northeastern strike in both peb-
ble dikes and North Lily dikes, and the inclusion of North
Lily clasts in pebble dikes and the inclusion of pebble dike
material in North Lily dikes. With these observations, we con-
clude that pebble dikes are related to the porphyry of North
Lily. In regard to question (3), argillization and silicification
of rhyolite host rocks accompanied pebble dike formation. In
most cases, the emplacement of a pebble dike resulted in the
loss of CaO, Na2O, MgO, and Fe2O3 from the rhyolite host,
with the simultaneous creation of smectite and kaolinite clay
and minor jasperoid. Both δ18O and δD are lower in altered
host rocks than they are in equivalent fresh host rocks. All
alteration associated with pebble dikes is very low grade. In
Brecciation caused by
fluid expansion response to question (4), stable isotope analysis of altered
rhyolite host rocks and pebble dike clasts determined that
pebble dikes formed in the presence of heated groundwater
that had experienced isotopic exchange with host rocks. This

Figure 17. Model of pebble dike formation. All labels defined


C in frame A apply to subsequent frames. Pebble dike size and
the thickness of the Tintic Quartzite are greatly exaggerated for
Possible hydrothermal illustrative purposes. A – Fracture-hosted groundwater is heated
explosion crater by the rising porphyry of North Lily. Pre-existing fractures and
Pebble dike
faults (oblique-slip denoted by the combination of strike-slip
and normal fault symbology) begin to reopen to accommodate
the rise of magma. Water temperature rises to near the boiling
Alteration surrounds point. B – When favorably oriented pre-existing fractures open to
pebble dike the surface, confining pressure on the hot water drops, allowing
it to flash into steam as it explosively expands. Surrounding rock
is brecciated during this phase change. C – Brecciated rock
is carried up the now open fractures and rounded as the hot
fluid expands and escapes to the surface. The hot fluid causes
alteration in the surrounding rock. Additional brecciation occurs
in the host rocks along pebble dike walls. Breccia material may
have escaped to the surface, forming a hydrothermal explosion
crater similar to those described in Morgan and others (2009).
Any evidence of such has since eroded away. Although shown
separately in this model, the events shown in frames B and C
likely occur simultaneously.

Comer, J.B., Inkenbrandt, P.C., Krahulec, K.A., and Pinnell, M.L.


38 The Nature and Origin of Pebble Dikes and Associated Alteration: Tintic Mining District (Ag-Pb-Zn-Au), Utah

water likely fell as rain or snow at high altitude over the in the summer of 2011 (Andover Mining Corp., 2011;
Tintic region during the Eocene to Oligocene and was Boden and others, 2013). Drilling ventures were guided
subsequently incorporated into the groundwater system. by pebble dikes, which were used as “vectors to potential
Heating of this water during the intrusion of the porphyry porphyry mineralization” (Andover Mining Corp., 2011,
of North Lily, to a temperature range of 100° to 170°C, p. 9). The utilization of pebble dikes in exploration tech-
resulted in the hydrothermal fluid that created pebble niques indicates the recognition of their potential role
dikes. as guides to Tintic mineralization. As discussed by Sil-
litoe (1985), pebble dike features are commonly found
Finally, to answer question (5), pebble dikes may have in association with porphyry Cu-Mo deposits. With this
been important to the ore-generating processes in the association in mind, it is possible to hypothesize that
Tintic mining district. In the midst of mostly imperme- the pebble dikes of the Tintic mining district occupy the
able rocks, pebble dikes could have served as conduits Pb and Zn-rich fringes of a yet undiscovered porphyry
for the transmission of ore fluids from the stocks to Cu system. Unfortunately, the full results of the explora-
more reactive rocks, provided that pebble dike forma- tion venture have been delayed by the 2014 bankruptcy
tion occurred prior to mineralization. Pebble dikes were filing of the Andover Mining Corporation. Given further
formed by and were pathways for hydrothermal fluids. research, pebble dikes have the potential to be identified
This is seen in the systematic decrease of the concen- as worthwhile indicators of, and guides to, deep porphy-
trations of CaO, Na2O, and other mobile elements and ry-type mineralization, especially when they are found in
stable isotope ratios in host rocks near the pebble dikes association with the Pb-Zn replacement bodies typical to
(figure 14), a trend that would only be possible if hydro- the distal portions of many porphyry-type deposits.
thermal fluids were using pebble dikes as pathways. On
the other hand, these hydrothermal fluids were shown
to be phreatic, low temperature, and not magmatic, and ACKNOWLEDGMENTS
therefore likely incapable of transporting ore. In addi-
tion, the dikes we examined have low concentrations of We thank Stephen Nelson, Michael Dorais, David Tingey,
ore metals (figure 13). Pebble dikes in our study area are and Kevin Rey for their guidance and assistance in the
mostly barren (Farmin, 1934; Kildale, 1938; Morris and laboratory. We thank Ilya Bindeman for providing D and
Lovering, 1979). Regarding the relationship of pebble O isotope analyses. We thank the BYU geology field camp
dikes to ore, Lindgren and Laughlin (1919) placed the students of 2013 who collected some of the samples
pebble dikes in the pre-ore stage of the Tintic’s Ag-Pb- used in this work. Finally, we thank Ken Krahulec and
Zn-Au deposits, as did Farmin (1934) and Morris and Jeff Keith, whose careful reviews made this paper signifi-
Lovering (1979). Lovering (1949) found that some peb- cantly better.
ble dikes are mineralized by late barren stage solutions
and only locally contain ore. Morris and Lovering (1979,
p. 81) state that some pebble dikes contain shoots of ore, REFERENCES
especially at depths of 300 to 450 m, and even mention
mineralized pebble dikes in the Burgin, Eureka Standard, Allen, T., 2012, Mafic alkaline magmatism in the East
and Apex Standard No. 2 mines. Unfortunately, the dikes Tintic Mountains, west-central Utah—Implications
we studied are far above that level of mineralization and for a late Oligocene transition from subduction to
the low temperature fluids that permeated our dikes may
extension: Provo, M.S. thesis, Brigham Young Uni-
be the cooler, distal manifestations of ore transport pro-
cesses. Despite the absence of ore, the close association versity, 60 p.
of pebble dikes to replacement ore bodies and jasperoid Andover Mining Corporation, 2011, Andover Ventures
masses in the North Lily and Copper Leaf mines indi-
corporate update November 2011: http://www.sil-
cates that near-surface-level pebble dikes can at least be
utilized as guides to the Tintic mining district’s valuable verminers.com/_resources/news/Andover_Corp_Up-
Ag-Pb-Zn-Au resources in the North Lily region (Kildale, date_November_2011.pdf (accessed September
1938; Morris and Lovering, 1979). 2014).

The process of using pebble dikes as guides to ore in Best, M.G., Barr, D.L., Christiansen, E.H., Gromme, S.,
the North Lily region was explored by an Andover Mining Deino, A.L., and Tingey, D.G., 2009, The Great Ba-
Corporation joint venture with Rio Tinto’s Kennecott Utah sin Altiplano during the middle Cenozoic ignimbrite
Copper Corporation, which attempted to identify a buried flareup—insights from volcanic rocks: International
porphyry Cu target at Big Hill, only ~1 km southwest of Geology Review, v. 51, No. 7-8, p. 589-633.
the main pebble dike swarm (figure 3). This area was
chosen as a potential target because it has similar Pb-Zn Boden, T., Vanden Berg, M., Krahulec, K., and Tabet, D.,
mineralization to the strong Pb and Zn-rich halo of the 2013, Utah’s Extractive Resource Industries 2012:
Bingham Canyon porphyry Cu deposit. Six ‘high-priority’ Utah Geological Survey Circular 116, 35 p.
targets, identified by aeromagnetic, induced polarization,
and magnetotelluric surveys in conjunction with geo- Brasier, M.D., 1993, Towards a carbon isotope stratig-
chemical sampling, were explored using rotary drilling raphy of the Cambrian System—Potential of the

UGA Publication 45 (2016)—Resources and Geology of Utah’s West Desert


Johnson, D.M. and Christiansen, E.H. 39

Great Basin succession: Geological Society of Lon- Keith, J.D., Whitney, J.A., Hattori, K., Ballantyne, G.H.,
don Special Publication 70, p. 341-350. Christiansen, E.H., Barr, D.L., Cannan, T.M., and
Hook, C.J., 1997, The role of magmatic sulfides
Bryner, L., 1961, Breccia and pebble columns associ-
and mafic alkaline magmas in the Bingham and
ated with epigenetic ore deposits: Economic Geol-
Tintic mining districts, Utah: Journal of Petrology,
ogy, v. 56, p. 488-508.
v. 38, no. 12, p. 1679-1690.
Bush, J.B., Cook, D.R., Lovering, T.S., and Morris, H.T.,
Keith, J.D., Tingey, D.G., Hannah, J.L., Nelson, S.T.,
1960, The Chief Oxide-Burgin area discoveries,
Moore, D.K., Cannan, T.M., MacBeth, A.P., and
East Tintic district, Utah; A case history: Economic
Pulsipher, T., 2009, Provisional geologic map of
Geology, v. 55, p. 1116-1147.
the Tintic Mountain Quadrangle, Juab and Utah
Byrne, K., and Tosdal, R.M., 2014, Genesis of the Late Counties, Utah: Utah Geological Survey Open-File
Triassic Southwest Zone breccia-hosted alkali por- Report 545, 19 p.
phyry Cu-Au deposit, Galore Creek, British Colum-
Kildale, M.B., 1938, Structure and ore deposits in the
bia, Canada: Economic Geology, v. 109, p. 915-
Tintic district, Utah: Stanford, Ph.D. thesis Stan-
938.
ford University, 219 p.
Christiansen, E.H., Baxter, N., Ward, T.P., Zobell, E.,
Chandler, M.R., Dorais, M.J., Kowallis, B.J., and Kim, C.S., 1992, Magmatic evolution of ore-related in-
Clark, D.L., 2007, Cenozoic Soldiers Pass volcanic trusions and associated volcanic rocks in the Tin-
field, Central Utah—Implications for the transition tic and East Tintic mining districts, Utah: Athens,
to extension-related magmatism in the Basin and Ph.D. thesis University of Georgia, 190 p.
Range Province: Utah Geological Association Pub- Kim, C.S., 1997, Petrology of productive intrusions and
lication 36, p. 123-142. comagmatic volcanic rocks in the Tintic and East
Cook, D.R., 1957, Ore deposits of the main Tintic min- Tintic mining districts, Utah, USA—Petrography,
ing district: Geology of the East Tintic Mountains mineralogy, and intensive variables: Geoscience
and ore deposits of the Tintic mining districts, Utah Journal, v. 1, no. 3, p. 123-135.
Geological Society Guidebook no. 12, p. 57-79. Kowallis, B.J., Christiansen, E.H., Blatter, T.K., and
Craig, H., 1961, Isotopic variations in meteoric waters: Keith, J.D., 1995, Tertiary paleostress variation
Science, v. 133, no. 3465, p. 1702-1703. in time and space near the eastern margin of the
Basin and Range province, Utah, in Proceedings
Farmin, R., 1934, “Pebble dikes” and associated miner- of the Second International Conference on the Me-
alization at Tintic, Utah: Economic Geology, v. 29, chanics of Jointed and Faulted Rock: Vienna, Aus-
p. 356-370. tria, Rotterdam, A.A. Balkema, p. 297-302.
Grant, J.A., 2005, Isocon analysis: A brief review of the Krahulec, K., and Briggs, D.F., 2006, History, geology,
method and applications: Physics and Chemistry of and production of the Tintic mining district, Juab,
the Earth, v. 30, p. 997-1004. Utah, and Tooele Counties, Utah Geological Asso-
Hamaker, S., and Harris, R.A., 2007, Fault-related ciation Publication 32, p. 121-150.
ground-water compartmentalization in the East Tin- Lindgren, W., and Loughlin, D.G., 1919, Geology and ore
tic mining district, Utah: Utah Geological Associa- deposits of the Tintic Mining District, Utah: U.S.
tion Publication 36, p. 405-423. Geological Survey Professional Paper 107, 282 p.
Hildreth Jr., S.C., and Hannah, J.L., 1996, Fluid inclu- Le Bas, M.J., and Streckeisen, A.L., 1991, The IUGS
sion and sulfur isotope studies of the Tintic min- systematics of igneous rocks: Journal of the Geo-
ing district, Utah—Implications for targeting fluid logical Society, v. 148, p. 825-833.
sources: Economic Geology, v. 91, p. 1270-1281.
Lovering, T.S., 1949, Rock alteration as a guide to ore–
Hintze, L.F., and Kowallis, B.J., 2009, Geologic history East Tintic district, Utah: Economic Geology Mono-
of Utah: Provo, Utah, Brigham Young University Ge- graph 1, 65 p.
ology Studies Special Publication 9, 225 p.
McCallum, M.E., 1985, Experimental evidence for flu-
Johnson, D.M., 2014, The nature and origin of pebble
idization processes in breccia pipe formation: Eco-
dikes and associated alteration: Tintic mining dis-
nomic Geology, v. 80, p. 1523-1543.
trict (Ag-Pb-Zn), Utah: Provo, M.S. thesis, Brigham
Young University, 72 p. McKean, A.P., 2011, Volcanic stratigraphy and a kin-

Comer, J.B., Inkenbrandt, P.C., Krahulec, K.A., and Pinnell, M.L.


40 The Nature and Origin of Pebble Dikes and Associated Alteration: Tintic Mining District (Ag-Pb-Zn-Au), Utah

ematic analysis of NE-trending faults of Allens Saltzman, M.R., Runnegar, B., and Lohmann, K.C.,
Ranch 7.5’ quadrangle, Utah County, Utah: Provo, 1998, Carbon isotope stratigraphy of Upper Cam-
M.S. thesis Brigham Young University, 104 p. brian (Steptoean Stage) sequences of the eastern
Great Basin: Record of a global oceanographic
McKean, A.P., Kowallis, B.J., and Christiansen, E.H.,
event: GSA Bulletin, v. 110, no. 3, p. 285-297.
2011, Kinematic analysis of northeast-trending
faults of the Allens Ranch 7.5’ quadrangle, Utah Savin, S.M., and Epstein, S., 1970, The oxygen and hy-
County, Utah: Utah Geological Association Publica- drogen isotope geochemistry of ocean sediments
tion 40, p. 89-116. and shales: Geochimica et Cosmochimica Acta, v.
34, p. 43-63.
Moore, D.K., Keith, J.D., Christiansen, E.H., Kim, C.S.,
Tingey, D.G., Nelson, S.T. and Flamm, D.S., 2007, Sharp, Z., 2007, Principles of stable isotope geochem-
Petrogenesis of the Oligocene East Tintic volcanic istry: New Jersey, Pearson Education, Inc., 344 p.
field, Utah: Utah Geological Association Publica-
Sheppard, S.M.F., and Taylor, Jr., H.P., 1974, Hydrogen
tion 36, p. 163-180.
and oxygen isotope evidence for the origins water in
Morgan, L.A., Pat Shanks III, W.C., and Pierce, K.L., the Boulder Batholith and the Butte ore deposits,
2009, Hydrothermal processes above the Yellow- Montana: Economic Geology, V. 69, p. 926-946.
stone magma chamber—Large hydrothermal sys-
Sheppard, S.M.F., Nielsen, R.L., and Taylor, Jr., H.P.,
tems and large hydrothermal explosions: Geological
1969, Oxygen and hydrogen isotope ratios of clay
Society of America Special Papers, v. 459, 95 p.
minerals from porphyry copper deposits: Economic
Morris, H.T., 1964, Geology of the Eureka Quadrangle, Geology, v. 64, p. 755-777.
Utah and Juab Counties, Utah: U.S. Geological
Sillitoe, R.H., 1985, Ore-related breccias in volcanoplu-
Survey, Bulletin B-1142-K, scale 1:24,000.
tonic arcs: Economic Geology, v. 80, p.1467-1514.
Morris, H.T., 1979, Geologic map of the East Tintic min-
Smith, A.G., and Pells, P.J.N., 2008, Impact of fire on
ing district, Utah and Juab Counties, Utah: Geo-
tunnels in Hawkesbury sandstone: Tunnelling and
logical Survey Professional Paper 1024, Plate 1,
Underground Space Technology, v. 23, p. 65-74.
scale 1:9,600.
Sturchio, N.C., Muehlenbachs, K., and Seitz, M.G.,
Morris, H.T., and Lovering, T.S., 1961, Stratigraphy of
1986, Element redistribution during hydrothermal
the East Tintic Mountains, Utah: U.S. Geological
alteration of rhyolite in an active geothermal sys-
Survey Professional Paper 361, 145 p.
tem—Yellowstone drill cores Y-7 and Y-8: Geochim-
Morris, H.T., and Lovering, T.S., 1979, General geology ica et Cosmochimica Acta, v. 50, p. 1619-1631.
and mines of the East Tintic mining district, Utah
Taylor, B.E., Eichelberger, J.C., and Westrich, H.R.,
and Juab Counties, Utah: U.S. Geological Survey
1983, Hydrogen isotopic evidence of rhyolitic mag-
Professional Paper 1024, 203 p.
ma degassing during shallow intrusion and erup-
Morris, H.T., and Mogensen, A.P., 1978, Tintic mining tion: Nature, v. 306, p. 541-545.
district, Utah: Brigham Young University Geology
Taylor Jr., H.P., 1968, The oxygen isotope geochemistry
Studies, v. 25, p. 33-45.
of igneous rocks: Contributions to Mineralogy and
Norman, D.K., Parry, W.T., and Bowman, J.R., 1991, Petrology, v. 19, p. 1-71.
Petrology and geochemistry of propylitic alteration
Taylor Jr., H.P., 1974, The application of oxygen and hy-
at southwest Tintic, Utah: Economic Geology, v. 86,
drogen isotope studies to problems of hydrothermal
p. 13-28.
alteration and ore deposition: Economic Geology, v.
Proctor, P.D., 1985, Preliminary geologic map of the Al- 69, p. 843-883.
lens Ranch Quadrangle, North Tintic District, Utah
Taylor Jr., H.P., 1977, Water/rock interactions and the
County, Utah: Utah Geological and Mineral Survey
origin of H2O in granitic batholiths: Journal of the
Open-File Report 69, 18 p., 2 pl., scale 1:24,000.
Geological Society, v. 133, p. 509-558.
Reed, M., and Spycher, N., 1984, Calculation of pH and
Utah Geological Survey and New Mexico Geochronology
mineral equilibria in hydrothermal waters with ap-
Research Laboratory, 2007, 40Ar/39Ar geochronol-
plication to geothermometry and studies of boiling
ogy results for the Tintic Mountain and Champlin
and dilution: Geochimica et Cosmochimica Acta, v.
Peak quadrangles, Utah: Utah Geological Survey
48, p. 1479-1492.
Open-File Report 505, 26 p.

UGA Publication 45 (2016)—Resources and Geology of Utah’s West Desert


Johnson, D.M. and Christiansen, E.H. 41

Walsh, S.D.C, and Lomov, I.N., 2013, Micromechani- Zhao, Z.F., and Zheng, Y.F., 2003, Calculation of oxygen
cal modeling of thermal spallation in granitic rock: isotope fractionation in magmatic rocks: Chemical
International Journal of Heat and Mass Transfer, v. Geology, v. 193, p. 59-80.
65, p. 366-373.
Wright, A.E., and Bowes, D.R., 1963, Classification of
volcanic breccias—A discussion: Geological Soci-
ety of America Bulletin, v. 74, p. 79-86.

Comer, J.B., Inkenbrandt, P.C., Krahulec, K.A., and Pinnell, M.L.


42 The Nature and Origin of Pebble Dikes and Associated Alteration: Tintic Mining District (Ag-Pb-Zn-Au), Utah

UGA Publication 45 (2016)—Resources and Geology of Utah’s West Desert

You might also like