0% found this document useful (0 votes)
135 views302 pages

Bartlett. Petrology and Genesis of Carbonate-Hosted Pb-Zn-Ag Ores, San Cristobal District, Department of Junin, Peru. Oregon State University-PhD

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
135 views302 pages

Bartlett. Petrology and Genesis of Carbonate-Hosted Pb-Zn-Ag Ores, San Cristobal District, Department of Junin, Peru. Oregon State University-PhD

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 302

AN ABSTRACT OF THE THESIS OF

Mark W. Bartlett for the degree of DOCTOR OF PHILOSOPHY

in Geology presented on April 26, 1984


title: PETROLOGY AND GENESIS OF CARBONATE-HOSTED Pb,iig ORES,

ISTOBAL DISTRICT.. DEPRTMENT OF

Abstract approved: Redacted for privacy


Cyrus W. Field

The SanCristobal District is located in the Cordillera


Occidental 110 kin east of Lima, Peru. The deposits consist of poly-
metallic vein and stratiform replacement (manto) oressituated in
rocks of Devonian through Early Jurassic age. Vein-type ores are
localized by steeply-dipping normal faults which trend transverse to
an asymmetrical anticline, whereas manto deposits are found only in
limestones of the Triassic-Jurassic Pucara Group at, or near, inter-
sections with large veins.
A hydrothermal convection cell was initiated by intrusion of an
intermediate to silicic quartz porphyry stock and dike system between
5 and 7 m.y. ago. Hydrothermal alteration produced a pervasive
quartz-sericite mineral assemblage in the intrusive and adjacent
country rocks. Vein mineralization took place in three distinct
episodes, each separated by intervals of brecciation, and produced
mineral assemblages consisting of (1) pyrite-woiframite-quartz, (2)
chaicopyrite-sphalerite-galena-sulfosalts, and (3) manganosiderite-
barite-marcasite. Alteration of the limestones took place prior to,
or concomitant with, mineralization and formed a distinctive man-
ganosiderite-specular hematite assemblage and later jasperoids. Mas-
sive replacement ore exhibits a paragenetic sequence, mineralogy, and
trace metal cOntent similar to those of the veins.
Fluid inclusion and sulfur isotopic data suggest a temperature
H range of 300-450°C for initial quartz-sericitf c alteration and 180-
300°C for deposition of the vein and manto ores. The salinity of the
hydrothermal fluids evolved from the quartz porphyry was between 30
and 50 weight percent NaCl equivalent but was substantially lower (4-
8 wt. % t4aCl equiv ) during mineralization, which suggests dilution
by meteoric waters The mean composition of hypogene sulfides
from vein and manto ore is 4 6 per mil and values range from 24 to
82 per mu Sulfide-sulfur was derived through mixing of magmatic
sulfur with reduced evapori tic sulfate-sulfur from the Pucara Group
Sulfide rnnerals were precipitated predominantly through a com-
bination of decreasing temperature and increasing pH, caused by
fluid-walirock reactions which consumed H+, under conditions of mod-
erate to high CO2 fugacity The intimate spatial and temporal asso-
ciation of vein and manto deposits, combined with supporting geologic
and geochemical evidence thus define an epigenetic hydrothermal ori-
gin for the stratiform manto deposits of San Cristobal
®Copyright by Mark W. Bartlett
April 26, 1984
All Rights Reserved
Petrology and Genesis of Carbonate-Hosted Pb-Zn-Ag Ores,
San Cristobal District, Departient of Junin, Peru
by

Mark W. Bartlett

A THESIS

submitted to
Oregon State University

in partial fulfillment of
the requirements for the
degree of
Doctor of Philosophy
Completed April 26, 1984
Commencement June 1984
APPROVED:

Redacted for privacy

Redacted for privacy


nt 0

Redacted for privacy

Date thesis Is presented April 26, 1984

Thesis typed by Mark W. Bartlett


ACKNOWLEDGEMENTS

This project was undertaken as part of a joint research effort


in behalf of Centromin Peru under the guidance of Dr. Cyrus W. Field.
I would like to thank both Julio Pastor and Cy Field for initiating
this project and making the necessary arrangements. Partial funding
of this work was provided by grants from Conoco Minerals Company, the
Economic Geologists Corporate Research Fund and Sigma Xi The field
work was fully supported by Centromin Peru.
Fidel Vera introduced me to the geology of the San Cristobal
District and provided me with a number of samples from his personal
collection. Fidel, and his wife Dora, also ensured that my stay in
Peru was productive and enjoyable. Victor Ampuero and Gil Davila
assisted me with sampling and mapping, and helped me find my way
through the labyrinth of mine workings. My sincere regards go to all
the other geologists, engineers, and their families at San Cristobal
and Mahr Tunel who graciously accepted my attemps at speaking spanish
and who allowed me to play goalie in the weekly game of fubito.
I am deeply indebted to Cy Field and Bill Purdom for their
critical reviews of the manuscript. Also I am grateful to the many
graduate students and faculty members in the department of Geology,
far too numerous to mention individually, who have supplied an un-
limited resource of ideas and encouragement. In particular, Ron
Senechal, Dr. H. E. Enlows, and Dr. E. M. Taylor for their useful
comments and suggestions throughout the project. Thanks go to Tom
Horning who provided his support and friendship during our under-
graduate and graduate studies. Therese Belden did an untold number
of favors for me, for which I will never be able to repay her.
Finally, I am indebted to my wife Gail who put her own career as
a geologist aside during the course of my graduate education, and who
provided endless love, support, and encouragement over the last four
years.
TABLE OF CONTENTS

Page

INTRODUCTION. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
History... . . . . . . . . . . . . . . . ........ . . . . . . . . . . . . . . . . . . . . . . . 1
Iermi nol ogy. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
Previous Investigations ...... ........ 6
PurposeandMethodsoflnvestigation ....... ............ 10

REGIONAL GEOLOGY AND TECTONIC FRAMEWORK............. ...... .. 12

STRATIGRAPHIC UNITS. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
Excelsior Phyllite....... . . . ..... . ... .. ... .. . .. .. ..... .. 17
MituGroup ...... . ... . .. .. .. ... . . .. ..... .. .. ... . ...... . .. 19
Pucara Group. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
Aramachay Formation... ....... . . . ..... . . ......... . . . 22
Condorsinga Formation...... . . . . . . . . . . . . . . . . . . . . . . . . 23
Goyllarisquisga Formatlon.......... ............ ....... .. 25
MachayGroup. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

IGNEOUSROCKS ................................................ 27
Carahuacra Quartz Monzoni te. ........ . .. ..... .. ... .. . 27
Petrography and Petrochemistry.. .. .. .. .. ........ 29
Alteration Mineralogy and Geochemistry 35
Chumpe Quartz Porphyry. . . . . . . . . . . . . . . . . . . ........ . * . . . . . 36
Petrography and Petrochemistry. .. .. . . 39
Alteration Mineralogy and Geochemistry ......... 41
Hydrothermal Al teration Zones..... ..... .. . ....... 46
Age and Correlation . . . . . . . . * ....... . ....... . . . ......... 53

STRUCTURAL GEOLOGY . . . ...... . . . . ...... . . . . . . . . . . . . . ........ . 58


Folds. ......... .. .... .. .. ... ....... ... ............... .. 58
Faults...... . . ....... . . . ........... . . . . . . . . . . . . . . . . . . . . 64
Joint Patterns. . . . . .......... . . . . . . . . . . . ...... . . . . . . . . . 66
Breccias . . . . . . . . . . . . . . . . ....... . . . . . . . . . . . . . . . . ....... . 67

MINERAL DEPOSITS. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
Vein-Type Ore. . . . . . . . . . . . . ....... . . . ..... . . . . . ......... 73
Structural and Lithologic Controls.. .......... 74
Mineralogy and Trace Metal Geochemistry ....... 77
Paragenesis. . . . . . . . . . . . . . . . . . . . ............... . * 94
Lateral and Vertical Mineral Zonation ......... 98
Metal-Metal Associations.... ... ..... . .......... 100
Alteration...... ........ . . . ........ . . . * . 103
Oxidation and Secondary Enrichment ..... ...... ...... 107
Page

Manto-type Ore.......................................... 107


Structural and Lithologic Controls................. 110
Mineralogy and Trace Metal Geochemistry............ 118
Paragenesi s . . . . ............. . . . . . . . . . . . . . . . . . . . . . . . 138
Lateral and Vertical Mineral Zonation.............. 143
Metal-Metal Associations...... . . . . . .. . . . . . . . . . . . . .. 145
Altera ti on . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
Oxidation and Secondary Enrichment.................. 152

ALTERATIONOFIHEPUCARAGROUP ......... 154


Fe-Mn Metasomati sin ..... . . . . . . . .......... . . . . . . . . . . . . . . 155
Sf1 icificatf on. ...... * . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
Argillic Alteration of Non-Calcareous Rocks............. 170

FLUID INCLUSION STUDIES. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175


C humpe Quartz Porphyry. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
H omogeni za ti on Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
Sal mi ty Data.. . . . . . . . . . . . ....... . . . . . . . . . . . . . . . . . . 190
Vein-Type Ore . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
Hoinogenization Data.. . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 196
Salinity Data...................................... 197
Manto-Type Ore... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
H omogeni za ti on Data . . . . . . . . . . . . ..... . . . . . . . . . . . . . . . 200
Salinity Data...................................... 201
Pressure Corrections. ..... . . ........ . . . . . . . . . . . . . . . . . . . . 201
Chumpe Fluid Inclusions...... . . . . . . . . . . . . . . . . . . . . . . 202
Vein andManto Fluid Inclusions ...... .............. 205
Interpretation of Data.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206

SULFUR ISOTOPES..... ...... .................................. 215


Source of Sulfur.. . . . . . ..... . . . . . . ....... $ ........ . . . . . 220
Isotopic Temperature Estimates........ . . . ......... . ..... 226

CHEMISTRY OF THE ORE-FORMING FLUIDS ...... .................... 229


Tempera ture and Pressure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
Oxygen,Sulfur,andCO2Fugacity ....... ..... 233
pH ...... . . . . . . .......... . . . $ . . . ....... 244
Sulfide Deposition...... ............. . * . . . ......... . . . . . 246

FEATtJRESOFEXPLORATIONIMPORTANCE. ........... . ......... ..... 250

SUMMARY AND CONCLIJSIONS ........ ..... ....... 253

B IBLIOGRAPHY ...... . . . . . . . . . . . . . . . . . . . . . .......... . . . . . . . . . . . . 261

APPENDIX. . . . . * . .............. . . . . . . . . . . . . . . . . . . . . . . . $ . . . . . . . . 273


List of Figures

Figure Page

1. Location map of central Peru. 2

2. Generalized geologic map of central Peru. 13

3. Generalized geologic map of the Vauli Dome Region, 15


central Peru.
4. Generalized stratigraphic column of the Yauli Dome 18
Region.

5. Contact of the Excelsior Phyllite and the Carahuacra 28


Quartz Monzonite.
6. Chlorite pseudomorph after pyroxene in the Carahuacra 32
Quartz Monzoni te

7. AKF diagram for the Chumpe Quartz Porphyry and the 45


Carahuacra Quartz Monzoni te.
8. Well-developed quartz-sericite alteration above the 50
Chumpe Quartz Porphyry Dikes, adjacent to Chumpe
Glacier.
9. Ridge above Pal vaine Lake showing pronounced color 51
change from white to black at the contact of quartz-
sericitic and propylitic-argillic alteration zones.
10. Map showing the trend of late Tertiary plutonic rocks 57
in central Peru.
11. Geologic map of the San Cristobal District, Peru. 59

12. Cross-section and interpretation of pinch and swell 61


structures in the Yauli Dome Region.
13. Well-developed chevron folds in the Excelsior 62
Phyllite near the axial crest of the Chumpe
Anticline.
14. Altered tuff bed containing a well-developed bedding 63
plane fault.
15. Back-scattered electron photographs showing typical 85
textural relations between sulfosalt and major ore
minerals.
Figure Page

16. Back-scattered electron photograph and elemental dot 88


map images of a tetrahedrite crystal.

17. Specimen of colloform siderite containing a layer of 93


colloform sphalerite from late-stage mineralization
of the San Cristobal Vein.

18. Parageneticsequence of mineral deposition in vein- 95


type ore.

19. Log-log distribution of lead versus copper concentra- 104


tions in vein-type ore.

20. Log-log distribution of zinc versus copper concentra- 105


tions in vein-type ore.

21. Outcrop of altered limestone showing three vertical 112


veins cutting perpendicular to bedding.

22. Outcrop of a small vein in the Catalina Volcanics. 113

23. Outcrop of the vein shown in Figure 22 at the inter- 114


section with the Pucara Group limestones.

24. Limestone bed cut by numerous veinlets containg man- 116


ganese oxide.

25. Early, manto-type mineralization. 126

26. Manto-type mineralization. 130

27. Paragenetic sequence of mineral deposition in manto- 139


type ores.

28. Sketch of relationship between vein structures and 159


alteration.

29. Photograph of area depicted in Figure 28. 160

30. Sketch of alteration zones developed adjacent to vein 161


and below impermeable tuffaceous layers.

31. Silicified (jasperoidal) limestone outcrop showing 169


preservation of bedding and closely spaced joint
fractures.
Figure Page

32. Argillically altered and mineralized tuffaceous lime- 172


stones.
33. Secondary(?) fluid inclusions in magmatic quartz 178
phenocrysts of the Chumpe Quartz Porphyry.
34. Fluid inclusions in vein and manto sphalerite. 180

35. Histogram of fluid inclusion homogenization tempera- 188


tures in quartz from the Chumpe Quartz Porphyry.
36. Histogram of fluid inclusion salinities in quartz 192
from the Chumpe Quartz Porphyry.
37. Diagram of the NaC1-H20 binary system. 193

38. Histogram of fluid inclusion homogenization tempera- 198


tures from vein sphalerite, vein quartz, manto
sphalerite, and manto quartz.
39. Sunnary of sul fur isotope data for the San Cristobal 219
and Huaripampa Mines.
40. Log f(CO2)-log f(O,) diagram showing stabilities of 231
iron-bearing mineraTs at 250, 1500, and 250°C.
41. Log f(S2)-log f(02) diagram showing stabilities of 234
iron-bearing minerals at 150°C with log f(CO2) = 1,
2, and 3.
42. Log f(S2)-log f(02) diagram showing stabilities of 239
minerals formed during massive sulfide mineralization
at 250°C.
List of Tables

Table Page

1. Monthly production of copper, lead, zinc, silver, and 4


tungsten from the San Cristobal District mines, 1981.
2. Yearly mineral reserves of combined and tungsten ore. 5

3. Theories of origin for stratiform ore bodies. 7

4. Modal analyses of the Carahuacra Quartz Monzonite. 30

5. Chemical analyses of the Carahuacra Quartz Monzonite. 33

6. CIPW normative minerals for the Carahuacra Quartz 34


Monzoni te.

7. Trace metal analyses of the Carahuacra Quartz 37


Monzoni te.

8. Modal analyses of the Chumpe Quartz Porphyry. 40

9. Chemical analyses of the Chumpe Quartz Porphyry. 43

10. CIPW normative minerals for the Chumpe Quartz 44


Porphyry.
11. Trace metal analyses of the Chumpe Quartz Porphyry. 47

12. Emission spectrographic analyses for selected ele- 48


ments in the Chumpe Quartz Porphyry.
13. Age of mineralization or alteration of igneous rocks 56
associated with mineral deposits in central Peru.
14 Emission spectrographic analyses for selected ele- 78
ments in vein-type sulfides.
15. SEM (EDA) analyses of sphalerite, tetrahedrite-ten- 79
nantite, stannite, and an unknown In mineral from
vein-type ore.
16. Microprobe analyses of sphalerite and pyrite in vein- 80
type ore.
17. Microprobe analyses of late-stage carbonate from 81
vein-type ore.
Table

18. Assays of copper, lead, zinc, and silver for selected 101
samples of vein-type ore.
19. Correlation matrix for copper, lead, zinc, and silver 102
assays of vein-type ore.
20. Emmission spectrographic analyses for selected ele- 120
ments in manto.-type sulfides.
21. Microprobe analyses of pyrite and marcasite in manto- 121
type ore.
22. Microprobe analyses of sphalerite in manto-type ore. 122

23. Microprobe analyses of late-stage carbonates in 123


manto-lype ore.
24. Assays of copper, lead, zinc, and silver for selected 146
samples of manto-type ore.
25. Trace metal analyses for samples of sub-economic ore 147
and altered limestone in the Carahuacra, Toldorrumi,
and Moises Mantos.

26. Correlation matrices for assays and geochemical anal- 148


yses of manto-type ore.
27. Micrcprobe analyses of unaltered carbonate, diagenet- 165
Ic dolomite, and Fe-Mn metasomatized carbonate.
28. Summary of homogenization temperatures and salinities 186
for fluid inclusion.
29. Sulfur isotope per mu values (ô34S) for sulfide and 218
sulfate mineral concentrates from the San Cristobal
and Iuaripampa Mines.
30. Delta values of sulfide-sulfide mineral pairs, in- 227
cludng isotopic temperature estimates and fluid
mci usion homogenization temperatures.

Al. Sample locations, descriptions, and summary of analy- 273


tical work.
PETROLOGY AND GENESIS OF CARBONATE-HOSTED Pb-Zn-Ag ORES,

SAN CRISTOBAL DISTRICT, DEPARTMENT OF JUNIN, PERU

INTRODUCTION

The San Cristobal District is located on the east flank of the

Cordillera Occidental at latitude 11°43' S. longitude 76°O5' W.,

approximately 110 km east of Lima, Peru, as illustrated in Figure 1.

The mineralized area encompasses 65 square kilometers and lies imme-

diately south of the well known Morococha District. Access to San

Cristobal from Lima is gained by proceeding easterly on the Central

Andean Highway to Mahr Tunel, and then south through Yauli, to San

Cristobal, by mine development roads.

The central Peruvian Andes are extremely rugged in the vicinity

of San Cristobal. Altitudes range from 4400 m within the Carahuacra

Valley to about 5300 m on the surrounding ridge tops. The district

has been glaciated and contains 1io short, slowly retreating glaciers

on the south sides of the larger peaks. Moraines, lakes, swamps, and

grasslands cover parts of the area, but outcrops are generally well

exposed elsewhere.

History

The San Cristobal Mine and the adjacent Carahuacra Mine have

been exploited on a small scale for silver-rich oxide ores since the

16th century (Mclaughlin, 1945; Geologic Staff, Cerro de Pasco

Corporation, 1948; Lyons, 1968). The Cerro de Pasco Corporation

acquired land in the district and began exploitation of the San


2

10°

120

14

Figure 1. Location map of central Peru (mine legend is AT=Atunsulla,


CACasapalca, CL=Colquijirca, CP=Cerro De Pasco, CZ=
Cobriza, HC=Huachocolpa, JL=Julcani, MOMorococha, SVSan
Vicente, YA=Yauricocha).
3

Cristobal Vein (Siberia Vein) in 1928. Production of copper, lead,


zinc, silver, and a small amount of gold has continued to the pres-.
ent, interrupted only by World War II. In 1967, production of tung-
sten (W03) began after the construction of a concentrator for wolf-
ramite (Vera, 1977).
Centromin Peru, a Peruvian government mining company, replaced
the Cerro de Pasco Corporation on January 1, 1974; they currently
operate the mines at San Cristobal, Huaripampa, and Andaychagua.
Ores of copper, lead, zinc, silver, and tungsten are extracted ata
rate of nearly 55,000 tons per month; of this approximately 20,000
tons per month are from veins and the remainder are from underground
and open pit man to ore bodies. See Table 1 and Table 2 for produc-
tion, reserve, and grade values. Total cumulative production figures
are unknown, but 24 million tons are estimated to have been mined
since 1929.

Terminology

Much confusion has arisen during the last 30 years over the
terminology used to describe ore deposits hosted in sedimentary rocks
and to describe the genetic processes which have formed them. A

deposit is considered strata-bound if it is confined to a single


stratigraphic unit even though the mineralization may or may not be
conformable with the bedding. A stratiform deposit is a subtype of
strata-bound deposits, in which the mineralization is strictly coex-
tensive with one or more sedimentary layers (AGI, 1980).
Table 1. Monthly production of copper, lead, zinc, silver, and tungsten from the San Cristobal
District mines, 1981.

Grade %
Orebody Ore milled (tons) Cu Pb Zn W03 Ag (oz/t)

San Cristobal 11,500 0.69 0.78 4.80 0.13 4.38


Andaychagua 8,000 0.90 6.70 4.90
Iluaripampa (underground) 31,000 1.13 7.17 2.45
Huaripampa (open pit) 4,000 0.90 1.50 5.50

Total combined' 43,000 1.08 6.50 3.20


Total tungsten2 11,500 0.69 0.78 4.80 0.13 4.38

1 Combined ore refers to all production exclusive of ores that contain woiframite.
2 Tungsten ore refers to production from the San Cristobal Vein which contains woiframite.
Table 2. Yearly mineral reserves of combined and tungsten ore.

Grade %
Year Type Tons Cu Pb Zn W03 Ag (oz/t)

1977 Combined 7,287,400 0.2 1.3 7.8 4.7


Tungsten 292,260 1.2 1.2 4.3 1.11 4.3

Total 7,579,660

1978 Combined 7,942,520 0.2 1.1 7.5 4.2


Tungsten 772,900 0.8 0.9 3.3 0.42 3.6

Total 8,715,420

1979 Combined 9,300,230 0.2 1.1 7.8 3.9


Tungsten 1,131,260 0.7 1.0 3.0 0.34 4.0

Total 10,431,490

1980 Combined 9,499,240 0.2 1.0 7.4 3.5


Tungsten 1,102,440 0.7 1.0 2.9 0.28 3.9

Total 10,601,680

1981 Combined 8,466,420 0.2 1.1 7.4 3.5


Tungsten 856,360 0.7 1.1 3.3 0.28 4.3

Total 9,322,780

Reserve values from San Cristobal Mine Division, Centromin Peru.


The term manto is of Spanish origin and literally means blanket

Its usage should be reserved for tabular bedded ores with extensive

lateral continuity and short vertical dimensions within a plane of

bedding. However, the term as used by most geologists has been

somewhat generalized to mean any ore deposit of tabular to oval

shape, with a lateral dimension longer than its vertical extent

(al though not restricted to one sedimentary bed). The latter defini-

tion will be used in this thesis.

Theories of origin of stratiform ore bodies have been summarized

by Snyder (1967). He states, "A deposit is syngenetic if formed by

processes similar to and simultaneously with the enclosing rock;

epigenetic if introduced into a pre-existing rock. A diagenetic

origin implies deposition of metals with the host sediments but with

recrystallization, rearrangement, and limited migration." (Snyder,

1967, p. 2). These theories and their respective modes of origin are

summarized in Table 3. The definitions agree with the usage of the

Glossary of Geology (AGI, 1980), and are used as such in the thesis.

Previous Investigations

The first geologic studies of the region were conducted in the

early 1900's but were primarily concerned with description of the

stratigraphy and biostratigraphy. Mclaughlin (1924) summarized this

previous work and presented a compilation of the stratigraphic sec-

tion with the probable ages of the units. Detailed stratigraphy and

geologic cross sections were published by Harrison (1940) for a large


7

Table 3. Theories of origin for stratiform ore bodies

Type of Deposit Mode of Origin

Detrital
Syngenetic Metals in solution in sea water
Metals derived from volcanic exhalations

Recrystallization of materials in situ


Diagenetic Minor migration, metals from host
Minor migration, metals from volcanic exhalations

Cold meteoric water


Hydrothermal
Epigenetic Heated meteoric water
Metals from known igneous source
Metals from unknown igneous source
Connate water

(modified after Snyder, 1967, p. 2)


part of the Department of Junin. Based on this mapping Harrison
interpreted the geologic history of the region. Mclaughlin and Moses
(1945) and Geologic Staff, Cerro de Pasco Corporation (1948) provided
the first geologic summary of the mines owned by Cerro de Pasco in
this region. Early production figures and the colorful history of
the corporation in Peru are contained in these writings. Other
stratigraphic studies pertaining to the Mesozoic section are provided
by Jenks (1951), Terrones (1949), and Szekely and Grose (1972).
Additional detailed geologic summaries of the San Cristobal
District include work by: Harrison (1943), Tosi (1956), Amstutz
(1961), Petersen (1965), Lyons (1968), Lujan (1970), Pastor (1970),
Dunin-Borkowski (1975), Kobe (1977), Daiheimer (1980), Lepry (1981),
Campbell (1982), Campbell and Rye (1982; 1983). Tosi (1956) publish-
ed descriptive geology of the Carahuacra Mine and stated that the
manto ore bodies were the result of replacement by low temperature
hydrothermal waters derived from a nearby intrusive. This work was

subsequently fol lowed by Lyons (1968) report on the same mine. He

agreed that the ore deposits were of a replacement type, but that the
genesis of the ore deposit was difficult to associate with either the
Carahuacra stock or to the crosscutting vein system.
Petersen (1965) briefly mentioned the district in his broad
review of the ore deposits in central Peru, and stated that the manto
ore bodies are hydrothermal replacement ores. Lujan (1970) described
the mineralogy, paragenesis and textures of the ores of the San
Antonio Manto (the southern portion of the Huaripampa Mine) in an
9

unpublished report for Stanford University. He also ascribes the


deposit to hydrothermal replacement, based on the mineralogy, al tera-
tion, and textural evidence. Detailed descriptive work of the vein
mineralization, paragenesis, and associated alteration for the San
Cristobal Mine are discussed by Pastor (1970). Fluid inclusion and
stable isotope work have recently been completed for the San
Cristobal Vein system by Campbell (1982) and Campbell and Rye (1982;
1983). Geologic mapping has been accomplished on both a regional and
local scale by Harrison (1943), Kobe (1977), and Lepry (1981). Lepry

also discussed the structural evolution of the Vauli Dome region


which includes the San Cristobal District. More detailed mapping
(1:10,000 and 1:5000 scales) of the manto ore body outcrops was
completed by Daiheimer (1980).
Reinterpretation of the geneses of some Peruvian manto-1'pe ore
bodies was suggested by Amstutz (1961). Dunin-Borkowski (1975) im-
plies that many of the ore deposits located in the basal Pucara
limestone are the result of syngenetic mineralization, which may have
been remobilized by later hydrothermal events or by diagenetic pro-
cesses. Kobe (1977) stated that the depositional environment of the
host rocks may favor the precipitation of heavy metal sulfides from
seawater. Tentative conclusions regarding the geneses of the nianto
ore bodies in the San Cristobal District were presented by Dalheimer
(1980); he believes that mineralization might be related to syngenet-
ic volcanism during deposition of the Pucara limestone. Further
detailed geochemical work has yet to be presented to substantiate his
10

thesis.

Purpose and Methods of Investigation

Previous geologic studies of the manto-type ore bodies in the

district have been of a somewhat cursory and generalized nature.

Because these manto-type ores are stratiform by definition, they may

have been formed by one of several genetically distinct processes,

such as epigenetic hydrothermal replacement, or syngenetic/diagenetic

sedimentary mineralization. The primary purpose of this investi-

gation has been to resolve some of the ambiguities which have arisen

among these genetic models of sulfide deposition. This has been

accomplished, in part, by detailed mapping of the distribution of

rocks, ores, and surface and subsurface structures. The author's

interpretations are supported by systematic petrographic, miner-

alogic, and chemical studies of samples collected during field work.

Various methods of study were used to resolve the complexities

of the problem. Thin sections, polished thin sections, X-ray dif-

fraction, electron microprobe, isotopic, and SEM energy dispersive

analysis techniques were used to identify specific lithic, mineralog-

ic, textural, and chemical variations of host rocks and ores. Chemi-

cal studies include major oxide, trace elements, assays, fluid inclu-

sion properties, isotopic abundances of sulfur, and radiogenic iso-

tope age-dating techniques.


11

These methods have collectively served to define, (1) specific


lithic, chemical, and structural controls of mineralization, (2) geo-
logic and absolute ages of intrusive emplacement and alteration,. (3)
provenance of sulfur, (4) chemical nature and source of the ore-
forming fluids and, (5) continuity of mineralization from vein-type
ore to manto-type ore.
Thus, as a consequence of these topical investigations, it has
been possible to develop a model for the formation of these strati-
form/stratabound sulfide deposits, as elaborated in the chapters that
follow, based on these geologic and geochemical methods.
12

REGIONAL GEOLOGY AND TECTONIC FRAMEWORK

The San Cristobal District lies within the Central Andean Meso-
zoic Bel t as defined by Petersen (1965) and depicted in Figure 2.
This lithologic belt strikes northwesterly along the highest part of
the central Peruvian Andes from approximately g0 130 south lati-
tide. It is characterized by thick accumulations of shelf carbonate
and clastic marine sediments which have been moderately to strongly
folded into northwest-trending structures (Petersen, 1965). Paleo-
zoic basement rocks, when exposed, crop out in the cores of anticlin-
al structures, whereas Tertiary rocks are preserved in synclinal
folds. Small to moderate sized stocks and volcanic vents of Tertiary
age were intruded along the axes of major anticlinal structures.
The Eastern Paleozoic Belt consists predominately of Paleozoic
(pre-Permian) shales, with thin sandstone interbeds which have been
folded and locally metamorphosed to phyllites, slates and low-grade
schists (Harrison, 1943; Petersen, 1965). These pre-Permian rocks
are overlain unconformably (discordantly?) by Permian volcanic flows
and continental clastic sedimentary rocks (Szekely, 1969). The East-
ern Paleozoic Belt is intruded by stocks and batholith-sized intru-
sives of Paleozoic to Tertiary age (Petersen, 1965; Megard, 1978).
To the west, the Cenozoic Volcanic Belt of Tertiary to

Quaternary volcanic rocks is exposed. They were intruded by small


subvolcanic stocks (Petersen, 1965) and have been slightly deformed
by Cenozoic tectonism (Harrison, 1943; Noble and others, 1979).
13

'76°
nina

p K i M\\Afl Mz \ Q
Pz

'-'., banuco
TQv Mz
Pz

7z\:\
\ç\
X\/)IZ
A
t.>) \\oOxaPon
<
/j\

\ \grreasco
ç

V42\ N
rKTi ,
TQv c Pz OSan Rai

Mz Fig 3 _ 12!_....1 VA

01
1.

- v
.7 ç
0

.
\J '\
I
TQv
i3OHuancayo
"N
t-
c.
Mz
<A V-'7 TQ v
CZKTI\
-
Mz\
-N- '- -

Mz Ps',L..\
> 4 *GPZ
10
Ayacucho
IOOfA< TQ v
Kilometers

Figure 2. Generalized geologic map of central Peru (modified after


Petersen, 1965; stratigraphic legend is PzPaleozoic base-
ment rocks, Mz=Mesozoic sedimentary rocks, pKi=pre-
Cretaceous intrusives, KT1=Cretaceous-Tertiary intrusives,
Tc=Tertiary Casapalco Fm, TQv=Cenozoic volcanics, Q
Quaternary sediments; mine legend after Figure 1.
14

The tectonic evolution of the Peruvian Andes is highly complex


and involves at least four major orogenic cycles (Caldas, 1983).
These orogenic events include the Precambrian, Caledonian, Hercynian,
and Andean cycles. Of these, only the later two have been recognized
in the area encompassed by Figure 3.
Locally, the Hercynian cycle affected sedimentary rocks depos-
ited during the Early and Middle Paleozoic Time. Two phases of
deformation, the Eohercynian and the Late Hercynian, had markedly
different effects on the rocks. Eohercynian deformation is charact-
erized by polyphase structures, which include primary folding and
development of schistosi' trending 1. 300 W. crossed by a second, N.

600 E. trend of folds (Megard and others, 1971; Martinez and others,
1972). Metamorphism to the lower greenschist facies was concomitant
with folding. Late Hercynian deformation in central Peru resulted in
general uplift accompanied by NW-SE trending block faulting (Cal das,
1983).

The Andean cycle began with deposition of Mesozoic sediments in


two elongate belts, controlled by basement faults parallel to the

western margin of Peru (Caldas, 1983). The western bet, also known
as the Coastal Mesozoic Bel t (Petersen, 1965), is characterized by
subaerial and marine volcaniclastic rocks which exceed 7000 m in
thickness. Marine clastic and carbonate rocks as much as 6000 in
thick are found in the eastern belt (Central Andean Mesozoic Belt).
The coastal batholith was emplaced between these two elongate sedi-
mentary basins during Late Jurassic and Cretaceous time.
Ti( 4 Pm P
Pm
II Km

/ \\
I.'
( 0
\" I Mahr Tunel
JP\p Jp
Casapalca \ Kg Kg \l<'
\
IYauii
Tc Km I Tc
-ll°40' Pm

jfl Tertiary Intrusives


\\\ De
"Km
Km
Km

Casapalca Formation TIT

\\
J
Kg Ti K \\rKg
Machay Group \ Corhuoci Kg \
Goyllarisquisgo Formation \ -
Tc Pm
"ZN
Triassic-Jurassic
Pucora Group
E1Permian Mitu Group
\\ Kmf\\

Kg
Jp
Sn CristobaN.
Ti
De
Ahd"aychagua
Km

Devonian Excelsior Phyllite Ti De


Kg
- Thrust Fault \\'K Km Pm Km
m

Figure 3. Generalized geologic map of the Yauli Dome region, central Peru (after Harrison, 1939).

I-I
16

Major deformational events of the Andean Cycle include (1) the


Late Cretaceous-Early Eocene "Peruvian phase", (2) the Upper Eocene-
Lower Oligocene "Incaic phase", (3) the Miocene-Pliocene "Quechua
phase" and, (4) Quaternary deformation. These names were proposed by

Steinmann (1929) and have been retained in most recent literature.


The Peruvian phase resulted in broad folding and general uplift
of the Andes to possibly 1000 m above sea level (Caldas, 1983).
Incaic deformation is responsible for most of the northwest-trending
folds, thrust faul ts, and compression-induced features in central
Peru (Noble and others, 1979). The Quechua phase folded Miocene
continental sedimentary rocks. Post-Quechua phase deformation gener-

ally is restricted to epirogenic uplift of the Andes and longitudinal


faul ting which gave rise to Pliocene-Quaternary volcanism in southern
Peru (Caldas, 1983). Reactivation of faults produced during earlier
deformational phases has continued to the present.
A detailed discussion of these events and their specific contri-
butions to the development of structures in the San Cristobal Dis-
trict is provided in another chapter. The reader should refer to the
works of Hamil ton (1969), James (1971), Cobbing and Pitcher (1972),
Megard (1978), Noble and others (1979), Zeil (1979), and Caldas
(1983) for a more detailed analysis of the tectonic evolution of
Peru.
17

STRATIGRAPHIC UNITS

Sedimentary rocks, which range from Early Devonian to Late


Cretaceous in age are exposed in the Chumpe Anticline; a northwest-
trending, asymmetrical fold. This anticline is part of a larger
regional feature known as the Yauli Dome (see Fig. 3). A strati-
graphic column showing generalized thicknesses of the units in the
Vauli Dome Region is presented in Figure 4. Tertiary plutonic rocks
have been intruded along the axis of the Chumpe Anticline and they
cross-cut all of the sedimentary units. The stratigraphic sequence
includes the Devonian Excelsior Phylifte, the Peniian Mitu Group, the
Triassic-Jurassic Pucara Group, the Lower Cretaceous Goyllarisquisga
Fonflation, and the Cretaceous Machay Group. A brief description of
each unit is included in this chapter along with a more detailed
description of the Pucara Group because of its importance In under-
standing the stratigraphic control of the manto-lype mineralization.
Excelsior Phyll ite

The Excelsior Phyllite is the oldest rock unit exposed in the


district. It crops out in the core of the Chumpe Anticline and in
the Yauli Dome where it has an exposed strike length of approximately
18 km and an width of 1 to 2 km, and is considered to be the basement
for much of central Peru. The Excelsior Phylifte is composed of dark
gray to black shales which have been metamorphosed locally to slate
and finely crystalline, strongly foliated phyllite. In addition to
the slate and phyllite, Lepry (1981) has described lens-shaped beds
1500

300

METERS

Figure 4. Generalized stratigraphic column of the Vauli Dome Region


(modified after Lepry, 1981).
19

of fossiliferous and recrystallized limestone up to 15 m thick in the


core of the Chumpe Anticline. Northwest-trending veins of "bull"
quartz cut the phyllites within the anticline.
Thickness measurements of the Excelsior Phyllite It? central Peru
are largely uncertain because the basal contact of the phyllite is
not exposed. Moreover, it is easily eroded and tends to form topo-
graphic lows. Harrison (1943) estimated a thickness of 2000 m for
the Excelsior Phyllite near Tarma, 50 km east of San Cristobal.
Lepry (1981) uses this thickness as a minimum estimate of the section

in the Yauli Dome. The Lower Devonian age of the phyllite was deter-
mined by Harrison (1943) based on brachiopods collected at Tarina.
Observations by Lepry (1981), Pastor (1970) and the author,
indicate that many outcrops of the phyllite have been chloritized and
sericitized and these alteration assemblages will be discussed in a
subsequent chapter. A distinct angular unconformity exists between
the Excelsior Phyllite and the overlying Mitu Group.
Mite Group

The Mitu Group of Permian age rests unconformably on the


Excelsior Phyllite. It is divided into two intertonguing and
contrasting ii thol ogies comprised of sedimentary and vol canic facies,
known as the Mite Formation and the Catalina Volcanics, respectively.
The Mitu Formation is composed of arkosic sedimentary rocks
represented by conglomerates, sandstones, and thin-bedded sil tstones.
The conglomerates contain large clasts of phyllite, bull quartz,
quartzite, and rounded volcanic rock fragments. Extreme lateral
20

variations in thickness and composition of both facies have been


noted by Lepry (1981) in the Yauli Dome region. At the San Cristobal
Mine, the MItU Group consists almost wholly of the Catalina Volcanics
member, but the clastic member is present immediately to the north
and south of the mine.
The Catalina Volcanics consist of several thick porphyritic
flows ranging from andesite to dacite in composition and containing
well-developed flow breccias. Lepry (1981) measured a 300 in thick-
ness for the Catalina Volcanics at San Cristobal, but elsewhere the
thicknesses vary widely throughout the region. Pastor (1970) meas-

ured a 700 m thickness for the Mitu Group southwest of San Cristobal.
The flows and continental sedimentary rocks are brownish red to
maroon in color, and are variably altered to minerals of the propy-
litic assemblage. Much of the alteration at San Cristobal was caused
by the hydrothermal system developed after the intrusion of the
Chumpe stock. This alteration will be discussed in a later section.
A distinct angular unconformi1' marks the boundary between the
Mitu Group and the overlying limestones of the Pucara Group. This
unconformity was noted by Johnson and others (1955) and Lepry (1981)
along the northeast flank of the Vauli Dome. The unconformity also
is well-exposed along the southwest flank of the Chumpe Anticline in
the Moises Manto area.
21

Pucara Group

The Triassic-Jurassic Pucara Group originally was named by


McLaughlin (1924) and represents deposition of the first regionally
extensive carbonate unit in the Mesozoic Period. Thickness of the
group is extremely variable throughout central Peru, and ranges from
a thick eastern facies of 2900 m to a much thinner western facies of
630 m (Jenks, 1951). Megard (1968) and Grose and Szekely (1968)
subdivided the Pucara Group into three formations; the Upper Triassic
Chambara Formation, the Lower Liassic Aramachay Formation and the
Middle to Upper Liassic Condorsinga Formation.
The Pucara Group according to Lepry (1981) is represented exclu-
sively by the Condorsinga Formation along the west flank of the
Chumpe Anticline. However, Szekely and Grose (1972) have suggested
the presence of a 108 m thick section of the Aramachay Formation in
the Yauli-San Cristobal area, although this unit is absent both to
the north and south. Moreover, based on their measured and described
section of the Aramachay Formation and especially the presence of six
green volcaniclastic beds, it is probable that the upper portion of
the Aramachay Formation is present in the Carahuacra Mine. The

absence of the Chambara and lower Aramachay Formations may be the


resul t of non-deposition and/or post-Chambara erosion (Szekely and
Grose, 1972).
22

The Condorsinga Formation conformably overlies the Aramachay


Formation at the Carahuacra Mine, and is in disconfomable contact
with the underlying Permian Mitu Group just southeast of San
Cristobal. Stratigraphic thickness varies from 400 m at Morococha to
approximately 113 m at Carahuacra (Lepry, 1981; Szekely and Grose,
1972).

Aramachay Formation

Szekely and Grose (1972) characterized the Aramachay Formation


in the Yaul i-San Cristobal area as consisting dominantly of black
shaly limestories which grade upward into cherty, silty sandstones and
limestones. A basal pebble conglomerate to coarse-grained calcareous
arenite is present at the contact with the Mitu Group. This unit is
present at the base of both the Aramachay and Condorsinga Formations
where they overlie the Mitu Group and it contains well-rounded clasts
of bull quartz and metamorphic and igneous rock. Thin to medium-
bedded micritic limestones containing abundant chert nodules are
interbedded with several thin (<1 m) water-laid volcaniclastic tuff
beds. These tuffaceous beds are abundant in the Carahuacra-
Huaripampa mine workings, and are visually conspicuous by their lack
of mineralization in the limestone replacement ores.
The geologic origin of the tuffs is controversial. The tuffs
are regionally extensive markers and characterize the Aramachay
Formation throughout central Peru. Kobe (1977) has discussed their
distribution and petrography. They are typically graded from lapilli
tuff at the base to a fine-grained vitroclastic tuff at the top. The
23

tuffaceous material has been diagenetically altered to chlorite,


smectite, zeolite, and carbonate with a characteristic apple-green
color. The coarse lapilli tuffs exhibit numerous devitrified clasts
with rel ict porphyri tic and pumiceous textures in a groundmass of
fine devitrified glass shards. Plagioclase feldspar crystals are now
mostly altered to carbonate, whereas the ferromagnesian minerals have
been converted entirely to chlorite. Complete gradations exist from
pure tuffaceous beds to carbonate beds having minor tuffaceous corn-
ponents. Their regional distribution and graded bedding suggest that
these tuffs accumulated by water-laid airfall deposition rather than
by local submarine volcanism during the time of Pucara carbonate
sedimentation deposition. They may have originated from volcanism in
the Coastal Mesozoic Bel t 100 km to the west.
Condorsinga Formation

The Condorsinga Formation consists of thin to thick-bedded mi-


cr1 tic 1 irnestones containing several ool i tic beds. Chert nodul es,
thin interbeds of volcanic tuff and tuffaceous limestone, and occa-
sional thin calcarenite beds are common near the base of the orma-

tion. Fossil debris is recognizable in some thin sections, but many


of the original textures have been destroyed by carbonate and silica
al teration adjacent to the man to ore bodies.
Terrones (1949) noted three distinctive units within the upper
portion of the Pucara (Potosi) limestone in the Morococha District.
They are the Sacracancha Trachyte, the Montero Basalt, and the
24

Churruca Limestone Breccia. Lepry (1981) used these marker beds to


delineate the extent of the Condorsinga Formation throughout the
Yauli Dome region. However, only the Montero Basalt and the Churruca
Limestone Breccia are present in the San Cristobal District, and only
as discontinuous beds.
The Montero Basalt is found approximately 150 rn above the Mitu-
Pucara contact in the Moises Manto area. It is about 10 to 15 m
thick, and has a lateral extent of about 900 m. Vesicles (now amyg-
dules) in the upper part of the flow are filled with calcite, and the
base exhibits highly altered pillow and flow structures. These

structures are suggestive of a basalt flow rather than of a sill-like


basalt intrusion.
The Churruca Limestone Breccia lies stratigraphically above the
Montero Basalt, and consists of several breccia units which contain
angular to subrounded clasts of limestone (much of which is oolitic)
cemented bymicrite. Ithas been suggested by Terrones (1949) that
these breccias are of intraformational origin, produced by storm and
wave action against a subaerial carbonate bank.
Detailed stratigraphic studies by Terrones (1949), Jenks (1951),
and Szekely and Grose (1972) do not mention dolomitic beds in the
Pucara, although '1dolomite" is common at both Morococha and San
Cristobal. These "dolomites° are not laterally extensive and will be
shown to be of hydrothermal origin in a later chapter.
25

Petersen (1965) and Haapala (1953) have documented the presence


of evapori tic minerals at the base of the Pucara Group and within
sedimentary rocks of the Mitu Group. Bedded anhydrite is common in
the Pucara Group at Morococha (Field, pers. comm., 1982; Rye and
Ohmoto, 1974). However, the author was unable to locate any evapori-
tic minerals in either formation at San Cristobal, although they may
have been removed by groundwater solution or hydrothermal alteration.
Goyl 1 an sgui sga Formation

The Lower Cretaceous Goyllarisquisga Formation disconformably


overlies the Pucara Group in the San Cristobal District (Wilson,
1963). It is composed of medium to thin-bedded pebble conglomerates,
sandstones, and silty sandstones which are buff to reddish brown in
color. Well-developed sedimentary structures such as ripple marks
and fine crossbedding are present in many of the sandstone beds.
Wilson (1963) measured a thickness of 217 m for the Goyllarisquisga
Formation in the district. He concluded that this formation had
originated in a deltaic to shallow marine environment of deposition
on the basis of sedimentary structures, local coal seams, and minor
carbonate beds that are present in the upper part of this formation.
Coal used for fuel at the smel ter in nearby La Oroya is mined from
the Goyllarisquisga Formation near the town of Cerro de Pasco.
Machay Group

Rocks of Middle to Late Cretaceous age of the Yauli Dome are

represented by limestones of the Machay Group. This group contains


three very distinct lithologic members which have been given forma-
tional status (Benavides, 1956). They consist of the Chulec,
Pariatambo, and the Jumasha Formations. The total thickness of this
group in the Yauli Dome region has been measured at approximately 730
in according to Lepry (1981).
The Chulec Formation contains 350 m of thin to medium-bedded
stratified marl s, sandy 1 imes tones and limes tones. The cl as tic frac-
tion of these sedimentary rocks decreases upward from the base of the
formation. Thus, the upward gradation from sandstones to limestones
suggests a change from a shal low to a deeper water marine environ-
ment. The Pariatambo Formation consists of 80 m of dark gray to
black, thin-bedded shales interbedded with light colored limestones.
Chert nodules are common in the upper part of this formation, and
they are bel ieved to be indicative of a deeper marine environment.
The Jumasha Formation consists of 300 m of medium to thin-bedded
light colored limestones and dolomites which are believed to be
indicative of a return to shallow water carbonate deposition (Lepry,
1981).
27

IGNEOUS ROCKS

Igneous rocks of Tertiary age are well exposed in the San

Cristobal and Morococha Districts. Intrusive activity was much more

intense in the Morococha area and produced lithologies which vary

from quartz diorite to quartz monzonite in composition. At San

Cristobal, two intrusive phases have been recognized; these are the

Carahuacra Quartz Monzoni te and the Chumpe Quartz Porphyry. The

Chumpe intrusive is part of a chain of late Tertiary stocks which

extend from Huancavelica north to Cerro de Pasco (Eyzaguirre and ot-

hers, 1975). Intrusive stocks at San Cristobal and Morococha were

emplaced along the axes of large anticlinal folds on the west side of

the Yauli Dome.

Carahuacra Quartz Monzonite

The Carahuacra stock is located adjacent to the Carahuacra Mine

and was intruded into the Excelsior Phyllite. It has been described

by Lyons (1968) as an elongate, northwest-trending quartz monzonite

porphyry with dimensions of approximately 1100 in by 850 in. Contacts

with the Excelsior Phyllite are relatively sharp, but the foliation

of the phyllite has been sheared and bent parallel to the intrusive

contact (Fig. 5) suggesting forceful emplacement. Alteration of the

phyllite is confined to the shear zone and consists of some bleaching

and silicification. Small veinlets with selvages of kaolinite, cal-

cite, and possibly sericite were observed to crosscut both the phyl-

lite and the stock. These veinlets are related to the San Cristobal
!
-- 1p#

!r:
'

Figure 5. Contact of Excelsior Phyllite (left) and Carahuacra Quartz


Monzonite (right). Note vertical foliation in phyllite
and crude columnar jointing perpendicular to contact in
quartz mozonite.
29

Vein system and were not derived from the Carahuacra stock.
A chill zone 10 to 15 m thick surrounds the stock is indicated
by a slightly finer crystalline texture and a poorly developed colum-
nar jointing.
The intrusive has a characteristic green color on fresh surfaces
owing to intense chloritjzation of the mafic minerals. Phenocrysts

of white plagioclase feldspar and dark green pyroxene (now completely


altered to chlorite) are abundant Outcrop surfaces display a pitted
texture caused by differential weathering of the mafic minerals.
They also show a light orange color due to oxidation of iron.
Petrography and Petrochemistry

The Carahuacra stock exhibits excel lent senate and porphyri tic
textures. The mineralogy, given by modal analyses in Table 4, con-
sist of phenocrysts of plagioclase and pyroxene set in a finely
crystalline hypidiomorphic granular groundmass of plagioclase feld-
spar, pyroxene, quartz, and orthoclase. Sphene, apatite, and titani-
ferous magnetite are the primary accessory minerals.
Phenocrysts of zoned plagioclase feldspar (An3520) generally
are euhedral to subhedral and range from 2.0 to 4.0 mm in length.
Smaller phenocrysts commonly are grouped together in a glomeropor-
phynitic texture, with individual glomerocrysts consisting of four to
five crystals. These smaller phenocrysts are slightly more sodic in
composition (An3020) and are less markedly zoned.
Table 4 Modal analyses (volume percent) of the Carahuacra Quartz Monzonite

Periphery ----------------------------------------------- Interior


MB-39-81 MB-40-81 CH-1 CH-2 CH-3 CH-4 MB-41-81

Groundmass1 44 0 55 0 29 2 32 0 31 6 34 6 40 0

Quartz 13 6 10 0 23 0 24 0 22 4 23 6 1]. 0

Plagioclase 30 8 22 0 30 2 27 2 29 8 31 2 35 4
phenocrysts

Pyroxene2 7 6 10 0 11 6 9 6 9 8 6 2 10 0
phenocrysts

Chlorite 20 10 40 30 36 30 26
matrix

Accessory3 2 0 2 0 2 0 2 2 0 8 1 4 1 0
minerals

%plagalt4 45 50 50 180 200 300 300

1 Groundmass refers to a fine-grained mixture of K-feldspar and Na-plagioclase microlites


2 Pyroxene phenocrysts are pseudomorphically replaced by chlorite with or without calcite
3 Accessory minerals include magnetite and sphene, now mostly altered to leucoxene
4 Percentage of plagioclase feldspar altered to calcite, white mica, albite, epidote, and chlorite,
with calcite as the dominant alteration mineral
See Appendix, Table Al for sample descriptions and locations
31

Pyroxene constitutes the other phenocrystic phase, but it has


been entirely replaced by chlorite minerals and calcite. Identifica-
tion is based on the morphology of the chlorite pseudomorphs (see
Fig. 6) after pyroxene which range from 0.5 to 3.0 mm in length.
The groundmass consists of subequal amounts of lath-shaped plag-
foclase feldspar (An15.10.,) and anhedral orthoclase mixed with a
smaller amount of quartz. The quartz Is distinguished from the
feldspar by its clearer appearance and higher relief. Fine-grafned
chloritic masses are distributed through the groundmass and probably
replaced a ferromagnesian mineral (pyroxene?). All groundmass con-
stituents are less than 0.09 mm in size, making accurate modal counts
difficult.
Whole rock major oxide analyses were obtained for seven samples
of the Carahuacra Quartz Monzonite (Table 5). CIPW normative min-
erals have been calculated for all samples and are provided in Table
6. Most significant is the contrast between the modal and normative
mineral compositions. Modal classification based on the IUGS system
(Streckeisen, 1973) places the Carahuacra intrusive in the quartz
monzonite field. Normative calculations however, indicate a composi-
tion very near the alkali granite field. This discrepancy is probab-
ly the resul t of intense propyl i tic alteration which has strongly
enriched the pluton in Na20. The large values ofNa20, up to 5.08
percent, compared to the low values of both K20 (1.76-3.45 percent)
and CaO (0.79-2.92 percent) is unusual. Average granites and quartz
monzonites usually contain subequal amounts of both K20 and Na20 and
32

-
' ,.-...-' L _________
1p a

::
Pl(
-

.7

c:; -F

(
c
4
4 -,
%_) F
0.

- -- * L' ' . I
-

,-

Figure 6. Chlorite pseudomorphs after pyroxene in the Carahuacra


Quartz Monzonite.
Table 5. Chemical analyses of the Carahuacra Quartz Monzonite.

Gil-i CH-2 CH-3 Ct-l-4 MB-39-81 MB-40-81 MB-41-81 Average

Si02 % 69.45 69.91 69.72 69.81 69.77 68.51 69.39 69.50

hO2 0.45 0.45 0.45 0.42 0.47 0.48 0.43 0.45

A1203 16.03 16.07 16.20 15.68 16.51 16.41 16.30 16.17

Fe203 1.41 1.05 1.35 1.27 1.08 1.56 1.47 1.31

FeO 1.61 1.21 1.54 1.45 1.24 1.79 1.68 1.50

MnO 0.03 0.02 0.02 0.01 O.03 0.08 0.12 0.04

MgO 2.21 2.17 2.17 1.47 2.09 2.04 2.50 2.09

CaO 1.99 2.19 1.95 2.92 2.22 0.79 1.14 1.89

Na20 4.75 4.34 4.73 4.37 4.71 4.76 5.08 4.68

1(20 1.97 2.51 1.76 2.50 1.77 3.45 1.77 2.28


0.11 0.08 0.12 0.10 0.12 0.12 0.11 0.11

Total 100.01 100.00 100.01 100.00 100.01 99.99 99.99 100.02

S.G. 2.62 2.65 2.63 2.64 2.59 2.59 2.63 2.62

Alt. prop prop prop prop prop prop prop

Analyses by Peter R. Hooper, Washington State University, Pullman, Washington, 1982.


Prop = propylitically altered rock
See Appendix, Table Al, for sample descriptions and locations.

(A)
Table 6. CIPW normative minerals for the Carahuacra Quartz Monzonite.

CH-1 CH-2 CH-3 CH-4 MB-39-81 MB-40-81 MB-41-81 Average

Quartz % 26.55 27.08 27.96 26.26 27.79 22.63 26.60 26.41


Orthoclase 11.65 14.85 10.41 14.79 10.47 20.41 10.47 13.2-9

Albite 40.15 36.68 39.98 36.93 39.81 40.23 42.93 39.55


Anorthite 9.16 10.35 8.90 13.84 10.24 3.14 4.94 8.65
Hypersthene 6.63 6.08 6.43 4.60 5.89 6.45 7.63 6.24
Corundum 2.72 2.41 3.24 0.70 3.08 3.68 4.21 2.86
Apatite 0.26 0.19 0.28 0.24 0.28 0.28 0.26 0.26
Magnetite 2.04 1.52 1.96 1.84 1.57 2.25 2.13 1.90
Ilmenite 0.86 0.86 0.86 0.80 0.89 0.91 0.82 0.86

Total 100.02 100.02 100.02 100.00 100.02 99.98 99.99 100.12

()
35

low CaO, whereas granodiori tes and diori tes usual ly have subequal
Na20 and CaO with much less K20 (Nockolds, 1954).
Alteration Mineralogy and Geochemistry

The Carahuacra Quartz Monzonite has been weakly altered to a


propylitic mineral assemblage. Dominantalteration minerals are
calcite, chlorite, white mica, epidote, clay, and leucoxene. Pheno-

crysts of plagioclase feldspar have been partly converted to calcite,


white mica, and some epidote, whereas those of pyroxene have been
entirely altered to chlorite and calcite. Small crystals of titani-
ferous magnetite show excellent rinds of leucoxene, and sphene has
been entirely converted to leucoxene. Epidote is not abundant, which
indicates that calcite was the most stable calcium mineral during
alteration. Calcite alteration is not restricted to the cores of
plagioclase feldspar phenocrysts, but is found as patchy aggregates
throughout these crystals. It is also disseminated throughout the
matrix as fine aggregates and small veinlets. Groundmass plagioclase
and orthocl ase appear less al tered than the phenocrysts, but they
have a somewhat cloudy appearance owing to finely disseminated clay
particles.
The intensity of alteration increases toward the central part of
the intrusion. This is shown by the increase in calcite alteration
of the plagioclase feldspar phenocrysts (see Table 4). This altera-
tion was probably caused by a deuteric hydrothermal process during
cooling and crystallization of the magma.
35

Trace element concentrations for silver, copper, molybdenum,


lead, and zinc were determined for seven samples of the quartz mon-
zonite, as shown in Table 7. The values are close to the average for
intermediate and silicic rocks (Turekian, 1971) but they are somewhat
erratic. Copper concentrations range from 40 ppm in less altered
samples near the outer contact to 2 ppm in the samples from the core
of the intrusive. Molybdenum concentrations are relatively constant
and less than 2 ppm in all samples. Silver values range from less
than 0.3 ppm to 0.7 ppm. Both lead and zinc concentrations are
highly variable and sporadic. Lead values range from 13 to 40 ppm
and those for zinc range from 35 to 130 ppm. With the exception of
copper and possibly lead, the concentration of metal does not appear
to correspond to alteration intensity.
Chumpe Quartz Porphyry

The Chumpe Quartz Porphyry consists of a series of small plugs


and subparal lel dikes which have been intruded into the Excel sior
Phyllite along the axis of the Chumpe Anticline. Various rock names
such as alaskite and granite porphyry have been applied to this rock.
The term quartz porphyry is preferred, because the rock contains
large phenocrysts of quartz and its primary mineralogy is somewhat
obscure due to intense alteration. Pastor (1970) named a small plug,
which crops out just south of the San Cristobal Vein, the Chumpe
Intrusive. This lithic designation has been extended here to include
the cogenetic quartz porphyry dikes which are present both northwest
and southeast ofthe plug. Primary minerals include phenocrysts of
37

Table 7. Trace metal analyses of the Carahuacra Quartz Monzonite.

ppm

Sample Ag Cu Mo Pb Zn

CH-1 0.7 40 1 25 60

CH-2 0.4 13 <1 30 80

CH-3 0.3 13 1 40 60

CH-4 0.4 17 <1 18 35

MB-39-81 <0.3 10 <1 14 80

MB-40-81 <0.3 2 2 15 80

MB-41-81 <0.3 2 2 13 130

Analyses by Chemical and Mineralogical Services, Salt Lake City, Utah,


1982.
h
quartz and plagioclase and orthoclase feldspar set in an aphanitic
groundmass. The ground mass consists of predominantly quartz, seri-
cite and clay minerals. Contacts with the Excelsior Phyllite are
sharp and have a near vertical attitude in all outcrops.
Intrusive breccias are present in many places where the dikes
were forcefully injected into the phyilite. Some portions of the
breccia contain up to 80 percent angular fragmental debris consisting
of ci asts of phyl it te and bull quartz. Many of the phenocrysts of
magmatic quartz and feldspar have been shattered into numerous angu-
lar shards.
In outcrop, the dikes have widths of up to 30 m and they are
typically bright white to light green in color. Supergene alteration
of pyrite to jarosite on some weathered surfaces has imparted a
yellow brown color to the rock. The most characteristic feature of
the rock is the quartz phenocrysts which remained unaltered. These

make up between 8 and 15 percent of the total modal mineralogy.


The dike system has been intersected by the underground workings
of the San Cristobal and Polvarin Veins, and it is well exposed above
the Polvarin Mine on the surface. At both localities, there are two
dikes about 20 to 30 m thick and approximately 20 to 50 m apart. They
are somewhat irregular in shape and highly altered.
Petrography and Petrochemistry

This quartz porphyry is characterized by phenocrysts of euhedral


quartz and highly altered feldspar, up to 3.5 nm in length, set in an
aphanitic groundmass of quartz, sericite, and clay minerals which at
one time may have been a glassy groundmass (unaltered samples were
not found in the San Cristobal area). Primary mineralogy included
quartz, orthoclase (sanidine?), plagioclase feldspar, biotite, and
interstitial glass. Modal analyses are listed tn Table 8. Altera-
tion products include quartz, sericite, chlorite, pyrite, and clays.
Jarosite and alunite are found locally and were probably formed by
supergene alteration and oxidation of pyrite.
Quartz phenocrysts of magmatic origin are generally euhedral,
but many have been embayed and resorbed. They contain minute
crystals of zircon, rutile, apatite, and mica, and appear to be
clustered together in 2 or 3 crystal units. Although highly
variable, phenocrysts or glomerophenocrysts of quartz make up approx-
imately 10 percent of the rock.
Because the phenocrysts of fel dspar are so completely al tered,
their original composition is indeterminate. Al teration products
include sericite, quartz, and clay minerals. It is probable that
both potassium and plagioclase feldspar were present before altera-
tion, and thus the original host may have had a composition between
quartz latite and rhyodacite.
Table 8. Modal analyses (volume percent) of the Chumpe Quartz Porphyry.

MB-13-81 MB-14-81 MB-36-81 MB-37-81 MB-43-81 MB-44-81 MB-78-81 Polv R-2 SCR-3

Quartz 10.0 8.4 8.0 8.4 23.0 4.0 15,2 13.4 9.8
phenocrysts
Feldspar 30.0 21.6 24.6 22.0 21.6 21.4 23.2 24.2 25.2
phenocrys ts

Mica 2.0 2.2 2.6 0.8 3.0 2.6 4.8 2.0 3.4
phenocrysts
Groundmass 56.2 62.4 62.8 68.8 43.6 66.2 55.4 53.2 59.8
Pyrite 1.0 1.0 2.0 tr 1.8 2.4 2.0
Alteration 0.8 4.2 --- 8.8 4.8
veinlets
Jarosite/ --- -- 6.0 ---
Alunite

1 Feldspars totally replaced by quartz and sericite; original composition indeterminable.


2 Mica is now 2M1 muscovite, but was probably biotite originally.
3 Fine-grained mixture of secondary quartz and sericite.
4 Veinlets mostly contain secondary quartz, but mica, chlorite, and pyrite are common.
See Appendix, Table Al for sample discriptions and locations.
41

Biotite also has been totally altered and pseudomorphically


replaced by sericite. Its recognition as a primary mineral is based
tentatively on granular aggregates of sphene within the sericite
pseudomorphs.

Pyrite is present both as finely crystalline disseminations and


in veinlets with quartz. It was added during hydrothermal altera-
tion, and constitutes from 1 to 2 percent of the host.
Alteration Mineralogy and Geochemistry

The Chumpe Quartz Porphyry has been pervasively altered along


veins and in all surface outcrops to a quartz-sericite mineral as-
semblage. The sericite has been identified as 2M1 muscovite from X-
ray diffraction patterns. Muscovite pseudomorphically replaces bio-
tite in large crystal forms while microcrystalline sericite replaces
all feldspars and much of the groundmass.
Other al teration products found in the quartz-sericite assem-
blage include secondary quartz, kaolinite, and minor chlorite. Orig-
inal quartz phenocrysts were largely unaffected by alteration, al-

though they contain numerous fractures, abundant fluid inclusions,


and secondary vein quartz. Finely crystalline secondary quartz along
with sericite has been added to the groundmass, and as fracture
fillings that form veinlets. Small amounts of kaolinite are dispers-
ed throughout the rock. Chlorite was found in only one sample (MB-
43-81) of the quartz porphyry, where it is present both in veinlets
and as disseminations throughout the rock. This chlorite, which
imparts a light green color to the host, is believed to be a retro-
42

grade alteration product of sericite.


Pyrite is the only sulfide mineral found in the quartz porphyry.
On the basis of textural occurrence and abundance, this pyrite is
interpreted to have formed by sulfur metasomatism of iron-rich phases
such as magnetite.
Jarosite and alunite, as replacements of disseminated pyrite,
were found in several samples from surface outcrops above Polvarin
Lake. Both sulfate minerals formed as the result of secondary oxida-
tion of pyrite during leaching of the surface outcrops.
Examination of the major oxide chemistry and the normative
mineral calculations (Tables 9 and 10) reveal a rather lypical pat-
tern for intense quartz-sericite alteration. Wholesale removal of
Na20, CaO, and MgO by the hydrothermal fluids has produced an altered
rock that is enriched in Si02, Al203, and K20. Normative corundum
values range form 8.7 to 11.65 weight percent. The distribution of
the samples on an AKF diagram is shown in Figure 7. Note how the
samples group near the muscovite region. With the exception of
chlorite-rich sample MB-43-81, nearly all of the original iron has
been conserved as pyrite, because iron-bearing silicate phases are
absent in hand specimens and thin sections of this rock. The major
oxide data for all samples of the quartz porphyry (Table 9) are
generally similar and imply enrichments of Si02 and K20 at the
expense of MgO, CaO, and Na20. These chemical effects are consistent
with the observed intense quartz-sericite alteration, as are the
normative analyses, except that calculated high contents of ortho-
Table 9. Chemical analyses of the Chumpe Quartz Porphyry.

MB-13-81 MB-14-81 MB-36-81 MB-37-81 MB-43-81 MB-44a-81 MB-44b-81 Polv R-2 SCR-3 AVE.

Si02 73.75 77.99 75.89 77.85 73.58 77.94 78.19 76.71 74.81. 76.30

1102 0 52 0 44 0 44 0 44 0 34 0 49 0 55 0 30 0 51 0 45

A1203 17.60 14.45 15.45 15.11 16.06 15.69 14.58 14.77 17.17 15.65

Fe203 0.86 0.77 1.49 0.34 1.62 0.11 0.50 1.47 0.70 0.87

FeO 0.99 0.88 1.70 0.39 1.86 0.13 0.57 1.69 0.80 1.00

MnO 0.02 0.00 0.03 0.01 0.07 0.01 0.01 0.08 0.06 0.03

MgO 0.50 0.35 0.34 0.38 1.12 0.36 0.35 0.45 0.41 0.47

CaO 0.09 0.02 0.02 0.03 0.03 0.02 0.03 '0.02 0.03 0.03

Na20 0.51 0.65 0.18 0.72 0.70 0.48 0.57 0.47 0.32 0.51

1(20
5.04 4.38 4.36 4.69 4.55 4.70 4.59 3.98 4.99 4.59
0.13 0.05 0.10 0.06 0.08 0.06 0.07 0.07 0.20 0.09

Total 100 01 99 98 100 00 100 02 100 01 99 99 100 01 100 01 100 00 99 99

SG 269 268 269 271 265 247 242 275 261


Alt qtz-ser qtz-ser qtz-ser qtz-ser qtz-ser qtz-ser qtz-ser qtz-.ser qtz-ser
+ chl
Analyses by Peter R. Hooper, Washington State University, Pullman, Washington, 1982.
Qtz-ser quartz-sericite alteration.
See Appendix, Table Al for sample location and descriptions.
C..)
Table 10. CIPW normative minerals for the Chumpe Quartz Porphyry.

MB-13 MB-14 MB-36 MB-37 MB-43 MB-44a MB-44b Polv R-2 SCR-3 Average

Quartz 50.78 56.90 57.32 55.37 49.82 57.03 57.03 57.52 53.65 55.05
Orthoclase 29.81 25.91 25.79 27.74 26.91 27.80 27.15 23.54 29.52 27.13
Albite 4.31 5.49 1.52 6.09 5.92 4.06 4.82 3.97 2.70 4.32
Anorthite 0.41 0.23 0.55 0.24 0.37 0.29 0.31 0.36 1.15 0.43
Flypersthene 1.53 1.03 2.07 1.22 4.24 1.54 1.41 2.67 1.19 1.85
Corundum 11.41 8.71 10.63 8.93 10.11 9.91 8.77 9.81 11.65 10.00
Apatite 0.31 0.12 0.24 0.14 0.19 0.14 0.17 0.17 0.47 0.22
Magnetite 1.25 1.12 2.16 0.49 2.35 0.16 0.73 2.13 1.02 1.27

Ilmenite 0.99 0.84 0.84 0.84 0.65 0.93 1.05 0.57 0.97 0.85

Total 100.35 100.35 101.12 101.06 100.56 101.86 101.17 102.87 102.32
45

K F

Figure 7. AKF diagram for the Chumpe Quartz Porphyry and Carahuacra
Quartz Monzonite.
46

clase are probably actually representative of sericite.


Geochemical determinations of copper, lead, zinc, silver, and
molybdenum for nine samples of the Chumpe Quartz Porphyry are shown
in Table 11. Three samples also were analyzed by semiquantitative
spectrographic methods for trace elements and metals (Table 12).
With the exception of abundant pyrite, the porphyry is unmineralized
from the surface to depths of approximately 200 m. However, Pastor

(1970) has shown a strong correlation between woiframite mineral-


ization and proximity to the porphyry. All metal values are equal to
or less than the crustal average for low-calcium granites (Turekian,
1971). The apparent and sporadic depletions of copper, lead, molyb-
denum, and zinc in a few samples are not unexpected considering the
high intensity of quartz-sericite alteration imposed on the quartz
porphyry. An interesting resul t obtained from the spectrographic
analyses is the unusually high values (10 to 20 ppm) for tin, which
are well above the crustal average of 3 ppm. Anomalously high tin
values also were detected in the ores of both the vein and manto
deposits. This relationship may indicate a genetic link between the
Chumpe Quartz Porphyry and the hydrothermal system which deposited
the ores.
Hydrothermal Alteration Zones

Pastor (1970) has described the effects of hydrothermal altera-


tion on the Excelsior Phyllite and Catalina Yolcanics adjacent to the
main veins at San Cristobal. It is apparent from a reconnaissance
47

Table 11. Trace metal analyses of the Chumpe Quartz Porphyry.

ppm
Sample Ag Cu Mo Pb Zn

MB-13-81 <0.3 12 1 10 40

MB-14-81 <0.3 7 <1 45 45

MB-36-81 0.6 4 2 25 85

MB-37-81 <0.3 10 2 25 85

MB-43-81 <0.3 45 <1 18 75

MB-44a-81 <0.3 4 <1 30 16

MB-44b-81 <0.3 <3 2 15 8

Polv R-2 <0.3 8 <1 25 170

SCR-3 <0.3 6 <1 18 20

Analyses by Chemical and Mineralogical Services, Salt Lake City, Utah,


Table 12. Emmission spectrographic analyses for selected elements in
the Chumpe Quartz Porphyry.

Sample (ppm)
Element MB-13-81 MB-43-81 MB-44-81

Ag <0.5 <0.5 <0.5

B 100 50 10

Cu 20 70 5

Mn 200 1000 70

Pb 15 20 50

Sn 10 15 20

Zn 1000 200 200

Spectrographic analyses performed by Tom Hancock, Specomp Services,


Boise, Idaho, 1982.
Detection limits are: 0.5 Ag, 10 B, 5 Cu, 10 Mn, 10 Pb, 200 Zn.
49

survey conducted in the area above Polvarin Lake that alteration is


not confined just to vein sel vages, but is pervasively widespread
throughout all rock exposures in this area.
Two types of hydrothermal al teration can be readily dis tin-
guished in outcrops on the basis of megascopic examinations. The
first, and most conspicuous, is the quartz-sericite alteration as-
semblage. This zone is extremely well-developed adjacent to and
above the Chumpe Quartz Porphyry plug and dike system. The Excelsior
Phyllite along the ridge above and to the south of the Polvarin
Glacier has been converted to a mixture of quartz, sericite, kao-
unite, and pyrite. This alteration produced a conspicuous change
from an original dark black to a bright white coloration (see Fig.
8). The zone of quartz-sericite alteration is roughly elliptical in
shape and extends approximately 1 km along the crest of the Chumpe
Anticline. Attendant with alteration is an apparent decrease in
foliation. Pastor (1970) noted this textural change in the phyllites
along and adjacent to the veins. In addition, few distinct veins are
observed in the zone of quartz-sericite alteration. The mineraliza-
tion of this zone appears to be largely restricted to minor dissemi-
nations of pyri te.
Surrounding the quartz-sericite zone is a larger zone of pro-
pylitic-argillic alteration (chlorite zone of Pastor, 1970). The

contact between zones is gradational, but the color change is quite


distinct and sharp (Fig. 9). Much chlorite is present in the
propyl I tical ly al tered phyl ii te. Leaching of magnesium and iron from
50

Figure 8. Well-developed quartz-sericite alteration above the


Chumpe Quartz Porphyry dikes adjacent to Chumpe Glacier.
Note the lateral change to darkly colored phyllite and
volcanics on the right side of photo.
51

Iy. .-',

': ;-;

-c )
$1

Figure 9. Ridge above Polvarin Lake showing pronounced color


change from white to b'ack at the contact of quartz-
sericite and propylitic-argillic alteration zones.
the chlorite to form sericite is the apparent cause of the color
change from black to white that is observed upon entering the quartz-
sericite zone (Pastor, 1970). Veinlets and small veins are common
throughout the propylitic-argillic zone. They contain abundant
specular hematite, quartz, and pyrite with lesser amounts of sphaler-
ite and chalcopyrite.
Hydrothermal alteration appears to be related to the Chumpe
Quartz Porphyry stock and dike system. This conclusion is based on
the presence of alteration zones that concentrically envelope and
overlie the intrusive. Fluid inclusion and mineralogic evidence in
support of this conclusion will be presented in a later chapter.
It is important to note that the intensity of alteration within
the phyllite decreases rapidly with distance away from the intrusive.
Permeability of the phyllites is the apparent controlling factor.
However, the Catalina Volcanics which overlie the Excelsior Phyllite
show pervasive propylitic alteration throughout the district. The

volcanic rocks contain moderately abundant chlorite, epidote, cal-


cite, and clay minerals; a "fresh" sample could not be found in the
district Apparently the volcanic rocks were much more susceptable
to chemical alteration because of their composition and (or) high
permeability related to numerous fractures. Veins which traverse the
the volcanic rocks served as major channeiways for the transport of
hydrothermal fluids to these rocks, and these fluids then spread
laterally through flow breccias where present
53

Additional work should be undertaken to delineate the extent of


hydrothermal alteration at San Cristobal. Good descriptive petrogra-
phy has been completed by Pastor (1970) for the principal alteration
types. However, more extensive mapping, sampling, and concomitant
chemical and petrographic analyses would provide a better under-
standing of the origin and distribution of this hydrothermal system.
Age and Correlation

Radiometric K-Ar age determinations of the two intrusives at San


Cristobal had not been undertaken prior to this study. Lyons (1968)

postulated that the Carahuacra Quartz Monzonite was an eroded vol-


canic neck which once fed the flows of the Permian Catalina
Volcanics. This conclusion was based on compositional similarities
with the Catalina Volcanics and the fact that the stock appears to
have been folded toward the southwest during Tertiary time. A

Tertiary age of intrusion for this stock was inferred by Lepry


(1981), who found crosscutting relationships between the intrusive
and the Catalina Volcanics. This discovery made untenable the
Permian age proposed by Lyons (1968).
A whole rock K-Ar age date of 43.5 + 1.6 m.y. was obtained on
the least altered sample collected from the Carahuacra Quartz Monzon-
ite. This age is considerably younger than the Permian age postu-
lated by Lyons (1968). If the stock was a Permian volcanic vent,
then it is possible that radiogenic 40Ar was lost during a later
hydrothermal alteration event Although Lyons (1968) has shown that
the intrusive was affected by Tertiary folding, it is possible that
54

the radiometric age reflects the true magmatic age. Noble and others

(1979) have described the "Incaic" tectonic event, which produced the

major folds and thrust faults in central Peru. Dating of this event

by Noble and others (1979) indicates that most of the deformation and

uplift associated with the Incaic pulse took place between 43 and 41

m.y. ago. This compressional event is apparently related to the

change in direction of absolute motion of the Pacific lithospheric

plate relative to the South American plate (Noble and others, 1979).

Thus, the geologic and radiometric evidence suggests that the

Carahuacra stock was intruded and crystallized inmiediately prior to

the Incaic pulse of compressive deformation.

Unaltered samples of the Chumpe Quartz Porphyry were not avail-

able for analysis. However, a sample with abundant hydrothermal

sericite (2M1 muscovite) located adjacent to the San Cristobal Vein

was analyzed and gave an alteration date of 5.4 0.3 m.y. Mineral-

ization is assumed to be directly related to hydrothermal alteration

of this quartz porphyry on the basis of mineral zonations, fluid

inclusion data, and by analogy to numerous districts elsewhere.

Therefore, this age represents the minimum age of intrusion, altera-

tion, and mineralization. It is likely that intrusion and crystalli-

zation may have occurred 1 to 3 m.y. before this date depending on

the depth of magma emplacement.


55

Petersen (1965) has suggested that most, if not all, of the

magmatic hydrothennal Cu-Pb-Zn-Ag deposits of central Peru are less


than 12 m.y. in age. K-Ar age dates of mineralization or alteration
of igneous rocks associated with hydrothermal deposits in central
Peru are summarized in Table 13. The linear trend of late Tertiary
plutonic rocks, which extends fri Huachocolpa northward to Cerro de
Pasco (Noble and others, 1979; Mckee and others, 1975; and Eyzaguirre
and others, 1975), is illustrated in Figure 10.
The age of 5.4 m.y. obtained for the Chumpe Quartz Porphyry is
part of this late Tertiary magmatic event, and it strongly supports
Petersen's (1965) observation that all polymetallic hydrothermal ore
deposits of central Peru are late Tertiary in age. In fact, this
trend of youthful plutonism and associated hydrothermal mineraliza-
tion may be projected as much as 200 miles north of Cerro de Pasco
and into the Cordillera Blanca Batholith on the basis of radiometric
age determinations reported by Giletti and Day (1968), Landis and Rye
(1974), McKee and others (1979), and unpublished data of my own

undertaken with colleagues.


Table 13. Age of mineralization or alteration of igneous rocks associated with mineral deposits
in central Peru.

District Age (m.y.) Method Source

Cerro de Pasco 14-15 K-Ar Silberman and Noble, 1977

Colqui 10-11 K-Ar Kamilli and Ohmoto, 1977


Morococha 7.3-8.2 K-Ar Eyzaguirre and others, 1975
San Cristobal 5-6 K-Ar This study
Yauricocha 6.7-7.2 K-Ar Giletti and Day, 1968

Julcani 10 K-Ar Petersen and others, 1977


Fluachocolpa 4-8 K-Ar Mckee and others, 1975
57

C)

0 2
G Cf

I____
0 Km 50

0 YauIi Dome
V
C)

?
A

Figure 10. Distribution of Tertiary plutonic rocks, central Andes,


Peru (modified after Megard, 1978; symbols as given in
Figure 1).
58

STRUCTURAL GEOLOGY

The structural development of the Yauli Dome region has played


an important role in the localization of mineral deposits throughout
the San Cristobal District. Lepry (1981) has divided the structural

elements of the region into two groups. The older of these was
produced during the Hercynian deformation of middle Devonian to
earliest Mississippian age, and it affected only the Excelsior
Phyllite. Andean deformation of Early Tertiary to Recent age affects
all the rock units of the region, and this event provides the local
control for all mineralized structures within the district
Compressional forces characterize the Andean deformational style
in central Peru. These forces produced the major north-northwest
trending thrust faults and folds (Lepry, 1981). Noble and others
(1979) discuss the Incaic" tectonic event of early Tertiary time,
and they attribute these the compressional forces to a change in
absolute motion of the Pacific lithospheric plate approximately 43
m.y. ago. The development of these structures and their present
forms in the district are discussed in this chapter.
Folds

The San Cristobal District is situated on the southwest flank of


the northwest-trending Chumpe Antici me. Dominant structural fea-
tures and local geology of the district are summarized in Figure 11
(after Vera, 1977). Although not evident from this figure, the
Chumpe Anticline is a doubly plunging fold. The axial apex of the
59

'
/
" Pcv
CARAHUACRA
MANTO
Kg
''
.

I
fjp
Km
i:::. uaripampa I
/ / Pcv

65 Open Pit 0e
/ / Tert quartz porphyry
Pcv (Chumpe)
Kg / Ten. quanz monzonite
/ " / (Corahuacra)
/ jp?Y /
Cret. Machay Group
600 Pcv 600 Cret. Goyliarisquisga
/
In.- Jun. Pucara Group
De
Perm Catalina Vaicanics
Dcv Excelsior Phyllite
Tiqp_g
Pcv

Km
:ein
I
6
KqI Tiqp De :
I JP San Cristobal
I Mine
Pcv Tiqp :. 2 Vein
400
U
Aiticiine

o
D
I 78°J
I I P0/varin
I I LOI(9ç
I I Pcv
TOLDORRUMI CO3,
MANTO I I' 1KM
f I

I
::. 750 De
Km J
/w:.°° 1"Tiqp
Kg 650

/ / Jp Pcv

/ /

(
()
\\ ; 400
PCV ('P6o0
4, i
Tiqp<\
-.-.'.,, I
( Andoychoqua
, Dc 0 700
MOISES I Dc Mine
U
MANTO I

80,ndaycn0gua Vein
65
Km I J_..-r /
-T1q

/ Kg

Figure 11. Geologic map of the San Cristobal District, Peru (modified
after Vera, 1977).
fold is located under the Chumpe Glacier. The exposed strata gen-
erally dip 550 to 75° Sw. at the Huaripampa Mine, and these dips
gradually decrease to approximately 350 to 450 SW. to the southeast
along the anticline in the Moises Manto area. The southwest flank of
the anticline is slightly oversteepened.
Lepry (1981) ascribes the formation of both the Chumpe Anticline
and the Yaul i Dome, to extreme contrasts in competency between the
ductile Excelsior Phyllite basement and the overlying more competent
volcanic and sedimentary rocks. Under compressive stress, the phyl-
lite basement was "squeezed into a narrow, inverted trough between
the expanses of volcanic rock, producing tight, pinched anticlinal
zones separated by broad synclinal warps.", according to Lepry (1981,
p. 66) and as depicted in Figure 12.
The axial zone of the Chumpe Antici me is highly deformed and
exhibits well-developed chevron folds and kink bands as illustrated
in Figure 13. Intrusion of the Chumpe Quartz Porphyry, and its
associated dikes, occurred along the axial zone of maximum deforma-
tion.
The overlying layered Mesozoic rocks were folded predominately
by flexural slip (Lepry, 1981). Much of the slip was taken up on
thin tuffaceous beds within the Pucara Group, and along the contact
between the Pucara Group and Mitu Groups. Nearly all of the tuffs
show extensive bedding plane shear and deformation; especially near
contacts with the Mitu Group as shown in Figure 14.
6000M1 SW NE 6000M
C ERRO
CHUMPE JERUSALEM ULTIMATUM PACUSH
ANTICLINE THRUST ANTICLINE 5000

1,111*1 Pm 4000

3000

De
2 2000

1000

Figure 12. Cross section and interpretation of pinch and swe1l structures in the Yauli
Dome Region (taken from Lepry, 1981, p. 67).
62

Figure 13. Well-developed chevron folds in the Excelsior Phyllite


near the axial crest of the Chumpe Anticline.
63

\. 'I

-Figure 14. Altered tuff bed approximately 1 m thick containing a


well-developed bedding plane fault. Note closely spaced
joints in silicified limestone adjacent to tuff. Photo
taken in the Nuaripampa Manto area.
64

Faul ts

Five large thrust faults have been mapped by Lepry (1981) along
the southwest flank of the Yauli Dome. They trend 200 to 35° NW. and
dip 35° to 75° SW. All have displacements of less than 1.5 km and
lengths of less than 16 km. However, thrust faults were not observed
in the San Cristobal District.
Numerous transverse faul ts cut the Chumpe Antici me (see Fig.
11). They consist of two varieties; conjugate transverse faults with
possible strike-slip separation, and transverse faults with normal
displacement oriented perpendicular to the fold axis. These trans-.

verse faults host all of the vein ores within the district.
Lepry (1981) noted that the conjugate faults tend to develop in
incompetent lithologies such as the Excelsior Phyllite, whereas nor-
mal transverse faults are best developed in brittle rocks such as the
Catalina Volcanics. Large faults such as the San Cristobal,
Prosperidad, and Andaychagua show a distinct change in attitude as
they traverse from the Catalina Volcanics to the Excelsior Phyllite.
The San Cristobal and Prosperi dad Faul ts delineate a graben
structure in the axial apex area of the Chumpe Antici in Faul ts
north of the axial apex (including the San Cristobal fault) dip
steeply to the southeast (550 to 85°), whereas those south of the
axial apex dip northwest (60° to 85°). All faults and joints in the
district exhibit this pattern. Movement along the faults typically
is normal, but the Aridaychagua Faul t may have reverse motion (Pastor,
1970). Large amounts of strike-slip motion are not evident along any
[I

of the faults.
Formation of the transverse faults appears to be related to a
shear stress imposed at an angle of less than 300 to the maximum
compressive stress that folded the anticline (Lepry, 1981). These
forces produced en echelon tension gashes which preceded actual
strike-slip motion. Longitudinal faults developed at the hinge of
the anticline in response to tensional stress.
There is an apparent problem with the timing of observed de-
formational features and the intrusion of the Chumpe Quartz Porphyry
plug and dikes. The dikes were intruded into the axial crest of the
Chumpe Anticl me approximately 7.0 to 5.4 m.y. ago. In contrast the

existence of transverse faults developed approximately 40 m.y. ago.

Why the Chumpe magma did not invade these fractures remains an un-
resolved problem. It is possible that the stress regime present
during the intrusive event might have prevented movement of magma
into these transverse faul ts. Lepry (1981) does not address this
question, but he does indicate that the transverse faul ts were re-
activated in a normal sense by doming associated with Late Tertiary
igneous activity. The Chumpe dikes are cutby normal faults which
would indicate recent movement along these structures. Pastor (1970)
provides evidence for the periodic brecciation of vein ores during
mineralization, and he also ascribes this brecciation to reactivation
of the faults by doming during episodic intrusion of the Chumpe
parental magma at depth. It would seem from this evidence that much
of the transverse fault movement is not concomitant with folding of
the anticline, but rather is related to Late Tertiary doming during
igneous activity and/or uplift of the Andes.
Joint Patterns

Dip joints are extremely wel 1-developed along the southwest


flank of the Chumpe Anticline. All lithologies above the Excelsior
Phyllite show closely spaced joints which parallel the major trans-
verse faults. The attitudes of these joint sets are directly related
to their location along the flank of the anticline. Joints in the
Huaripampa Mine area are closely spaced (1 to 5 cm apart; see Fig.
14) and have strikes of N. 40-45° E. and dips of 60-75° SE. The

vein system in this area has an attitude of H. 500 E. and 45 to 650

SE. Joints in the Toldorrumi Manto area, one kilometer south of San

Cristobal, strike N. 45_600 E., and dip 75-90° N. whereas the as-
sociated vein system strikes N. 35-50 E. and dips 75-90° H. Farther
to the south, the attitude of the joints also mimics the vein trend,
and the dip angle decreases to 50_600 1
These patterns indicate that the joint system was produced by
the same stress regime that produced the major transverse faul ts.
The attitude of the joints also indicate convergence toward the axial
apex of the Chumpe Anticline.
It is important to note that although the major faults appear to
die out as they enter the Mesozoic section, the joint patterns con-
tinue with apparent persistence. Thus, the carbonate rocks behaved
in a brittle fashion and the rocks must not have been subjected to
67

much confining pressure.


Breccias

The development of the breccias in the Pucara Group is important


because much of the manto-type ore is essentially contained within
them. Several types of breccias can be distinguished in the district
including (1) intraformational breccias formed by wave action during
deposition of the limestone, (2) tectonic breccias formed both by
faults which intersect the limestones normal to strike and by bed-

ding-plane faults within the limestones, (3) hydrothermal solution


breccias (hydrothermal karst breccias), and (4) breccias produced
through hydrofracture enlargement along fractures with the vein and
joints.
The origin of the intraformational breccias has already been
discussed, and the Churruca Limestone Breccia is such an example.
Volumetrically, the tectonic breccias are much more abundant and
constitute the most important host rock for the manto ores. Trans-

verse faults (veins) such as the Virginia, Vein 722, San Cristobal,
Polonia, Catalina, Prosperidad, and Andaychagua intersect limestones
of the Pucara Group on the southwest fi ank of the Chumpe Anticl me.
These faul ts have produced zones of brecciation up to 10 m wide in
the limestone host rocks. The Catal ma Vein split into three seg-
ments and produced three extensive breccia zones as it entered the
Pucara limestone in the Toldorrumi Manto.
Lyons (1968) has discussed the formation of a 110 m thick brec-
cia zone located at the base of the Pucara where it oven ies the Mitu
Group in the Carahuacra Val ley. This breccia narrows to 20 m and
eventually pinches out over a distance of several hundred meters
toward the north. With increasing proximity to the hanging wall, the
well-developed breccia textures diminish by gradation into a crackle
breccia and eventually into non-fractured rock. According to Lyons
(1968), the formation of this extensive breccia was related to thrust
faul ting along the base of the Pucara Group.
Evidence derived from the present study supports a tectonic
origin for the breccias at Huaripampa, and relates their formation to
bed-on-bed slippage during concentric folding of the anticline. The

Huaripampa Mine is located in an extensive zone of brecciation at the


base of the Pucara Group. Here, the footwall contact with the Mitu
Group dips steeply (70° to 85° W). Also, the Carahuacra intrusion is
located just 300 m to the east and into the anticline from this
contact. It appears that the Carahuacra Quartz Monzonite may have
acted as a buttress which helped oversteepen the west flank of the
fold. Because the limestones of the Pucara Group were unable to
adjust soley by slippage on bedding planes, they were crushed and
brecciated against the buttress of the intrusive.
Several other breccias were produced by bedding-plane movement
during flexural-slip folding. Donath and Parker (1964) have de-
scribed cataclasis or brecciation of more brittle layers in rocks of
dominantly one lithology (such as limestone). Several laterally
1*

extensive breccia horizons are situated between two volcaniclastic


tuff beds. These tuffs exhibit severe deformation by bedding-plane
shear. The limestone bed located between the two tuff layers acted
in a brittle fashion and was brecciated during flexural-slip folding.
Hydrothermal solution collapse mechanisms have been used to
explain breccias in the Morococha District (Mark and others, 1942).

Anhydrite beds located in the base of the Pucara Group were supposed-
ly leached by hydrothermal solutions, which produced caverns. Col-

lapse of the cavern roof allowed propagation of the breccias upward


into the limestone section. Haapala (1949) inferred a post-ore
tectonic event to explain the Morococha breccias. However, some
solution of the anhydrite must have taken place, because the sulfide
minerals have incorporated the heavy 34S isotope from the evaporite
(Field and others, 1983).
Although many of the breccias at San Cristobal are of tectonic
origin, there is some evidence for dissolution of breccia fragments
and conduit walls by hydrothermal solutions. Enlargement of breccf a

zones adjacent to vein and joint fractures through a process of


hydrofracturing also is indicated. These mineralogical and deforma-
tional features (including cavern development that result from dis-
sol ution by hydrothermal solutions) have been termed "hydrothermal
karsts" (Dzulynski, 1976). Actual caverns have not been excavated
during mining in the Huaripampa Mine, but smal 1 voids up to 0.5 m are
sometimes found. Clasts within the breccia are extremely angular in
most areas, but rounded fragments showing distinct dissolution and
70

replacement textures frequently are found adjacent to open spaces.


Detailed description of the dissolution, replacement and hydrofrac-
turing processes are discussed subsequently under ROCK ALTERATION.

Typical processes of karst formation invol ving cool meteoric


groundwater were not responsible for any of the brecciation within
the district. Features indicative of paleokarsts include (1) resid-
ual material such as clay or stratified dolornitic sands from solu-
tional disintegration of carbonate, (2) evidence of local discon-
forniities such as fossil soils and organic debris, (3) stalagmites
and stalagtites of carbonate material in caverns, and (4) caverns.
None of these features can be found in the district. Also, the
presence of evaporite minerals in the basal Pucara Group would mdi-
cate a hot, dry climate, rather than the humid, wet climate necessary
for karst development.
There remains one other possible mechanism for the formation of
the breccias. Middleton (196k) has discussed solution breccias pro-
duced in limestones that contain significant concentrations of evap-
orite minerals, and demonstrated an excel lent correlation between
brecciated surface outcrops and anhydrite-bearing limestones in the
subsurface. It is possible that some breccias observed seen in the
mines and mantos (especially the thin, laterally extensive mantos)
may be the resul t of sol ution of thin evapori tic horizons. Permis-

sive indications of this process are indicated by the rather heavy


values of sulfide-sulfur in ores of the mantos and veins.
71

In conclusion, formation of the breccias appears to have been


the result of several events acting at different geologic times.
Structural preparation began during the mid-Tertiary event of folding
and faulting (Incaic period), and subsequent enlargement of breccias
originated through hydrothermal sol utlon and hydrofracturing during
mineralization in late Miocene time.
72

MINERAL DEPOSITS

Ore mineralization at the San Cristobal District occurs within


numerous northeast-trending veins that crosscut the Chumpe Anticline,
and within mantos that replace limestones of the Pucara Group. These

two morphologically distinct ore types are here designated as vein-


type and manto-type deposits, respectively.
Vein-type mineralization is distributed over an area of more
than 30 km2, with the largest individual veins having lengths of up

to 4 km as shown in Figure 11. At present, only the Virginia, San


Cristobal, and Andaychagua Vein systems are mined, although smaller
veins may contain appreciable mineral resources for the future.
Manto-type mineralization is distributed somewhat erratically
over a distance of 8 kilometers along the southwest flank of the
Chumpe Anticline. Major surficial exposures of manto ore have been
named the Santa Aqueda, San Antonio, Toldorrumi, and Moises Mantos
from north to south respectively (Fig. 1]). The Santa Aqueda and San
Antonio mantos actually are part of the same manto system, and will
be referred to collectively in this thesis as the Carahuacra Manto.
The Carahuacra Manto is by far the largest manto in the district.
Both Centromin Peru and Volcan Mines Company operate mines in this
manto. The Huaripampa Mine is located in the southern half of the
Carahuacra Manto; production from both large underground mines worked
by room and pillar methods and open pit workings constitutes almost
two thirds of the total ore mined by Centromirt Peru in the district.
73

Surficial oxidized ores from both the vein and manto systems
were exploited first by the Inca civilization and later by the
Spaniards. The Cerro de Pasco Corporation consolidated the district
claims in 1928, and began production from the San Cristobal Vein.
The manto ores were recognized as potential reserves in 1948 by the
corporation, but were considered uneconomic (Geologic Staff, Cerro de
Pasco Corporation, 1948). In the early 1960's, dril 1 ing indicated
substantial reserves of ore amenable to underground room and pillar
mining. Open pit mining was initiated after depletion of the near-
surface underground ores. The following chapter is devoted to de-
scription of the salient features of both ore types.
Vein-Type Ore

Nearly all of the faults present within the San Cristobal


District are mineralized to some extent The San Cristobal Vein was
the first commercial vein to be exploited, and it has been the
largest producer of copper, lead, zinc, silver, and tungsten in the
district. Development and mining continues on ten levels; the 500 m
level at an elevation of 4,593 m constitutes the deepest working in
the mine. Vein mineral ization extends to the northeast approximately
4 km, although workings have been developed only for a distance of
two km to a point beneath Chumpe Lake. Mineralization continues at
depth, but the tenor and extent of the ore has yet to be determined
by deep drilling.
74

Pastor (1970) has divided the San Cristobal Vein into sections
based on host-rock lithology and vein character. From west to east
these are (1) the Volcanic section which includes vein material
hosted by the Catalina Vol canics, (2) the Contact section which is
the area immediately adjacent to the Catalina Volcanics-Excelsior
Phyllite contact, (3) the Phyllite section that includes vein mater-
ial hosted by the Excelsior Phyllite, and (4) the Dike section where
the vein transects two large dikes of the Chumpe Quartz Porphyry.
These section names will be referred to when specific mineralogical
and textural trends are different from the vein as a whole.
Structural and Lithologic Controls

Localization of the vein ores was controlled predominantly by


pre-existing structures developed during folding and subsequent up-
lift of the Chumpe Antici me. Normal faul t movement in Late Tertiary
time appears to have been largely responsible for the present form of
the veins, al though minor strike-slip movement is evident in parts of
the mine.

According to Pastor (1970), host rock lithology was the most


important structural control to ore deposition and resulted from
contrasts in competency between host rock types during faul t dis-
placement. The Excelsior Phyilite and Catalina Volcanics host the
vast majority of ore grade mineralization. These two lithologies, as
would be expected from their composition, behaved quite differently
during faulting. Where the veins cut the Catalina Volcanics, the
faults typically are well-defined and exhibit clean fractures with
75

little or no development of gouge. Mineralization in these rocks


shows much crustification and cavity filling typical of open space
deposition. Faultdevelopmentin the Excelsior Phyllite was dif-
ferent. The phyllite behaved in a more ductile fashion and tended to
shear along multiple planes of foliation and bedding. Considerable
gouge was produced and in places prevented circulation of the ore-
bearing fluids. Vein walls are ill-defined, and the larger veins
show innumerable stringer zones and splits (Pastor, 1970). Vein size
is highly variable, and ranges up to 10 m in width. Individual vein
structures are characteristic in the Catalina Volcanics, whereas
multiple vein and stringer zones are common in the Excelsior
P hy 11 i te.

Pastor (1970) attributed a large increase in the size of the San


Cristobal Vein to an increase in brecciation within the Contact sec-
tion of the San Cristobal Vein. He attributed this increase in vein
size to brecciation by bedding plane slippage with folding of the
anticline along the Catalina Volcanics-Excelsior Phyllite contact.
Mineralization actually extends laterally into the contact zone from
the veins at this point
As mentioned earlier, the San Cristobal Vein and the Prosperidad
Vein delineate a graben structure developed transverse to the Chumpe
Anticline. In general, all veins (faults) north of the axial apex of
ChumpeAnticline (centered under the Polvarin Glacier) dip to the
south, whereas those south of this point dip to the north. Provided

these two veins maintain their angles of dip (55° SE. and 60° NW.,
76

respectively) they would converge at a point approximately 1.2 km


below the surface. Thus, it is possible, to imagine a source area
and convective plumbing system for the veins at this depth.
The atti tudes of faul ts show a marked change as they traverse
from the Catalina Volcanics into the Excelsior Phyllite (see Fig.
11). Those in the Catalina Volcanics are 'pically N. 45-65° E., but
change abruptly to almost east-west as they enter the phyllite. The

San Cristobal Vein shows this refraction effect best. Pastor (1970)
concluded that the refractive change in direction of the veins was
due to contrasts in competency contrasts between the two lithologies.
Dips on all the veins remain relatively constant at 50-65° throughout
the transition from volcanic to phyllite host rock.
Pastor (1970) has demonstrated that movement along these faults
continued during and after mineralization. In fact, much of his
evidence for paragenesis of the vein mineralization is based on
crosscutting relationships developed during reactivation of the
faults. Renewed movement during mineralization allowed deposition of
more ore minerals in a fracture that had been previously healed.
These movements enlarged the vein widths and also increased the tenor
of the ore. It appears that renewed brecciation also changed the
character of mineralization. This effect will be discussed further
in a later section, but it is possible that resurgence and subsidence
of a magma body at depth may have been responsible for reactivation
of the faults and paragenetic changes in vein mineralogy.
77

Mineralogy and Trace Metal Geochemistry

The mineralogy of the San Cristobal Veins has been discussed by


Geologic Staff, Cerro de Pasco Corporation (1948), Pastor (1970),
Vera (1977), and more recently by Campbell and Rye (1982). The

principal minerals of economic interest include woiframite,


chalcopyrite, sphalerite, galena, tetrahedrite/tennantite, and pyr-
argyrite/polybasite. Pyrite, quartz, barite, marcasite, mangano-
siderite, and rhodochrosite are the dominant gangue minerals.
Present in minor and microscopic amounts are specular hematite,
bornite, covellite, chalcocite, bismuthinite, stannite, magnetite,
stromeyerite, acanthite, native silver, native copper, and several, as
yet unidentified mineral phases.
The following discussion includes a summary of Pastor's (1970)
work on the San Cristobal Vein system, and additional information
obtained by the author regarding trace mineral phases and the distri-
bution of trace metals in the vein ores. Trace metal distributions
were determined through 13 emmission spectrographic analyses on both
mineral separates and whole rock ores to identify general composi-
tional trends. An additional 45 analyses utilizing a scanning elec-
tron microscope equipped with an energy dispersive analyzer (EDA) and
the electron microprobe (wavelength dispersive) were performed to
help locate and identify specific trace metal anomalies and trace
minerals. Tables 14-17 summarize the spectrographic, EDA, and micro-
probe resul ts, respectively.
Table 14. EmIssion spectrographic analyses for selected elements In vein-type sulfides.

ppm
Sample Mineralogy1 Ag As 81 Cd Co Cu Ni Pb Sb Sn W Zn

MB-1-81 si, py, gn 150 5000 N 200 20 150 10 20,000 100 100 N >10,000
M8-9-81 si, cp, tet 150 5000 500 >500 N >20,000 N 200 1,000 1,000 N >10,000
MB-26-81 si, py, gn, mc 10 1500 N 100 N 100 15 10,000 500 500 N >10,000
MB-26-81 py sep 15 1500 N N N 50 30 300 200 150 N >10,000
MB-33-81 sl, gn 700 N N 100 N 300 15 >20,000 700 50 N >10,000
148-34-81 sl, gn 700 7000 100 200 N 300 15 >20,000 300 70 N >10,000
SCM-2 py Sep 5 N N N N 70 30 30 N 10 200 3,000
SCM-4 sI, cp, tet 1500 10000 500 500 N >20,000 N 5,000 >10,000 700 N >10,000
SCM-i cp sep 300 N 30 N N >20,000 N 150 100 300 N 5,000
SCM-9 gu sep 1000 N N N 300 5 >20,000
14 1,500 N N 1,000
SCM-b sl sep 30 N N >500 N 300 5 3,000 100 20 N >10,000
IIPA 14-11 si sep 50 3000 N >500 300 N 200 200 >1000
14
N >10,000
14-2 sl, gn, carb 200 2000 N 1000 N 1,000 N 10,000 500 20 N >10,000

lower Detection LImit 0.5 200 10 20 10 5 5 10 100 50


10 200

Analyses performed by Tom Hancock, Specomp Services, Boise, Idaho, 1982-1983.


1 Mineralogy in order of abundance excluding trace minerals; sep - mineral separate, si sphalerite, py = pyrite, yn =
galena, cp = chalcopyrite, tet tetrahedrite, mc marcasite, carb = carbonate
N not detected, < detected, but below limit of determination, > = greater than value shown.
Table 15. SEN (EDA) analyses of sphalerite, tetrahedrite-tennantite, stanniteO), and
unknown In mineral from vein-
type ore.

Sphalerite
Sample Zn Fe Cu Mn S Total
MB-9-81 70.99 0.76 1.50 0.00 26.74 99.99
SCM-4 71.73 .24 0.42 0.00 27.61 100.00
SCt4-4 69.90 1.11 0.44 0.06 27.89 100.00
SCN-4 69.49 1.22 1.56 0.00 27.73 100.00
SCN-4 71.51 0.66 0.38 0.00 27.46 100.01
SCM-4 69.87 1.48 0.46 0.00 28.19 100.00
Tetrahedri te-Tennanti te
Cu As Sb Ag Fe Zn Sn S Total
118-9-81 45 21 15 61 4 90 1 97 3 01 5 71 ND 23 58 99 99
118-9-81 43.11 5.78 13.71 3.78 2.21 7.15 ND 24.26 100.00
SCM-i 45.93 13.22 7.56 MD 0.92 9.35 0.47 22.55 100.00
Stannite-Stannoidite?
Cu Fe Zn Sn S Total
SCM-i 39.51 6.76 13.54 12.67 27.51 99.99
SCM-I 37.68 8.12 16.33 10.47 27.39 99.99
SCM-i 40.47 6.47 9.86 14.38 26.84 98.02
SCM-I 42.73 7.23 9.45 13.63 26.97 100.01
Unknown In Mineral
Cu In S Total
SCM-i 34.61 37.80 27.59 100.00
SCM-i 34.31 37.81 27.88 100.00

Analyses performed by Pete Romans, U S Bureau of Mines, Albany, Oregon, 1983


Note: Energy dispersive analyses were done with a standardless correction program
and recalculated to 100 percent.
Therefore, they are semiquantitative data
See Appendix, Table Al for sample location and description.

-4
0
Table 16. Microprobe analyses of sphalerite and pyrite in vein-type
ore.

Sphalerite
Sample1 Zn Fe Mn Cu Cd S.. Tota1

MB-1-81 61.47 4.81 0.05 < 32.69 99.00

MB-1-81 56.94 9.84 0.51 0.41 < 33.05 100.70

MB-12-81 62.50 3.28 < < 32.86 98.60

MB-12-81 65.61 0.67 < 32.68 98.90

MB-21-81 66.40 < 32.67 99.00

MB-21-81 66.39 0.42 < 32.83 99.60

MB-32-81 62.70 4.57 < 32.90 100.10

SCM-4 A 65.35 1.22 < 33.34 99.90

SCM-4 B 63.33 1.66 0.88 < 33.14 99.60

SCM-4 C 66.10 0.74 < 32.82 99.60

SCM-4 D 65.62 0.90 < 33.04 99.50

Pyrite

Fe Co Cu As Ni S Total

MB-1-81 46.03 < 53.69 99.70

MB-1-81 45.80 0.07 < 52.36 98.20

MB-12-81 45.97 < 53.29 99.20

MB-21-81 45.47 < 53.80 99.20

Analyses performed by the author under the supervision of Michael


Shaffer at the University of Oregon, Eugene, Oregon, 1983.
Detection limits are:
Sphalerite Fe = 0.05%,. Mn = 0.04%, Cu 0.11%, Cd = 0.18%.
Pyrite Co = 0.06%, Cu = 0.11%, Ni = 0.08%, As = 0.05%.
= less than the detection limit.
1 Letter designation in the sample number indicates the paragenetic
order in compositionally zoned crystals, with A representing the
earliest crysta1 see Appendix, Table Al for sample location and
description.
S determined by stoichiometry.
81

Table 17. Microprobe analyses of late-stage carbonate from vein-type


ore.

Sample1 CaO MgO FeO MnO CO2 Total

MB-12-81 A 0.80 2.69 55.06 2.91 39.09 100.54


MB-12-81 B 0.70 2.88 52.78 2.92 37.84 97.12
MB-12-81 C 0.89 4.76 49.14 4.05 38.51 97.34
MB-12-81 D 2.01 2.41 36.59 17.33 37.37 95.70
MB-21-81 A 0.29 0.17 47.76 13.79 38.22 100.23
MB-21-81 B 0.27 0.30 40.10 20.92 38.09 99.69
MB-21-81 C 0.33 0.68 38.23 21.18 37.56 97.98
MB-21-81 D 1.69 1.53 29.23 27.38 37.88 97.70
MB-32-81 A 0.19 0.31 39.49 20.64 37.48 98.12
MB-32-81 B 0.61 0.63 37.12 22.85 38.07 99.27
MB-54-81 0.31 2.06 41.82 16.96 38.64 99.79
MB-65-81 A 0.47 1.36 44.28 14.21 37.33 98.27
MB-65-81 B 0.32 1.58 44.17 13.07 37.07 96.13
MB-65-81 C 0.40 1.75 41.02 15.79 37.15 96.11

Average 0.65 1.81 42.25 15.52 37.96 98.19

Analyses performed by the author under the supervision of Michael


Shaeffer, at the Department of Geology, University of Oregon, Eugene
Oregon, 1983.
1 Letter designation in the sample number indicates the paragenetTc
order in compositionally zoned crystals, with A representing the
earliest crystals; see Appendix, Table Al for sample location and
description.

CO2 determined by stoichiometry.


Pyrite is by far the most abundant sulfide mineral in the veins.
It occurs in massive bands up to 3.0 m thick adjacent to the vein
walls (Pastor, 1970). These bands, where unbrecciated, often contain
vugs lined with cubic and pyritohedral crystals of pyrite up to 5 cm
in length. The amountof pyrite decreases laterally away from the
dike and phyllite sections of the San Cristobal Vein (Pastor, 1970).
Pyrite also is found as disseminations in the Chumpe Quartz Porphyry
and in the altered wallrock. Spectrographic analyses of two pyrite
concentrates (Table 14 samples MB-26-81 and SCM-2) reveal trace
amounts of nickel. Reported values for copper, lead, tungsten, and
zinc are high and invariably due to contamination. Sample MB-26-81,

located in a vein near the Carahuacra Manto, contains 1500 ppm


arsenic. This value may be the result of contamination or may possi-
bly indicate the presence of arsenopyrite. All microprobe analyses
for arsenic in veinpyrites were below the detection limit (Table
16). Lujan (1970) has reported arsenopyrite in the San Antonio
Mantos near the intersection of crosscutting veins, but this mineral
has not been found in the vein-type ore. Microprobe analyses of
four vein pyrites (Table 16) do not reveal any detectable amounts of
cobalt, copper, arsenic, or nickel except for sample MB-i-Si which
has 700 ppm cobalt. This sample is from a vein in the manto ores.
Pyrite in the manto ores does show high concentrations of cobalt, and
these will be discussed in a later section.
Quartz is ubiquitous in the veins and is associated primarily
with pyrite-wolframite mineralization in the Phyllite section. It Is
generally massive and milky in color, but clearer crystals sometimes
are found within vugs.
Wolframite is found as tabular crystals up to 6 cm in length,
generally enclosed in quartz. Its distribution has been directly
linked to proximity with the Chumpe Quartz Porphyry in the dike
section (Pastor, 1970). Linares (1973) reported 46 chemical analyses
of wolframite crystals from the San Cristobal and several other
veins. The content of MnWO4 was found to vary between 2 and 32
weight percent; this identifies the mineral as ferberite, the Fe-
rich wolframite. The ratio of MnO/FeO was found to increase with
depth.

Specular hematite is common in veinlets and mantos located


within the central portion of the district. It is not abundant in
the larger veins such as the San Cristobal or Andaychagua. Crystals
general ly are smal 1 (<1 mm) and occur as fel ty masses, lining vein
wall 5.
Chalcopyrite is the dominant copper-bearing mineral in the
veins. It is found as thin bands having a massive texture and having
been deposited paragenetically between early quartz and pyrite and
later sphalerite. According to Pastor (1970), the percentage of
chalcopyrite is greatest in the Contact section of the San Cristobal
Vein, and it diminishes rapidly to the west. Polished sections of
chalcopyrite reveal the presence of two exsolution(?) products or
intergrowths. Also, a band of complex mineralogy consisting predomi-
nantly of sulfosalt minerals was found between chalcopyrite and dark-
brown sphalerite. Identification of these mineral phases was diffi-
cult under reflected light. Accordingly, three samples were analyzed
spectrographically; two samples (MB-9-81 and SCM-4) were admixtures
chalcopyrite, sphalerite, and sulfosalts, and the third (SCM-9) was a
separate of pure chalcopyrite. These three samples showed anomalous
concentrations of bismuth (up to 500 ppm) and tin (greater than 1000
ppm) in addition to high silver, antimony and arsenic values (see
Table 14). EDA analyses confirmed the presence of argentiferous
bismuthinite (6 percent Ag). Figure iSa shows the lath-shaped bis-
muthinite crystals in chalcopyrite, and Figure 15b illustrates
another larger bismuthinite crystal extending out from chalcopyrite
into a zoned band of tetrahedrite/tennantite. The tetrahedrite
crystals in Figures 15b and 15c are generally zoned from an antimony-
rich core (lighter color) to an arsenic-rich rim (darker color),
- although some reversals in composition are visible. Compositional

data for tetrahedrite/tennantite are shown in Table 15.


Dark-colored (presumably Fe-rich) sphalerite was deposited after
the bulk of chalcopyrite mineralization. It exhibits exceedingly
fine compositional bands which are commonly interlayered with argent-
iferous galena and mixed sulfosalt bands. Emission spectrographic
analyses again revealed high concentrations of tin, silver, arsenic,
and antimony concentrations (Samples MB-9-81 and SCM-4, Table 14). A

polished thin section showing both light and dark colored bands of
Figure 15. Back-scattered electron photographs showing typical
textural relations between sulfosalt and major ore
minerals.

(a) Argentiferous bismuthinite needles (bright white)


in chalcopyrite (darkest gray). A thin layer of
tetrahedrite (light gray) separates chalcopyrite
from sphalerite (dark gray).

(b) Argentiferous bismuthinite needle (bright white)


crossing from chalcopyrite (dark gray) into a zoned
tetrahedrite crystal (light gray). Sphalerite
(medium gray) mantles the tetrahedrite. Black areas
are either holes in the polished section or crystals
or quartz.

(c) Tetrahedrite crystal (light gray) containing small


needles of argentiferous bismuthinite (bright white
in core) surrounded by sphalerite (dark gray).

(d) Two mixed sulfosalt bands consisting of tetrahedrite


(light gray) in sphalerite (dark gray). Bright
white areas are galena and dark black regions are
holes in the polished section.
rA :

J
I

C1500P

I
: '. ;:
:'%'
"4.

Figure 15
87

sphalerite was analyzed with the SEM (Table 15) to determine the
compositional range of the bands. Although iron is believed to be
the suspected pigmenting agent, the range of values was low, between
0.24 and 1.71 weight percent Fe. However, other microprobe analyses
document iron contents as high as 9.84 weight percent (Table 16,
sample MB-i-81) in the darker colored sphalerite. This same sample

also contains 0.51 weight percent manganese and 0.41 weight percent
copper. Previous studies by Pastor (1970) have noted the presence of
chalcopyrite exsolution(?) blebs in much of the darker colored sphal-
en te. The col or and composi tion of sphal en te changes with para-
genesis from a dark, more Fe and Cu-rich variety to a clear or ruby
colored Fe-poor variety, although numerous reversals in composition
were noted. Therefore, copper in solid solution or minute exsolution
blebs of other minerals may also contribute to the rhythmic color
banding. Sample (SCM-i) with bands of chalcopyrite, sphalenite, and
sulfosalts was analyzed by EDA to identify the origin of the anoma-
lous tin values. The sulfosalt bands seemed the most likely source
for the anomalies, and subsequent analyses found them to contain not
only tin, silver, antimony, arsenic, copper, and zinc, but also
indium as well. This sample is illustrated in Figure 15d, and the

mineral bands are clearly displayed at X80 magnification. A small


crystal of tetrahednite/tennantite was further examined at Xi000
magnification (Figure 16a). A series of backscatter elemental photo-

graphs are given in Figure 16 b-f that show the distribution of


specific elements in this sample. The tetrahedrite crystal in the
Figure 16. Back-scattered electron photograph and elemental
dot map images of a tetrahedrite crystal.

(a) A small crystal of tetrahedrite (light gray) from


outlined area in Figure 15d. Sphalerite (dark gray) at
top, stannite(?)(medium dark gray) and an unknown indium
mineral (lighter gray). Elemental image dot maps are
given for (b) zinc, (c) copper, (d) antimony, (e) tin,
and (f) indium.
.i -1 .'.
-
'L
J r ?
-
t' ?- .
.
-. I
.. 3
'..d
I
I .
'. ,',

t,
r,
,
p
f
l 7
C
'C ,
_________ ,
,.
.7!
I
:
b
, d .3
C ' \.
t,-z .'
-
f 8) .. 1 4
1 C C' . .
3. .
. p
.
p.
4II
______ '3 ,
I.
;
.1 :t43
'
, .
'...
,-.,
.
1
.U-..3iC-I3
2ti. -

;, :3
-'
4 r
4_ _\ ,Lr
S
'-,:;)
' 4
. 1,
I
I
I 4,
t
'
-
p!
;;'
3
53 .t.. L4
t
4
,- 3 -
7 .'-3 ..-,
I
7.. -
C '. -
#
-
.3 '
:4 -
i'.__
7. c.
Il
2
. .t
center of Figure 16a is coated and partly replaced by a Cu-Sn min-
eral. Furthermore, a Cu-In mineral appears to have exsolved from, or
coated the Cu-Sn mineral. Compositional analyses of the Cu-Sn min-
eral are shown in Table 15, and are consistent with the formula
Cu8(Fe,Zn)3SnS12, known as the minimal stannoidite. However, a

Zn/(Zn+Fe) ratio of 2/1 is too high for stannoidite according to


Yamanaha and Kato (1976). Additionally, the amounts of S and Sn are
lower than that required for the stannite-kesterite (Cu2FeSnS4-
Cu2ZnSnS4) solid solution series.
Although EDA analyses of the Cu-In mineral fits the formula
(Cu,In)S, a mineral name with this composition could not be found.
Roquesite (CuInS2) has an atomic ratio of Cu/In equal to 1/1 while
the mineral in this sample has a Cu/In ratio of 2/i. It is possible
that the SEM (EDA) analyses for S and Sn are low for both the Cu-Sn
and Cu-In minerals, but the tetrahedrite analyses are exactly what
they should be. The mineralogy of these two phases may be compli-
cated by further intergrowths or exsolution, or there could be two
new minerals present in this sample.
A crystal of galena in chalcopyrite from this sample (SCM-i) was
also analyzed. The crystal contains at least four distinct mineral
phases which apparently resul ted from exsolution. These include
argentiferous galena, which has exsolved needles composed of a Pb-Bi-
Ag sulfide, probably benjaminite (P. Romans, written comm., 1983),
and a matrix that contains two phases; one that is sil ver-rich and
another that is lead-poor. The silver-rich phase contains >50 weight
91

percent Ag, probably argentite, and the lead-poor phase is possibly a


Ag-Bi-Pb sulfide.
This cursory examination of several ore samples has revealed
heretofore unknown mineralogical complexities in ores of the San
Cristobal veins. Further chemical and mineralogical investigations
of these phases utilizing the electron microprobe will be continued
as time and facilities permit in the future.
Gal ena is found typically as bands mixed with the dark sphal -
erite and as wel 1-formed cubes in the paragenetical ly later ores.
Pyrargyrite and tetrahedrite are present as exsolution intergrowths
in galena. Pyrargyrite forms small rod-shaped exsolution blebs, many
of which show brilliant red internal reflections under reflected
light. Tetrahedri te, with or without chalcopyri te, is al so found as
larger irregularly-shaped blebs. Spectrographic analyses of galena
separates (Table 14) show high values of silver (up to 1000 ppm),
that are probably attributable to tiny inclusions of pyrargyrite.
According to Pastor (1970), both bornite and covellite are found
as hypogene minerals associated with chalcopyrite and sulfosalt min-
eralization, but they are not important contributers to copper pro-
duction at the mine. One sample from the Triunfo mine (east end of
the San Cristobal Vein) contained massive tetrahedrite intergrown
with bornite, chalcopyrite, covellite, and acanthite. This assem-

blage appears to be the result of exsolution, but a supergene origin


cannot be rejected on the basis of available mineralogical or tex-
tural evidence.
92

Manganosiderite is a major gangue mineral found in virtually all


veins. It generally has a colloform habit and is zoned in color and
composition from brownish-black siderite to white and pink rhodo-
chrosite (Fig. 17). However, a rhombohedral habit was observed in
samples from some veins and vug-like fillings. Microprobe analyses
were obtained for 15 samples of late-stage carbonate (Table 17). In

all cases where paragenesis could be determined accurately lie, zoned


crystals or growth bands) the earliest carbonatewas usually a sid-
erite containing minor amounts of manganese. Manganese contents vary
from 2.91 up to 27.38 weight percent, and the average of all car-
bonate analyses provides a composition of 42.25 percent FeO and 15.52
percent MnO. Values for MgO and CaO were less than 2.0 and 1.0
weight percent, respectively. Although rhodochrosite is not found in
the San Cristobal Vein proper, and that from the Andaychagua Vein to
the south was not analyzed, it is likely that a complete continuum
of compositions from FeCO3 to MnCO3 is present in the district.
Samples of pure rhodochrosite from the manto ore deposits were
analyzed. Much of the late-stage carbonate is botryoidal and con-
tains layers of barite, sphalerite, breccia fragments, and marcasite.
Barite crystals up to 5 cm in length are oriented perpendicular to
growth zones in the carbonate. This sulfate also is found as larger
tabular crystals up to 12 cm in length in open vugs where carbonate
was not deposited. Small cockscomb crystals of marcasite coat open
vugs.
93

Figure 17. Specimen of colloform siderite containing a layer of


colloform sphalerite (dark black) from late-stage
mineralization of the San Cristobal Vein. Note tabular
cyrstals of barite oriented perpendicular to growth
zones.
94

Masses of colloform sphalerite commonly are found with the


massive vein carbonate. In transmitted light, these bands exhibit
layer of both isotropic and birefringent minerals that suggest an
intergrowth of sphalerite and wurtzite forming a schalenbiende tex-
ture. Nonetheless, X-ray diffraction analysis failed to reveal the
presence of wurtzite.
Chalcocite, native silver and native copper are found only on
the upper levels of the mine workings. The native metals and sulfide
formed by supergene oxidation and replacement, respectively, of the
primary hypogene sul f ides.
Paragenesis

The paragenetic sequence of mineral deposition for the vein ores


has been wel 1 documented by Pastor (1970) and Vera (1977). Further
work by Campbell and Rye (1982) and this author generally support the
previous investigations and serve to further refine the paragenetic
sequence. To summarize, Pastor (1970) divided the paragenetic his-
tory of mineralization into three periods of mineral deposition based
on both structural disruption of the ores and on seemingly abrupt
changes in the character of mineralization. The paragenetic sequence
for the veins in the San Cristobal District is graphically portrayed
in Figure 18.
The first period of mineralization includes the assemblage py-
rite-quartz-wolframite + specular hematite. Pyrite was the first
mineral to be deposited and it is by far the most common sulfide in
the veins. Much of the pyrite appears to have partially replaced
TIME 4
T°C 300 200 150

Mineral First Second Third Supergene

specu'ar hematite -
pyrite
i
I
I
woframIte
-
quartz
I--- I
chalcopyrite - - - -
sphalerfte
Ferich Fepoor
I
tetrahedrite-
tennantite - - - -
bornite I

stannite
(In,Cu)S
bismuthinite
galena I

polybasite-
pyrargyrite
I

manganosiderite
barite
I ----I
marcasite
coveflite
I
I
chalcocite
native Ag
I
native Cu

Brecciatjon1
Figure 16. Paragenetc sequence of mineral deposition in vein-type ores (thickness of lines Is proportional to mineral
abundance).
fragments of phyllite and walirock in the Phyllite section. In
contrast, it was deposited as open space fillings, In the Volcanic
section. Brecciatfon of the pyrite band occurred prior to the de-
position of quartz and woiframite. Symetrical bands of quartz and
wolframite surround fragments of pyrite and fill fractures in the
pyrite band. Minor amounts of pyrite continued to be deposited with
the later quartz and wolframite. Increasingly larger amounts of
quartz were deposited toward the end of the first period of minerali-
zation.
Pastor (1970) does not mention specular hematite as a common
mineral in the San Cristobal Vein. Although it is found in minor
quantities in many of the large veins, it is an abundant constituent
of smaller veins and veinlets in the central graben structure. In

the San Cristobal veins, specular hematite is associated with pyrite


and quartz as an apparently early-formed mineral. Some veinlets
exhibit a lining of specular hematite adjacent to the walirocks, with
pyrite, quartz, chalcopyrite and sphalerite filling the central
areas. The paucity of specular hematite in the large veins may
possibly be. explained through differential sulfidization by the
fluids which deposited pyrite. Reactions of reduced, sulfur-rich
fluids on specular hematite may have converted the Fe-oxide to pyrite
by a process similar to that described by Naldret and Kullerud
(1965). By virtue of their size and mineralogy, the larger veins
allowed more fluid flow and remained open longer, thereby possibly
penitting the complete conversion of hematite to pyrite. The smal-
97

ler veins commonly show only one period of mineralization. Pyrite


has been found replacing specular hematite in the manto ores. A more
thorough discussion of this problem and its bearing on the chemistry
of the ore fluids will be presented later.
The second period of mineral deposition is responsible for most
of the economic mineralization. Chalcopyrite, sphalerite, galena,
and tetrahedrite with trace quantities of the other minerals men-
tioned previously, were deposited during this interval. Again, the

beginning of the second period of mineralization is marked by the


brecciation of previously formed minerals. Pyrite was replaced by
chalcopyrite along fractures, and quartz accompanied the chalco-
pyrite. The major episode of chalcopyrite deposition occurred early,
although minor amounts are found cutting all minerals of the second
period. The change from predominately chalcopyrite mineralization to
sphalerite is marked by bands of tetrahedrite/tennantite mixed with
minor amounts of bismuthinite and other sulfosalts. Thereafter,
minor bands consisting of sul fosal ts and other copper-bearing min-
erals were deposited with the sphalerite and galena. Exsolution(?)
blebs of chalcopyrite commonly are seen in dark sphalerite. The

color and composition of sphalerite changes with time from an early


dark brown Fe-rich variety to a light, transl ucent red and Fe-poor
type. Argentiferous galena was deposited in bands with the early
sphalerite; it becomes the dominant phase toward the end of the
second period of mineralization.
Renewed brecciation and precipitation of barite, manganosider-
ite, and marcasite mark the beginning of the third mineral izatlon
period (Pastor, 1970). Colloform (botryoldal) manganosiderite con-
taining some rhythmic bands of colloform sphalerite are found filling
voids and fractures (see Fig. 17). Small breccia fragments contain-
ing sphalerite, galena and chalcopyrite are present in the carbonate.
Layering in the manganosiderite varies considerably in both color and
composition. Barite was deposited contemporaneously with carbonate.
Small amounts of marcasite coat all exposed surfaces and appear to
be the last mineral precipitated in the veins.
Lateral and Vertical Mineral Zonation

Pyrite was deposited in the first phase of mineralization and it


is the most abundant sulfide in veins of the district (Pastor, 1970).
It is distributed along the entire length of the San Cristobal Vein,
although its abundance decreases away from the Dike and Phyllite
sections, but there is no apparent change in the proportion of pyrite
with depth. The other veins always contain some pyrite, but the
abundances again decrease with distance from the Chumpe Quartz Por-
phyry.
Wolframite is associated predominately with pyrite and quartz in
the earliest phase of ore deposition (Pastor, 1970). Its distribu-
tion appears to be related to proximity to the Chumpe Quartz Porphyry
dikes and the concentration of wolframite decreases rapidly away from
these intrusives. This implies a thermal or chemical control on
deposition of the woiframite. Sporadic occurrences of woiframite are
found elsewhere, but only where the veins are hosted in Excelsior
Phyllite. Abundances of wolframite as presently exposed in the mines
appears to decrease with depth, but drilling west of the Chumpe
Quartz Porphyry dike on the 500 level may find a continuation of
tungsten mineralization at depth. Wolframite is found in other
veins, but only within the phyllite and/or in proximity to the dikes.
Chalcopyrite was deposited early in the second phase of mineral-
ization. It is present in most veins located within the central
graben structure. Copper values gradually diminish laterally outward
toward the contact between the Pucara Group and Catalina Volcanics.
Two areas of high chalcopyrite concentration in the San Cristobal
Vein have been described by Pastor (1970). The first is located
within the Phyllite section below the 320 level just west of the dike
section. The other is located in the Contact section. In general,
the chalcopyrite content of the veins decreases north and south of
the central graben structure.
Sphalerite is found throughout the length of the veins. How-

ever, its abundance in the San Cristobal Vein increases away from the
Dike and Phyllite sections. It is the dominant ore mineral in the
Volcanic section. Galena, although associated with sphalerite, has
its highest concentrations within the Dike and Phyllite section.
Both sphalerite and galena increase to the north and south of the
graben structure.
100

To summarize, the veins exhibit a pattern of district-wide


mineral zonation, similar to that found in other large hydrothermally
mineralized districts, and such is the case in the Morococha District
nearby to the north. Woiframite, pyrite and chalcopyriteare as-
sociated within a central region, centered on the Chumpe Quartz
Porphyry dikes. Their abundances decrease away from this central
zone where sphalerite and galena replace them as the major ore com-
ponents of the veins.
Metal-MetalAssociations

Twelve samples from the San Cristobal Vein were assayed for
copper, lead, zinc, and silver (Table 18). These values were then
plotted on log-log probability graphs and correlation coefficients
were calculated. This method was chosen to help reveal the presence
of linear and curvilinear trends between the metals. The correlation
matrix for all combinations of metals is given in Table 19. Corre-
lation coefficients were calculated on the base 10 logarithm of the
assay values in ppm. Although an insufficient number of samples were
assayed for a good statistical comparison, the data reveal some
potentially interesting correlations.
The correlation coefficient for lead to silver is 0.4494, which
indicates a moderately good covariant association of these metals,
either in solid solution or as a separate mineral phases. This
relationship is consistent with the petrographic data, which shows
small amounts of pyrargyrite exsolved from galena. The low coeffi-
cient of 0.0698 for copper versus silver, and 0.0794 for zinc versus
101

Table 18. Assays of copper, lead, zinc, and silver for selected
samples of vein-type ore.
0/
/0

Sample Ag(oz/t) Cu Pb Zn

MB-1-81 5.78 0.04 6.00 18.00


MB-2-81 37.00 0.11 24.00 11.60
MB-3-81 8.38 0.26 0.51 12.70
MB-9-81 11.30 14.20 0.07 14.80

MB-10-81 8.18 0.43 2.00 1.92

MB-12-81 1.17 0.05 0.33 30.10


MB-21-81 5.26 8.20 1.88 1.20

MB-32-81 6.69 0.06 7.20 22.10

MB-33-81 8.21 0.09 5.20 37.20

MB-34-81 16.20 0.03 18.00 6.30

MB-47-81 14.30 1.38 0.60 9.40

MB-59-81 1.58 1.58 0.46 2.90

Assays determined by Centromin Peru, 1981.


See Appendix, Table Al for sample location and description.
102

Table 19. Correlation matrix for copper, lead, zinc, and silver
assays from vein-type ore.

Cu 1.00
Pb -0.6445 1.00
Zn -0.4739 0.0271 1.00

Ag 0.0698 0.4494 0.0794 1.00


Cu Pb Zn Ag
103

silver, suggests that although large quantities of tetrahedrite are

associated with both copper and zinc mineralization, it is not highly

argen ti ferous.

Figures 19 and 20 illustrate the distribution of lead and zinc

relative to copper. High negative coefficients of -0.6445 and

-0.4739 respectively indicate a decrease in copper concentration with

increasing lead and zinc. This is expected both from paragenetic

studies, and from the observed regional zonation of minerals in the

veins.

The correlation between zinc and lead is near zero (0.0271)

which implies a non-linear trend between these two metals. However,

Pastor (1970) has demonstrated that values of lead decrease as those

for zinc increase within the San Cristobal Vein. Insufficient sample

representation is probably the cause of this apparent lack of corre-

1 a.ti on.

Al teration

Alteration of the wall rocks adjacent to the San Cristobal Vein

has been adequately discussed by Pastor (1970) and in this thesis

under Hydrothermal Al teration Zones. In summary, there are three

distinct zones of alteration developed outward from the vein into

wallrocks of the Excelsior Phyllite and two into the Catalina

Volcanics. From the vein outward, these are (1) the silicified zone

(2) the sericitic-argillic zone, and (3) the chioritic zone. Only

the latter two are developed in the Catalina Volcanics.


.10E407

.IOE+0E

10000.00

1000.00

100.00
IOC
LEAD (PPM)

Figure 19. Log-tog distribution of lead versus copper concentrations in vein-type ore.
.10E407

.10E406

10000.00

1000.00

100.00

ZINC (PPM)

FIgure 20. Log-log distribution of zinc versus copper concentrations in vein-type ore.
106

The silicified zone is approximately two meters thick and con-


tains abundant secondary quartz with minor sericite and disseminated
pyrite. Siliciflcation decreases westward from the phyllite section.
The sericitic-argillic zone is between three and five meters wide,
and resulted in the conversion of primary clay minerals into sericite
and kaolinite. Sericitization also decreases westward into the
Catalina Volcanics, where it attains a thickness of only two to three
meters. Alteration of primary plagioclase feldspar in the volcanics
to sericite and kaolinite is complete over this interval. The

chioritic zone is between 7 and 14 meters wide in the phyllite, and


up to 47 meters wide in the volcanics. Chlorite is the most abundant
mineral in the phyllite, whereas the volcanic rocks contain calcite
and epidote in addition to chlorite. This assemblage would be better
termed propylitic. Although Pastor (1970) indicates a maximum thick-
ness of 47 meters for the propylitic zone adjacent to the San
Cristobal Vein, it is apparent that the Catalina Volcanics are propy-
litically altered throughout the district.
Pastor (1970) has suggested that much of the calcium, iron, and
magnesium leached from the wall rocks was fixed as carbonate through
the addition of carbon dioxide from the mineralizing fluids. A

considerable portion of the cations leached from these rocks, as well


as the Chumpe Quartz Porphyry, also was carried outward from the
thermal source to the Pucara limestone, where they reacted with the
carbonate and were fixed as compositionally variable manganosiderite.
107

Oxidation and Secondary Enrichment

Some secondary enrichment of copper and silver is evident in the


veins, although glaciation has removed any significant zones of
enrichment. Chalcocite, covelifte, argentite, and native silver are
the most coninon secondary minerals encountered in the upper oxidized
portions of the veins.
According to Pastor (1970), chalcocite forms coatings on sphal-
erite and chaicopyrite above the 270 level. Native silver, although
not abundant in the San Cristobal Vein, is found in veiniets, pock-
ets, and as coatings on hypogene sulfide minerals and in chalcocite.
Large quantities of native silver were mined in the early stages of
development of the Andaychagua Vein. Covellite is found as thin
replacement rinds on chalcopyrite in many of the old surficial vein
prospec ts.

Man to-Type Ore

The mantos occur within the basal 1 imestones of the Triassic-


Jurassic Pucara Group. Al though they are most common within 75 to
100 m of the underlying contact with the Mitu Group, there are a
number of small deposits distributed erratically throughout the for-
mation, especially in the Carahuacra area. These ore bodies general-
ly conform to the definition of a manto, but many show pipe-like
features, particularly near their intersections with veins. The man-

tos are irregularly shaped and highly controlled by local structures,


but most are generally thin and tabular. Similar smaller deposits in
Peru have been described by Johnson and others (1955) in the Atacocha

District and by Bodenlos and Ericksen (1955) in the Cordillera

Blanca.

The Carahuacra Manto is located in the Carahuacra Valley im-

mediately north of the San Cristobal Mine (see Fig. 11, p. 59). It

is comprised of several manto zones which have been named the

Carahuacra, Santa Aqueda, and San Antonio Mantos from north to south,

respectively. These three manto zones will be collectively referred

to here as the Carahuacra Manto. Centromin Peru operates the

Huaripampa Mine which is located in the southern portion of the

Carahuacra Manto, whereas Volcan Mines Company operates a large open

pit and underground mine in the northern half of this manto. Two

additional manto outcrops are located south of the San Cristobal Mine

and have been named the Toldorrumi and Moises Mantos, respectively

(Fig. 11).

Each manto zone is composed of one or more, 1 to 2 m thick ore

deposits of variable lateral dimensions (true mantos) which are

separated from the next overlying manto by a thin (0.5-1.5 m) bed of

volcaniclastic tuff or tuffaceous limestone. These tuffs generally

are altered to a kaolinitic clay, whereas the unmineralized tuffa-

ceous limestones are altered to kaolinite and manganosiderite.

Mineralization in the Carahuacra Manto extends discontinuously

within the Pucara Group for a strike length of nearly 3000 m, and

down dip to a known depth of 320 m. The stratigraphic thickness of

the manto zone varies from 5 to 100 m. Manto zones are roughly
109

elliptical in cross-section but may become pipe-like near large


cross-cutting veins.
Little development, other than a few diamond drill holes, has
been completed on the Toldorrumi and Moises Mantos. At Toldorrumi
Manto two veins intersect the limestone. Both were tested by small
drifts into both vein and manto-1'pe mineralization, but the metal
concentrations proved to be subeconomic. In the Moises Manto, an old
working named the Santa Rosa adit contained a 20 m length of
"bedding plane vein" (manto) replacement mineralization located ap-
proximately 20 m above the contact with the Catalina Volcanics.
Metal values of 10 ounces silver, 0.4 percent lead, 5.1 percent zinc,
8.1 percent manganese and 0.01 ounces gold were encountered over a
0.5 m thickness (Laverty, 1949). However, grade and tonnage of this
tnanto is presently subeconomic.
The following discussion of mineralogy and paragenesis was based
on numerous polished sections of hypogene ore collected from within
the Carahuacra Manto at the Huaripampa Mine. Surficial examination
and mapping in the Carahuacra, Toldorrumi, and Moises Mantos contri-
buted to the overall understanding of alteration, structural control,
and ore deposition. On the basis of my observations, all three manto
zones appear to have formed contemporaneously by a geologically
common process. Thus, the following discussion and facts apply
collectively to all manto deposits within the study area.
110

S truc tural and Li thol ogic Controls

Structural controls of ore deposition include (1) well-developed


transverse faults, (2) closely spaced dip joints, (3) breccias de-
veloped along transverse and bedding-plane faults, (4) impermeable
tuffaceous beds, (5) primary carbonate porosity, and (6) secondary
porosity and permeability developed by hydrothermal solution and
recrystallization of the limestones. The most obvious lithologic
control is, of course, the chemical composition of the carbonate host
rock.
The development of faul ts, breccias, and dip joints has been
discussed previously. Here, the principal consideration is how these
structures relate to the migration of the hydrothermal fluids from
their source to the Pucara Group limestones and to the localization
of ore. Discussion relevant to the extent of alteration and min-
eralization adjacent to these structures is provided in the following
chapter.
With few exceptions, mantos are found exclusively adjacent to
and/or extending from structural dislocations that also contain vein-
type mineralization. Wumerous examples of this relationship can be
found throughout the district, but they are more readily discerned
where mineralization and alteration are least intense. Both the
Toldorrumi and Moises Mantos contain excellent exposures which are
otherwise obscured by mining activities and alluvium in the

Carahuacra Valley. Furthermore, surficial outcrops containing man-


ganosiderite have been oxidized to a black ferromanganese gossan,
111

which permits one to readily distinguish between mineralized-altered


and barren-unaltered carbonate host rocks, respectively.
Examination of Figure 11 (P. 59) documents the spatial proximity
of the man tos to major transverse veins. In general, the size of a
manto varies proportionally to the size or frequency of nearby veins
in the limestone host. In the Carahuacra area, a number of large
veins that belong to the Virginia Vein system have produced an ex-
ceptionally large manto zone. Important also in this example, is the
fact that brecciation was intense and produced a zone of high porosi-
1' and permeability. In the Toldorrumi Manto, the Catalina, Polonia,
and a strand of the Prosperidad Vein enter the Pucara Group. The

limestone here is not associated with laterally extensive breccias,


so mineralization was primarily confined to the cross-cutting vein
structures. The fracture control of mineralization and alteration in
this manto is illustrated in Figure 21.
Farther to the south, the size of veins and mantos decreases, as
does the density of mineral ized veins per unitarea. Although not
evident at the scale of Figure 11, smaller vein structures are re-
sponsible for individual mantos in the Moises area. One such vein
crossing the Catalina Volcanics and entering the Pucara Group is
shown in Figure 22. This vein maintains its attitude upon entering
the limestones, but the displacement decreases rapidly. Figure 23
shows the vein with its attendant zone of alteration in the basal
part of the Pucara Group.
112

1
, \

figure 21. Outcrop of altered limestone showing three vertical


veins (splits of the Polvarin Vein, black) cutting
perpendicular to bedding. A]teration of the lime-
stone to manganosiderite is complete. Note hammer
for scale.
113

. :

:'
-
5k.ii S'

- -

-. _'_I
Figure 22. Outcrop of a small vein in the Catalina Volcanics.
Small adits in the background are located in the
basal Pucara Group.
114

'
" '\

Figure 23. Outcrop of the vein in Figure 22 at its intersection


with the Pucara Group limestones. Vein cuts perpen-
dicular to bedding. Note black manganese oxide gossan
in vein.
115

Dip joints played a key role in allowing the ore-bearing solu-


tions to migrate outward from major structures. Each joint became a
small veinlet with a proportionally smaller alteration zone. The

'pical densil' of these joints as well as their irregular nature is


illustrated in Figure 24.
The intersection of major veins with bedding-plane breccias
produced the laterally extensive mantos, which represent about 80
percent of the mineralization in the Pucara Group. The remaining 20
percent is restricted to the transverse veins. Obviously, a breccia
represents a zone of high permeability that allows the fluids to
chemically react with a substantially larger surface area. The

relationship between transverse veins and laterally extensive brec-


cias is well displayed in the Moises Manto. Here the most extensive
manto is located approximately 70-80 m above the Catalina Volcanics-
Pucara Group contact. Continuations of the Andaychagua Vein and
several smaller veins cut through the Pucara Group for 80 m without
producing a manto-like body, although small, irregular pipe-like
bodies are present adjacent to the veins, until they reached a two
meter thick zone of breccia. The fluids then migrated laterally
for several hundred meters along the breccia, replacing the breccia
fragments and portions of the adjacent non-brecciated limestone.
Al teration of the limestone beds on either side of the breccia was
confined to less than 1 m except where joints were present. Also,
volcaniclastic tuff layers located above brecciated limestone beds
helped channel the fluids laterally by acting as an impermeable
116

Figure 24. Limestone bed cut by numerous veinlets containing


manganese oxide. Note bedding is parallel to hammer
head and limestone is altered to reddish-brown
manganosi den te.
117

barrier to fluid movement (see Fig. 30 in the following chapter).


Porosity and permeability of the limestone appears to be im-
portant in the formation of the mantos, because all mineralized areas
are located adjacent to and in structures which increased these
factors. Primary porosity, although difficult to assess after min-
eralization, may have been important in the localization of some
mantos. Many of the limestone beds in the Pucara Group are oolitic.
Primary porosity in recently deposited ooliths Is about 20 percent,
but decreases significantly after compaction and diagenesis. Never-

theless, many of the altered carbonate rocks examined in thin section


show remanent oolitic textures. These oolite beds may have contri-
bute significantly to the lateral migration of fluids and to the
formation of the manto ores. This hypothesis, while not examined in
detail, should be investigated more fully in future studies of the
district.
Secondary porosity and permeability produced by alteration of
the limestone were important controls in the localization of mantos,
but to a lesser extent than structurally induced porosity. The

conversion of limestone to dolomite involves a 12 to 14 percent loss


of volume (more for manganosiderite), which may result in a propor-
tional increase in porosity (Lovering, 1969). The increase in poros-
ity during alterationwould allow the ore solutions access not only
to the wall rocks adjacent to breccias and other structural channel-
ways, but also to the breccia fragments themselves. Accordingly, the
surface area available for chemical alteration increases dramatical-
118

ly.
The most important lithologic control to ore deposition in the
mantos is the chemical composition of the host rock; specificall.y the
limestones. As has been shown by numerous studies involving ore
deposits hosted by sedimentary rocks, limestone is a particularly
favorable lithology for the localization of massive sulfide ore.
Furthermore, the dominant mechanism of ore deposition in hydrothermal
deposits hosted in carbonates is by replacement Limestone (CaCO3)

acts as a tremendous chemical buffer for acidic solutions. Burnham

(1967) has shown that hydrothermal fluids rich in H and HC1 will be
neutralized upon entering limestone, causing precipitation of sulfide
minerals. Selective replacement of the Pucara Group is observed
where ore-bearing fluids encountered both limestone and volcani-
clastic tuff beds. The limestone was selectively replaced both by
massive sulfides and gangue minerals, whereas the tuffs underwent
alteration to clay minerals with little replacement by sulfides.
Moreover, limestone beds that contained significant proportions of
tuffaceous sediment underwent differential replacement causing segre-
gation of the tuffaceous component into clay pods interstitial to the
massive sulfide replacement ore.

Mineralogy and Trace Metal Geochemistry

The mineralogy and paragenesis of the manto ore bodies has been
discussed by Lyons (1968), Lujan (1970), and in several unpublished
reports for Cerro de Pasco Corporation. The principle minerals of
119

economic interest are sphalerite and galena. Pyrite, marcasite,


manganosideri te, quartz, ban te, specul ar hemati te, magneti to, rhodo-
chrosite, and kaolinite are the dominant gangue minerals. Present in
minor and microscopic amounts are chalcopyrite, tetrahedrite-
tennantite, pyrargyrite-proustite, pyrrhotite, and native silver.
Otherminerals foundby previousworkers, butnotconfirmed in the
present study, include jamesonite, stibnite, gypsum, argentite, and
arsenopyri te.
The following discussion is primarily concerned with ore tex-
tures, para genesis, and trace metal geochemistry. A sunrnary of both

Lyons (1968) and Lujan's (1970) work is included, but the interpreta-
tion of the paragenesis and geochemistry differs considerably. Trace
metal distributions were determined through 10 emmission speCtro-
graphic analyses both on whole rock ores and sulfide mineral
separates. Specific major and minor elemental trends in ore and
gangue minerals were determined through 52 rnicroprobe analyses on 8
samples representative of the major ore types. These analyses are
suninarized in Tables 20-23.
A general discussion of the spectrographic analyses given in
Table 20 is warranted to provide an overview of the trace metal
distributions. Silver values range from 20 to 3000 ppm, equivalent
to an average of about 3 to 4 ounces per ton of ore. The extreme

value of 3000 ppm silver reflects both high concentrations of silver-


bearing sulfosalts (polybasite) and secondary enrichment through
supergene processes. Arsenic concentrations range from nil to 7000
Table 20. Emission spectrographic analyses for selected elements in manto-type suifides.

ppm

Sample Mineralogy1 Ag As Bi Cd Co Cu Mn Ni Pb Sb Sn Zn

MB-6-81 py, sl, carb 70 7,000 N 150 15 500 1500 30 2,000 100 70 >10,000

MB-18-81 py, sl 70 1,500 N 500 N 700 700 N 1,500 N 100 >10,000


MB-29-81 si, py, mc 50 200 N 500 N 300 5000 5 500 500 10 >10,000

MB-38-81 si, gn, py, 3000 <200 N 500 N 1500 >5000 5 20,000 7000 N >10,000
rhodo
MB-77-81 si, carb 100 N N >500 N 700 5000 5 500 1000 N >10,000

HPA M-4 sl, gn, carb 30 N N >500 N 30 5000 5 15,000 N N >10,000

HPA M-6 sl, py 20 N 50 500 N 500 1000 5 2,000 N N >10,000

HPA M-10 si, py 20 3000 N 300 N 3000 200 15 500 N 300 >10,000
M-5 si, carb 30 N N >500 N 200 1500 N 200 N 200 >10,000
M-5 sl sep. 30 N N >500 N 500 700 N 1,500 N 300 >10,000

Lower Detection Limit 0.5 200 10 20 10 10 5 5 10 100 10 100

Analyses performed by Tom Hancock, Specomp Services, Boise, Idaho, 1982-1983.


1 Mineralogy in order of abundance, excluding trace minerals; sep = mineral separate, py = pyrite,
Si = sphalerite, carb = carbonate, mc = marcasite, gn = galena, rhodo = rhodochrosite, N = not de-
detected,< = detected but below lower limit of determination.
See Appendix, Table Al for sample location and description
121

Table 21. Microprobe analyses of pyrite and marcasite in manto-type


ore.

wt. %
Sample Fe Co Cu Ni As S Total

MB-18-81 46.08 < 52.88 98.90

MB-18-81 45.72 < 53.39 99.10

MB-18-81 46.50 < 52.97 99.40

MB-29-81 47.28 0.15 0.07 53.73 101.20

MB-29-81 46.43 0.08 0.09 53.01 99.60

MB-29-81 46.11 0.06 53.51 99.60

MB-38-81 46.80 0.09 < < 53.34 100.20

MB-74-81 46.15 < 0.16 53.78 100.00

HPA M-11 46.92 0.08 0.36 53.00° 100.30

HPA M-11 47.25 0.10 0.75 52.90 101.00

HPA M-11 45.53 0.10 1.39 52.35 99.30

M-4 46.22 0.09 53.76 100.00

M-4 44.30 0.08 < 54.00 98.30

Analyses performed by author under the supervision of Michael


Shaffer at the University of Oregon, Eugene, 1983.
Detection limits are :Co = 0.06%, Cu = 0.11%, Ni = 0.08%, As 0.05%.
= less than the detection limit.
Note: analyses are for the iron disulfide, marcasite, In sample MB-
29-81; sulfur calculated by stoichiometry.
122

Table 22. Microprobe analyses of sphalerite in manto-type ore.

wt. %
Sample Zn Fe Mn Cu Cd S Total

MB-18-81 56.89 10.24 < 32.78 99.90


MB-18-81 56.50 10.64 0.17 0.61 33.68 101.60
MB-29-81 A 64.51 2.39 < 32.51 99.40
MB-29-81 B 66.31 0.12 32.90 99.30
MB-29-81 C 65.98 0.12 < 0.52 32.78 99.30
MB-29-81 D 66.52 0.24 < 32.27 99.00
MB-29-81 E 66.26 0.23 < 0.24 32.73 99.40
MB-29-81 F 65.92 0.61 < 32.61 99.10
MB-29-81 G 65.08 1.70 < 32.27 99.00
MB-29-81 H 66.80 0.13 < 32.41 99.30
MB-29-81 1 63.40 4.38 < 33.01 100.70
MB-38-81 64.93 1.17 0.09 33.30 99.40
MB-38-81 65.52 1.10 0.08 32.96 99.60
MB-74-81 66.50 0.37 0.16 32.87 99.80
MB-74-81 66.98 0.24 < 32.64 99.80
MB-74-81 65.50 0.52 0.76 32.66 99.40
MB-74-81 66.59 0.27 0.19 32.75 99.80
MB-77-81 66.57 0.33 32.74 99.60
MB-77-81 65.24 0.18 0.58 0.27 32.44 98.70
MB-77-81 67.01 0.85 32.67 100.50
MB-77-81 64.94 0.27 0.42 31.94 97.50
HPA M-11 A 64.63 3.25 0.13 32.85 100.80
HPA M-11 B 61.32 5.95 0.32 33.05 100.60
HPA M-11 C 64.21 2.07 0.14 32.57 98.90
HPA M-11 D 64.66 2.40 0.13 32.90 100.10
HPA M-11 E 61.48 6.24 32.77 100.40
HPA M-11 F 62.18 4.28 0.80 32.29 99.50
HPA M-11 G 63.93 3.84 32.75 100.50
M-4 65.19 1.08 32.73 99.00
M-4 64.87 2.09 32.87 99.80
M-4 61.22 7.10 0.39 32.80 101.50

Analyses performed by the author under the supervision of Michael


Shaffer at the University of Oregon, Eugene, 1983.
Detection limits are: Fe = 0.05%, Mn = 0.04%, Cu = 0.11%, Cd = 0.18%.
= less than the detection limIt.
Note: letter designation in sample number indicates paragenetic order
in zoned crystals, with A earliest; sulfur calculated by stoichiometry.
123

Table 23. Microprobe analyses of late-stage carbonates in manto-type


ore.

wt. %
Sample CaO MgO FeO MnO CO2 Total

MB-20-81 0.32 1.20 44.86 14.63 38.11 99.12


MB-20-81 0.42 0.19 47.26 13.08 37.60 98.56
MB-38-81 1.95 3.07 0.66 51.62 37.31 94.61
MB-77-81 A 4.21 1.70 35.88 17.02 37.02 96.51
MB-77-81 B 2.11 1.20 41.02 15.88 37.96 98.18
MB-77-81 C 4.00 2.43 37.18 16.59 38.90 99.22
MB-77-81 0 1.28 0.86 38.74 18.93 37.45 97.38
MB-77-81 E 2.42 0.66 37.90 18.27 37.20 96.56

Analyses performed by the author under the supervision of Michael


Shaffer, at the department of Geology, University of Oregon, Eugene,
1983.

Note: Letter designation in sample number indicates paragenetic order


in zoned crystals, with A earliest.
CO2 calculated by stoichiometry.
124

ppm with the higher values in samples containing abundant pyrite.


Much of the arsenic is believed to be contained in the atomic struc-
tore of pyrite as confirmed by microprobe analyses listed in Table
21. Copper values are generally low (<3000 ppm) which reflect the
lack of chalcopyrite, except as a minor constituent in sphalerite.
Most important, however, are the anomalous tin values for six sam-
ples. Tin anomalies have been recorded for the Chumpe Quartz
Porphyry (Table 12) and in ores from the vein-type mineralization
(Table 14). These anomalies may provide a genetic link between the
processes of hydrothermal alteration and mineralization operative in
the intrusion, veins, and mantos, respectively. Values for antimony
generally reflect the presence of sulfosalts. Further geochemical
information will be considered individually during subsequent discus-
sions of the specific mineral phases.
The pyrite and sphalerite represent 95 percent of the sulfides
localized in the mantos. They are found in virtually all manto
orebodies, except in oxidized rocks near the surface.
Pyrite occurs as granular aggregates, veinlets, and as dissemi-
nations. It generally is found as euhedral pyritohedrons which vary
from 0,1 to 2.0 mm in diameter. The pyrite is commonly intergrown
with sphalerite and quartz in open spaces and vugs. Larger crystals
generally are found in the vugs. Pyrite observed in polished sec-
tions is generally idiomorphic except where it has been corroded and
replaced by sphalerlte and galena. Sphalerite commonly replaces the
cores of pyrite crystals to form "atoll" textures, as shown in Figure
125

25a, although many other textural modes of replacement also were


observed. Where pyrite is present as massive aggregates, inclusions
of pyrrhotite are commonly found in the central portions of the host
crystal (Fig. 25b). These p,yrrhotite blebs are small (< 0.01 mm) and
rounded, and they are recognized by low relief and bireflectance.
There is little apparent orientation of the blebs. Itis possible
that pyrrhotite and pyrite were deposited simul taneously, or more
likely, that pyri te repi aced earl ier formed pyrrhoti te. Some pyrrho-

tite was also found in pyrite from a vein which traverses the manto.
Lujan (1970) states that much of the pyrite in the manto is dis-
tinctly anisotropic. Although some pyrite was observed to be aniso-
tropic, the majori1r examined in this study was found to be isotrop-
ic. Lujan suggested that the anisotropism might be the resul t of
strain caused by the pyrrhotite inclusions, or by large amounts of
arsenic in the crystal structure of pyrite. Pyrite in some polished
sections was found to oxidize with time, revealing intricate patterns
of growth zonation (Fig. 25c) and multiple generations of pyrite
deposition. rnvestigation of this phenomenon with the microprobe
showed variable concentrations of arsenic and cobalt up to 1.39 and
0.10 percent, respectively (Table 21, sample HPA M-11). The color
and intensity of oxidation appears to be the resul t of increasing
amounts of arsenic in the pyrite. Other samples of pyrite showed
detectable amounts of cobalt and arsenic, whereas copper and nickel
were not detected in any sample. A study of the distribution of
arsenic in the manto pyrites would be of interest in future research
126

Figure 25. Early manto-type mineralization.

(a) Core replacement of pyrite by sphalerite forming


an texture.
"atoll11

(b) Inclusions of pyrrhotite in massive pyrite.

(c) Growth zones in pyrite crystals as revealed by


oxidation.

Cd) Specular hematite blades surrounded and incipi-


ently replaced by sphalerite.
,
t 4
1
d
I 4
4'
4
;N
:!Z *-
,1_ s- I
F
0
*
t
-
4
I- i
,-_j_
-
is
) 4
S
-
-
S
,
I - r
_
4 r
( r ;
* .
:r
1
7
1
If
: f
:
:4-
-

i r :
- ¼
- S J -
3 1
-.
128

at the mine. The mechanism of pyrite replacement in the limestone


host appears to be nearly volume for volume, although much of the
pyrite elsewhere has been deposited in open spaces. Bands of pyrite
were observed to have replaced and preserved sedimentary structures
in the Pucara Group limestone at several locations within the mine.
Samples containing these textures have led several geologists to
reason that the pyrite was deposited syngenetically with the lime-
stone. However, the author was unable to find any extensive areas of
finely laminated pyrite indicative of syngenetic deposition. Pyrite
and sphalerite more conronly replace entire limestone beds, without
preserving the limestone textures.
Quartz was deposited with pyrite and specular hematite early in
the history of mineralization. Sillcification of the Pucara Group
along faults and breccias by microcrystalline quartz was followed by
sulfide mineralization and precipitation of euhedral quartz crystals
in solution cavities. Crystal sizes range from 0.1 cm in massive
sulfide ore to 10 cm -in solution cavities. Although most pyrite and
quartz appear to have been deposited simultaneously, with little
evidence o4replacement between the two minerals, some quartz is
found enclosing small discrete pods of massive pyrite. Sphalerite
replaces quartz and pyrite, although all three minerals when oc-
cupying vugs within early formed massive ore are intergrown with
little obvious replacement
129

Specular hematite is found throughout the manto ore, but it is


generally more abundant away from the centers of massive sulfide
mineralization. The exact paragenetic relation of specular hematite
to the other minerals Is ambiguous. In the mantos, it is present
both as massive felty aggregates of fine-grained rosettes with quartz
and pyrite, and as disseminated clusters associated with coarse-
grained manganosiderite. Specular hematite in polished sections
exhibit characteristic euhedral needle or blade-like crystals which
range up to 1 cm in size (Fig. 25d). It appears to be replaced by
pyrite, in many cases as pseudomorphic fonns (Fig. 26a), associated

with quartz-pyrite mineralization. More conunonly, it is replaced by


massive pyrite and sphalerite which may rarely contain unrepl aced
relict blades of the oxide. Minute inclusions of hematite are found
concentrated in growth zones of quartz. Where the hematite occurs
with manganosiderite, it appears that this oxide precipitated in
cavities or solution voids of massive sulfide ore, and was later
surrounded by coarse-grained manganosiderite. Some replacement of
hematite by manganosiderite is indicated by resorption and in-filling
of blades in an atoll-like texture. In most places where mangano-
siderite and specular hematite are intergrown, the orientation of the
oxide blades is independant of the rhombohedral cleavage in the
carbonate, thus suggesting an early paragenesis. Specular hematite
is pseudomorphically replaced by magnetite, especially at locales
where the oxide is in direct contact with manganosiderite (Fig. 26b).
In fact, where hematite is enclosed by pyrite, sphalerite, or quartz,
130

Figure 26. Manto-type mineralization.

(a) Pyrite pseudômorphs after specular hematite in


quartz.

(b) Crystals of specular hematite in sphalerite, and


with magnetite pseudomorphically replacing hema-
tite at the contacts with siderite.

(c)Curved blades of wurtzite(?) pseudomorphically


replaced by sphalerite.

(d) Chalcopyrite replacing earlier formed sphalerite


and specular hematite. Siderite has replaced all
minerals, and magnetite has pseudomorphically re-
placed blades of specular hematite.
-
F,
I
-
-
' 1
1'
;};
U
0 .
;:F :: : ,
.
1''L4 I
, 7 .
'I -
4 4
-
I -
-
- -.-,
-,
------- ---
- - -
4 ,_j I
4 ---
-
--'p s- --- 4 -
-
4 - - -;
I
132

it shows little if any replacement by magnetite (see Fig. 26b). This

relationship indicates that the replacement process took place after


the deposition of sphalerite, and probably simultaneously with, the
formation of late-stage manganosiderite.
Magnetite is not found as a primary hypogene phase except as
pseudomorphs after specular hematite. Discrete euhedra of this in-
termediate iron oxide were not found in any location within the
Huaripampa Mine.

Sphalerite is the main ore-bearing sulfide constituent of the


mantos. Three major varieties of sphalerite can be distinguished on
the basis of composition, texture, distribution, mineral association,
and paragenetic position. The earliest, and quantitatively most
abundant type, is an iron-rich, brown to black sphalerite found
associated with massive pyrite. Crystals range from 0.1 to 6.0 nm in
size, with the larger ones occupying vugs with quartz and pyrite.
Microscopically, the sphalerite conmnonly contains oriented blebs of
chalcopyrite, especially near grain contacts with pyrite. These

bl ebs may be of exsol ution origin, but it is more likely that they
were formed with replacement of iron-rich sphalerite by copper-
bearing solutions, and thus are the result of processes illustrative
of the uchalcopyrite disease11 as described by Barton (1978).' rnclu-
sions or exsolution blebs of tetrahedrite are commonly found adjacent
to chalcopyrite or scattered randomly throughout the sphalerite. It
is apparent from polished thin sections of sphalerite that the darker
(iron-rich?) regions host most of the inclusions of chalcopyrite and
133

tetrahedrite. Growth zonations are well defined by variations in


color and composition. Individual crystals are typically zoned from
dark brown or black to light yellow or amber red from the core
outward, although numerous reversals in color and composition can be
seen. Variable substitutions of Fe+2 for Zn+2 presumably caused the

major color variations in this type of sphalerite. Microprobe


analyses of a sample of banded sphalerite (sample HPA M-11, Table 22)
show excellent correlation of iron content with color.
The second variety of sphalerite found in the mantos has been
called "ruby" sphalerite by mine geologists because of its light red
to yellow color. This type of sphalerite is paragenetically younger
than the early iron-rich variety, and is found in association with
galena, rhodochrosite and pyrite. It is usually coarsely crystalline
and contains abundant, large fluid inclusions.
The third variety of sphalerite is distinguished by its collo-
form habit and association with late-stage colloform siderite. The

col loform texture has been cited by many geologists in the past as
evidence for the transport and deposition of the ores as colloids.
However, Roedder (1968b) has concluded from his studies that most
colloform sphalerite actually grew as small fibrous crystals project-
ing into a supersaturated ore fluid. The colloform habit originates
by growth from a large number of closely-spaced points of nucleation.
Cal loform sphalerites from the Carahuacra Manto typically exhibit
alternating bands of fine-grained isotropic and anisotropic sphal-
erite in a texture referred to as "schalenblende". X-ray diffraction
134

analyses of several samples revealed weak diffraction peaks indica-


tive of wurtzite, and strong peaks for sphalerite. Curved plates of
wurtzite, that are pseudomorphically replaced by sphalerite, as. de-
termined by X-ray diffraction, are shown in Figure 26c. The trans-
formation of wurtzite to sphalerite may possibly be dependent on
temperature as demonstrated by Styrt and others (1981) from their
recent work on the mineralogy of black smokers" on the East Pacific
Rise. Compositions of the three major varieties of sphalerite, based
on seven samples, were determined by microprobe analyses that are
listed in Table 22. Early black sphalerites have iron contents which
range from 1.08 to 10.64 weight percent (samples MB-18-81, M-4, and
HPA fl-li). These samples also contained measurable amounts of man-
ganese and copper, although lacking systematic variations with iron
content. Late ruby sphalerite (samples MB-38-81 and MB-74-81) are
typically low in iron (0.24 to 1.17 percent), but all contain trace
amounts of manganese and are devoid of detectable copper. Colloform
sphalerites (samples MB-29-81 and MB-77-.81) have iron contents which
range from 0.12 to 4.38 percent. However, these exhibit little
correlation of color with iron content. In addition, the only de-
tectable concentrations of cacknium are found in these samples. Un-

fortunately, the cause of color banding incolloform sphalerites


remains unresol ved from microprobe analyses of the present study.
Similar negative results were obtained by Roedder and Dwornik (1968)
from their investigation of sphalerite-bearing ores from Pine Point,
Northwest Territories, Canada.
135

Chalcopyrite is not a major constituent of the manto ores,


except locally in and near the intersections with major veins. Trace
amounts are found in association with early iron-rich sphalerite. and
adjacent to exsolution blebs of tetrahedrite in sphalerite. Grain
size is variable from 0.1 to 0.5 nm. A unique texture is illustrated
by Figure 26d wherein sphalerite and chalcopyrite have partially
replaced a crystal of specular hematite that was later converted to
magnetite during carbonate precipitation.
Galena is not abundant in the manto ores, as confirmed by the
average grade of 1.0 percent lead listed in Table 1, and most of this
metal is derived from lead-rich veins in the Huaripampa Mine. None-

theless, galena is found locally in pods near the intersection of


veins with the mantos, and as a minor constituent with massive sphal-
erite ore. Crystals of galena range from 0.1 to 3.0 mm in size.
Those of the manto ores rarely contain exsolved phases such as tetra-
hedrite and other sulfosalts. Galena extensively replaces pyrite and
some of the early iron-rich sphalerite. It in turn may be sparingly
replaced by iron-poor sphalerite, but in general both sphalerite and
gal ena appear to have been deposited simul taneously (a fact docu-
mented by sulfur isotope equilibria).
Manganosiderite is a major gangue mineral found in all manto
ores. It is present in three distinct forms. The first is a fine-
grained and iron-rich variety, which formed through reaction of the
original calcite crystals in the limestone with fluids rich in iron
and manganese. This metasomatic event does not represent discrete
136

precipitation of manganosiderite, but rather chemical replacement of


the original carbonate material. The other two varieties were de-
posited rather late in the formation of the ore deposits. They are
represented by botryoldal and crystalline forms. The botryoidal
variety is most typically found in late solution cavities located
between massive sulfide ore and altered limestone. It is intergrOwn
with barite, colloform sphalerite and marcasite, and is commonly
coated with a layer of finely crystal line marcasi to. Large botry-
oidal masses, up to 15 cm across, are commonly found in solution
cavities. The coarse-grained crystalline variety is a component of
the veins which crosscut massive ore. Zoned rhombohedral crystals up
to 5 cm in diameter were observed in vuggy veins. The crystalline
manganosiderite typically fills solution voids within the massive
sulfide ore. Apparently the depositlonal environment for the differ-
ent types of manganosiderite type changed with time from the crys-
talline variety to the botryoidal or colloform type. Compositional

data for manganosiderite were obtained through microprobe analyses


and are listed in Table 23. In general, this carbonate contains
minor amounts of CaO and MgO with typical values of less than 4.21
and 3,07 percent, respectively. In contrast, contents of FeO and MnO
are highly variable. Compositions of FeO range from 0.66 to 47.26
percent with the larger values obtained from early-formed crystals or
the cores of zoned crystals. Those of MnO vary between 13.08 and
51.62 percent; the higher value (sample MB-38-81) represents a fairly
pure rhodochrosite. On the basis of these data, It is likely that a
137

compositional continuum from pure siderite to rhodochrosite would be


found with additional sampling and microprobe analyses. Sample MB-

77-81 is illustrative of a zoned colloform manganosiderite In which


analysis A represents the earliest formed carbonate and E the latest
Analyses for vein-type carbonates were presented earlier in Table 18.
From the standpoint of texture and composition, the late-stage car-
bonates of the San Cristobal Vein are identical to the late-stage
carbonates of the mantos.
White to honey-colored crystals of barite are associated with
the late-stage minerals as thin plates ranging from 0.5 to 3 cm in
length. Larger and thicker crystals up to 6 cm in length have been
found in vugs. Some are intergrown with finely crystalline layers of
botryoidal manganosiderite that exhibit textures similar to those
previously described for carbonates in the San Cristobal Vein.
Barite, atone location in the Moises Manto, was observed to have
been replaced by quartz.
Marcasite is a minor component of small cross-cutting veinlets
in the massive manto ores. It is also found as fine-grained, cocks-.
comb crystals which coat surfaces in most open spaces and as anhedra
in thin bands within botryoidal manganosiderite. Marcasite replaces
nearly all minerals including galena, pyrite, specular hematite, and
sphalerite. In one sample (MB-29-81), large needle-like marcasite
crystals were replaced by colloform sphalerite-wurtzite. Massive
amounts of marcasite are found in the upper portions of the
Huaripampa Mine and in the open pit associated with vuggy jasperoidal
138

limestone where it is believed to be of supergene origin after prim-


ary hypogene sulfides. This supposition is supported by samples
collected from the mine and displayed in the San Cristobal Mine
office, which have stalagtitic form with a central tube.
Paragenesi s

Based on the preceding textural evidence and cross-cutting rela-


tionships exhibited in the ores, as well as data from fluid inclu-
sions and the electron microprobe, the paragenetic sequence can be
divided into three periods of hypogene mineralization that were
preceded by a period of hydrothermal al teration. Limestones of the
Pucara Group, prior to sulfide mineralization, were altered by iron
and manganese-bearing solutions which converted the original car-
bonate into a finely crystalline, iron-rich manganosiderite. This
carbonate was subsequently replaced by microcrystalline quartz and
fine-grained specular hematite. The pre-mineralization period pro-
duced the bulk of the jasperoids, ferroan manganosiderite, and mas-
sive specular hematite present in the district. The paragenetic
sequence of mineral deposition is summarized and graphically
illustrated in Figure 27.
The first period of mineralization is marked by replacement of
the manganosiderite and specular hematite by first pyrrhotite and
then pyrite and quartz. Pyrrhotite is preserved now only in large
masses of pyrite located near the intersections of cross-cutting
veins and the mantos. Pyrite apparently was deposited during several
intervals separated by periods of corrosion as determined by growth
TIME
T°C 290 220

Mineral First Second Third Supergene

specular hematite
quartz - - - - -
pyrrhotite
pyrite
chalcopyrite
tetrahedrite-
tennantl te
Ferich FeDoor
sphalerite
- -
galena
pyrargyrlte-
polybasite
Is. aLt.
manganosiderite-
rhodochroslte
after hm.
magneti te

barite
marcas)te
native silver
covellite
earthy hematite
manganese oxides

Figure 27. Paragenetic sequence of mineral deposition In manto-type ores (thickness of lines is proportional to minerals
abundance).

(,.)
0
140

zoning and compositional changes in arsenic content. Quartz was

deposited continuously from the pre-mineralization period to the end

of the first period of mineralization, and minor amounts of both

quartz and pyrite may have been deposited subsequent to the first

period.

The second period of mineralization began with the precipitation

of dark brown, iron-rich sphalerite. This sulfide extensively re-

placed massive pyrite, specular hematite, and quartz, and fills

solution voids in the ore. The earliest sphalerite is typically

homogeneous in composition with an iron content of near 10.0 weight

percent Further dissolution of previously deposited minerals and

the altered limestone allowed precipitation of sphalerite in open

spaces as suggested by the presence of banded and growth-zoned

euhedral crystals. The iron content of sphalerite, as determined

from microprobe analyses, decreased with paragenesis, al.though num-

erous reversals in iron content are present. Trace quantities of

tetrahedrite and chalcopyrite are dispersed throughout the Iron-rich

massive sphalerite, and commonly are found in the darker, more iron-

rich bands within zoned crystals of sphalerite. The oriented chalco-

pyrite blebs in sphalerite may represent replacement by fluids con-

taining either small amounts of copper or fluids at temperatures

below that necessary for the precipitation of hypogene chalcopyrite.

An example of chalcopyrite that replaces both sphalerite and hema-

ti te, and pseudomorphical ly repl aced by magnetite is shown in the

photomicrograph represented by Figure 26d.


141

Galena is found in minor amounts with the earliest sphalerite,

but always as a product of subsequent repl acement. Pyri te is also

replaced by galena along fractures and corrosion cavities. In addi-

tion, galena is present as euhedral crystals on iron-rich sphalerite

at several locations in the Huaripampa Mine. Although it contains

few inclusions of the sulfosalts, small blebs of tetrahedrite, pyr-

argyrite, and chalcopyrite are located at grain contacts with sphal-

erite. With decreasing iron content, the sphalerite and associated

galena assume mutual boundaries Indicative of simultaneous deposition

and equilibrium. The largest proportion of galena in the mantos was

deposited with sphaierites containing between 1.0 and 4.0 percent

iron. As the iron content decreases to less than 1.0 percent, the

color of the sphalerite changes from brown to light yellow and red,

and this iron-poor sphalerite embays and cross-cuts galena. During

the second period mineralization, deposition of galena and sphalerite

probably alternated or overlapped as a consequence of episodic intro-

ducton of the ore-bearing fluids. This repetitive process would

account for the compositional banding and reversals observed in the

sphalerite.

The third period of mineralization is the most complex in terms

of paragenesis. Gangue minerals, including manganosiderite-rhodo-

chrosite, barite, and marcasite, were the predominant phases de-

posited, but the evidence also suggests periodic deposition of minor

amounts of the ore minerals. Manganosiderite, with marcasite, was

deposited in veins which cross-cut the massive ore and in vugs within
142

the ore. Barite was deposited at this time or slightly later, but
before the majority of botryofdal carbonate. A fundamental textural
change took place during this last stage of mineral ization, and
caused typically crystalline minerals such as sphalerite and siderite
to assume a botryoidal or colloform habit Continued solution of the
limestone produced large vugs which were subsequently filled by large
botryoidal masses of manganosiderite containing bands, or layers, of
colloform sphalerite-wurtzite. Sample MB-29-81 contained pyrite
replaced by tetrahedral sphalerite, which was then coated and re-
placed by large crystals of marcasite that in turn were replaced by
colloform sphalerlte. Crystallization of barite ceased before that
of most botryoidal manganosiderite. Compositional changes in man-
ganosiderite are reflected in increasing amounts of manganese with

time, and with distance from the source of the ore-forming solutions.
These progressed to the formation of nearly pure rhodochrosite.
Minor influxes of the metals occurred periodically, which produced
pockets of galena, sphalerite, pyrite, and minor amounts of sulfo-
salts with rhodochrosite.
As previously mentioned, magnetite pseudomorphically replaces
specular hematite, and only where in contact with manganosiderite.
From paragenetic relationships, most replacement took place during
deposition of the manganosiderite. However, this relationship is
based tenuously on the recognition and timing of carbonate influx,
which could have occurred earlier in the paragenetic sequence and
near the end of the second period of mineralization.
143

The alteration of the volcaniclastic tuffs to dickite is unclear


in terms of age relationships. Inferences based on the conditions of
formation for dickite would favor either the first or third periods
of sulfide mineralization. Formation of kaolinite is related to acid
conditions, or to the presence of excess H+ and MCi. The presence of
marcasite and wurtzite are indicative of acid conditions, but they
might also form during other conditions of sulfide deposition and/or
later oxidation of massive sulfides.
Subsequent to the third period of mineralization, an episode of
oxidation and minor secondary enrichment of the ore deposits took
place. This event was marked by the enrichment of manganese through
the oxidation of manganosiderite and rhodochrosite and the production
of numerous oxides of manganese and iron. Further impl ications of
this effect are discussed later under Secondary Oxidation and
Enrichment.

Lateral and Vertical Mineral Zonation

Mineral zonations within the individual mantos somewhat obscure


because of the complex relationships that exist between the mantos
and the cross-cutting vein structures. However, some general mm-

eralogic trends can be recognized within an individual tnanto zone.


These trends appear to be dependent on (1) the proximity and numbers
of crosscutting veins, (2) distance from the Pucara Group-Mite Group
contact, and (3) distance from the hydrothermal center of mineraliza-
tion for the dis.trict.
144

Observational data is derived primarily from exposures in the


Carahuacra Manto at the Huaripampa Mine. Manto deposits to the south
have not been adequately tested in the vertical dimension, but in-
formation such as a] teration patterns and chemical data from sur-
ficial outcrops of these mantos implies similar trends.
The Carahuacra Manto, as mentioned previously, is centered on
the Virginia Vein system with mineralization extending laterally
along the Pucara Group-Mitu Group contact and outward and upsection
from the veins. Lateral mineralogic trends with increasing distance
from the intersections of veins with 1 imestones hosting the manto
ores are defined by decreases in the (1) abundance of pyrite, (2)
iron content of sphalerite, (3) FeO/MnO ratio of the manganosiderite,
(4) abundance of gal ena, and increases in the (5) abundance of mar-
casite, (6) abundance of barite, (7) abundance of silver-bearing
sulfosalts, and (8) specular hematite/sulfide ratio.
Vertical mineralogical trends are less well defined, but with
increasing depth generally show decreases in the (1) amount of specu-
lar and earthy hematite, (2) abundance of marcasite, (3) grade of
silver, and increases in the (4) total sulfide/oxide ratio, (5) iron
content of sphalerite, (6) abundance of pyrrhotite inclusions in
pyrite, (7) abundance of chalcopyrite inclusions in sphalerite, and
(8) FeO/MnO ratio of manganosiderite.
145

Metal-Metal Associations

Fifty three samples collected from the Carahuacra, Toldorrumi,


and Moises Manto were analyzed for copper, lead, zinc, and sil ver.
Of these, 13 were assays of high-grade ore listed in Table 24 and the
remainder were trace metal geochemical analyses given in Table 25.
Data from these analyses have been plotted on log-log probability
graphs and correlation coefficients have been calculated using the
logarithmic value of each concentration. Correlation matricies for
all data groups and a correlation matrix for all low-grade samples
collectively are listed in Table 26.
Some background information is neâessary before meaningful corn-
parisons can be made. Samples listed as Huaripampa high-grade are of

massive ore collected from subsurface mine workings. Huaripampa low-


grade samples represent altered rocks containing sub-economic concen-
trations of ore metals which were also collected underground. All
samples from the Toldorrurni and Moises Mantos are partly oxidized
specimens of low-grade ore collected from Fe-Mn metasomatized lime-
stone with the exception of MB-68-81, which is an unaltered limestone
(see Appendix, Table Al, for sample locations and descriptions). It
is also important to note that the correlation coefficients may be
affected by the location of samples with respect to the source of
mineralization. Samples from the Huaripampa Mine are intensely al-
tered and mineralized and represent locations nearer the inferred
intrusive source of the metals, whereas those from the Toldorrumi and
Moises Mantos are more distant
146

Table 24. Assays of copper, lead, zinc, and silver for selected
samples of manto-type ore.

ppm
Sample1 Cu Pb Zn Ag

MB-4-81 2200 240000 116000 141

MB-6-81 600 3700 83000 128

MB-18-81 1000 3300 203000 94

MB-22-81 600 2300 91000 10

MB-23-81 100 12300 9600 60

MB-24-81 3700 1500 289000 30

MB-26-81 400 52000 87000 43

MB-28-81 300 1000 260000 36

MB-29-81 1200 4000 414000 91

MB-38-81 2100 184000 94000 2297


MB-46-81 600 6200 131000 26

MB-71-81 10300 6100 45000 41

MB-75-81 200 260000 5800 685

Assays determined by Centromin Peru, 1981.


I See Appendix, Table Al for sample location and. description.
147

Table 25. Trace metal analyses for samples of sub-economic ore


and altered limestone in the Carahuacra, Toldorrumi,
and Moises Mantos.

ppm

Sample1 Manto Cu Pb Zn Ag

MB-7-81 CA 35 175 340 6.2


MB-20-81 CA 30 110 12500 0.3
M8-25-81 CA 125 18 26500 0.3
MB-31-81 CA 17 275 370 3.4
MB-69-81 CA 15 70 410 0.6
I(PA R-1 CA. 35 1320 7710 28.2
HPA R-2 CA 7 1050 3420 10.5
HPA R-3 CA 10 360 545 1.3
HPA R-5 CA 5 60 120 0.8
HPA R-7 CA 3 360 860 5.0
HPA R-9 CA 13 615 2130 14.1
I(PA R-11 CA 16 4410 6310 252.0
HPA R-14 CA 14 45 605 1.1
HPA R-15 CA 25 850 865 10.8
R-3 CA 5 265 165 6.2
R-5 CA 140 230 1880 2.5
R-6 CA 12 20 255 0.3
R-8 CA 5 50 1380 0.7
R-9 CA 40 70 60 2.1
R-10 CA 5 80 1250 0.4
R-12 CA 6 40 650 0.4
R-13 CA. 710 820 10450 23.5
R-14 CA 110 2010 3580 4.1
MB-48-81 T 6 200 25 1.7
MB-49-81 7 5 120 32100 1.6
M8-50-81 1 160 250 4910 4.5
MB-54-81 T 25 595 12250 1.8
M8-55-81 1 20 1130 3590 2.8
MB-56-81 T 440 915 63000 30.1
MB-57-81 T 14 2820 11050 5.8
M8-51-81 M 8 40 12 0.3
MB-52-81 H 650 190 5200 24.1
M8-58-81 H 4 1430 3920 5.1
M8-60-81 H 4 1230 1850 1.4
M8-62-81 H 6 615 530 3.3
M8-64-81 M 160 320 1650 44.5
M8-65-81 M 125 835 11550 260.0
M8-66-81 H 110 1750 12010 110.4
M8-67-81 H 85 418 17500 283.0
MB-68-81 H 8 515 1005 6.3

Analyses performed by Chemical and Mineralogical Services, Salt


Lake City, Utah, 1981.
I See Appendix, Table Al for sample location and description.
CA Carahuacra; I Toldorrumi; H Moises.
148

Table 26. Correlation matrices for assays and geochemical analyses


of manto-type ore.

Cu 1.00 Huaripampa High-grade (n=13)


Pb -0.0617 1.00
Zn 0.3566 -0.3501 1.00
Ag 0.0138 0.7206 -0.1033 1.00

Cu Pb Zn Ag

Cu 1.00 Huaripampa Low-grade (n23)


Pb 0.2059 1.00
Zn 0.4878 0.3482 1.00
Ag 0.2066 0.8858 0.2359 1.00

Cu Pb Zn Ag

Cu 1.00 Toldorrumi Low-grade (n7)


Pb 0.2556 1.00
Zn 0.4442 0.2917 1.00
Ag 0.8282 0.4976 0.4772 1.00

Cu Pb Zn Ag

Cu 1.00 Moises Low-grade (n=10)


Pb -0.1334 1.00
Zn 0.5128 0.7058 1.00

Ag 0.7469 0.3784 0.8485 1.00

Cu Pb Zn Ag

Cu 1.00 Combined Low-grade (n=40)


Pb 0.1991 1.00
Zn 0.5010 0.4335 1.00

Ag 0.4961 0.7175 0.4502 1.00

Cu Pb Zn Ag
149

In general, lead and silver exhibit strong positive correlation


in both the high and low-grade sample suites (0.7206 and 0.8858,
respectively) of the Huaripampa Mine, but they decrease as a function
of ei ther distance from the center of mineralization or 1 eachi ng of
the surficial outcrops. The correlation coefficient for all low-
grade samples is 0.7175. Correlations between both zinc and silver
and copper and silver are higher in the lower-grade ores, but they
decrease as the grade of ore increases. Correlations between copper
and lead appear to be unimportant in all sampl e groups.
The foregoing discussion serves mainly to identify possible
geochemical trends for this reconnaissance suite of samples. Greater
geological and statistical significance for the apparent positive
correlations between lead and silver, and possibly between copper and
silver, zinc and silver, and copper and zinc, would require a larger
carefully selected suite of samples. Moreover, because of lateral
and vertical variations in the mineralogy and geochemistry of the
ores, it is unlikely that mixed sample suites obtained from geo-
graphically and geologically distinct ores can provide meaningful
correlations on a more local level.
Al teration

Three distinct 1'pes of hydrothermal alteration are recognized


in the limes tones of the Pucara Group surrounding man to ore bodies
and fracture-controlled channelways. These are (1) sideritization
produced by intense Fe-Mn metasomatism, (2) sf licification, and (3)
150

argillic alteration of non-calcareous rocks. Products of these al-


teration types are manganosiderite, jasperoid, and kaolinite
(dickite) respectively. The Fe-Mn metasomatized limestone is more
volumetrically abundant than ,Jasperoids and argillitized rock.
Argillic alteration constitutes a small percentage because of the
relative paucity of non-calcareous rock types in the Pucara Group.
Al teration at San Cristobal occurred just prior to and concomitant
with sulfide deposition, in contrast to that described by Lovering
(1949) for the Tintic District, Utah, where a significant time inter-
val preceded ore mineralization.
The center of al teration and mineral ization is located in the
Carahuacra Valley, with the amount and intensity of alteration de-
creasing both north and south of this area. A number of factors
controlled the distribution and intensity of alteration. These in-

clude (1) the type and amount of porosity and permeability in the
limestone, (2) the location with respect to the hydrothermal center,
(3) the number of fluid channelways (fractures or veins) entering the
limestone, and (4) the stratigraphic distance into the Pucara Group
above its contact with the Catalina Volcanics.
The primary product of Fe-Mn metasomatism is manganosiderite,
which formed through the introduction of fluids rich in Fe2, Mn2,
and minor amounts of Mg2+. These same fluids removed Ca2+ to solu-
tion from carbonate host rocks of the Pucara Group. The widths and
shapes of the alteration zones are proportional to the size and shape
of the nearby fluid channeiways. rn most areas the transition from
151

altered to unaltered limestone is sharp and distinguished by a dis-


tinct change in color and reactivity of the rock to MCi. Mangano-

siderite always forms an outer zone of alteration that envelopes, the


ore bodies and walls of the fluid channeiways.
Jasperoidal limestone is the product of silicification. It is
found only in the areas of most intense alteration and mineral iza-
tion. Jasperoids are abundant in the Carahuacra Manto area, minor in
extent in the Toldorrumi Manto, and almost non-existent farther to
the south in the Moises Manto. Fine-grained silica comprising the
jasperoids was introduced along the same channelways that produced
the manganosiderite. Replacement of the earlier formed mangano-
siderite ranges from pervasive (100 percent) in large bodies of
jasperoid to selective (less than 5 percent) in outlying areas.
Surficial oxidation of unreplaced manganosiderite gives the jasperoid
a dark gray to black color. Bedding typically is well preserved.
Argillic alteration is confined to interbedded volcaniclastic
tuffs and small cross-cutting dikes of unknown composition and
origin. With the exception of minor pyrite as disseminations and in
crosscutting veinlets, the dike rocks and tuffs are essentially
unmineralized. The alteration mineralogy typically consists predomi-
nantly of kaolinite with minor quartz and trace amounts of sericite.
It is likely that argillic alteration formed during the first stage
of massive sulfide replacement in response to fluids characterized by
low pH and high sulfur fugacity.
152

Oxidation and Secondary Enrichment

Surficial exposures of manto ore are distinctively characterized


by a well-developed black, ferromanganese-oxide gossan. This black
oxide capping serves to delineate mineralized outcrops from the
surrounding barren limestones. Oxidation of most hypogene sulfide
minerals is complete in surface exposures. Depths of oxidation are
highly variable because of recent glaciation. The thickest zone of
oxidized ore is located in the northernmost extension of the
Carahuacra Manto. Lyons (1968) has described significant enrichment
of silver values in the surficial oxidized ore, and secondary enrich-
ment In the form of native silver, argentite,and pyrargyrite, in
fractures as deep as 180 m below the zone of oxidation. The

Huaripampa Mine area has minor quantities of secondarily enriched


silver that is present in the form of fracture-controlled occurrences
of the native metal, but most of the zone of oxidation was physically
removed by glaciation.
The zone of oxidation, where present, is composed of a friable
or sometimes silicified mixture of ferromanganese oxides. Haederle

(1966) reported the major secondary manganese mineral to be braunite


(3Mn203.MnSiO3). Other secondary minerals of the gossan include
goethite, earthy hematite, manganite, and pyrolusite, which were
derived predominantly through oxidation of pyrite, rhodochrosite, and
manganosiderite. Sphalerite may be found in many outcrops, and its
presence is usually marked by coatings of hydrozincite. Galena and

chalcopyrite have been leached, but remnants in silicified outcrops


153

show rinds of cerrusite and covellite respectively.


Phendler (1957) conducted an evaluation of the manganese re-
sources of San Cristobal in both the oxidized and hypogene ores.
Average grades from outcrops of the Toldorruini and San Antonio
(Carahuacra) Man tos were 11.63 and 18.6 percent manganese, respec-
tively. Subsurface ore in the San Antonio Manto averaged 14.92
percent manganese, and total reserves at that time were calculated to
be 3,200,000 tons of 8.12 percent manganese. Phendler also noted
that the surficial outcrops were generally enriched by as much as 80
percent with respect to subsurface hypogene ore. He found an ap-
parent increase in manganese to the north, and away from the central
mineralized part of the district.
154

ALTERATION OF THE PUCARA GROUP

Three distinct 1'pes of rock alteration are recognized in lime-


stones of the Pucara Group adjacent to manto ore bodies of the San
Cristobal District. These include (1) Fe-Mn metasomatized carbonate,
(2) siliciffcation, and (3) argillic alteration of non-calcareous
rocks. The products of both mineralization and alteration are simi-
lar to those described in the East Tintic District, Utah (Lovering,
1949; Morris and Lovering, 1979) and in the Gilman District, Colorado
(Radabaugh and others, 1968; Lovering and others, 1978). The major
difference between the al teration in these two districts and the San
Cristobal District is the extensive replacement of the limestone by
manganosiderite, rather than by dolomite. The processes which formed
the various lypes of alteration and mineralization can be regarded as
one of simultaneous solution and replacement of the limestone,
rather than of strictly low-temperature (<300°C) deposition from
hydrothermal fluids rich in dissolved constituents.
Based on cross-cutting and replacement relationships, the prob-
able paragenetic order of alteration events relative to one another
is as listed above, namely early Fe-Mn metasomatism, silicification,
and late argillization. Some apparent discrepancies exist because
of overlap and repetition of the alteration events. However, the al-
teration process can be described as one involving fronts of advanc-
ing alteration that moved outward from the hydrothermal source along
zones of high porosity and permeability (fractures), and formed
concentric alteration envelopes or aureoles in the carbonate host
155

rocks adjacent to the manto ore bodies.


Fe-Mn Metasomatism

Dolomitizatlon of limestone country rocks adjacent to replace-


ment deposits of lead and zinc have been recorded In numerous dis-
tricts elsewhere and has been summarized in a classic paper by Hewitt
(1928). According to Lovering (1969), conversion of calcite to
dolomite involves the simultaneous solution of calcite and precipita-.
tion of dolomite by fluids having a high Mg2/Ca2 ratio. This
process may be referred to as magnesium metasomatism. Hover, the
source of fluids and magnesium sufficient to produce the large scale
conversion of limestone to dolomite as recorded in many districts,
particularly in the Colorado Mineral Belt, has been a continuing
enigma. Whereas al teration of igneous rocks can provide magnesium
for dolomitization, it requires the complete alteration of approxi-
mately 10 volumes of igneous rock to form one volume of dolomite
(Hewitt, 1928). Thus, it is likely that dolomitization takes place
through the heating and circulation of groundwaters adjacent to
igneous intrusions by mechanisms suggested by Lovering (1969). Al-

though Lovering (1969) impi ied that high Mg2/Ca2 ratios are neces-
sary for dolomite formation, experimental studies by Rosenberg and
Hol land (1964) on the system CaC12-MgCT2-0O2 have shown that at
higher temperatures (>250°C) a high Mg2/Ca2 ratio is not a pre-
requesite for the formation of dolomite. In fact, a chloride-rich
hydrothermal fluid need have only Mg2/Ca2 ratios of between 0.075
156

and 0.30 to favor dolomite (Rosenberg and Holland, 1964). However,

these authors caution extrapolation of their data below 250°C.


It is apparent that both hydrothermal and primary dolomites are
rare in the San Cristobal District. The presence of manganosiderite
implies that the chemistry of alteration was different, and involved
abundant iron and manganese with smaller amounts of magnesium in the
fluids. The source of iron and manganese is undoubtedly of hydro-
thermal origin and was derived by leaching of both elements from the
Chumpe Quartz Porphyry and the Excelsior Phyllite.
The conversion of limestone to manganosiderite is governed by
the reaction

Fe2 + Mn2 + Mg2 + CaCO ----- > FeMnMg(CO3)3 + 3Ca2 (1)

or if pure siderite is considered

Fe2 + CaCO ----- > FeCO3 + Ca2 (2)

These reactions will proceed to the right if sufficient Fe2 is


present. However, problems regarding the actual concentrations of
Fe2+, Mn2+, Mg2+ and Ca2+ are unresolved. Provided the formation of
siderite is analogous to that of dolomite as determined experimental-
iy by Rosenberg and Holland (1964), then the amounts of Fe2, Mn2,
and Mg2 in the fluid are relatively low compared to that of Ca2.
Little experimental work has been published on the system Fe-Mn-0O2,
thus the analogy may not be entirely similar. Nonetheless, it is
possible to calculate the equilibrium constant for reaction (2) and
157

thereby determine the theoretical Fe2+/Ca2+ ratio at a given tempera-


ture. Using the data of Robie and others (1978) and Kraupskopf
(1979), the calculated Fe2/Ca2 ratios for equation (2) are 0.0044,
0.088, and 0.0133 at 25°, 1500, and 250°C, respectively. Any value
greater than these will cause the reaction to proceed from left to
right and form sideri te at the expense of calcite. Al though these
are theoretical values derived from thermochemical data, the exten-
siveFe-Mnmetasomatismof the limestone walirock adjacent to the
veins and mantos attests to the introduction of large quantities of
Fe2 and Mn2 from the hydrothermal fluids.
Fe-Mn metasomatism of the Pucara Group limestones occurred early
in the paragenetic sequence, and prior to or concomitant with the
onset of massive sulfide deposition. The intensity of metasomatism
was dependent on the quantity of fluid, the concentrations of Fe2+

and Mn2+, and the type and amount of porosity and permeability avail-
able in the host rock. Alteration varied from narrow selvages along
closely spaced joints to selective replacement of entire brecciated
limestone beds. The zone of manganosideri te repl acement, adjacent to
large mantos in the Carahuacra area, may extend outward from 5 to 60
m although the distance depends chiefly on the density of the joint-
controlled channelways.
Textural changes of the limestone during conversion to mangano-
siderite included recrystallization to a hypidlotopic-granular mass
of rhombic crystals with an attendant slight increase in the size of
carbonate crystals nearest the ore bodies and solution channeiways.
158

At most locations, destruction of the original sedimentary fabric is


complete, although "ghosts" of fossil fragments and ooliths are
common in the outer portion of the alteration aureoles. However,

larger textures such as breccias and bedding laminae commonly are


preserved.
Distinct color anomalies produced during surficial oxidation of
the manganosiderite help delineate altered from unaltered limestone.
Manganosiderite with high concentrations of manganese oxidize to a
black color, whereas the rocks with increasing concentrations of iron
and magnesium grade from dark, reddish brown to yellow. Fresh lime-
stone is typically light blue in color. The transitional boundary
between fresh and altered limestone is typically sharp and less than
3 cm in less altered areas, whereas the change near large mantos may
grade over a distance of 3 to 4 m.
As stated previously, migration of the hydrothermal fluids was
controlled by zones of high permeability such as faults, breccias,
joints, and porous interbeds of calcareous sandstone. Several small
mantos and fault zones in the Moises Manto were mapped in detail to
Illustrate both the structural control for solution migration and the
resultant alteration patterns adjacent to mineralized veins and man-
tos. Detailed sketches of these areas are presented in Figures 28
and 30.
An area at the north end of the Moises Man to (N 8565, E 11840)
is depicted in Figure 28. Here, small veins which cut perpendicular-
ly across the strike of the limestones have produced narrow (< 1 in)
$ -
_e * . -I -
$1 I -
- J__ _ C I
I I C, I I -
iu 'i IWA
IaIII
am iima
I
I$U 'tt a
III II
mmmi
III
IIU I II!
5'
III IIIItk i1k 'II I
Ililkik air 1' I I I
.a t
ii
II
I 'ii 'I
160

Figure 29. Photograph of area depicted in Figure 28. Note


backpack for scale and unaltered limestone (white)
on right side of photo.
_______________
-'
.-.--

I a 4 1¼' 1 I p
.
.1 ( b 4 -
(I.

.' I $

',
,. '. wwrn -

www wuu- _
- £
RU1¼
' I

I
'

'
p
-:
- -
N
'

Figure 30. Sketch of an outcrop In the Noises tianto Illustrating alteration zones adjacent to vein
and below Impermeable tuffaceous layers. Note diversion of vein mineralization Into
"bedding plane vein" (manto) below upper tuff bed.

1.I
0i
I'
162

brecciated fault zones showing little apparent offset of the lime-


stone beds. Mineralization in the vein consists of inanganosiderite,
sphal en te, and ban te. Other sul fides may have been present origin-
ally, but have since been removed by oxidation. The veins traversed
upsection until encountering two bedding-plane breccias which merge
laterally to form a 1. to 2 m thick manto. Numerous joints, spaced 4
to 6 cm apart, parallel the trend of the vein and contain lessor
amounts of ore mineralization. As portrayed in Figure 28, alteration
of the limestone adjacent to the veins is proportional to the size of
the vein. The manto extends southward for 300 m from this location,
and along which several other veins were observed to intersect this
replacement deposit Interestingly, the manto maintains its relative
thickness laterally, and increases in size only where cross-cut by
other veins. The same area depicted by sketch in Figure 28 is repre-
sented by photograph in Figure 29.
Another important structural control operative in the district,
namely the channeling of solutions by a relatively impermeable bar-
rier is illustrated in Figure 30. At this location, a narrow vein in
the Moises Manto (10-15 cm) cuts upward through the limestone sec-
tion. Adjacent to it on either side is a brecciated zone 30 cm wide
containing highly altered and partially mineralized clasts of lime-
stone. The vein itself is easily discerned from the enclosing brec-
cia by its mineralogy which includes sphalerite, manganosiderite, and
large tabular crystals of barite up to 5 cm in length. Interbedded
with the limestone in this area are two volcaniclastic tuff beds 10
163

to 15 cm in thickness. The thinner of the two tuffs is lower in the


section and cut by the vein. However, the vein does not cut the
upper tuff, al though some shearing is evident in the tuff and a smal 1
joint is present in the limestone. There is little displacement of
either the limestone or tuff. Clasts of the lower tuff were not
recognized in the fault zone, and apparently this bed was completely
obliterated although the upper tuff remained intact Upon tracing
the barite mineralization upward to its contact with the upper tuff
bed, the vein was seen to spread laterally iniediately under the tuff
to form a "bedding plane" vein or impounded horizon of replacement
Moreover, the vein retains its textural character for several tens of
meters before disappearing under soil cover. Both tuffs apparently
impounded some fluid movement as indicated by the conversion of
limestone to manganosiderite beneath each volcanic layer. Further-
more, some fluid must have permeated these horizons as a thin layer
of metasomatized limestone is present ininediately above each tuff.
Two unresolved problems are relevant to the foregoing interpre-
tation, and a may be inferred from an examination of Figure 30.
First, why is the breccia zone adjacent to the fault so large given
the lack of apparent displacement? Second, why doesn't the breccia
zone continue upward through the second tuff? The answer to both of
these questions probably relates to the mechanism by which the brec-
cias formed or became enlarged. Two possible mechanisms that serve
to explain these enigmas are described below, although there is
little evidence to support either.
164

Phillips (1972) published an hypothesis regarding hydraulic


fracturing and the formation of breccias during the hydrothermal
mineralization of sedimentary rocks. He suggested that high fluid
pressures in a fracture could extend the length of a fracture and
force fluids into the pore spaces of the adjoining rock. A sudden

decrease in pressure In the vein might then cause the expansion of


pore fluids and create a burst toward the vein to thereby form a
breccia. Another plausible mechanism, termed "chemical brecciation",
has been suggested by Sawkins (1969). This process is known as
"alkali reactivi1" by materials engineers and might produce a brec-
cia in carbonate rock. According to Sawkins (1969, p. 616),
"aggregate material containing either opal, acid or inter-
mediate volcanic glass, cristobalite, tridymite, chalcedo-
ny, or fine-grained dolomite is mixed with alkali-contain-
ing Portland cement....producing abnormal expansion and
cracking of mortar or concrete"
By analogy, he suggested that hydrothermal fluids rich in alkalies
upon mixing withfine-grained silica (jasperoid) or dolomite might
cause similar expansive reactions and thereby produce breccias. The

source of the al kal ies is presumed to be the hydrothermal sol utions


rich in NaC1. Regardless of the mechanism causing brecciation, the
features illustrated in Figure 30 accurately reflect the field rela-
ti on ships.

Chemical changes imposed on the limestone host with hydrothermal


alteration were determined by means of 14 microprobe analyses as
listed in Table 27. Whole rock chemical analyses were not performed.
Sample MB-68-81 is a relatively fresh oolitic limestone located
27 Plicroprobe analyses of unaltered carbonate, diagenetic dolomite, and Fe-Mn metasoma-
tized carbonate

Oxide 1 2 3 4 5 6 7 8 9

CaO 56 08 56 03 56 44 56 62 31 57 0 28 0 29 0 41 0 38
MgO 026 031 030 013 2107 105 030 137 148
FeO 0 11 0 00 0 00 0 00 0 16 41 76 46 93 40 84 44 68
MnO 0 15 0 00 0 00 0 22 0 00 17 20 13 34 17 70 14 53
CO2 44 46 44 31 44 62 44 71 47 88 37 62 37 58 37 81 38 31

Total 101 06 100 65 101 36 101 68 100 68 97 92 98 45 98 13 99 39

10 11 12 13 14
CaO 0 41 0 35 0 39 0 39 0 65

MgO 0 09 0 19 2 40 3 24 5 83
FeO 50 54 47 59 43 31 35 38 27 59
HnO 10 95 13 10 14 50 20 49 25 99
CO2 38 17 37 76 38 46 38 22 39 90

Total 100 16 99 98 99 06 97 72 99 97

1-2 oolites in unaltered limestone, sample M8-68-81, 3-4 calcite veinlets In unaltered limestone,
sample 148-68-81, 5 dlagenetic dolomite which replace oolites sample MB-68-81, 6-8 Fe-Mn metasoma-
tized limestone, sample 148-20-81, 9-11 Fe-Mn carbonate sample 118-25-81, 12-14 Fe-Mn carbonate sample
MB-54-81.
Analyses (weight percent) performed by the author under the supervision of Michael Shaffer at the
University of Oregon, Eugene, 1983, CO2 calculated by stoichiometry

L __
166

within a laterally extensive breccia. Analyses 1 and 2 are from the


ooltths in the limestone and 3 and 4 from calcite veinlets that
cross-cut the sample. They collectively represent typical concentra-
tions of CaO, MgO, FeO, and MnO for calcite. Analysis 5 is from a
diagenetic dolomite crystal in the same sample and it was found to
contain almost stoichiometrically appropriate concentrations of CaO
and MgO. Analyses 6 through 14 represent Fe-Mn metasomatized car-
bonate. It is apparent from these data that the limestone has under-
goneintense leaching of CaO, with a concomitant influx in FeO and
MnO and smaller amounts of MgO. The content of FeO ranges from 27.59
to 50.54 percent whereas that of MnO varies from 10.95 to 25.99

percent. Concentrations of MgO, although low, show an increase over


those obtained for pure calcite. With increasing distance from
mineralization (samples MB-20-81 to MB-25-81 to MB-54-81), there is a
decrease in the FeO/MnO ratio, whereas the relative amount of MgO
increases. Because of insufficient data, it is uncertain whether or
not there is a gradational change in composition between altered and
unaltered limestone. However, the field evidence based on mineralog-
ical data and color changes suggest a relatively sharp contact be-
tveen the two.

Extensive Fe-Mn metasomatism of limestone has not been described


elsewhere for similar deposits. However, Radabaugh and others (1968)

do mention that the ore deposits at Gilman, Colorado contain siderite


that in part forms an enveloping shell surrounding these replacement
deposits.
167

Sfl icification

Silicification of the limestone began subsequent to sideritiza-


tion and prior to ore mineralization. The hydrothermal fluids appar-
ently utilized the same series of fissures and permeable channeiways
that gave rise to the earlier Fe-Mn carbonatization, because there
are no occurrences of jasperoid without an outer shell of mangano-
siderite. Jasperoids are typically found in wall rocks located at
the intersections of large mineralized veins and limestones of the
Pucara Group, and as envelopes that surround large manto orebodies
in brecciated limestone. Moreover, their form generally reflects the
shape of the fluid conduit; the form being somewhat tabular at the
intersections of faults with thin mantos, and pod-like adjacent to
massive replacement ore. Silicification was especially intense in
the Carahuacra Manto area where large masses of jasperoid enclose the
major orebodies. In the smaller Toldorrumi and Moises Mantos,
silicification was confined predominantly to vein walls and breccias
located near the base of the Pucara Group. The bodies of jasperoid
range from less than 2 m to well over 100 m in diameter. Volumetric-
ally, they are much less abundant than the Fe-Mn carbonates that also
replace the limestone. Zoning in the larger replacement deposits
consists of a massive sulfide core surrounded by successive shells of
jasperoid and Fe-Mn carbonatized limestone.
Outcrops of the jasperoids are typically black, dense masses
that are considerably more resistant to erosion than either the
altered limestone or sulfide ores. However, their relief is subdued
because of recent glaciation. Bedding is typically well-preserved as
illustrated by Figure 31. Moreover, the closely spaced joints and

fractures which contributed to the development of the jasperoid are


also shown in this figure.
Petrographic examination the jasperoids demonstrates that they
are composed of minute crystals of clear quartz which may contain
small laths of specular hematite. Crystal sizes increase slightly
toward vugs and solution cavities. Unreplaced crystals of mangano-
siderite commonly are scattered throughout the jasperoid, and they
become more abundant near the contacts of jasperoid with limestone.
Oxidation of the Fe-Mn carbonate and minor disseminated sulfide, to
form a Fe-Mn oxide gossan, is responsible for the darker colors of
the jasperoids.
Intense silicification has produced a quartzite at the inter- -

sections of the Catalina and Polonia Veins with a basal quartz are-
nite of the Pucara Group in the Toldorrumi Manto. Silica has selec-
tively replaced the groundmass without apparent reaction or dissolu-
tion of the primary quartz grains. These particular outcrops also
contain abundant chlorite and specular hematite as disseminations in.
the fine-grained quartz matrix. The disseminated chlorite probably
was formed during the al teration event characterized by Fe-Mn meta-
somatism through reaction of the iron and magnesium-rich fluids with
169

Figure 31. Silicified (jasperoidal) limestone outcrop showing


preservation of bedding and closely spaced joint
fractures. Photo taken west of'the Huaripampa Mine.
170

clay minerals in the tuffaceous sediments.


There is some evidence to suggest later periods of silicifica-
tion in the genesis of the manto deposits. Late-stage barite crys-
tals in both the Moises and Toldorrumi Mantos have been replaced by
finely crystalline quartz. In addition, minor quantities of vein
quartz fill late fractures in distal manto deposits. Although some

may be of supergene origin, the presence of small fluid inclusions


containIng minute vapor bubbles suggest a late-stage hydrothermal
origin for this quartz.
Argillic Alteration of Won-Calcareous Rocks

Two varieties of non-calcareous rocks are present interbedded


with, or cross-cutting limestones of in the Pucara Group. These

consist of volcaniclastic tuffs and several thin, narrow dikes of


unknown composition and origin that are exposed in the workings of
the Huaripampa Mine. Tuffaceous limestones are generally found
stratigraphically above and below the thicker beds of tuff. In the
areas of most intense hydrothermal al teration and mineral ization,
these siliceous rocks have been altered to an assemblage of kaolinite
(dickite), quartz, and commonly sericite, and in association with
minor quantities of disseminated pyrite and manganosiderite. How-

ever, other sulfide minerals are confined exclusively to cross-


cutting veinlets.
171

A unique texture, which consists of pods of essential ly pure

dickite generally surrounded by massive sulfide ore, has formed in

the tuffaceous limestones and is illustrated in Figure 32a. They

probably formed during replacement of the limestone as a result of

the segregation of non-reactive aluminous clay into pods as the

replacement front moved through the limestone host. In most areas,

the pods have their maximum abundance at the leading front of mm-

eralization or near the upper part of a limestone bed. The location

of these pods with respect to the replacement front, and extending

outward from a fracture in the limestone, is illustrated by Figure

32 b.

The cross-cutting dikes have been pervasively argillized and

usually contain sericite. Disseminated pyrite appears to have selec-

tively replaced the primary mafic minerals. The dikes are most

likely related to the Carahuacra Quartz Monzonite, and thus presum-

ably were emplaced prior to mineralization. Because of the extensive

alteration imposed on these dikes, it is impossible to determine

their original chemical or mineralogical composition. Alteration of

the dikes is similar to that found in the Catalina Volcanics that

occupy the footwal 1 of the Huaripampa Mine.

According to Meyer and Hemley (1967), argillic alteration

generally is associated with fluids of low pH and high sulfur

fugacity. Therefore, it is inferred that this argillic alteration of

the tuffs and dikes in the Huaripampa Mine was imposed during the

first period of mineralization because the bulk of the pyrite was


172

Figure 32. Argillically altered and mineralized tuffaceous limestone.

(a) Photograph of working face in the Huaripampa Mine


showing numerous clay pods of dickite in massive
sulfide ore. Note the increase in number of clay
pods towards upper contact of manto with tuff.

(b) Sample of altered limestone with clay pods develped


near massive sulfide replacement front. Note also
the fracture control of mineralization.
173

,,

Figure 32.
174

deposited at this stage. Moreover, it is also likely that most of


the solution cavities in limestone formed during this early stage,
thereby allowing a partof the sulfide minerals to be deposited as
open space fillings in much of the deposit.
175

FLUID INCLUSION STUDIES

Fluid inclusions have been a source of interest and discussion


ever since the classic paper by Sorby (1858). However, only in the
last 30 years have they been extensively used as indicators of the
fluid chemistry and temperatures associated with plutonic processes,
and the evaluation of base metal ore deposits.
Arguments against fluid inclusion data are based on the assump-
tion that the inclusions may have leaked, and hence do not represent
the original fluid. Recent studies by Roedder and Skinner (1968)
have shown leakage to be a relatively rare phenomena and that most
assumptions regarding fluid inclusions generally are valid. Fluid
inclusions have since proved to be an invaluable aid to modern in-
vestigations of ore genesis and petrology.
Fluid inclusion work had not been attempted until recently on
the vein and manto ores of the San Cristobal District. Campbell and
Rye (1982) have completed the first study of fluid inclusions in
minerals from the San Cristobal Vein. They found inclusions in
augelite (Al2PO4(OH)3), quartz, and sphalerite and reported homogeni-
zation temperatures which ranged from 1600 to 210°C in sphalerite to
400°C in augelite. Salinities of the fluids in these inclusions were
fairly constant and ranged from 4-8 weight percent NaCl equivalent.
The present investigation has involved a detailed study of
inclusions from both vein and manto-type ores. However, emphasis was
directed toward inclusions in the manto ores because of the ambiguous
origin of these deposits. Fluid inclusions also were found and
176

measured in magmatic phenocrysts of quartz from the Chumpe Quartz


Porphyry. These inclusions were highly variable In composition and
type, and apparently are different from those reported by Campbell
and Rye (1982). In fact, they have provided more complete evidence
on the source and composition of the ore fluids. The purpose of
examining fluid inclusions from these areas was to (1) characterize
the temperature and composition of the mineralizing fluids, (2)
define the pressure or depth at which the ore deposits formed, (3)
delineate lateral continuity and variation of ore depositional temp-
eratures between the vein and manto deposits, and to (4) provide an
independant geothermometer to check calculated sulfur isotopic temp-
eratures. Resul ts of this study agree well with the data of Campbel 1
and Rye (1982) and provide additional information on the genesis of
the manto deposits.
A total of 40 polished fluid inclusion plates were prepared from
approximately 30 samples. Unfortunately, many of the plates did not
yield usable inclusions because of the inherent visual difficulties
in working with dark sphalerites or milky quartz. In general, excel-
lent inclusions were found in quartz phenocrysts from the Chumpe
Quartz Porphyry and light colored sphalerites from both vein and
manto ores. However, some good inclusions also were found in dark
sphalerite.
177

Following the proceedures described by Roedder (1967; 1976) and


Hollister and others (1981), each sample was examined visually before
heating or freezing in order to characterize the inclusion typeand
mode of origin. Primary, pseudosecondary, and secondary inclusions
were located using the criteria of Roedder (1967), although itwas
coninonly impossible to arrive at an unambiguous mode of origin for
many of the inclusions.
Primary f1uid inclusions, as defined by Roedder (1967), form
contemporaneously with the host mineral and usually are found along
growth zones or in crystal irregularities. The fluids trapped in
these inclusions represent the conditions at the instant of trapping
and thus provide geochemical constraints on the P-T-X properties of
the fluid during formation of the host mineral. Inclusions produced
at a later time, during episodes of fracturing and annealing, are
termed secondary fluid inclusions. These may or may not be related
to the fluids which produced the alteration and/or mineralization.
Fluid inclusions formed In fractures of a growing crystal are termed
pseudosecondary, and may represent the conditions present during
crystal growth.
Three main types of fluid inclusions, that roughly correspond to
inclusion types I, II, and III of Nash (1975, p. 1449), were observed
in samples from San Cristobal. Type I inclusions are the most common
in all samples, and consist of between 60 and 80 percent aqueous
phase with the remainder being a bubble of low-density vapor, as
shown in Figure 34. These inclusions represent entrapment of an
178

Figure 33. Secondary(?) fluid inclusions in magmatic quartz pheno-


crysts of the Chumpe Quartz Porphyry. Scale for a-c is
as shown in a; scale for d-i is as shown in d.

(a) Type I inclusion adjacent to zircon crystal. Homog-


inized to liquid phase at 360°C.

(b) Type 11 inclusions in a fracture plane. Homogen-


ized to vapor phase at approximately 380°C.

Cc) Type III inclusion with halite daughter crystal.


Homogenized to liquid phase at 323°C. Halite
dissolution temperature was 412°C (salinity is
approximately 47 wt. percent NaC1).

(d) Type III inclusion with halite daughter crystal.


Homogenized to liquid phase at 365°C with halite
dissolution at 379°C (salinity approximately 44.5
wt. percent NaCl).

(e) Type III inclusion with large halite daughter


crystal. Homogenized to liquid phase at 348°C.
Salinity undetermined.

(f) Type III inclusion with very large halite daughter


crystal. Homogenized to liquid phase at 360°C.
Salinity undetermined.

(g) Type III inclusion with anhydrite(?) or molycite(?)


daughter crystal. Homogenized to liquid phase at
384°C with little visible solution of daughter crys-
tal.

(h) Same as g above but with crossed polarizers.

(i) Type II inclusion with apparent unidentified daugh-


ter crystal (anhydrite or molycite?). Homogenized
to vapor phase at approximately 374°C.
!i '
' 4
- ;44

:!.$ØWi

A. :
C

DF I5ji
F

G I

-4
Figure 33
180

Figure 34. Fluid inclusions in vein and manto sphalerite. Homog-


enization temperature (Th, to the liquid phase) is given
for each inclusion. Scale for all photos is shown at
bottom of figure.

(a) Primary inclusions in vein sphalerite (SCM-10).


Th = 195°C.

(b-c) Large primary(?) inclusions in manto sphalerite


(MB-74-81). Th = 274° and 260°C.

(d) Two primary inclusions in manto ore trapped between


two sphalerite crystals (M-8). Note growth zones
parallel to triangular inclusions. Th = 249°C.

(e) Large primary inclusion in manto sphalerite (MB-74-


81). Th = 265°C.

(f) Primary inclusion in manto sphalerite (MB-38--81).


Th = 273°C.

(g) Primary(?) inclusion in manto sphalerite (MB-74-81).


Th = 263°C.

(h) Large flat inclusion in manto sphalerite showing


some evidence of necking down (MB-74-81). Th =
244°C.

(i) Primary inclusion in manto sphalerite (MB-38-81).


Th = 271°C.

(j) Primary inclusion in manto sphalerite (MB-74-81).


Note possible daughter crystal (insoluable) near
bottom right edge of vapor bubble. Th = 257°C.
*4,

, -. I
4..

b
..
,.1 B
A
Lc'
:

4.

L
0
44

'

100 J I.-.
G I

1-
Figure 34.
182

original homogeneous aqueous brine at elevated temperatures. Type II


and type III inclusions were found only in the Chumpe Quartz Porphyry
samples. Type II inclusions are dominated by a vapor phase and
contain less than 25 percent aqueous brine of low salinity (Fig. 33b
and 33i). Inclusions of this type represent trapping of a high-
temperature H20 vapor which later condensed during cooling. Type III
inclusions contain three phases; a highly saline (saturated) brine, a
low density vapor bubble, and usually one daughter mineral (Fig. 33c-
h). These inclusions were formed by trapping of a saturated aqueous
brine, which became oversaturated as it cooled, causing the precipi-
tation of a daughter crystal. The most common daughter mineral
encountered was halite, which indicates that NaC1 was the dominate
constituent of the brine. Apparently there was insufficientK in
solution to form discrete crystals of syl vite, although potassium
chloride probably is present as a minor constituent Other daughter
products seen infrequently include a long rectangular birefringent
mineral that may be anhydrite(?), or a form of iron chloride, and a
rhombohedral birefringent mineral that may be a carbonate. These
inclusions will be discussed subsequently.
The homogenization temperature (Th) is the temperature at which
the contents of the inclusion homogenize as measured on the heating
stage. The trapping temperature (Tt) is the true temperature of the
fluid when it was enclosed during mineral growth (Cunningham, 1977).
Thus, the homogenization temperature, without a correction for pres-
sure, is only the minimum trapping temperature. Accordingly, there
183

is usually a difference between these two temperatures, which results


from pressure gradients (either lithostatic or hydrostatic) that may
have existed during the formation of the fluid inclusion. Methods of
determining the necessary pressure correction for temperature are
discussed later.
The fol lowing sections are arranged to provide first, descrip-
tive information of the inclusions, followed by homogenization and
salinity data. Interpretation of these data are summarized at the
end of the chapter. The terminology, methodology, and intrinsic
assumptions used in fluid inclusion studies have been thoroughly
discussed by Roedder (1967; 1972; 1976; 1977), Nash (1973), and
Cunningham (1977). The suggested terminology of these authors will
be used in the following discussion.
Chumpe Quartz Porphyry

The Chumpe Quartz Porphyry originally ws sampled for the pur-


pose of characterizing the alteration mineral assemblage. However,

numerous fluid md usions were observed in magma tic quartz pheno-


crysts during routine petrographic examinations, which lead to the
preparation of polished plates of samples containing abundant quartz.
Systematic sampling was not undertaken because inclusions were not
known to be present in the porphyry prior to this investigation.
However, important compositional and temperature data were obtained,
which may reflect the P-T-X conditions present during hydrothermal
alteration of the quartz porphyry and subsequent mineralization of
the veins.
184

The quartz phenocrysts were found to contain fluid inclusions of


various types and sizes. The shapes of the inclusions are highly
variable, but round and negative crystal forms are the most prey-
alent (Fig. 33, d-h). Many phenocrysts contain numerous fracture
planes filled with secondary inclusions, although areas of singular
or multi-inclusion patches also are common. It was not possible to
conclude with certainty that any of the inclusions were of primary
magmatic origin because of the large number of secondary fractures
present. Some primary(?) inclusions were found adjacent to minute
crystals of zircon and apatite included in the quartz (see Fig. 33a).
These also may have formed in secondary fractures which have been
subsequently annealed. Solitary inclusions may be of primary origin,
or they may represent isolated remnants of well-healed fractures. It
is also possible that many inclusions were reopened and filled with
a secondary fluids during episodic events of fracturing. However,

fluid inclusions present in the phenocrysts of quartz should repre-


sent part of the alteration and cooling history, albeit somewhat
fragmentary and incomplete.
Homogenization Data

Six polished plates from four samples were prepared. Of these,


two samples contained usable inclusions (Polv R-2 and MB-78-81).
Each plate was broken into small chips containing one to three quartz
phenocrysts that would fit into the sample chamber without creating a
large thermal gradient. The homogenization temperatures of 78 fluid
185

inclusions were determined in 25 individual runs on the heating


stage. The results of these runs are surnarized in Table 28, and a

histogram of the homogenization temperatures is shown in Figure 35.


The inclusions that appeared to have the lowest homogenization
temperatures were measured first, where possible, to avoid stretching
or decrepitation. The data for individual inclusions were generally
reproducible to + 1.5°C, as measured in repeat runs of the same
inclusion. The accuracy of the data decreased slightly with increas-
ing temperature because of poorer visibility. All homogenization
temperatures for type I Inclusions were measured precisely when the
bubble visually disappeared. Although vapor bubbles did tend to move
into the dark regions of total reflection as they were heated, it was
possible with the aid of an incident light beam, to successfully
measure many of these inclusions (a fiber optics illuminator was not
available to the author). As described by Roedder (1971), homogeni-
zation temperatures of type II inclusions (gas-rich) are extremely
difficult to measure precisely owing to the small amount of fluid
present in the inclusion. Accordingly, homogenization temperatures
of type II inclusions have an accuracy of approximately + 10°C. Two

samples decrepitated within the temperature range used in this study


(< 450°C).
Fluid inclusions which contain daughter minerals (type III) are
more complex because of an increased number of phases. The homogeni-

zation temperature for type III inclusions is the point at which only
one phase exists in the inclusion. However, specific phase transi-
Table 28. Summary of homogenization temperatures and salinities for fluid inclusions.

Type of Type of Number Homogenization1 Equivalent2


Sample Chip I Deposit Mineral Inclusion Tested Temperature °C MaCi (wt. %) Notes

118-11-81 3 vein qtz S(?)-I 8 257-261 5.5


118-11-81 4 vein qtz p-I 9 305-314 6.4-6.5
MB-21-81 1 vein si P-I 1 197 6.5
MB-21-81 2 vein si 5(7)-I 1 167 6.8
118-21-81 3 vein si S(?)-I 1 159 6.8
MB-38-81 4 manto sl p-I 7 250-278 7.1
MB-74-81 1 manto si p-I i 250 6.2
1413-74-81 2 manto si P-I 6 242-272 5.8-6.4
148-74-81 5 manto si P-I 6 231-277 4.3-7.2
MB-74-81 6 manto si p-I io 244-285 4.0-8.0
118-78-81 1 intrusive qtz S-I 9 380-395 6.4-14.9(8)
118-78-81 4 intrusive qtz S-I 2 315,365 9.0(1)
MB-78-81 4 intrusive qtz S-Ill 4 311-348 29.5-30.0 Tm(halite)=145-155°C
MB-78-81 5 intrusive qtz S-I 1 400 13.5
1413-78-81 5 IntrusIve qtz S-Il 8 373-381 2.7-2.9 Th"Th 1-V(V)
1413-78-81 6 intrusive qtz S-I 7 379-405 20.0-20.2 Te(Ice)=-42.8°C (CaC12)
Poiv 13-2 1 intrusive qtz S-lI 6 365-375 lID Th"Th L-V(V); Dx?
Poiv 13-2 1 intrusive qtz S-Ill 1 380 27.0? Dx"anhydrite?
Te(ice)-43.5°C.
Poiv R-2 2 intrusive qtz S-Ill 3 289-309 41.0-49.0 Tm(haiite)346-435°C.
Tm(syivite?)116°C.
Poiv R-2 3 intrusive qtz S-I 4 327-410 8.8-16.7(2)
Poiv R-2 3 intrusive qtz S-I 1 345 26.5? Te(ice)=-45.0°C(CaCi2?)
Tm(lce)=-35.0°C.
Poiv R-2 5 intrusive qtz S-I 10 370-419 11.2-16.0(7)
Polv 13-2 6 intrusive qtz S-I I ND 15.9
Poiv 13-2 6 intrusive qtz S-I 6 387-412 10.4-15.9(5)
co
0i
Table 28 con't

Type of Type of Number Homogenization Equivalent


Sample Chip # Deposit Mineral Inclusion Tested Temperature °C NaC1 (wt. %) Notes

Poiv R-2 6 intrusIve qtz S-Ill 7 305-358 33.0-39.0(6) Tm(halite)234-410°C.


Poiv R-2 7 intrusive qtz S-Ill 2 338, 360 44.5, 45.5 Tm(haflte)°374, 390°C.
Poiv R-2 12 intrusive qtz S-Ill 4 318-348 32.0-47.0 Tm(haltte)=196-412°C.
SCM-3 1 vein qtz 5-1 5 260-269 6.8-7.1
SCM-l0 I vein 51 p-I 6 194-197 8.0
SCM-lU 2 veIn si P-I 9 180-186 7.5-7.6
SCM-ID 4 veIn Si p-I 1 171 6.0
SCM-10 7 vein sl P-I 1 192 6.8
HPA 11-4 1 manto si P-I 10 270-274 5.6-6.2
HPA 14-6 2 manto si P-I 4 238-241 6.8
HPA 14-6 3 manto si P-I 6 255-274 6.1
HPA 11-6 4 manto qtz P(?)-I 4 272-290 4.0
HPA 11-6 4 manto Si P-I 4 235-236 7.0
HPA 11-6 5 manto si P-I 1 238 6.0
11-2 1 vein Si P-I 8 192-208 7.6
11-8 1 manto si P-I 2 245-246 6.1
11-8 4 manto si P-I 1 268 7.0

1 homogenizatIon temperature is given for vapor to liquid phase transition (Th°Th L-V(V)).
2 Number in parenthese equals the number of salinity measurements if different from number of samples tested.
Abbreviations: qtz°quartz; si=sphaierite; P=prlmary inclusion, S°secondary Inclusion, I=type I Inclusion, II°type II inclu-
sion, llI°type Ill inclusion; N0°not determined; Dx=daughter crystal, 1°liquid, V=vapor, Te=eutectic melting temperature,
Th=homogenization temperature, Tm=melting (dissolution) temperature.
See Appendix, Table Al for sample locations and descriptions.

I-.
co
20

U,

0
U,

C.,
C

0
I.-

E
z

ri

280 320 360 400

Temperature (°C)

Figure 35. Histogram of fluid inclusion homogenization temperatures


in quartz from the Chumpe Quartz Porphyry.
tions are also important, and temperatures for these are given separ-
ately in Table 28. For example, Th L-V (L) is the homogenization
temperature, of the two phases liquid and vapor, into the liquid
phase. Likewise, Tm (halite) is the dissolution temperature of the
halite daughter mineral.
Measured fluid inclusion homogenization temperatures for quartz
phenocrysts in the Chumpe intrusive ranged from 305° to 434°C. In-
clusions which contained less than 23.3 weight percent NaC1 equiva-.
lent homogenized in a relatively narrow range from 3600 to 410°C,
whereas type III inclusions saturated with NaCl (halite daughter min-
eral present) varied from 305° to 434°. Type II gas-rich inclusions
are common in all samples, but those measured grouped tightly around
375°C. As mentioned previously, these inclusions are highly (but not
conclusively) indicative of boiling, or trapping of a vapor phase.
Boiling probably occurred over a wide range of temperatures and
additional analyses of type II inclusions might substantiate this
hypothesis.
Decrepitation during heating runs was observed in only two
inclusions, and took place approximately 25° to 40°C above their
respective homogenization temperatures. The apparent cause of de-
crepitation, in at least one case, was the presence of a fracture
near the surface of the inclusion. Problems with leakage or stretch-
ing of inclusions were not encountered in any of the other analyses,
as supported by repeat measurements on most groups of inclusions.
190

Type III inclusions containing daughter minerals not observed


elsewhere were found in sample Polv R-2, chip 1. These included a
prismatic, birefringent daughter mineral which may be anhydrite,
molycite (FeC13), or lawrencite (FeC12) (T. Theodore, 1983, pers.
commun.) and are illustrated in Figure 33g and 33h. Additionally,
this daughter crystal was found in a gas-rich inclusion (Fig. 33i)
which suggests subsequent reopening or leakage of the inclusion.
Homogenization temperatures of both inclusions were between 3700 and
3800C, although an error of + 20°C is likely for the vapor-rich
inclusion. There was little apparent dissolution of this daughter
mineral at the vapor-liquid homogenization temperature, even though a
temperature of 4000 was maintained for an hour. A carbonate-bearing
type III inclusion was found in a thin section of the same sample.
Unfortunately, carbonate-bearing inclusions were not found in
polished inclusion plates. Identification as a carbonate is support-
ed by the extremely high birefringence of the crystal (at a thickness
of < 3 microns) and its rhombohedral shape. One inclusion may have
contained a daughter crystal of sylvite(?), in addition to halite,
that dissolved at 1160C and did not reappear upon cooling. This
mineral was isotropic and had a cubic shape.
Salinity Data

Freezing point depression temperatures were measured on 43 in-


clusions used for homogenization runs to determine their salinity.
An additional 22 salinity determinations were made by comparing the
solution temperatures of halite daughter minerals with the data of
191

Sourirajan and Kennedy (1962). The results are summarized in Table

28 and illustrated by histogram in Figure 36. The accuracy of the

freezing point depression temperatures is approximately+1.O°C as

determined from multiple runs on the same inclusion, although most

melting temperatures were reproducible to O.20C. Salinities from

freezing point depression measurements of fluids less saline than the

eutectic melting temperature of ice were calculated using the data of

Potter and others (1978). Inclusions which contained saturated NaC1

solutions (23.3 to 26.3 weight percent) formed crystals of NaCl2H20

(hydrohalite) upon freezing as expected from the NaCl-H20 binary

system illustrated in Figure 37.

The salinity measurements and daughter crystals suggest that the

fluids were dominantly NaCl-rich brines, although eight inclusions

did exhibit melting temperatures far below that expected for a pure

saturated NaC1 solution. Additionally, these inclusions also showed

eutectic melting points that were extremely low (Te = _300 to -45°C).

These low temperatures suggest a mul ticomponent brine that may con-

tain KC1, MgCl2, and/or CaC12 in addition to NaC1. According to

Crawford (1981), the addition of KC1 (maximum of 5 %) to a NaC1-H20

brine would lower the eutectic point only 2.1°C from -20.8°C to

-22.9°C. However, the addition of CaC12 or MgC12 would depress the

eutectic temperature to -52.0° and -35°C, respectively. Because some

inclusions in the present study show Te points as low as -45°C, it

was assumed that the fluids were dominated by NaC1 and CaCl2, rather

than KC1 and MgC12, although all four halides probably are present.
192

20

U)
C
0
U,
=12
C.)
C

.0
E
z

0 10 20 30 40 50

Salinity (wt.% NaCI equiv.)

Figure 36. Histogram of fluid inclusion salinities in quartz


from the Chumpe Quartz Porphyry.
+1 11 I)
NaC I + SOLUTION

0
0

Ld

-I NCI 2H20 + SOLUTION


(ii

LU NaCI .2H20+NaCI
I
-2

-3O°L_
- -
0 60 61.9 65 100
H20 Na CI

WEIGHT PERCENT NaCI

Figure 37. The system sodium chloride-water for temperatures below + 10°C (from Roedder, 1962).
I.-.

(A)
194

Quantitative chemical analyses of the fluid were not attempted. The

total salinity of these multicomponent brines was estimated by extra-


polation of the relative sal mi ties for the system CaC12-NaC1-H20
from the data of Yanatieva (1946). High CaC12 compositions would
support the presence of anhydri te as a daughter mineral , if suffi-
dent S042 is present in the fluid.
Type I inclusions generally showed behavior typical of a non-
saturated NaC1 brine, with the exception of the previously discussed
CaC12-bearing inclusions. Salinities ranged from 7.0 to 20.0 weight
percent MaCi equivalent in inclusions with homogenization tempera-
tures between 3600 and 4100C. Most type I inclusions exhibited
eutectic mel ting temperatures near -20.8°C. The eutectic mel ting
temperature (Te) is exceedingly difficult to measure accurately be-
cause of the small amount of liquid formed at this temperature.
Salinities of type II gas-rich inclusions were found to be low
(< 2.0 weight percent NaCl equivalent) as would be expected from a
vapor derived by boiling of a NaCl brine. These inclusions also were
difficult to freeze, probably because of the lack of nucleation
points and paucity of fluid.
Type III inclusions show extremely complex phase behavior upon
freezing. Roedder (1971) has indicated that most of these phase
changes are not readily amenable to interpretation because of poor
compositional controls. He further suggests that freezing these
inclusions is worthwhile only if the inclusion contains a significant
KC1 content Nonetheless, several type III inclusions were frozen
195

to determine if any significant information could be obtained.


Because the inclusions are saturated with NaC1, it was necessary to
hol d the temperature at -80.0°C for several minutes to freeze them.
A coating of NaC12H20 (hydrohalite) formed on all halite cubes upon
freezing, which protected them from further reaction with the liquid.
Hydrohalite persisted to +0.1°C, its melting point, on warming. The

ice that formed on freezing melted between _28.00 (metastable eutec-


tic) and the eutectic point (-20.8°C).
From the available evidence, it would appear that the Chumpe
Quartz Porphyry was subjected to hydrothermal fluids which varied
appreciably in composition and temperature. The mean homogenization

temperature of 375°C is in good agreement with the stability of the


quartz-sericite alteration mineral assemblage. It is probable that
these fluids were responsible for most of the alteration present in
the quartz porphyry and adjacent country rocks.
Vein-Type Ore

As mentioned previously, the three main periods of vein mineral-


ization are represented by 1) pyrite-quartz-wolframite (2) pyrite-
chal copyri te-sphal en te-gal ena and (3) ban te-carbonate-marcasi te.
Polished plates representative of all three mineralization periods
were prepared, but usable inclusions were found only in quartz and
sphalerite. Very small inclusions (<3 microns) were seen in the
late-stage carbonate and sphalerite assemblage, but measurements on
these were unobtainable. Five samples contain workable inclusions.
196

Samples MB-11-81 and SCM-3 represent the quartz-wolframite stage,


whereas MB-21-81, SCM-1O, and M-2 are samples of sphalerite from the
second stage of mineralization (Table 28). All were collected from
the main San Cristobal Vein, with the exception of M-2.
Quartz is typically milky white in color because of numerous
extremely small inclusions. Two samples of quartz yielded 23 usable
fluid inclusions. Of these, 10 inclusions appeared to be primary and
the remainder are probably secondary or pseudosecondary in origin.
All are small liquid-rich type I mci usions which contain a vapor
bubble that occupies 25-35 percent of the inclusion volume. Type II
and type III inclusions were not observed in any of the polished
plates.
The vast majority of the vein sphalerite is inordinately dark
and finely banded. Although some inclusions were observed in these
sphalerites, most proved to be so small and poorly illuminated that
measurements were unobtainable. The paragenetically later, amber-
colored sphalerites contain excellent, although not abundant, type I
inclusions. A total of 28 inclusions from 3 samples were measured.
All contain a vapor bubble which occupies between 20 and 30 percent
of the inclusion volume as shown in Figure 34a.
Homogenization Data

Primary fluid inclusions in crystals of quartz homogenized in a


temperature range between 2900 and 3150C, whereas secondary inclu-
sions had lower homogenization temperatures which clustered between
2500 and 265°C as shown in Figure 38. It is apparent that these
197

secondary inclusions are related to a later episode of fracturing,


and correlation of these to another period of mineralization will be
discussed later.
Primary inclusions in sphalerite had homogenization temperatures
which ranged from 159° to 208°C, but the majority were between 185°
and 205°C. The data for both vein quartz and sphalerite are pre-
sented in Table 28, and a histogram of homogenization temperatures is
given in Figure 38.
Results from this study are in excel lent agreement with the data
of Campbell and Rye (1982) for the same vein. Sample M-2 is from
Vein 722 in the Huaripampa Mine. Homogenization temperatures in this
vein appear to correlate well with those in the Main San Cristobal
Vein. Evidence for periodic episodes of boiling are lacking in the
fluid inclusions examined. Apparently, there was sufficient hy-
drostatic/lithostatic pressure to prevent boiling of the hydrothermal
fluids.
Salinity Data

Freezing point depression measurements were determined for all


inclusions in quartz and sphalerite. The salinity data are given in
Table 28, and are listed in equivalent weight percent JaCl corre-
sponding to the recorded freezing point depression. Deviations from
expected freezing behavior were not encountered in any of the inclu-
sions and all showed eutectic mel ting temperatures near that expected
of ideal NaC1-H20 fluids.
198

'I,

0
U)

C.)
C
1I

0
a)
-o

170 2)0 250 290 330


Temperature (°C)

Figure 38. Histogram of fluid inclusion homogenization temperatures


from vein sphalerite, vein quartz, manto sphalerite and
manto quartz.
199

Fluid salinities in both quartz and sphalerite varied slightly


between 5.0 and 8.0 weight percent NaCl equivalent. Estimates of the
concentration of other ions would be highly speculative.
Man to-Type Ore

Quantitative data for 63 inclusions from five samples of the


manto ore in the Huaripampa Mine are presented in Table 28. Efforts
to locate inclusions in other samples were frustrated by the extreme-
ly dark color of the sphalerite. All data in Table 28 are from
primary inclusions; secondary inclusions were abundant but most
showed significant evidence of necking. mci usions (< 3 microns)
were observed in the late carbonate-sphalerite assemblage, but deter-
minations were unobtainable because of the small size and problems
with ice nucleation. All inclusions in sphalerite are the type I
variety, and are representative of the second period of sulfide
mineralization in the mantos (Fig. 34, b-j).
Crystals of quartz contain few inclusions of sufficient size for
study. Furthermore, because the deposition of quartz appeared to
have been episodic over time, its exact paragenetic relationship to
the ore is uncertain. Measurements were obtained on four inclusions
found in quartz (HPA M-6, chip 4, Table 28) that was deposited con-
temporaneously with or slightly later than pyrite.
200

Homogenization Data

Homogenization temperatures for primary fluid inclusions in


sphalerite ranged from 230° to 285°C. This range represents the
depositional temperature of 95 percent of the sphalerite in the
Carahuacra P4anto. The other 5 percent was deposited in the late
carbonate stage at lower temperatures. Type II inclusions were not
observed in any of the polished plates of the manto ores. All homog-
enization temperatures were, on repeat analyses of the same inclu-
sion, reproducible to + 1.00C. Leaking, decrepitation, or stretch-
ing of the inclusions were not observed in any of the heating anal-
yses.

The mean homogenization temperature for manto sphalerite inclu-


sions is 260°C. The difference in homogenization temperatures be-
1ieen vein and manto sphalerites is significant and may be related to
chemical depositional controls which are discussed subsequently.
The four homogenization temperatures for quartz varied from 268°
to 290°C. These temperatures overlap those of the sphalerite, but
are higher than the mean homogenization temperatures for sphalerite.
Al though only four determinations were made, it is apparent that the
quartz was deposited prior to, or contemporaneously with the early
sphaleri te.
201

Salinity Data

All inclusions in sphalerite and quartz of the mantos have


Tm(ice) temperatures of _2.40 to -5.1°C, which correspond to a sa-
lini ties in the range of 4.0 to 8.0 weight percent equivalent NaC1.
Accuracy of the mel ting point depression deteniiinations is + 0.4°C.
Although this range of salinities is small, there is considerable
variation in the salinities of inclusions from a single sample.
Deviations from expected freezing point behavior were not observed in
any of the inclusions.
Pressure Corrections

Pressure corrections for homogenization temperatures are neces-


sary to obtain the true filling temperature of an inclusion. If a
true estimate of the pressure during ore deposition can be made, the

depth of mineralization can be determined. A number of methods have


been used in the past to arrive at both the pressure and depth of
mineralization. Roedder and Bodnar (1980) have reviewed these
methods and suggested additional means to estimate pressure and depth
of mineralization from the fluid inciLsion data, and they also review
the general principles and assumptions necessary to obtain these
pressure estimates.
The only case in which pressure corrections to fluid inclusion
data are unnecessary is when the fluid was on the liquid-vapor
(boiling) curve during entrapment (Roedder, 1971). At these points,
the vapor pressure of the fluid is equal to the total pressure of the
202

system. However, if the total pressure is greater than the vapor


pressure of the fluid at the moment of trapping, a temperature cor-
rection has to be added to the homogenization temperature.
Chumpe Fluid Inclusions

Fluid inclusions in phenocrysts of magmatic quartz from the


Chumpe Quartz Porphyry show the widest variations in composition,
temperature, and inclusion type. Inclusions indicative of boiling
are abundant (Fig. 33 b and h), but it was impossible to determine
cogenetic relationships with the highly saline inclusions. Estab-
lishmentof this cogenetic relationship is necessary to conclude,
with certainty, that the system was boiling, and that fluids of a
particular composition were the source of the low-density type II
inclusions. If the saline inclusions are cogenetic with the type II
gas-rich inclusions, then the total pressure on the fluid is equal to
the vapor pressure. Depending on the composition of the saline
inclusions, the vapor pressure can be determined using the data of
Sourirajan and Kennedy (1962) and Haas (1971) for the NaCl-H20 sys-
tem. According to Cunningham (1976), the vapor pressure estimates in
brines containing Na-K-Ca, are lower than those of the simple NaCl-
H20 system. This difference creates a small error (-i- 5%) when using
a simplified fluid composition. Fluids associated with hydrothermal
al teration of the Chumpe Quartz Porphyry show a considerable varia-
tion in salinity. Three major populations of salinities can be seen
in Figure 36 and which are clustered at <5, 8-20, and >25 weight
percent NaC1. Periodic boiling took place in these fluids and is
203

represented by the type II inclusions containing less than 5 weight


percent NaC1. Dissolution temperatures of halite in Type III inclu-
sions were both below and above the homogenization of vapor into
liquid. According to Roedder and Sodnar (1980), the difference be-
tween the solution temperature of halite and the homogenization
temperature of vapor to liquid (Tm(halite) >Th L-V (L)) is most
likely caused by pressure effects imposed on the system during trap-
ping. Ten inclusions, with compositions ranging from 39 to 49 weight
percent NaC1 equivalent, exhibited this behavior in the present
study. According to Cunningham (1976), an inclusion containing 40
weight percent NaC1 equivalent, which homogenizes at 3500C, requires
a minimum pressure of 110 bars to prevent boiling (extrapolated to
higher salinities using the data of Haas, 1971). For the maximum
salinity recorded in this study (48 weight percent NaC1 equivalent at
430°C), the minimum pressure would have to be approximately 250 bars.
These pressures are valid if the solution is boiling, but they could
be any greater value if the solution was not.
Minimum pressures were calculated for the same 10 inclusions
using the method for high-salinity, multiphase inclusions described
by Roedder and Bodnar (1980, p. 281-287). This method does not
require boiling, but it does assume that the difference between Tm
(halite) and Th L-V (L) is the result of pressure. Thus, a minimum
pressure of 800 bars was obtained for the inclusion which contained
49 weight percent NaC1 and had Tm (halite)=430°C, with Th L-V
(L)290°. This pressure is the highest calculated for any of the
204

inclusions; others had calculated pressures of less than 650 bars and
most were in the range of 300-400 bars. The validity of these pres-
sure calculations is questionable. They appear to be excessive based
on reconstruction of the local geologic section and the fact that
several inclusions (Fig. 33 e and f) contain large halite crystals
and very little liquid. This paucity of liquid phase may be indica-
tive of reopening and leakage in these inclusions.
For inclusions of moderate salinity (10-20 weight percent NaCl
equivalent), pressure determinations can only be estimated using the
solution-vapor curve and the data of Haas (1971) extrapolated to
higher temperatures. These inclusions necessitate minimum vapor
pressures of 100 to 200 bars.
The previous pressure determinations are based on the hydro-
static model developed by Haas (1971), who thoroughly considered the
limitations of his data and their application to fluid inclusion
geobarometry. These pressure estimates obtained for the San
Cristobal samples indicate a depth of formation of between 900 and
4800 m, based on this hydrostatic model.
Pressure corrections for fluid inclusion homogenization data
were extrapolated from the curves prepared by Potter (1977). Temper-

atures from these curves are added to the homogenization temperature

to obtain the filling temperature for the inclusions. The type III
inclusions require a 500 to 80°C temperature correction, whereas the
moderately saline type I inclusions require a smaller additive cor-
rection of only 100 to 15°C.
205

Vein and Manto Fluid Inclusions

Fluid inclusions in both the vein and manto-type ores are simi-
lar in composition. Homogenization temperatures for both types range
from 150 to 310°C, and salini ties never exceed 8 weight percent NaCl
equivalent. Neither ore type contains inclusions that exhibit evi-
dence of boiling, despite the high fluid temperatures and low salini-
ties. Therefore, sufficient hydrostatic or lithostatic pressure
existed at the time of mineralization to prevent boiling. These

fluids would be at, or below, the liquid vapor curve corresponding to


their composition. If they were on the curve, then a minimum hydro-
static pressure can be estimated using the data of Haas (1971). For

a fluid of 10 weight percent NaCl, the minimum pressure would be 4.4


bars (35 m depth) for a fluid trapped at 150°C and 92 bars (1050 m
depth) at a temperature of 3100C (Haas, 1971). Again, these are
minimum pressures; the fluids could have been trapped at any higher

pressure. This is the only pressure constraint which can be applied


to these two phase, liquid-vapor inclusions (Roedder and Bodnar,
1980).

Another common practice used to determine the pressure, or depth


of mineral ization, is to estimate the amount of overburden at the
time of mineral deposition, and then extrapolate pressure using a
hydrostatic or lithostatic model. This method may produce spurious

pressure estimates because of uncertainties in the reconstruction of


the geologic column at some time in the past. Most vein and geo-
206

thermal systems are probably open to the surface, so the system is


probably dominated by hydrostatic pressure. Deteminations of pres-
sure based on a column of water (or rock) require several assumptions
on the physical and chemical conditions of the system (see Roedder
and Bodnar, 1980).
Because of the assumptions needed to apply the hydrostatic model
of Haas (1971), and the fact that these pressures are minimums (since
no evidence of boiling could be found in the vein or mantos), only a
minimum temperature correction for pressure can be applied to the
vein and manto data. This correction amounts to 0°C at a 150°C
homogenization temperature, and at most 10°C for homogenization temp-
era tures near 3000C. If the true pressure of the system was 200
bars, the pressure correction for homogenization temperatures ob-
tained would be no more than 20°C.
Interpretation of Data

The fluid inclusion data reveal a pattern of both spatial and


temporal changes in fluid composition and temperature typically found
elsewhere in many large hydrothermal ore deposits, especially the
porphyry Cu-Mo deposits (Roedder, 1971). In accordance with metal
zonations and fluid inclusion homogenization temperatures, ore de-
posits of the San Cristobal District appear to be centered on the
Chumpe Quartz Porphyry which contains high-temperature, high-salinity
inclusions. An abrupt decrease in temperature and fluid salinity is
evident from comparisons of the data between inclusions in the quartz
porphyry (8-49 weight percent NaCl equivalent) and those in the vein
207

and manto ores (4-8 weight percent NaC1 equivalent). However, the
fluid inclusions in sphalerites of the mantos have significantly
higher homogenization temperatures than do those in the veins. The

discussion that follows is based on the preceeding observations, and


concerns (1) the chemistry and temperature of the fluids responsible
for quartz-sericite alteration in the Chumpe Quartz Porphyry, (2) the
abrubt change in composition and temperature of the hydrothermal
fluids between the intrusion and ores, (3) the discrepancies between
vein and manto sphalerite homogenization temperatures, and (4) the
hypothesized syngenetic origin for the manto ore deposits.
The presence of highly saline type III inclusions in hydrotherm-
ally altered porphyritic igneous rocks is common (Roedder, 1971,
1976; Cunningham, 1976). It has been shown through experimental
studies that aqueous fluids in contact with a c'rystal 1 izing magma
will be preferentially enriched in total chlorides (Burnham, 1967,

1979; Holland, 1972; Gammon and others, 1969; Cloke and Kesler,
1979). Therefore, the evolution of a highly saline aqueous phase is
considered by most researchers to be a normal consequence of magma
crystallization. Roedder (1971) has shown that numerous inclusions
from the porphyry copper deposit at Bingham, Utah, have homogenized

at temperatures approaching magmatic conditions, and he has sug-


gested that some of these saline inclusion formed as a result of
late-stage immiscibility with the siliceous magma.
As previously noted, fluid inclusions in the Chumpe Quartz
Porphyry are probably of secondary origin and related to repeated
periods of intense fracturing. The homogenization temperatures re-
corded therein are well below magmatic conditions (6000 to 1000°C),
even if a 100°C pressure correction is applied. These highly saline
fluids probably originated through imiscibility with the magma which
formed the exposed dike system, and were subsequently released during
fracturing of the main pluton, when the fluid pressure in the crys-
tallizing magma exceeded the lithostatic pressure of the overburden
and the carapace of the intrusion. This process has been described
in detail by Burnharn (1979). As the fluids are released, they mi-
grate upward through fractures and along conduits adjacent to pre-
viously emplaced dikes. They become trapped during cooling with the
precipitation of secondary quartz. The wide range of homogenization
temperatures (300°-450°C) and the large variety of inclusion types
(including the gas-rich type II) presumably reflect episodic frac-
turing and release of fluids from the magma.
Magmatic fluids which migrate appreciable distances from their
source are unlikely to be in equilibrium with their new surroundings.
The pervasive alteration of the Chumpe Quartz Porphyry attests to
this disequilibrium. Most investigators believe that the quartz-
sericite alteration assemblage results from the influx of heated
meteoric waters (groundwater) as the convecting hydrothermal cel 1
collapses in on its thermal source. The fluid inclusion evidence
from San Cristobal does not conflict with this premise. Moreover,
209

because 95 percent of the inclusions in the Chumpe Intrusive are of


the moderate-salinity type I variety, they may represent a mixture of
magmatic and meteoric waters. The mean homogenization temperature of
370°C is well within the field of mineral stabilities for quartz-
sericite alteration. However, Burnham (1967, p. 73) states:
"The cooling of these chloride-bearing hydrothermal
solutions below the solidus temperature of the magma, in
contact with the parent rocks or other quartzofel dspathic
rocks, will result in a marked reduction in the K+ + KC1
content and upon entering the stability field of muscovite
(about 640°C at 1000 bars fluid pressure), an even more
marked reduction in H+ + HC1. Inasmuch as the total chlor-
ide concentration in the fluid remains unchanged, there a
concomitant marked increase in the Na++ WaC1 and Ca +
CaC12 contents. These changes in aqueous phase composition
are brought about, of course, by reaction with the wall
rocks, and are reflected mineralogically in the deposition
first of orthoclase, with or without biotite, and then
muscovite, all mainly by replacement of plagioclase. Cool-
ing of these solutions from magmatic temperatures also re-
sults in the precipitation of abundant quartz; however, the
deposition of quartz extends to somewhat higher temperatures
than does enrichment in potassium-bearing minerals."
Thus, quartz-sericite alteration can be produced soley by magmatical-
ly derived hydrothermal fluids, and meteoric waters may only become
important after there is a decrease in the supply of magmatic water
or collapse of the magmatic hydrothermal system with diminishing
temperatures (Burnham, 1979).
Since the composition of the aqueous alteration fluid would be
controlled in a large part by wall rock reactions, it is necessary to
establish the contemporaneity of fluid inclusion formation with wall-
rock alteration (Roedder, 1971). This relationship is difficult to
establish unambiguously in the Chumpe intrusive, and the evidence is
circumstantial at best. However, compositions of the aqueous phase
210

in the fluid inclusions have been shown to be predominately chloride


brines with Na4, Ca2, H4, K+, and Mg24 as the major cations in
approximate order of abundance.
The district exhibits an elongate "bullseye" pattern of metal
and temperature zonations which trends parallel to the Chumpe

Anticline, and has its center over the Chumpe Quartz Porphyry. The

core contains the Chumpe intrusive that is characterized by dissemi-


nated pyrite mineralization and veins fil led with pyrite-quartz-
woiframite of the first period and highest temperature mineraliza-
tion. Outward from this inner zone, the mineral assemblage changes
to pyrite-chalcopyrite-sphalerite-sulfosalts. Farther outward, in
the periphery, the mineralization is dominated by galena and sphal-
erite and eventually gives way to just sphalerite. Campbell and Rye
(1982) have shown this progressive change in mineralization to be
representative of a temperature decline of 200°C from 300°-400°C to
less than 200°C. The data from the present study generally agree
with their results. Typically, the paragenetically later ore min-
eral s were deposi ted ci oser to the central zone as the hydrothermal
system waned, and this is apparently the case with galena in the San
Cristobal Vein.
The abrupt change in salinities between inclusions in the Chumpe
intrusive and the vein ores probably reflects the increase in the
meteoric water component of the hydrothermal fluids. From deuterium

and oxygen isotope work, Campbell and others (1983) have concluded
that the mineralizing solutions at San Cristobal were meteoric waters
211

which underwent isotopic exchange with a granitic intrusive at a very


low water to rock ratio.
As mentioned previously, there is a significant discrepancy
between the homogenization temperatures of fluid inclusions in the
vein and manto sphalerites. The mean temperature for Inclusions in
the vein sphalerite is 195°C, whereas those of sphalerite in the
manto show a mean homogenization temperature of 265°C. This result
is surprising in view of the spatial zonation of ores in the dis-
trict. However, a probable explanation of this observation is not
inconsistent with ore depositional theory.
Ore deposition has been postulated to take place in response to
a tremendous variety of processes that include boiling, dilution of
ore fluids by groundwater, temperature and chemical changes, etc. In
terms of the chemistry of the mineralizing fluids, it should be
noted that the salinity and apparent overall composition varies only
slightly between the vein and manto ores. Temperature changes caused

by addition of cooler meteoric water would not readily account for


the lower homogenization temperatures in the vein, which is closer to
the thermal source. However, a change in chemistry accompanied by
concomitant changes in the stabilities of the metal-chloride com-
plexes is possible.
Burnham (1967) has demonstrated from experimental studies that
an abrupt decrease tn the H+ and HC1 content of a hydrothermal solu-
tion, such as might occur upon encountering carbonate-rich rocks,
would lead to large scale precipitation of Fe, Zn, Cu, and Pb sul-
212

fides. The ore fluids in the vein would have been buffered to some

extent by the wallrocks until they were altered. to a stable assent-


blage and/or were armored by mineral deposition. The first period of
mineral deposition (quartz-pyrite-wolframite) produced most of the
wallrock alteration early in the depositional history of the vein.
Pyrite and quartz are paragenetically early in both vein and manto
ores, and the fluid Inclusion temperatures are roughly equivalent
(300-350°C in the veins; 280-300°C in the mantos). During the second
stage of mineralization, when sphalerite was deposited, the hydro-
thermal fluids may have been excessively hot to deposit sphalerite in
the veins, especially if there was not a means to alter the fluid
chemistry lie, through boiling or wallrock reactions). However,
because of a large reservoir of carbonate in the host rocks, sub-
sequent reactions caused a decrease in the H+ and HC1 content of the
fluids, which in turn promoted the precipitation of sphalerlte in the
mantos at the higher temperatures. Sphalerite was not precipitated
in the veins until temperatures decreased sufficiently to destabilize
metal-chloride complexes. This process will be discussed more
thoroughly in a later section.
Finally, the syngenetic origin of the mantos, as postulated by
others, derives little support from the fluid inclusion evidence
which suggests depositional temperatures of at least 300°C for the
early quartz-pyrite stage, and decreasing to 230°C for the majority
of the sphalerite in the deposit. These manto ores are vaguely
similar in form and mineralogy to the well-known Mississippi Valley-
213

type lead-zinc deposits, which are themselves epigenetic. According


to the data of Roedder (1976), homogenization temperatures for these
deposits are generally between 1000 and 150°C, and rarely as high as
200°C. Fluid inclusion salinities in the Mississippi Valley-type
deposits generally range from 15 to 20 weight percent NaC1 equivalent
in contrast to lower salinities (< 10 weight percent NaC1) measured
for the carbonate-hosted deposits at San Cristobal. Moreover, the

higher temperatures of homogenization for the manto deposits, in


addition to geologic evidence, would appear to exclude them from
consideration as Mississippi Valley-type ore deposits. Truly syn-
genetic ores (see definition, p.6) might be restricted to volcano-
genic massive sulfide deposits such as the Kuroko type, or to the
recently described "black smokers" found on the East Pacific rise.
Roedder (1976) has sunnarized the results of fluid inclusion studies
of Kuroko-type ores. In these deposits, which are intimately as-
sociated with submarine volcanism, the homogenization temperatures
range from 80° to 315°C and salini ties are generally less than 10
weightpercentNaCl. Styrt and others (1981) have reported hydro-
thermal fluids from the East Pacific Rise with temperatures of 300 to
350°C which are presently depositing sulfide minerals. Although
these data are similar to those of the manto ores, the deep and mid-
oceanic depositional environment of the "black smokers" contrasts
markedly to the carbonate-hosted manto ores of the Pucara Group, in
close proximity to silicic quartz porphyry intrusions. The geologic
evidence negates comparisons of the syngenetic volcanogenic massive
214

sulfide ores to those contained in the manto or vein deposits of San


Cristobal.
215

SULFUR ISOTOPES

Sulfur isotope abundances were determined for the principal


sulfide and sulfate phases in a number of samples to (1) determine
the probable source of sulfur in the ore minerals, (2) provide a
pressure independant geothermometer, (3) provide constraints on the
physicochemical conditions of mineralization, and (4) provide a chem-
ical means from which to suggest separate or cogenetic origins for
the vein and manto ores.
Campbell and Rye (1982) have completed the first study of sulfur
isotope abundances in the district, but focused most of their efforts
on the main San Cristobal Vein. They reported an average ô34S value
of 4.2 per mil for pyrite from the first mineralization period (pyr-
ite-quartz-wolframite). Their average 5S values for pyrite,
sphalerite, chalcopyrite, galena, and barite from the second mineral-
ization period are 6.2, 6.3, 5.8, 4.2, and 16.5 per mu, respective-
ly. Al though they did not report data for other veins or the manto
deposits, seven samples of sphalerite and galena from the San Antonio
and Santa Aqueda (Carahuacra) Mantos were analyzed. Average
values for sphalerite and galena were 6.58 and 3.70 per mil, respec-
tively (A. Campbell, writ. conriun., 1983).

Efforts in the present study have been directed primarily toward


the manto ores, although samples from several veins including the San
Cristobal Vein were analyzed. Detailed information concerning the
location, geology, and mineralogy of each sample.is given in Table
Al, of the Apperrdix. The order of samples listed in Tables 29 and
216

30, and Figure 39, approximately represents increasing lateral and


vertical distance from the Inferred (plutonic) center of the hydro-
thermal system and, thus, probable decreasing paragenetic age. This
order Is necessarily subjective because many samples contain minerals
of differing paragenetic position. Samples MB-35-81, M-2, MB-1-81,
and MB-12-81 are vein-type ores that contain minerals of the second
mineralization period. Massive manto ores are represented by samples
HPA M-4, HPA M-1, HPA M-10, MB-4-81, MB-80-81, and MB-38-81 and these

contain minerals from both first and second mineralization periods.


Four samples (MB-10-81, SCM-4, GBX-1, GBX-2) consist of barite from
the San Cristobal Vein, and they are representative of the third
mineralization period (late second-stage of Campbell and Rye, 1982).
The remaining two samples (MB-71-81 and MB-64-81) are of the super-
gene phases, marcasite and barite.
Sulfur isotope analyses have been obtained on 30 concentrates of
sulfide (pyrite 6, sphalerfte 10, galena 8, and marcasite 1) and
sulfate (barite 5) minerals separated from 16 samples. Mineral
separates were prepared using standard methods and each concentrate
was checked for purity using the binocular microscope. The concen-

trates were essentially monomineralic and purities of better than 98


percent were obtained.
Barite concentrates were prepared, prior to isotopic analyses,
at Oregon State University by Dr. Cyrus W. Field. Sulfate-sulfur of
the barite concentrates was reduced to hydrogen sulfide in a boiling
solution of hydrochloric-hydriodic-hydrophosphorous acid as described
217

by Thode and others (1961). Silver sulfides, derived from the reduc-
tion of sulfate minerals, and the other sulfide minerals were oxi-
dized to sulfur dioxide gas using a modification of the cupric oxide
combustion method (Grinenko, 1962) at facilities of the U. S.

Geological Survey. Sulfur isotopic analyses were performed by Cyrus


W. Field and Richard H. Fifarek at the Isotope Geology Branch of the
U. S. Geological Survey, Denver, Colorado, courtesy of Dr. Robert 0.
Rye.

All isotopic data are reported as per mil deviations ('534S)


obtained from the relationship

34s 0/00 = - 1)

where Rx and Rs represent the measured and assumed S/32S ratios of


the sample and standard, respectively. Positive and negative per mil
deviations indicate enrichment and depletion of s in the sample
relative to the sulfur standard (troilite phase in the Canyon Diablo
Meteorite, 0 per mu by definition). Analytical uncertainty is + 0.2

per mil or better, as determined from multiple preparations and


analyses of the 0.S.U. secondary standard.
Isotopic data for 30 mineral concentrates and brief descriptions
of the 16 samples are listed in Table 29. The distribution of all
4s per mu values, with horizontal lines connecting values from
coexisting mineral phases of a single sample, is graphically por-
trayed tn Figure 39. The mean 53S composition of 24 hypogene sul-
fides is 4.6 per mil, and the values range from 2.4 to 8.2 per mil.
218

Table 29. Sulfur isotope per mu values (6S) for sulfide and sulfate mineral concentrates
from the San Cristobal and Huarlparnpa Mines.

Sample1 Description2 ss °ioo

M8-35-81 San Cristobal Vein; 500 in; banded sphalerite and galena.: 51 6.47
On 3.07
M-Z Vein 722; 630 in; brecciated zphalerite and galena ore 51 6.34
cemented with quartz. On 3.13
HPA M-4 Manto ore; 580 at intersection of Vein 119; massive
in Si 5.59
ruby sphalerfte and galena. On 3.17
MB-12-81 Vein in Catalina Volcanics; 580 in; pyrite replaced by Py 5.47
sphalerite and galena. 51 5.99
On 3.58
HPA F4-1 Manto ore; 580 in; massive ore with pyrite replaced by Py 8.17
sphalerite and galena. Si 7.47
On 4 22
HPA M-l0 Manto ore; 730 m; massive ore with layers of pyrite Py 6.13
replaced by banded sphalerite Si 5 53
MS-4-81 Manto ore; 780 in; massive ore with pyrite replaced 51 6.09
by sphalerite and galena. On 2.36
MB-1-81 Vein cross-cutting manto; 580 in; pyrite replaced by Py
sphalerita and galena.. 51 5.34
Gm 2.57
MB-80-81 Manto ore; 780 in; massive, vuggy ore with intergrown Py 4.52
pyrite, sphalerite, and quartz. 51 4.52
MB-38-81 Manto ore; 540 in; massive ore with pyrite replaced Py 5.05
by ruby sphaierite and galena. Si 4.67
On 2.42
MB-ld-81 San Cristobal Vein; 170 in; vuggy, late-stage barite Ba 14.33
in brecciated pyrite-quartz-wolfrainite ore.
SCM-4 San Cristobal Vein; 170 in; late-stage barite veiniet Ba 17.38
cutting chalcopyrlte-sphalerite ore.
GBX-1 San Cristobal Vein; exact location unknown; large late- Ba 15.23
stage barite crystals on galena and sphalerite.
GBX-2 San Cristobal Vein; exact location unknown; large late- Ba 19.67
stage barite crystals.
MB-71-81 Manto ; supergene(?) marcasite in open pit. MC 7.47

M8-64-81 Moises Manto; finely crystalline barite in oxidized Ba 6.48


manto outcrop.
Analyses performed by Dr. Cyrus W. Field and Richard Fifarek at the U. S. Geological Survey
courtesy of Dr. Robert 0. Rye, 1983.
1 Sample data listed in approximate order of district zonation and/or paragenesis.
2 More complete sample locations and descriptions are provided in Appendix, Table Al..
219

S 34s %
0 5 10 5 20
I I

P48-35-el
I

P4-2 0- *Triass,c Jurassic


HPA P4-4 Q._._..j? Sea Water Sulfur

P48-12-SI 0_D2
HPA P4-I

I4PA P4-tO

P48-4-81

MO-i -$1

P48-So-Si

P48-38-el

MS-la-Si j.
SCM-4
i- Magmatic Sulfur I

GOX-1

GBX-2

P48-71-81 0
P48-64-el
I

O pyrite 0 marcasite
sphalerite barite
Q galena

I I
-5 0 5 0 IS 20

Figure 39. Summary of sulfur isotope data for the San Cristobal and
Huaripampa Mines, with samples listed in general order of
orebody zonation and/or paragenesis.
220

Analyses of the four hypogene sulfates range from 15.2 to 19.7 per
mu, and the mean value is 16.7 per mil. The.niean values for pyrite,
sphalerite, and gal ena are 5.7, 5.8, and 3.1 per mu , respectively.
The data are in good agreementwith those reported by Campbell and
Rye (1982) for the San Cristobal Vein and with the data of Field and
others (1983) for the Morococha District, Peru. Although the per mu

values are relatively close to the accepted 0 per mu value for


meteoritic sulfur, and thus for mantle sulfur (magmatic sulfides),
nonetheless they are distinctly heavier.
Source of Sul fur

According to Ohmoto and Rye (1979), the majority of the 634S


values for sulfides and sulfates in porphyry copper deposits range
between -3 and +1 and between +8 and +15 per mu, respectively. How-

ever, studies by Sakai (1968) and Ohmoto (1972) have shown that the
isotopic composition of sulfur in hydrothermal minerals is strongly
control led by the fugaci ty of oxygen and the pH of the ore-forming
fluid as well as by temperature. Given geologically reasonable
changes in the above parameters, and sulfur of 0 per mil composition,
Ohmoto (1972) demonstrated that sulfides could be produced with
compositions of between -26.5 and +4 per mil. Moreover, Ohmoto and

Rye (1979) state,


"magmatic fluids with 53S of between -3 and +7 per mu are
unlikely to produ sulfide minerals such as pyrite or
sphalerite, with S values lar,9er than approximately +8
per mu at temperatures above 200 C."
Therefore, ô34S values different from 0 per mu do not necessarily
221

indicate sulfur from non-magmatic sources, and conversely values near


o per mu do not necessarily indicate a magmatic origin for the
sulfur. However, Ohmoto (1972) further concluded that values
for minerals of greater than +4 to +6 per mu need to be explained
either as a mixture of magmatic sulfur and a source of heavy sulfur,
or by fractionation of a single source of heavy sulfur.
In accordance with the above discussion, a source of magmatic or
crustal sulfur alone is inconsistent with the observed S34S values
for the manto ores. Given a syngenetic-diagenetic origin for the
mantos as proposed by Kobe (1977) and Dunin-Borkowski (1975), the
isotopic composition of sulfur should reflect its original sedi-
mentary origin. Ohmoto and Rye (1979) give the range for S34S in
sedimentary sulfides as -70 to +70 per mil, and further state that
they are typically depleted in 53S, with a mean average of -17 per
mil. Bacteriological reduction of sea water sulfate can produce
sulfides of various isotopic compositions depending upon whether or
not the system is open or closed with respect to S042 and H2S. In
systems where reduction of sulfate is slow relative to supply, such
as in deep euxinic environments, sulfides are typically depleted by
40 to 60 per mu relative to sea water sulfate and result in sulfides
with S3'S values of -40 to -20 per mu (Ohmoto and Rye, 1979). In
this case the system is open with respect to the supply of sulfate.
However, in environments where sulfate is reduced faster than it is
supplied, such as in restricted shallow-marine or brackish-water
environments, the system is considered closed to sulfate. Under
222

these conditions, with continuous removal of hydrogen sulfide in this


system through precipitation of sulfide minerals, a wide range of
4s values (-5 to +70 per mu) can be produced (Ohmoto and Rye,
1979). However, if there are limited supplies of metal to form
sulfides, the 634S values will vary from -5 to approximately the '534S

value of the sea water sulfate (+20 0/00, today). Sulfides formed in
the above system will exhibit a wide range of 634S values and have a

maximum population of 534S values depleted by approxImately 25 per


mil relative to that of the original sea water sulfate (Ohmoto and
Rye, 1979). Rye and Ohmoto (1974) also have indicated that coexist-
ing minerals will show non-equilibrium relationships among their S34S
values. Incorporation and fractionation of sulfur that is pre-
dominately depleted in 34S cannot produce the observed sulfur iso-
topic range (+4 to +8 per mil) in sulfides at San Cristobal.
Furthermore, the narrow range of ô34S values exhibited by these
sulfides in the mantos would appear to exclude sedimentary sulfide-
sulfur as a possible source of sulfur in the manto ore deposits.
The only logical remaining source of 34S-enrlched sulfur is that
of sea water or marine evaporites. The Pucara Group contains a
section of marine evaporite (anhydrite) 10 m thick at Morococha, 20
km to the north. Sporadic occurrences of anhydrite are also present
In the San Cristobal District (Campbel 1 and Rye, 1982; F. Vera, 1981,
pers. commun.). The isotopic composition of oceanic sulfate-sulfur
and that of marine evaporites has varied throughout geologic time and
within the range between +10 and +30 per mil. Sulfates deposited
223

from sea water in the Triassic have relatively narrow values ranging
from +11 to +14 per mu (Claypoole and others,. 1980). Early Jurassic
sulfates from western Idaho have values of +14.4 to +14.9 per
mu according to Field (1984, oral commun.), and those for two
samples of sedimentary anhydrite from the Pucara Group in the
Morococha District are +13.1 and +13.8 per mu, respectively (Field
and others, 1983). Experimental work by Ohmoto and Rye (1979) sug-
gest that sulfide minerals with &34S values ranging from +30 to as
low as -15 per mu can be produced from the reduction of sea water
sulfate (or evaporite) at temperatures of 250 °C or greater in the
presence of reduced iron (Fe2+) according to the reaction:

+ 8Fe2 + 10H ----- > HS + 8Fe3 + 41120

This type of reaction is important in the formation of Kuroko-type


volcanogenic massive sulfide deposits, where sea water sulfate is
considered to have been the source of reduced sulfide-sulfur (Ohmoto
and Rye, 1979). For the reaction to proceed, there must be a source
of reduced iron and sulfate-sulfur, excess hydrogen ions, and a
minimum temperature of approximately 250°C. Geologic evidence for
such conditions to have prevailed at San Cristobal is abundant.
Petrographic investigations have documented the presence of sericite,
kaolinite, and sulfide pseudomorphs after ferromagnesian minerals in
both the Chumpe Quartz Porphyry and the Catalina Volcanics. Intense
sericitic alteration of the porphyry and the surrounding country
rocks attests t the strong acidity of these hydrothermal fluids.
224

Fluid inclusion homogenization temperatures measured in phenocrysts


of quartz in the porphyry and in the ore minerals indicate minerali-
zation temperatures of greater than 200 °C. Moreover, the Pucara
Group is known to contain sulfate-sulfur with values of +13 to
+14 per mu, and Field and others (1983) have conclusively estab-
lished that heavy sulfate-sulfur was incorporated into hypogene sul-
fides at Morococha, Peru. Campbell and Rye (1982) have recently pro-
posed a similar origin for the heavy sulfur in the San Cristobal
Vein. The sulfide ores in both districts are unquestionably products
of epigenetic hydrothermal mineralization.
The preceeding discussion, based on sulfur isotopic and other
evidence, has presented a plausible argument for the leaching of
sedimentary sulfur from anhydrite in the Pucara Group, and with
reduction of this sulfate-sulfur to sulfide-sulfur and its subsequent
incorporation into epigenetic hydrothermal sulfides. This hypothesis
would appear to be viable given the large size of the San Cristobal
hydrothermal system, reactions invol ving marine sedimentary host
rocks, and the 34S-enriched isotopic compositions of sulfur in both
the sulfides and sulfates. The extent to which other sources of
sulfur contributed to the total sulfur in the system is unknown, as
it would be extremely difficult In this particular case to dis-
tinguish a magmatic sulfur component with certainty. Mineral concen-
trates have been made for both disseminated and vein pyrite from the
Chumpe Quartz Porphyry, but analyses have not been completed as yet.
However, the distribution of values around a mean value of 4.6
225

per mu would suggest mixing of two sulfur sources. Experimental


work by Ohmoto and others (1976) has shown that values of +8 to
+27 per mu are typical of pyrite formed by the reduction of sulfate
at high temperatures in the presence of reduced iron compounds. As

the mean value for San Cristobal is somewhat enriched in


relative to that expected for "magmatic" sulfur and somewhat depleted
in s with respect to that for reduced sea water sulfate, possible
mixing of sulfur from these two sources would result in an inter-
mediate value dependent upon the relative proportions of these iso-
topically distinct components. If it is assumed that the composition
of the magmatic component was 0 per mu, and that of the evapori tic
component was +14 per mu, then subeqüal (50:50) proportions of the
two sources would result in a value of + 7 per mil for total sulfur
in the system, provided fractionation of sulfur did not take place
during reduction of sulfate.
Two samples (MB-64-81 and MB-71-81) are representative of super-
gene sulfate and sulfide, respectively, on the basis of geologic,
mineralogic, and textural criteria. The 634S value of barite in
sample MB-64-81 is 6.8 per mil, is considerably lighter than the mean
value of 16.7 per mu for hypogene barite. This anomalously light
value may be the result of a unidirectional (non-equilibrium) and
quantitative oxidation of hypogene sulfide-sulfur under supergene
conditions as described by Field (1966). Possibly a similar mechan-
ism may apply for the marcasite in sample MB-71-81.
226

Isotopic Temperature Estimates

For equilibrium isotope exchange reactions, the amount of frac-


tionation between the isotopes of an element that equilibrate between
two compounds is inversely proportional to temperature and approaches
unity at high temperatures. Temperature estimates are based on the
isotopic differences between two minerals (A and B), which are ex-
pressed as delta (AAB) values, and are determined from the measured
per mil values for each mineral. The equation

= 6345A°/OO 34SB°/oo - 1,000 in c

relates the delta value () to the fractionation factor (ct of the


mineral pair. Provided the variation in the fractionation factor
with temperature are known, either from theory or experiment, the
delta value can be used as an estimate of the temperature of mineral
deposition. This validity of this method, of course, depends upon
the existence of isotopic equilibrium between the two mineral phases.
Numerous experimental data are now available for the fractionation
factors between various minerals and those derived from a recent
critical summary by Ohmoto and Rye (1979) are used herein.
The delta (A) values and calculated isotopic temperatures esti-
mates for sphalerite-galena, pyrite-galena, and pyrite-sphalerite
mineral pairs from San Cristobal are listed in Table 30. Tempera-
tures of 169° to 287°C were obtained on the sphalerite-galena pairs,
and they show a'remarkable agreement with four fluid inclusion
Table 30. Delta () values of sulfide-sulfide mineral pairs, including Isotopic temperature (°C) estimates', and fluid
Inclusion homogenization temperatures.

Sample Description AS1..Gn APy_Gn APy_Sl Temperatures(°C)

118-35-81 Vein 3 4 (188°) 18602


11-2 Vein 3.2 (202°) 203°
IIPA 11-4 Manto 2.4 (276°) 273°
118-12-81 Vein 2.4 (276°) 1.9 (460°) 0.5 (505°)
HPA ti-i tlanto 3 2 (202°) 3 9 (238°) 0 7 (384°)
IWA 11-10 Manto 0 6 (437°)
148-4-81 tlanto 3.7 (169°)
118-1-81 Vein in manto 2 8 (235°) 2 1 (424°) 0 7 (384°)
1111-38-81 Nanto 2 3 (287°) 2 6 (353°) 0 4 (5970) 261°

1 Isotopic temperature estimates calculated from the data of Ohmoto and Rye (1979). UncertaInty Is t 25°C for
sphalerite-galena, ± 20°C for pyrite-galena, and ± 40°C for pyrlte-sphalerite based on an analytical uncertainty
of ± 2 per mil for A values
2 Data from nearby, paragenetically similar sample
See Appendix, Table Al for sample locations and descriptions.
228

homogenization temperatures each based on averages of 10 to 20


analyses, and all uncorrected for pressure. The pyrite-galena and
pyrite-sphalerite pairs provided much higher and more variable temp-
era ture estimates ranging from 238° to 460°C and 384° to 597°C, re-
spectively. The disparity in temperatures between the latter two
mineral pairs and those for s'phalerite-galena are undoubtedly the
result of disequilibrium. Petrographic studies of polished sections
demonstrate that pyrite in all samples was consistently deposited
early and subsequently replaced by sphalerite and galena. In con-
trast, sulfide pairs of sphalerite and galena show with few excep-
tions textural relationships indicative of contemporaneous deposi-
ti on.

Because the isotopic fractionation between mineral pairs is


independent of pressure (C. W. Field, 1983, pers. commun.), observed
differences between the isotopic temperature estimates and fluid
inclusion homogenization temperatures, uncorrected for pressure, may
reflect the amount of lithostatic or hydrostatic pressure which
prevailed during mineral deposition. Although temperature estimates
from both methods are in excellent agreement, the isotopic tempera-
tures are slightly higher by 2° to 16°C. At 250°C, a mean tempera-
tore difference of 10°C gives a pressure correction of 110 bars for a
10 percent solution of NaC1 (extrapolated from Potter, 1977), which
corresponds to depths of approxImately 470 and 1250 m for lithostatic
and hydrostatic loads, respectively.
229

CHEMISTRY OF THE ORE-FORMING FLUIDS

The general chemical environment during hydrothermal al teration


and mineralization can be estimated through evaluatiofl of the mineral
assemblages in the ores and alteration zones, and in the context of
constraints imposed by the fluid inclusion and sulfur isotope data.
Determination of the approximate f(S2)-f(02)-f(CO2) conditions in the
hydrothermal fluid necessitates a sufficient number of coexisting
mineral phases representative of an equilibrium assemblage and infor-
mation concerning temperature, pressure, and composition of the hy-
drothermal fluids. Because the deposits are characterized by re-
placement, equilibrium conditions existed among few minerals. Al-
though good temperature information was obtained for much of the ore
stage of mineralization, little direct information is available for
the time interval before and after this period. Furthermore, the
concentrations ofseveral components had to be estimated because
direct geologic or experimental data are lacking, and thus they are
necessarily subjective. Nevertheless, it was possible to estimate
the general chemical conditions which prevailed during alteration and
mineralization.
Temperature and Pressure

Excel lent estimates of the temperature and pressure during mm-


eralization are known from fluid inclusion and sulfur isotope stud-
ies. Homogenization temperatures of fluid inclusions in quartz and
sphalerite ranged from slightly greater than 3000 to 190-200°C during

L
230

deposition of pyrite, sphalerite, and galena. Sulfur isotopic temp-


eratures from galena-sphalerite mineral pairs also support this range
of temperatures. Pressure estimates, based on fluid inclusion corn-
positions, Indicate a minimum pressure of 40 to 50 bars for 5 to 10
weight percent solutions of NaC1 at 250°C (Haas, 1971). Inclusions
of higher salinity in the Chumpe Quartz Porphyry require an minimum
pressure of 100 to 150 bars, or as much as 800 bars, as calculated by
the method of Roedder and Bodnar (1980). These pressures correspond
to depths of between 500 and 1500 m, assuming hydrostatic conditions,
and they agree reasonably well with reconstructions of the local
geologic column.
Temperature estimates for pre-mineralization alteration of the
Pucara Group limestones and for the late-stage carbonate period are
difficult to assess because data could not be obtained from fluid
inclusions or sulfur isotopes of suitable mineral pairs. However,

because the pre-mineralization period is defined by the conversion of


limestone to a mixture of manganosiderite and specular hematite, it
is possible to estimate the temperature of formation through consid-
eration of the relative stability of this mineral assemblage. Sid-
erite, hematite, and magnetite are stable at low temperatures, but
siderite requires increasingly higher CO2 fugacitles at higher temp-
eratures. Above 250°C, Co2 fugaci ties of greater than 100 bars are
required to form siderite as shown in Figure 40. Since the assem-
blage siderite-hernatite is present without magnetite, CO2 fugacities
must have been extremely high or the temperature was relatively low
231

Figure 40 Log f(CO2)-log f(02) diagram showing the stabilities


of iron-bearing minerals at 25°, 150°, and 250°C.
Calculated from the data in Robie and others (1978).
-
y/
/ //

0 /
00 I
/ //
233

(l00°-175°C). At 150°C, the CO2 pressure must be at least 400 bars


to prevent the formation of magnetite (Fig. 40). Although fluid
inclusions formed during this alteration period are lacking, there is
little evidence for extremely high CO2 pressures in any of the fluid
inclusions studied. However, the mineralogy suggests a temperature

range of 100-175°C and moderate CO2 pressure during this early al-
teration stage.
Sulfur, Oxygen, andf9 Fugacity

Mineralization in the mantos began with an early alteration


stage (siderite-hematite) at 100-175°C, followed by deposition of
massive sulfides at temperatures approaching 300°C. Temperatures
gradually declined during the later period of mineralization to below
180°C, at which time the late-stage carbonate mineral assemblage was
deposited (third period). Utilizing the methods of Barnes and
Kullerud (1961), Barton and others (1977), Helgeson (1969), Holland
(1959; 1965), and Garrels and Christ (1965), mineral stability dia-
grams have been constructed which depict the general chemical para-
meters such as f(CO2), f(o2), and f(s2) that existed during these
phases of hydrothermal activity (Fig. 40-42).
From the preceeding generalizations concerning the stability of
siderite and hematite, an inferred temperature of less than 175°C is
likely for the early period of alteration. Because the stabilies of
the iron minerals are dependent on the fugacity of both oxygen and
sul fur, a f(02)-f(s2) diagram was constructed for a temperature of
150°C and is shown in Figure 41. Iron sulfide minerals were not
234

Figure 41. Log f(S2)-log f(02) diagram showing stability fields


for iron-bearing minerals at 150°C. Dashed lines out-
line stability field of siderite at f(CO2) = 10, 100,
1000 bars. Note magnetite field is eliminated at f(CO2)
398 bars.
c'J
cc)
9-

-58 -54 -50 -46 -42

togf02
Figure 41. 01
236

deposited during the early alteration period, thereby fixing the


fugacity of sulfur to less than i017 bars at the magnetite-hematite-
pyrite join. Magnetite is absent from the early alteration mineral
assemblage, which necessitates a sufficiently high f(CO2) to prevent
its formation. The effect of increasing CO2 fugacity is illustrated
by plotting the stability field of siderite at 10, 100, and 1000 bars
CO2 (Fig. 41). The field of magnetite stability decreases rapidly,
with increasing f(CO2), until at io2.6 bars CO2 it disappears entire-
ly allowing the stable coexistence of siderite and hematite. Given

this fugacity of CO2. the fugacity of oxygen must be approximately


io46 io42 bars. If this alteration stage took place at tempera-
tures lower than estimated here, both the f(CO2) and f(02) could be
decreased further while still maintaining equilibrium betweensid-
en te and hemati to.
In accordance with the previous discussion, the. fugacity of CO2
plays a key role in the stability of siderite. It also governs
reactions which form caic-silicate minerals at the expense of cal-
cite. However, calculation of the absolute CO2 fugacity that existed
during mineral fzation is difficult without specific compositional
data regarding the hydrothermal fluids (ie. concentration of CO2).
The CO2 fugacity was estimated through examination of the stability
fields of the iron-bearing minerals at various temperatures (Fig.
40). This exercise revealed that extremely high f(CO2) values are
needed to stabilize siderite in the presence of hematite. Holland
(1965) noted that the thermochemical data for siderite may be in
237

serious error at 25°C, and additionally for extrapolations of this


data to higher temperatures. Siderite Is geologically a common
mineral in mesothermal vein deposits (1=250°C), which contradicts the
thermochemical calculations that show an extremely small field of
stability for siderite at this temperature (Holland, 1965). There-
fore, it is likely that the actual fugacily of CO2 is lower than that
calculated from the siderite-hematite thermochemical data, but it is
sufficiently high to prohibit the formation of typical skarn minerals
such as wollastonite, tremolite, or even rhodonite. Equilibrium
constants were calculated for the following reactions at 300°C:

1 cal + 1 qtz = 1 woll + CO2 log K = -2.30


1 rhod + 1 qtz = 1 rhdn + CO2 log K = -1.15
2cal +5mag+8qtz=1 trem+7CO2 logKl6.30
Because these minerals are absent from the mineral assemblage, the

fugacity of CO2 must have been greater than the equilibrium CO2

fugacity. Assuming unit activity for the solid phases and water, the
log f(CO2) is equal to -2.3, -1.15, and 2.33 bars for these three
reactions, respectively.
Einaudi and others (1981) have suggested a genetic link between
zinc skarns and zinc-lead carbonate vein deposits, which is charact-
erized by veins containing minor amounts of calc-silicate or mangano-
silicate minerals. Deposits at Uchucchacua, Peru (Alpers, 1980) and
Bluebell, Canada (Ohmoto and Rye, 1970) may represent this inter-
mediate link. The controlling factors appear to be temperature,
238

composition of the hydrothermal fluids, and pressure (Einaudi and


others, 1981; Titley, 1968). Replacement deposits of lead and zinc,
characterized by siderite and jasperoid, are favored by high pressure
and low temperature (Titley, 1968). Therefore, manganese silicates
might have formed in the manto deposits at San Cristobal, if the
temperature of the hydrothermal fluids had been slightly higher
(300°-400°c). Further speculations concerning the fugacity of CO2
are unwarranted without quanti tative chemical analyses from the fluid
mci usions.

With an increase in temperature to 250-300°C, at the beginning


of massive sulfide deposition, a concomitant increase in both f(S2)
and f(02) takes place, and siderite ceases to be an important phase
at geologically reasonable fugacities of CO2 (<1000 bars). At 250°C,
the maximum and minimum limits of oxygen and sulfur fugacity can be
determined by various phase boundaries for the coexisting mineral
assemblages found at San Cristobal, and these have been plotted in
Figure 42. The minimum oxygen fugacity is constrained by the C-CO2

boundary at to bars for 10 and 100 bars CO2 pressure,


respectively, and maximum oxygen fugacity is limited by the galena-
anglesite boundary at iO31.8 bars 02. Pyrrhotite, formed early in
minor amounts prior to or with pyrite, which places the minimum
sulfur fugacity at io135 to io25 bars. Because bornite is not
found in the manto ore, maximum sulfur fugacity is limited at io8.5
239

Figure 42. Log f(S2)-log (02) diagram showing stability fields for
minerals forme during massive sulfide mineralization at
250°C. Shaded region delineates probable chemical condi-
tions. Dotted lines are mole percent FeS in sphalerite
from data of B rton and others, 1977. Pyrite-pyrrhotite
boundary corre ponds to 20 mole percent FeS. C-CO2
boundaries dra n at f(CO2) = 1, 10, and 100 bars. Arrow
indicates prob ble trend in solution chemistry with time
or decrease in temperature.
uf uI I

I
I I
I

0.1 L_.. ___ ._:.J: ...- ..._ .. .


-.

py

hm

:
tog '02

Figure 42.
241

bars by the reaction:

5 cp + 1 r+1S2_..__>5py+1 bn

The stability field of chlcopyrite is limited in the magnetite and


hematite regions by reactions which also form bornite at the expense
of chalcopyrite. Sphalerite is stable throughout the area delineated
in Figure 42. Thus, the mineral assemblage defines a triangular
region of f(S2) and f(02) in which massive sulfide deposition took
place.
Delineation of a mor precise region of sulfur and oxygen fugac-
ity necessitates the prese ce of additional coexisting mineral phases
or data concerning the c mposition of the hydrothermal solutions.
Barton and others (1977) plotted the iron content of sphalerite with
respect to f(02) and f(s2) to show possible temporal trends in these
variables, in the OH vein at Creede, Colorado. According to Barton
and Skinner (1979), the i on content of sphalerite can only be re-
lated to these two variab es if sphalerite forms in equilibrium with
pyrite and/or pyrrhotite Microprobe analyses of manto and vein
sphalerites at San Crist bal show a moderate range of 15.67 to 0.18
mole percent FeS (Table 22). However, textural relationships between
sphalerite and pyrite conistently show evidence for disequilibrium
conditions, with sphalerite replacing pyrite. Moreover, sulfur iso-
topic fractionation factor and temperatures therefrom on the pyrite-
sphalerite mineral pair are highly variable and significantly
greater than those temperaures obtained from sphal en te-gal ena pairs
242

or those from the data fluid inclusions. Despite these observa-


tions, pyrite and sphalerte may have been in chemical equilibrium,
with respect to the actiity of reduced iron, during at least the
earlier episodes of sulfide deposition. In accordance with this
assumption, the mole percnt FeS in sphalerite was plotted on Figure
42 at 0.1, 1, 10, and 20 percent FeS with the pyrlte-pyrrhotite
boundary corresponding to 20 percent FeS using the data of Barton and
others (1977). There are wo ways to interpret this change in sphal-
erite composition given the limits imposed by the aforementioned
phase boundaries and the eneral trend of decreasing iron content in
the sphalerite with paragnesis. First, If temperature remained con-
stant at 250°C, the f(02)ff(S2) conditions may have increased with
time from near the pyrrhotite-pyrite-magnetite triple point along a
path on or above, and pa allel to the magnetite-hematite-pyrite
boundary. A second, and more likely, interpretation that would
result in a similar path y simple migration of the phase boundaries
with decreasing temperat re, provided the total concentration of
sulfur and oxygen remain d constant. Fluctuations in temperature
and/or composition of th hydrothermal fluids would produce th
observed compositional zoning and reversals in sphalerite.
One problem remains oncerning the paragenetic position of mag-
netite in the ores. Magnetite, in all cases, is seen only as pseudo-
morphs after hematite, andnormally only in the presence of siderite.
Within the constraints of Iikely fluid components, the conversion of
243

hematite to magnetite can e expressed by the following reactions:

1Fe2+ hm+1H2O=1mt+2H
3 m=2mt+1/202
1 hm 1 sid=lmt+1CO2

Equilibrium is therefore a function of the activity of Fe2, pH,


f(02), f(CO2), and temper ture, and the conversion of hematite to
magnetite is likely to ccur with either an increase in (1) the
activity of ferrous iron, (2) pH, and (3) temperature, or a decrease
in (4) f(02), or (5) f(CO2 . Minor amounts of pyrrhotite are found
associated with the earlie t pyrite. The incompatibility of pyrr-
hotite with a hematite ass mbl age suggests that some, if not all, of
the hematite was emplaced ither prior to, or subsequent to, sulfide
precipitation. If it is rgued that hematite preceded sulfide mm-
eralization, the most likely sequence of events is that hematite was
deposited at lower temperlatures, and was subsequently replaced by
sulfides. Because this p ocess involves an increase in both temp-
erature, and probably the activity of ferrous iron, a considerable
amount of magnetite should have formed at the expense of hematite
with precipitation of the assive sulfides. However, textural rela-
tionships in many samples imply that specular hematite was replaced
by massive pyrite and sp alerite with little or no formation of
magnetite (see Fig. 25d). It is likely that specular hematite may
have formed at two or more times during the course of mineralization;
the first took place dunn early alteration at low temperatures, and
244

the second, during sulfid mineralization after deposition of pyrr-


hotite at higher fugaci1ies of oxygen. This interpretation pre-
supposes that the fugacity of CO2 is an unimportant factor at this
temperature.

El

Determination of pH of the ore-forming fluids is somewhat


problematical given the lack of sufficient compositional data. How-

ever, the stabilities of ',arlous alteration minerals associated with


sulfide mineralization, s ch as sericite and kaolinite, are governed
by the following hydrolys s reactions (Helgeson, 1969, Table 11):

3 K-feld + 2H = 1 mus + 6 qlz + 2K (1)

2mus+2+3H2O=3kaol+2K (2)

Values of log K for equation (1) are 8.49, 8.12, and 7.82 at 2000,

2500, and 300°C, respectively. For equation (2) the log K values are
7.04, 6.20, and 5.44 at the same temperatures (Helgeson, 1969, Table
11). tf the solid phases and water are assumed to have unit activi-
ty, the stability of potasium feldspar, sericite, and kaolinite are
a function of the activity ratios of K to H. Fluids associated
with mineralization were lound to have salinities of between 4 and 8
weight percent NaC1 equivlent. The activity of K is approximately
0.1 molal, assuming the f uids contain a 6 weight percent chloride
solution (1 molal) with a Na/K ratio of 9/1. The calculated equilib-
riumpH values at 250°C fpr equations (1) and (2) are 5.06 and 4.01,
245

respectively. An increas? in temperature of 50°C will raise the pH


by 0.15 and 0.51, for reactlons (1) and (2), while a similar decrease
in temperature will lower the pH by an equivalent amount. Reactions

between fluids and wallrocks would likely buffer the pH of the hydro-
thermal fluids between thse values. Thus, it can be inferred that
the pH of the hydrotherml fluids entering the Pucara Group lime-
stones was between 4.01 a d 5.06 at a temperature of 250°C, provided
these assumptions are val d.
The pH of a hydrotheviial fluid in limestone is governed by the
reaction:

CaCO3 + 2 H = Ca2 + H20 + CO2 (3)

The log k value of this r action is 8.81 at 250°C (Helgeson, 1969).


Again, assuming the soli phases and water have unit activity, the
stability of calcite is a function of the activity ratio of Ca2 and

CO2 to H. To determine i4e equilibrium pH conditions, it is further


necessary to assume the ativity of calcium ion and the fugacity of
CO2. Given a high concentration of 4000 ppm Ca2 (0.1 molal) and a
more conservative value of 400 ppm (0.01 molal), the following pH
values were calculated at various fugacities of CO2:

log f(CO2)
Ca2 3 2 1 0 -i -2

.1 m 3.4 3.91 4.41 4.91 5.41 5.91

.01 m 3.9 4.41 4.91 5.41 5.91 6.41


246

An inference regarding t e fugacity of CO2 may be made from these


data provided an independ nt measure of pH is .known. The manto ores

contain considerable dick te (kaolinite) which requires an equilib-


rium pH of 4.01 or less to form at 250°C. Comparison of the values
listed above with the ca culated equilibrium pH for equation (2)
indicates a CO2 fugacityof between 10 and 100 bars with .1 molal
Ca2, and between 100 ard 1000 bars with .01 molal Ca2. Thus,

dickite may form under coiditions of high calcium ion activity and
low CO2 fugacity, or at low calcium ion activity and high CO2 fugaci-
ty. Because dickite is i timately associated with the massive sul-
fide replacement ores, it most likely formed during the early stages
of mineralization, and t slightly higher temperatures (300°C).
Moreover, a low pH is m re likely to be attained at the site of
sulfide deposition than in the presence of fresh limestone.
Obviously, these calculation are simplified and further specula-
tions are precluded by the lack of quantitative chemical data on the
abundance of other calciu species in the fluids and thermochemical
data for siderite.
SUlfide Deoosi tion

From the available e,idence, it is clear that the hydrothermal


fluid was a chloride-rich rine derived as an exsolved phase from the
crystallizing Chumpe Qurtz Porphyry, and in part from meteoric
and/or connate water. A considerable body of theoretical, experi-
mental, and observationa evidence (Anderson, 1973, Barnes, 1979;
247

Browne and Ellis, 1970; Burnham, 1979; Burnham and Ohmoto, 1981;
Helgeson, 1969, 1970; Holland, 1972; Skinner and others, 1967; and
Weissberg and others, 1979) suggests that sufficient concentratlons
of metal can be carried in a hydrothermal fluid in the form of
chloride complexes. The pH of the fluids is probably neutral to
slightly acidic at temperatures of 200° to 400°C, which enhances the
stability of these metal-chloride complexes.
Precipitation of metals (Me2) from chloride complexes follows
the reaction (Barnes, 1979):

MeC12+H2S ----->MeS+2H+2CF

Obviously, the reaction will proceed to the right, favoring deposi-


tion of sulfides, if there is an increase in (1) the concentration of
H2S or (2) pH, caused either by reactions with carbonate or feldspar
in the walirock or by boiling of fluids, or a decrease in (3) the
chloride ion concentration caused by the addition of strong C1 ion-
pairing cations like Ca2 or (4) temperature. Barnes (1979) has
quantitatively demonstrated the effectiveness of each of these mech-
anisms.

There is little evidence of boiling of the hydrothermal fluids


in either the vein or manto ores, although it may have been important
in that part of the deposit removed by erosion. Process (1), an
increase in the concentration of H2S, is difficult to evaluate with-
out evidence regarding the composition of the hydrothermal fluids or
of possible connate waters in the sedimentary rocks of the district
248

Compositional information may eventually be obtained from fluid in-


clusions to provide information on this and other problems referred
to earl ier. The most important processes at San Cristobal are (2)
and (4). For the vein ores, wallrock reactions which decrease the
effective hydrogen ion concentration include sericitic and kaolinitic
alteration. These reactions cause precipitation of sulfide minerals,
provided the fluid is able to react with the wallrock. As mentioned

earlier, the vein wallrocks may have become armored by the precipita-
tion of quartz during the first mineralization period, thereby in-
hibiting the deposition of second period minerals until temperatures
decreased sufficiently to allow sulfide precipitation by that mechan-
ism alone.. However, mixing of cooler meteoric waters with the hydro-
thermal fluids would also have the same affect.
The manto ores, by virtue of their host rock lithology (lime-
stone), precipitated sulfides continuously as the hydrothermal fluids
underwent reaction with calcite or even siderite, causing a decrease
in the hydrogen ion concentration in the reaction:

CaCO3 + 2 H ----- > Ca2 + H2CO3

This reaction also creates an abundant supply of Ca2+ ions which may

cause a decrease in the total chloride content of the hydrothermal


solutions as previously described in process (3), further promoting
sulfide deposition. Accordingly, massive sulfide replacement pro-
gressed outward from fluid channeiways into "fresh" carbonate along
zones of high porosity and permeability.
249

The processes outlined above provide an internally consistent


model to explain (1) the deposition and localization of vein and
massive replacement (manto) ore, (2) the difference in fluid inclu-
sion homogenization temperatures in vein and manto sphalerites, and
(3) the zonation of alteration and sulfide minerals.
250

FEATURES OF EXPLORATION IMPORTANCE

A potentially useful aspect of this investigation was to derive


information and criteria relevant to future exploration for both
additional reserves of ore in the San Cristobal District and undis-
covered Pb-Zn-Ag deposits in the central Andes of Peru. The follow-
ing list outlines both obvious and subt1e features which may be
associated with low-temperature, hydrothermal replacement deposits
hosted by carbonate rocks.
1. Hydrothermal replacement deposits are associated with sub-
volcanic, porphyritic stocks of intermediate to silicic composition,
which are generally less than 15 m.y. old.
2. Alteration of the intrusiverocks is characterized by assem-
blages that are typical of those described for districts hosting
porphyry copper deposits, and these include quartz-sericitic (phyl-
lic), argillic, and propylitic types of alteration.
3. Zones of high porosity and permeability such as faults,
joints, and breccias control the circulation and migration of ore-
forming fluids from the intrusions to the limestone.
4. Ore deposits are most likely found near the base of the
Pucara Group, since this formation contains the stratigraphical ly
lowest carbonate and thus most chemically reactive units in the rock
sequence of central Peru.
251

5. Deposition of ore in the limestone is locally controlled by


zones of high permeability such as faults, breccias, joints, and
possibly oolitic limestone beds.
6. Characteristic al teration features in the Pucara Group in-
dude jasperoids formed by intense silicification and Fe-Mn-rich
carbonates formed by metasomatic reactions. However, at higher temp-
eratures and lower f(CO2), alteration is likely to be characterized
by typical calcium or mangano-silicate skarn minerals.
7. The recognition of altered limestones is based on visible
color anomalies produced by the oxidation of ferroan dolomite and
manganosiderite. Ferroan dolomite oxidizes to a reddish-brown color
which becomes darker as the surficial alteration intensifies, whereas
manganosiderite is converted to a mixture of dark brown and black
iron and manganese oxides. The presence of sulfides creates a van-
ably colored gossan rich in both manganese and iron oxides, with or
without typical secondary minerals such as hydrozincite, cerrusite,
smithsonite, marcasite, native silver, and others. These gossans
will also tend to be porous and cellular as a consequence of the
dissolution of sulfide minerals.
8. Primary hypogene minerals which commonly are preserved in
oxidized deposits include barite, specular hematite, quartz, and
possibly rnagnetite.
9. Elements which serve as excellent geochemical indicators of
the mineralization may include, iron, manganese, arsenic, antimony,
silver, boron, barium, copper, lead, zinc, and cadmium.
252

10. In general, the FeO/MnO ratio in al tered carbonates in-


creases toward known mineral deposits and(or) toward major hydro-
thermal fluid channeiways such as faults, fractures, breccias, and
stratigraphically favorable horizons for replacement, and thus toward
possible or hidden deposits of ore.
11. The iron content of sphalerite increases toward the centers
of sulfide mineralization and with increasing depth.
12. Geophysical prospecting for blind ore bodies should concen-
trate on magnetic and gravity surveys. The deposits typically con-
tain abundant magnetite and hematite, and accordingly, they should
exhibit excellent magnetic anomalies, especially in carbonate rocks.
Particular care shoul d be used however, as the magnetic anomal tes
might also delineate the mafic lava flows that are interbedded with
the carbonates. Locally, gravity surveys might be useful in areas of
limestone alteration and permit the definition of large massive
sulfide deposits because of the extreme density contrast between
limestone and sulfide/oxide ore.
253

SUMMARY AND CONCLUSIONS

Mineral deposits of the San Cristobal District consist of poly-


metallic vein and stratiform replacement (manto) ores. Vein-type
deposits are localized in steeply-dipping and west-southwest trending
normal faults which are transverse to the Chumpe Anticline. Manto

deposits are found only in limestones of the Pucara Group, and their
distribution within these carbonate rocks appears to be controlled by
intersections with or close proximity to the larger mineralized
veins.
The sedimentary sequence includes rocks which range in age from
Devonian to Tertiary. The Devonian Excelsior Phyllite constitutes
the regional basement rock and is overlain nonconformably by the
Permian Mitu Group. Mesozoic sedimentary rocks are represented by
limestones of the Triassic-Jurassic Pucara Group, sandstones of the
Lower Cretaceous Goyllarisquisga Formation, and limestones of the Mid
to Upper Cretaceous Machay Group. Although the Pucara Group consists
predominantly of shallow water marine limestones, it also contains
intercalated volcaniclastic tuffs of airfall origin and minor flows
of basalt.
These country rocks were folded into a northwest-trending, asym-
metrical anticline, known locally as the Chumpe Anticline, during the
Incaic pulse of compressive deformation, approximately 40 to 43 m.y.
ago. Conjugate transverse faults with minor strike-slip motion
formed initially as a result of shear stresses during the compressive
deformation, but subsequently yielded to normal displacement with
254

general uplift of the Andes over the past 10 m.y. This normal fault
movement also produced a transverse graben centered over the apex of
the Chumpe Anticline that is bounded by the San Cristobal and
Prosperidad Faults, which subsequently localized major vein-type
ores. Concentric folding of the stratigraphic units formed local
zones of breccia situated between the Excelsior Phyllite and Mitu
Group, and between the Mitu and Pucara Groups. The Huaripampa Mine

(Carahuacra Manto) occupies a large breccia zone between the latter


two stratigraphic units. Other smaller, but laterally extensive
breccias within the Pucara Group were produced by several processes
including (1) bedding-plane fault movement, (2) hydrothermal solution
collapse, and possibly (3) hydraulic fracturing. Karstic-like solu-
tion breccias were not observed in the district.
Two small stocks were intruded into the sedimentary sequence in
separate magmatic events during the Eocene and Miocene. The

Carahuacra Quartz Monzonite was emplaced at 43 m.y. prior to the


Incaic pulse of compressive deformation, and it was subsequently
altered to an assemblage of propylitic minerals by an early, possibly
deuteric, stage of hydrothermal activity that did not produce visible
metallization. Although the evidence is tenuous, the Carahuacra
stock may have been tilted to the southwest during folding of the
Chumpe Anticline. The Chumpe Quartz Porphyry stock and related dike
system were emplaced between 5 and 7 m.y. ago along the axis of the
Chumpe Anticline. This stock is centered approximately at the anti-
clinal apex, and it is located within the transverse graben. Doming
255

associated with intrusion of this stock (or a larger pluton at depth)


may have been partly responsible for the later normal movement along
the transverse faul ts.
Subsequent hydrothermal activity produced elliptical zones of
al teration centered on the Chumpe Quartz Porphyry, and formed the
vein and manto ore deposits of the district. The central zone con-
tains a pervasively distributed quartz-sericite assemblage that is
associated with minor occurrences of disseminated.and vein pyrite,
whereas the outer zone contains minerals typical of argillic and
propylitic alteration and small amounts of pyrite. Planar zones or
selvages of alteration have formed by the reaction of fluids with the
wall rocks. These grade progressively outward from the vein-wall
rock contact through successive zones defined by silica, quartz-
sericite, and argillic-propylitic alteration. Chemical analyses of
samples from the quartz-sericite zone of alteration suggest additions
of Si02 and K20 with concomitant losses of CaO, Na20, MgO, and FeO.
However, these gains and losses cannot be quantified because un-
altered equivalents of the samples are lacking. Three distinct types
of rock alteration are recognized in the limestones. The earliest
was defined by a period of Fe-Mn metasomatism, which led to the
wholesale conversion of primary calcite of the carbonate host to
manganosiderite and specular hematite. This was followed by an
episode of silicification which formed the jasperoids. Minor amounts
of argillic alteration, typified by the presence of dickite, are
found in interbedded volcaniclastic tuffs. Apparently the hydro-
thermal fluids were either of low temperature (< 300°C), or main-
tained a sufficiently high fugacity of CO2 (> 100 bars), to inhibit
the formation of skarns and the associated caic- or mangano-silicate
minerals.
Hypogene vein mineralization took place over three successive
periods, separated by episodic brecciation. The first period was
marked by the deposi tion of pyri te, quartz, and wol frami te. Wolf-
ramite is spatially related to the quartz porphyry dikes, which may
reflect higher temperature conditions, whereas the pyrite and quartz
are also widely present in all veins. The second period mineral iza-
tion produced the bulk of the ore minerals which consists largely of
chalcopyrite, sphalerite, galena, and several types of sulfosalts.
The third period is represented predominantly by the gangue minerals
which are chiefly manganosiderite-rhodochrosite, barite, and marca-
site, and trace quantities of the ore-bearing sulfides. Hypogene re-
placement ores that comprise the limestone-hosted mantos contain
essentially the same minerals as do the veins, except for the absence
of woiframite and lesser amounts of chalcopyrite. Although the manto
ores are mineralogically and chemically similar to the veins ores,
they are marked by distinctly different textures, such as finer grain
size and laminated textures (formed by pseudomorphic replacement of
bedding) composed of polycrystalline mineral aggregates. However,

these textural distinctions become subjective at the intersections of


vein and manto ores. Most chemical and mineralogical differences
between vein andmanto ores can be attributed to a thermally imposed
257

district-wide zonation of metals and minerals, and not to the en-


vironment (host rock) or mechanisms (vein-filling versus replacement)
of ore deposition.
Fluid inclusion data from fractured quartz phenocrysts in the
Chumpe Quartz Porphyry indicate that sericitic alteration formed
between 325° and 450°C, by reaction of fel dspar with fluids having
salinitles of between 18 and 49 weight percent NaCl equivalent. The

dominant chloride species are inferred to have been NaCl, KC1, Cad2
and HC1. Evidence for periodic boiling of the fluids in the quartz
porphyry is abundant, and considered in terms of the NaCl-H20 system,
requires a minimum pressure of 100-250 bars (equivalent to 900-2400 m
hydrostatic load). Fluid inclusion data from sphalerite and quartz
in the veins and mantos Indicate mineralization temperatures of 1800_

300°C and fluid salinities of 4 to 8 weight percent NaC1 equivalent.


Because evidence of boiling is lacking from inclusions in the ores,
minimum pressures were between 20 and 100 bars (equivalent to 200-
1400 m hydrostatic load). Mixing of the dense magmatic brine derived
from the porphyry with cooler meteoric waters at a distance from this
intrusive center may account for the lower temperatures and salini-
ties of the ore fluids associated with vein and manto ores.
Sulfur isotope analyses of barites (- 16 0/00) and sulfides
(.. 5 0/00) provide evidence of fractionations that are compatible
with equilibrium isotope exchange reactions for sulfur in a hydro-
thermal system at moderate temperatures. The smal 1 range of
values among sulfides (2.4 to 8.2 0/00) and the lack of obvious
258

temporal or spatial variations in isotopic composition suggest a


system well-homogenized with respect to sulfur. However, these data
clearly show that the sulfides are anomalously enriched in 34S. This

enrichment is suggestive of a mixed source of sulfur, comprised of


magmatic ( 0 0/00) sulfide-sulfur and reduced evaporitic (- 14 °/oo)
sulfate-sulfur components, and this interpretation is consistent with
the geologic and isotopic evidence from the nearby Morococha
District. Isotopic temperature estimates (169°-287°C) determined
from coexisting sphalerite-galena mineral pairs are virtually identi-
cal to fluid inclusion homogenization temperatures (159°-285°C), and
thus document isotopic equilibrium between these minerals. Delta
values from pyrite-sphalerite and pyrite-galena pairs give excessive-
ly high temperatures indicative of disequilibrium, a relationship
that is also corroborated by paragenetic and textural observations.
Thermochemical calculations performed on phases comprising the
observed mineral assemblages specify temperatures of 100°-175°C and
CO2 fugacities of 200-600 bars for the early manganosiderite-specular
hematite stage of alteration. These early fluids were essentially
devoid of sulfide-sulfur. As temperatures increased to 300°C, at the
onset of massive sulfide mineralization, a concomitant rise in sulfur
and oxygen fugaci ties toOk place. A combination of decreasing CO2
fugacity and increasing temperature, ferrous iron activity, and sul-
fur fugacity, led to the replacement of early-formed specular hema-
tite by both pyrite and magnetite. The iron contents of sphalerite
reflect increasingly oxidized conditions during subsequent cooling,
259

with occasional compositional reversals caused by Influxes of more

reduced or hotter ore-bearing fluids. The pH of the ore-forming


fluid was buffered by walirock reactions at 4.01 to 5.06 in the
veins, and 4.41 to 5.41 in the carbonate-hosted mantos. Locally
higher pH values may have prevailed at the reaction front between
fluid and rock, and especially in unal tered rocks, because of dis-
equi 1 ibrium condi tions.
The ore metals were carried by the fluids as chloride complexes
in the presence of relatively low concentrations of reduced sulfur.
Precipitation of vein sulfides was accomplished by dilution and
diminishing temperature of the fluids, caused by the influx of cooler
meteoric waters, and by wal 1 rock reactions which decreased the H+
concentration. Massive replacement (manto) ores are inferred to have
been deposited in a similar manner through the reaction of fluids
with limestones of the Pucara Group.
In conclusion, the stratiform and strata-bound (manto) deposits
of the San Cristobal District are the result of epigenetic, hydro-
thermal replacement processes at moderate temperature and pressure.
They are intimately associated, both temporally and spatially, with
epigenetic vein mineralization and acquired their distinctive morph-
ology as a natural consequence of fluid migration outward from a
thermal source along avenues of high porosity and permeability.
Proponents of a syngenetic/diagenetic origin for these deposits must
demonstrate a plausible mechanism, and one operative at low tempera-
tures, which would account for the geologic and chemical data pre-
260

sented herein. Characteristics of particular relevance include the


(1) coninon temperatures (2000 to 300°C) of fomation; (2) spatial as-
sociation of manto and vein ores; (3) similar and homogeneous iso-
topic compositions of vein and manto ores; (4) similarities in min-
eralogy, paragenesis, and trace metal distributions, (5) anomalously
high iron and manganese content of the altered limestones; (6) sharp
contacts between mineralized and unmineralized rock; and (7) numerous
cross-cutting and replacement relationships. Determination of a
specific origin for any mineral deposit must be based on a variety of
mutually independent data to avoid spurious conclusions. The results
of this investigation virtually preclude a syngenetic/diagenetic ori-
gin for the manto deposits of the San Cristobal District
261

BIBLIOGRAPHY

Alpers, C. N., 1980, Mineralogy, paragenesis., and zoning of the Luz


vein, Uchucchacua, Peru: Unpub. B.A. thesis, Harvard Univ., 138
p.
American Geological Institute, 1980, Glossary of Geology, second
edition, R.L. Bates and J.A. Jackson, editors. 749 p.
Amstutz, G. C., 1961, Origin de los depositos minerales congruentes
en las rocas sedimentarias: Bol. Soc. Geol. del Peru, V. 36,
Lima, p. 5-30.
Anderson, G. M., 1973, The hydrothermal transport and deposi tion of
galena and sphalerite near 100°C: Econ. Geol., v. 68, p. 480-
492.

Barnes, H. L., 1979, Solubilities of ore minerals in Barnes, H.L.,


ed., Geochemistry of hydrothermal ore deposits: John Wiley and
Sons, New York, p. 404-460.
and Kul lerud, G., 1961, Equilibria in sulfur-contain-
ing aqueous solutions, in the system Fe-S-O, and their correla-
tion during ore deposition: Econ. Geol., v. 56, p. 648-688.
Barton, P. B. Jr., 1978, Some ore textures involving sphalerite from
the Furutobe Mine, Akita Perfecture, Japan: Mining Geol., V.
28, p. 293-300.
Bethke, P. M., and Roedder, E., 1977, Environment
of ore deposition in the Creede mining district, San Juan
Mountains, Colorado: III. Progress toward interpretation of the
chemistry of the ore-forming fluid for the OH Vein: Econ. Geol.,
v. 72, p. 1-25.
and Skinner, B. 3., 1979, Sulfide mineral stabil-
ities in Barnes, H. L., ed., Geochemistry of hydrothermal ore
deposits: John Wiley and Sons, New York, p. 278-403.
avides, V.,, 1956, Cretaceous system in northern Peru: American
Mus. Nat. History Bul 1., v. 108, p. 252-494.
enlos, A. J. and Ericksen, G. E., 1955, Lead-zinc deposits of
Cordillera Blanca and Northern Cordillera Huayhuash, Peru:
U.S.G.S. Bull. 1017, 166 p.
I3ogacz, K., Dzulynski, S., and Haranczyk, C., 1970, Ore filled hydro-
thermal karst features in the Triassic rocks of the Cracow-
Silesian region: Acta Geologica Polonica, v. 20, No. 2, p. 247-
269.

1973, Caves filled with


clastic dolomite and galena mineralization in disaggregated
dolomite: Geol. Soc. Poland Bull., v. 43, p. 59-76.
Browne, P. R. L. and Eli is, A. Je, 1970, The Ohaki-Broadi ands hydro-
thermal area, New Zealand: Mineralogy and related geochemistry:
Am. Jour. Sci,, v. 269, p. 97-131.
Burnham, C. W., 1967, Hydrothermal fluids at the magmatic stage:
in Barnes, H. L., ed., Geochemistry of hydrothermal ore de-
posits: Hol t, Rinehart, and Wilson, Inc., New York, p. 34-76.
1979, Magma and hydrothermal fluids: in Barnes, H.
L., ed., Geochemistry of hydrothermal ore deposits: John Wiley
and Sons, New York, p. 71-136.

and Ohmoto, H., 1981, Late magmatic and hydrothermal


processes in ore formation in Mineral Resources: Genetic under-
standing for practical appTications: National Academy Press,
Washington, D. C., p. 62-72.
Burress, R. C., 1977, Analysis of fluid inclusions in graphitic
metamorphic rocks from Bryant Pond, Maine, and Khtada Lake,
British Columbia: Thermodynamic basis and geologic interpreta-
tion of observed fluid compositions and molar volumes: Unpub.
Ph.D. thesis, Princeton Uni versi ty.

Caldas, J.., 1983, Tectonic evolution of Peruvian Andes: in Cabre, R.,


ed., Geodynamics of the Eastern Pacific Region, Caribbean and
Scotia Arcs: Am. Geophy. Union, Geodynamics Series, v. 9, p. 77-
81.

Campbell, A. R., and Rye, D. M., 1982, Fluid inclusion and stable
Isotope study of the San Cristobal Mine, Peru: Geol. Soc.
America, Abstracts with programs, v. 14, p. 458. -
Rye, D. M., and Petersen, U., 1983, The role of
temperature and water to rock ratio in the isotopic composition
of ore solutions: Geol. Soc. America, Abstracts with programs,
v. 15, n. 6, p. 538.
Claypoole, G. E., Holser, W. 1., Kaplan, I. R., Sakal, H., and Zak,
I., 1980, The age curves of sulfur and oxygen isotopes in marine
sulfate and their mutual interpretation: Chem. Geol., V. 28, p.
199-260.
263

Cloke, P. L., and Kesler, S. E., 1979, The halite trend in hydro-
thermal solutions: Econ. Geol., v. 74, p. 1823-1831.
Cobbing, E. J., and Pitcher, W. S., 1972, The coastal batholith of
central Peru: Geol. Soc. London Jour., v. 128, p. 421-460.
Cunningham, C. G. Jr., 1976, Petrogenesis and postmagmatic
geochemistry of the Italian Mountain Intrusive complex, eastern
Elk Mountains, Colorado: Geol. Soc. America Bull. v. 86, p. 897-
908.

1977, Fluid inclusion geothermometry: Geol.


Rundsch., v. 66, p. 1-9.
Daiheimer, M., 1980, Geologla del lado oeste del domo de Yauli---
especial atencion a los grupos Mite, Pucara, y sus yacimientos:
XX Convenclon de Geologos de Centromin-Peru, La Oroya, Ddiembre
1980, 37 p.
Donath, F. A. and Parker, R. B., 1964, Folds and folding: Geol. Soc.
America Bull., v. 75, p. 45-62.
Dunin-Borkowski, E., 1975, Control litologico y estratigrafico en la
ubicacion de los mantos con sulfuros de metales no ferrosos en
las capas calcareas del Peru central: Bol. Soc. Geol. Peru, Tomo
50, p. 25-52.
Dzulynski, S., 1976, Hydrothermal karsts and Zn-Pb sulfide ores:
Geol. Soc. Poland Bull., v. 46, no. 1-2, p. 217-230.
Einaudi, M. 1., Meinert, L. D., and Newberry, R. J., 1979, Skarn
deposits: Econ. Geology 75th anniversary volume, p. 317-391.
Eyzaguirre, V. R., Montoya, 0. E., Silberman, M. L., and Nobel, 0.
C., 1975, Age of igneous activity and mineralization, Morococha
District, central Peru: Econ. Geol., v. 70, p. 1123-1125.
Farrar, E. and Noble, 0. C., 1976, Timing of late Tertiary deforma-
tion in the Andes of Peru: Geol. Soc. America Bul 1., v. 87, p.
1247-1250.

Field, C. U., 1966, Sulfur isotopic method for discriminating between


sulfates of hypogene and supergene origin: Econ. Geol., v. 61,
p. 1428-1435.
and Gustafson, L. B., 1976, Sulfur isotopes in the
rphyry copper deposit at El Sal vador, Chile: Econ. Geol., v.
, p. 1533-1548.
264

and Moore, W. J., 1971, Sulfur isotope study of the "8"


limestone and galena fissure ore deposits of the U.S. Mine,
Bingham Mining District, Utah: Econ. Geol., V. 66, p. 48-62.
____________, Rye, R. 0., Dymond, J. R., Wheland, J. F., and
Senechal, R. G., 1983-(Metalliferous sediments of the East
Pacific In Shanks, W. C., III, ed., Cameron Volume on unconven-
tional iiTheral deposits: New York, Soc. Mining Eng., Am. Inst.
Mining Metall. Petrol. Eng., p. 133-156.
Gammon, J. B,, Borcsik, M., and Holland, H. D., 1969, Potassium-
sodium ratios in aqueous solutions and coexisting silicate
mel ts: Science, v. 163, p. 179-181.
Garrels, R. M., and Christ, C. L., 1965, Solutions, Minerals and
Equilibria, New York: Harper and Row., 450 p.
Geological Staff, Cerro De Pasco Corporation, 1948, Lead and zinc
deposits of the Cerro de Pasco Corporation in central Peru:
XVIII International Geological Congress, Part VII, Great
Britain, p. 154-186.
Giletti, B. J., and Day, H. W., 1968, Potassium-Argon ages of igneous
intrusive rocks in Peru: Nature, v. 220, p. 570-572.
Grinenko, V. A., 1962, Preparation of sulfur dioxide for isotopic
analysis: Zeits. Neorgan. Khimii, V. 7, p. 2478-2483.
Haapala, P. S., 1949, On Morococha breccias: Soc. Geol. Peru, v.
Jubilar, pt. 2, f. 2, 11 p.
1953, Morococha Anhydrite: Soc. Geol. Peru Bull., v.
26, p. 21-32.
Haas, J. L. Jr., 1971, The effect of sal mi ty on the maximum thermal
gradient of a hydrothermal system at hydrostatic pressure: Econ.
Geol., v. 66, p. 940-946.
Hamil ton, W., 1969, The volcanic central Andes--A modern model for
the Cretaceous Batholiths and tectonics of western North
America: in Mcbirney, A. R., ed., Proceedings of the Andesite
Conference: State of Oregon Dept. Geol. Mi Ind., Bull. 65, p.
175-184.

Harrison, J. V., 1943, The geology of the central Andes in part of


the province of Junin, Peru: Geol. Soc. London Quart. Jour., v.
99, p. 1-36.
265

Helgeson, H. C., 1969, Thermodynamics of hydrothermal systems at


elevated temperatures and pressures: Am. J. Sci., v. 267, p.
729-804.

1970, A chemical and thermodynamic model of ore


deposition in hydrothermal systems: Mm. Soc. America Spec.
Paper 3, p. 155-186.

Haederle, W., 1966, Unpub. report Cerro De Pasco Corporation, 4 p.

Hewitt, D. F., 1928, Dolomltization and ore deposition: Econ. GeOl,


v. 23, p. 821-863.

Holland, H. D., 1959, Some applications of thermochemical data to


problems of ore deposits. I. Stability relations among the
oxides, sulfides, sulfates and carbonates of ore and gangue
minerals: Econ. Geol., v. 54, p. 184-233.

_______________, 1965, Some appi ications of thermochemical data to


problems of ore deposits. II. Mineral assemblages and the
compositions of ore-forming fluids: Econ. Geol., v. 60, p.
1101 -1166.

1972, Granites, solutions, and base metal deposits:


Econ. Geol., v. 67, p. 281-301.

Hol 1 ister, L. S., Crawford, M. L., Roedder, E., Burress, R. C.,


Spooner, E. T. C., and Touret, J., 1981, PractIcal aspects of
microthermometry: in Hollister, L. S., and Crawford, M. L.,
eds., Fluid Inclusi6ii: Applications to Petrology., Mm. Assoc.
Canada short course Handbook, v 6, p. 278-304.

James, D. E., 1971, Plate tectonic model for the evolution of the
central Andes: Geol. Soc. America Bull., v. 82, p. 3325-3346.

Jehl, V., 1975, Le metamorphisme et les fluides associes des roches


oceaniques de l'Atl antique Word: These de docteur-ingenieur,
Univ. Nancy I, France.

Jenks, W. F., 1951, Triassic to Tertiary stratigraphy near Cerro de


Pasco, Peru: Geol. Soc. America Bull., v.62, p. 203-220.

Johnson, R. F., Lewis, R. W., and Abele, G., 1955, Geology and ore
deposits of the Atacocha District, Departamento de Pasco, Peru:
U.S.G.S. Bull. 975-E, p. 337-388.

Kamilli, R. J. and Ohmoto, H., 1977, Paragenesis, zoning, fluid


inclusion, and isotopic studies of the Finlandia Vein, Colqui
District, central Peru: Econ. Geol., v 72, p. 950-982.
266

Kelly, W. C., and Rye, R. 0., 1979, Geol ogic, fluid md USIOfl and
stable isotope studies of the tin-tungsten deposits of
Panasqueira, Portugal: Econ. Geol., v. 74, p. 1721-1822.
Kobe, H. W., 1977, El grupo Pucara y su mineralization en el Peru
central: Bol. Soc. Geol. Peru, Tomo 55-56, p. 61-84.
Krauskopf, K. B., 1979, Introduction to Geochemistry (2nd ed.), New
York: Mcgraw-Hill, 617 p.
Kulp, L. J., Amstutz, G. C., and Eckelmann, D. F., 1957, Lead isotope
composi tion of Peruvian gal enas: Econ. Geol., v. 52, p. 914-922.
Landis, G. P., and Rye, R. 0., 1974, Geologic, fluid inclusion, and
stable isotope studies of the Pasto Bueno tungsten-base metal
ore deposit, northern Peru: Econ. Geol., v. 69, no.7, p. 1025-
1059.

Laverty, M. A., 1949, Private report of the Cerro de Pasco


Corporation, 3 p.
Lepry, L. A., 1981, The structural geology of the Yauli Dome region,
Cordillera Occidental, Peru: Unpub. Masters Thesis, Univ. of
Arizona, Tucson, 1981, 99 p.
Linares, P. J. P., 1973, Geologia y minerlizacion de tungsteno en la
veta San Cristobal, Peru: Segundo Congreso Latinoamericano de
Geologia, p. 3839-3860.
Lovering, 1. S., 1949, Rock alteration as a guide to ore-East Tintic
District, Utah: Econ. Geol. Mono. 1, 64 p.
1962, The origin of jasperoid in limestone: Econ.
Geol., v. 57, p. 861-889.
1969, The origin of hydrothermal and low temperature
dolomite: Econ. Geol., v. 64, p. 743-754.
Tweto, 0., and Lovering, T. G., 1978, Ore deposits
of the Gilman District, Eagle County, Colorado: U.S.G.S. Prof.
P. 1017, 85 p.
Lujan, M. A., 1970, Paragenesis and al teration of the San Antonio
Mantos, Peru: Unpub. paper, Stanford Univ., Palo Alto,
CalIfornia, 1970, 38 p.
Lyons, W. A., 1968, The geology of the Carahuacra Mine, Peru: Econ.
Geol., v. 63, p. 247-256.
267

Mark, W. D., Faulkner, R. L., and Graton, L. C., 1942, Localization


of certain ore bodies at Morococha, Peru: in Newhouse, W. H.,
ed., Ore deposits as related to structural àtures: Princeton
Univ. Press, p. 239-241.
Martinez, C., Tomasi, P., Dalmayrac, B., Laubacher, G., and Morocco,
R., 1972, Carateres generaux des orogenes Precambriens,
Hercyniens etAndins au Peruou et en Bolivie: 24th mt. Geol.
Cong., Sec. 1.
Mckee, E. H., Noble, D. C., Petersen, U., Arenas, M., and Benavides,
A., 1975, Chronology of late Tertiary volcanism and mineraliza-
tion, Huachocolpa District, central Peru: Econ. Geol., v. 70, p.
388-390.

McLaughlin, D. H., 1924, Geology and physiography of the Peruvian


cordillera, Deparbnents of Junin and Lima: Geol. Soc. America
Bull., v.35, p. 591-632.
1945, Origin and development of the Cerro de
Pasco Corporation: Mm. Metal., v. 26, p. 512-519.
'
and Moses, J. H., 1945, Geology and the mining
region of central Peru: Amer. Inst. Mm. Eng. Mining and Metal.,
v. 26, p. 512-516.
Megard, F., 1968,Geologia del cuadrangulo de Huancayo: Servicto de
Geol.y Minera del Peru Bol. 18, 123 p.
1978, Etude geologique des Andes du Perou central:
Office de la Recherche Scientifique et Technique Outre-Mer, Mem.
86, 310 p.
Dalmayrac, B., Laubacher, C., Marocco, R., Martinez, C.,
Parades, J., and Tomasi, P., 1971, La chaine hercynienne au
Perou et en Bolivie, premiers resultats: Cah. ORSTOM, Ser. Geol.
III.
Meyer, C. and Hemley, J. J., 1967, Wall rock alteration, in Barnes,
H. L., ed., Geochemistry of hydrothermal ore depoTts: New
York, Hol t, Rinehart, and Winston, Inc., p. 166-235.
Middleton, G. V., 1961, Evaporite solution breccias from the
Mississippian of southwest Montana: Jour. Sede Petrology, v. 31,
p. 189-195.
Morris, H. 1. and Lovering, 1. S., 1979, General geology and mines of
the East Tintic Mining District, Utah and Juab Counties, Utah:
U.S.G.S. Prof. P. 1024, 203 p.
Nagell, R. H., 1960, Ore controls in the Morococha district, Peru:
Econ. Geol., v. 55, p. 962-984.

Naldrett, A. J., and Kullerud, G., 1965, Two examples of sulfuriza-


tion in nature (abs.): Econ. Geol., v. 60, P. 1563.

Nash, J. T., 1973, Geochemical studies in the Park City district-i,


Ore fluids in the Mayflower mine: Econ. Geol., v. 68, p. 34-51.

1975, Fluid inclusion studies of vein, pipe, and re-


placement deposits, Northwestern San Juan Mountains, Colorado:
Econ. Geol., v. 70, p. 1448-1462.

Noble, D. C., Mckee, E. H., and Megard, F., 1979, Early Tertiary
"Incaic" tectonism, uplift, and volcanic activity, Andes of
central Peru: Geol. Soc. America Bull., pt. 1, v. 90, p. 903-
907.

Noble, J. A., 1976, Metal logenic Provinces of the cordillera of


western North and South America: Mineral. Deposita, V. 11, p.
219-233.

Nockolds, S. R., 1954, Average chemical compositions of some igneous


rocks: Geol. Soc. America Bull., v. 65, p. 1007-1032.

Ohmoto, H., 1972, Systematics of sulfur and carbon isotopes in hydro-


thermal ore deposits: Econ. Geol., v. 67, p. 551-579.

and Rye, R. 0., 1970, The Bluebell mine, British Columbia


I. Mineralogy, paragenesis, fluid Inclusions, and the isotopes
of hydrogen, oxygen and carbon: Econ. Geol., V. 65, p. 417-437.

1979, Isotopes of sulfur and carbon in


Barnes, H. L., ed., Geochemistry of Hydrothermal Ore Deposits:
Holt, Rinehart, and Wilson, Inc., New York, P. 509-567.

Pastor, J., 1970, The mineralization in the San Cristobal mine:


Unpub. Masters Thesis, Univ. Arizonla, Tuscon, 117 p.

Petersen, U., 1965, Regional geology and major ore deposits of


central Peru: Econ. Geol., v. 60, p. 407-476.

Noble, D. C., Arenas, M. J., and Goodel 1, P. C., 1977,


Geology of the Julcani Mining District, Peru: Econ. Geol., v.
72, p. 931-949.

Phendler, R. W., 1957, Private report for the Cerro de Pasco


Corporation, 6 p.
269

Phillips, W. J., 1972, Hydraulic fracturing and mineralization: Geol.


Soc. London Jour., v. 128, P. 337-359.
Potter, R. W., 1977, Pressure corrections for fluid inclusion homog-
enization temperatures based on the vol umetric properties of the
system NaC1-H20: Jour. Research U.S. Geol. Survey, v. 5, no. 5,
p. 603-607.
Clynne, M. A., and Brown, D. L., 1978, Freezing point
depression of aqueous sodium chloride solutions: Econ. Geol.,
v. 73, p. 284-285.
Radabaugh, R. E., Merchant, J. S., and Brown, J. M., 1968, Geology
and ore deposits of the Gilman (Red Cliff, Battle Mountain)
District, Eagle County, Colorado, in Ridge, J. D., ed., Ore
Deposits of the United States, Grat-Sales Volume, ASI.M.E.
Publication: p. 641-664.
Robie, R. A., Hemingway, B. S., and Fisher, J. R., 1978, Thermo-
dynamic properiges of minerals and related substances at 298.15
K and 1 bar (10 Pascal s) pressure and at higher temperatures:
U.S.G.S. Bul 1. 1452, 456 p.
Roedder, E., 1962, Studies of fluid inclusions I: Low-temperature
application of a dual-purpose freezing-heating stage: Econ.
Geol., v. 57, p. 1045-1061.
1963, Studies of fluid inclusions II: Freezing data and
their interpretation: Econ. Geol., v. 58, p. 167-211.
1967, Fluid inclusions as samples of ore fluids: in
Barnes, H. L., ed., Geochemistry of hydrothermal ore deposft
(Chapter 12): New York, Hol t, Rinehart and Winston, p. 515-574.
1968a, Temperature, salinity, and origin of the ore-
forming fluids at Pine Point, Northwestlerritories, Canada,
from fluid Inclusion studies: Econ. Geol., v. 63, p. 439-450.
__________, 1968b, The noncolloidal origin of °colliform" textures
in sphaierite ores: Econ. Geol., v. 63, p. 451-471.
1971, FluId inclusion studies on the porphyry-1'pe ore
deposits at Bingham, Utah, Butte, Montana, and Climax,
Colorado: Econ. Geol., v. 66, p. 98-120.
1972, Composi tion of fluid md usions: U.S.G.S. Prof.
Paper 440-JJ, 164 p.
270

1976, Fluid-inclusion evidence on the genesis of ores


in sedimentary and volcanic rocks: In Wolf, K. H. (ed.) Hand-
book of strata-bound and strataform ore deposits. Elsevier,
Amsterdam, v. 2, p. 67-110.
and Bodnar, R. J., 1980, Geologic pressure determina-
tions from fluid inclusion studies: Ann. Rev. Earth Planet.
Sd. Letters, v. 8, P. 263-301.
and Dwornik, E. J., 1968, Sphalerite color banding: lack
of correlation with iron content, Pine Point, Northwest
Territories, Canada: Amer. Mi, v. 53, p. 1523-1529.
and Skinner, B. J., 1968, Experimental evidence that
fluid inclusions do not leak: Econ. Geol., v. 63, p. 715-730.
Romans, P., 1983, Iinpubl ished Ii. S. Bureau of Mines report # E5937P,
Albany, Oregon, 2 p.
Rosenberg, P. E., and Hol land, H. D., 1964, Cal ci te-dol omi te-magne-
site stabili1' relations in solutions at elevated temperatures:
Science, v. 145, p. 700-701.
Rye, R. 0., and Ohmoto, H., 1974, Sul fur and carbon isotopes and ore
genesis: A review: Ecori. Geol., v. 69, p. 826-842.
and Sawkins, F. J., 1974, Fluid inclusion and stable
isotope studies on the Casapalca Ag-Pb-Zn-Cu deposit, central
Andes, Peru: Econ. Geol., v. 69, p. 181-205.
Sakai, H., 1968, Isotopic properties of sulfur compounds in hydro-
thermal processes: Geochem. Jour (Japan), v. 2, p. 29-49.
and Yamamoto, M., 1966, Fractionation of sulfur isotopes
in and preparation of sulfur dioxide--an improved technique for
the precision analysis of stable sulfur isotopes: Geochem.
Jour. (Japan), v. 1, p. 35-42.
Sawkins, F. J., 1969, Chemical brecciation, an unrecognized mechanism
for breccia formation?: Econ. Geol., v. 64, p. 613-617.
Silberman, M. L., and Noble, 0. C., 1977, Age of igneous activity and
mineralization, Cerro de Pasco, central Peru: Econ. Geol., V.
72, p. 925-930.
Skinner, B. J., White, 0. E., Rose, H. J., and Mays, R. E., 1967,
Sul fides associated with the Sal ton Sea geothermal brine: Econ.
Geol., v. 62, p. 316-330.
271

Snyder, F. G., 1967, Criteria for origin of stratiform ore bodies


with application to southeast Missouri: in Brown, J. S., ed.,
Genesis of stratiform lead-zinc-ban te-fi uói9 te deposi tion car-
bonate rocks. Econ. Geol. Mono. 3, p. 1-13.
Stewart, J. W., Evernden, J. F., and Snelling, N. J., 1974,Age
determinations from Andean Peru: A reconnaissance survey: G. S.
A. Bull., v. 85, p. 1107-1116.
Styrtt, M. M., Brackmann, A. J., Hol land, H. D., Clark, B. C.,
Pisutha-Arnond, V., Eldridge, C. S., and Ohmoto, H., 1981, The
mineralogy and isotopic composition of sulfur in hydrothermal
sulfide/sulfate deposits on the East Pacific Rise, 21°N lati-
tude: Earth and Planet. Sci. Letters, V. 53, p. 382-393.
Sorby, H. C., 1858, On the microscopic structure of crystals, indi-
cating the origin of minerals and rocks: Geol. Soc. London
Quart. Jour., v. 14, pt. 1, p. 453-500.
Sourirajan, S., and Kennedy, G. C., 1962, The system H20-NaC1 at
elevated temperatures and pressures: Am. Jour. Sci., V. 260, p.
115-141.

Stelnmann, G., 1929, Geologia de Peru, Univ. Heidelburg.


Streckeisen, A. L., 1973, Plutonic rocks; classification and nomen-
clature recommended by the IUGS subcommission on the systematics
of igneous rocks: Geotimes, October 1973, p. 26-30.
Szekely, T. S., 1967, Geology near Huallacocha Lakes, central high
Andes, Peru: Am. Assoc. Pet. Geol. Bull., v. 51, no. 7, p. 1346-
1353.

1969, Structural geology, Cochas to Yauricocha,


central high Andes, Peru: Am. Assoc. Pet. Geol. Bull., V. 53,
no. 3, p. 553-567.
_______________, and Grose, L. T., 1972, Stratlgraphy of the carbon-
ate, black shale, and phosphate of the Pucara Group (Upper
Tniassic-Lower Jurassic), central Andes, Peru: Geol. Soc.
America Bull., v.83, p. 407-428.
Terrones, A. J., 1949, La estratigraphia del distrito minero de
Morococha: Soc. Geol. Peru, v. Jubflar, XXV Aniversario, Pt. II,
p. 1-15.
1958, Structural control of contact metasomatic
deposits in the Peruvian Cordillera: Am. Inst. Mining Metall.
Petroleum Eng. Trans., v. 211, p. 365-372.
272

Thode, H. G., Monster, J., and Dunford, H. G., 1961, Sulfur isotope
geochemistry: Geochim. Cosmochim. Acta, V. 25, P. 159-174.

Titley, S. R., 1961, Genesis and control of the Lynchberg orebody,


Socorro County, New Mexico: Econ. Geol., v. 56, p. 695-722.
1968, EnvIronment of deposition of the hydrothermal
metamorphic (pyrometasomatic) deposits (abs.): Econ. Geol., v.
63, p. 699.
Tosi, P. A., 1956, Geologia y mineralizacion en Carahuacra, Junin:
Soc. Geol. Peru Bul 1., v. 30, p. 375-384.
Turekian, K. K., 1971, Distribution of elements in the earth: in
Encyclopedia of Science and Technology, 2nd edition, Mcgraw-
Hill.
Urusova, 14. A., 1975, Volumetric properties of aqueous solutions of
sodium chloride at elevated temperatures and pressures: Neorg.
Khim., 20:3103-10 (in Russian) Transl. in Russ. Jour. Inorganic
Chemistry, v. 20 (11), p. 1717-21.
Vera, F., 1977, Geologia del distrito minero San Cristobal: Unpubl.
report, Centromin Peru, 15 P.
Weissberg, B. G., Browne, P. R. L., and Seward, T. M., 1979, Ore
metals in active geothermal systems in Barnes, H. L., ed.,
Geochemistry of hydrothermal ore deposits: John Wiley and Sons,
New York, p. 738-780.
Wilson, J. J., 1963, Cretaceous stratigraphy of the central Andes of
Peru: Am. Assoc. Petrol eum Geol. Bull, v. 47, p. 1-34.
Yamana, 1. and Kato, A., 1976, Mossbauer effect study of 57Fe and
1
Sn in stannite, stannoidite, and mawsonite: Amer. Mm., V.
61, p. 260-265.
Zeil, W., 1979, The Andes: A geological review: Berlin, Gebruder
Borntraeger, 260 p.
APPENDIX
Table Al. Sample locations, descriptions, and summary of ana1yticaI work (see abbreviations below)

Sample' Mine/Area Location Description Analyses2

MB-l-81 UPA (L 680, G 455-N) Small vein which crosscuts tuff bed in limestone; 1,3
contains py, si, gn, and trace cp.

MB-.2-81 SC (L 370, G 544-E) Vein in bx phyllite; contains py, qtz, sl, and 1
botryoidal sid.

118-3-81 SC (L 500, St 791) Vein in phyllite; contains py and banded sl-gn; 1


considerable shear evident in si and gn band.

MB-4-81 HPA 11408N-8240E Massive manto ore; contains py, sl, gn, and qtz,
(L 780+45, Cu 570) w/ many dickite clay pods.

MB-5-81 HPA 11409N-8243E Catalina Volcanics in manto footwall; arg alt w/


(L 780+45, Cu 570) vnits of py and si; some dissm. py and Si.

1-Samples prefixed MB- collected by the author in 1981; all others collected by Fidel Vera of Centromin
Peru in 1980.
2-Analyses, 1=assay, 2=Ag, Cu, Mo, Pb, Zn geochemical analysis, 3=emmlssion spectrographic analysis,
4=whole rock major oxide analysis.
Abbreviations: CA=Carahuacra, HPA=Huaripampa Mine, M=Moises Manto, P=Polvarin Mine, SC=San Cristobal
Mine, T=Toldorrumi Manto, Cc=crosscut, Ch=chlmney, Cu=cuerpo, G=galery, Llevel, St=stope, alt:
alteration, arg:arglllic, bar=barite, bx=breccla, calc=calcareous, carb=carbonate, cp=chalcopyrite,
chl=chlorlte, dissm=disseminated, frac=fracture, frag=fragment, gn=galena, hm=hematlte, mag=magnetite,
mc=marcasite, porph=porphyry, prop=propylltic, qtz=quartz, rhodorhodochrosite, ser=serlcite, sid
siderite, sIllc=silicified, spec=specular, sl=sphalerlte, vnit=veinlet, and wolf=wolframite

t3
(4
Table Al con't

Sample Mine/Area Location Description Analyses

MB-6-81 HPA ll409N-8243E Massive manto ore at Pucara-Catalina Volcanics 1

(L 780+45, Cu 570) contact; contains py and si.

MB-7-81 HPA 11408N-8245E Tuffaceous limestone; replaced by Fe-Mn carb; 2


(L 780+45, Cu 570) contains vnits and dissm py and sl.

M-8-8l SC (L 170, St 810) Vein in phyllite adjacent to Chumpe qtz porph


dike; contains py, qtz, and wolf.

MB-9-81 SC (1 170, St 033) Vein in phyllite; contains banded py, qtz, cp, sl. 1,3

MB-l0-81 SC (L 170, St T-64W) Vein in bx silic phyllite; contains py, qtz, wolf; 1

late frac contains gn, si, and bar.

MB-ll-81 SC (L 220+25, St 979) Vein in phyllite; contains py, qtz, and wolf.

MB-12-81 HPA 1l460N-8530E Small vein in Catalina Volcanics about 30 m from 1


(L 580, G 826-S) Pucara; contains si and gn.

MB-13-81 SC (1 430, St 928) Chumpe qtz porph dike, qtz-ser alt w/ dissm py 2,4

MB-14-81 SC (L 430, St 878) Chumpe qtz porph dike; qtz-ser alt w/ dissm py.

MB-15-81 HPA 11340N-8490E Massive botryoidal sid w/ coating of mc in vuggy


(L 630, Cc 342) alt limestone and manto ore.

MB-16-81 HPA 11366N-8490E Massive manto ore; contains py, sl, qtz, and bar.

MB-18-81 HPA 11411N-8240E Massive manto ore; contains py, al, and qtz showing 1,3
N)
(L 780+45, Cu 570) pseudomorphic replacement of limestone laminae
Table Al con1t

Sample Mine/Area Location Description Analyses

MB-19-81 tWA 11275N-8365E Igneous dike of unknown source which cut limestone; 2,3,4
(L 730, Cu 423) arg alt w/ dissm py and sl; some vnits.

MB-20-81 tWA 11286-8358E Limestone; bx silic WI Fe-Mn carb alt; replaced by 2,3
(1 730, Cu 423) si, spec hm, sid, mc, and mag.

MB-21-81 SC (1 500, St 598E) Vein in phyilite; contains bx py-qtz and banded 1


cp, sl, gn, and sid.

M8-22-81 HPA 11300N-8333E Limestone; bx silic WI Fe-Mn carb alt; replaced by 1


(L 730, Cu 423) spec hm, mag, Si, mc; vuggy.

MB-23-81 tWA 11290N-8338E Limestone; bx silic w/ Fe-Mn carb alt; replaced by 1


(1 730, Cu 423) spec hm, mag, si, sid, mc; vuggy.

MB-24-81 HPA 11255N-8395E Limestone; bx silic w/ Fe-Mn carb alt; replaced by 1


(1 730, Cu 423) spec hm, mag, si, sid, some clay pods and vugs

MB-25-81 HPA 11252N-8400E Limestone; frac, silic, WI Fe-Mn carb alt; replaced 2,3
(L 730, Cu 423) by si, spec, py from cross cutting fracs.

MB-26-Bl HPA 11400N-8235E Massive manto ore; contains py, gn, Si, mc, and bar. 1,3

MB-27-81 HPA 11287N-8466E Igneous dike of unknown source; bx, arg alt, and 2,4
(L 780+45, G 367) frags of phyllite and Catalina Vokanics; x-cuts manto.

MB-28-81 HPA 11400N-8233E Massive manto ore; si w/ abundant clay pods. 1


0. 780+45, Cu 570)

N)
Table Al con't

Sample Mine/Area Location Description Analyses

MB-29-8l HPA , 11400N-8233E Massive manto ore; botryoidai si and sid replac- 1,3
(L 780+45, Cu 570) ing mc crystals; filled open spaces in mantos.

MB-31-81 HPA 11400N-8234E luff and tuffaceous limestone; some hm vnits. 2


(L 780+45, Cu 570)

MB-32-81 HPA 11103N-8938E Vein in Catalina Volcanics, bx si and gn cemented 1


(V 722, L 580) w/ qtz and sid; approx. 30 m from Pucara.

MB-33-81 IWA 11008N-8858E Vein in Catalina Volcanics; bx sl and gn cemented 1,3


(V 755, L 630) WI qtz and sid.

MB-34-81 HPA l1016N-8893E Vein in Catalina Volcanics; banded si, gn, qtz, 1,3
(V 755, L 630) sid, and mc.

MB-35-81 SC (L 500, St 791W) Vein in phyllite; banded gn and dark si.

MB-36-8l SC (1 370, St 929) Chumpe qtz porph dike, qtz-ser alt w/ dissm py 2,4

MB-37-81 SC (L 370, Ch 841) Chumpe qtz porph dike; qtz-ser alt w/ dissm py.

MB-38-81 HPA 11815N-8187E Massive manto ore; py, ruby Si, gn, rhodo, and mc. 1,3
(L 540, G 813-N)

MB-39-81 CA 11960N-8675E Carahuacra qtz monz, moderate prop alt, located 2,4
(surface) at the contact with Excelsior Phyllite.

MB-40-81 CA 12000N-8680E Carahuacra qtz monz; moderate prop alt; located 2,4
(surface) 10 m from phyllite contact
Table Al con't

Sample Mine/Area Location Description Analyses

MB-41-81 CA 12125N-8500E Carahuacra qtz monz; moderate prop alt; located 2,4
(surface) near center of intrusive.

MB-42-81 P surface Chumpe qtz porph dike; contains frags of phyllite; 2


qtz-ser alt wI dissm. py.

MB-43-81 P surface Chumpe qtz porph dike; qtz-ser alt Wi chi vnits. 1,2,3,4

MB-44a-81 P surface Chumpe qtz porph dike; qtz-ser alt; minor alunite 1,2,4
and jarosite.

MB-44b-81 P surface Chumpe qtz porph dike; qtz-ser alt; minor alunite 2,3,4
and jarosite.

1413-45-81 P surface Chumpe qtz porph dike; contains phyilite and bull 2
qtz frags; qtz-ser alt.

MB-46-81 tWA 11412N-8231E Tuffaceous limestone; silic, Fe-Mn carb alt, replaced 1
(L 780+45, Cu 570) by py, si, and gn.

MB-47-81 I surface Vein in Catalina Volcanics; oxidized; contains py, cp, 1


si, sid, and spec hm.

MB-48-81 I 9480N-10935E Basal Pucara; calc ss; silic and replaced by sid, hm. 2

MB-49-81 I 9595N-10630E Basal Pucara; ss; silic, chi w/vnits of spec hm, bar, 2
andsi.

MB-50-81 I 9595N-10630E Limestone, alt to Fe-Mn carb 2,3


N)
-.4
Table Al con't

Sample Mine/Area Location Description Analyses

MB-51-81 T 9540N-10720E Basal Pucara; ss; intensely silic and replaced 1


by chi, and spec hm.

MB-52-81 I 9540N-10720E Basal Pucara; ss; bx, silic, WI vnits of spec hm, 1
sl, cp, qtz; large sid crystals in vugs.

MB-53-81 I 9540N-10720E Basal Pucara; ss; completely replaced by spec hrn, 1


sl, and qtz; silic.

MB-54-81 I 9480N-10720E Limestone; completely alt to Fe-Mn carb. 2

MB-55-81 I 9470N-10690E Limestone, alt to Fe-Mn carb and silic 2

MB-56-81 I 9485N-10625E Limestone; alt to Fe-Mn carb; x-cut by numerous 2,3


vnits containing si and sid.

MB-57-81 I 9530N-10460E Limestone, bx and alt to Fe-Mn carb, completely 1,2,3


oxidized to Mn-rich oxides.

MB-58-81 M Limestone; alt to Fe-Mn carb and cut by numerous 1,2,3


vnits containing sl and sid

MB-59-81 M Vein in Catalina Volcanics, 20 m from Pucara, 1


contains cp, si, qtz, and sid

MB-60-81 M 8600N-11730E Limestone; contains vnit of Mn-oxide WI selvage of 2


Fe-Mn carb alt grading into unalt limestone

MB-61-81 M 8370N-11940E Montero Basalt, flow in Pucara Group 2


Table Al con't

Sample Mine/Area Location Description Analyses

MB-62-8l M 8370N-12040E Limestone; silic, bx and replaced by Fe-Mn carb; 2


qtz in vugs around frags.

MB-63-81 M 8075N-l2500E Limestone; silic to jasperold; remenant bedding. 2

MB-64-8l M 7950N-12670E Limestone; bx, silic, and cemented wI qtz, bar, 2,3
si, and sid; somewhat oxidized.

MB-65-81 M 7950N-12670E Limestone; bx, vuggy, cemented and replaced by 2


sid, qtz, bar, Si, and Mn-oxides.

MB-66-81 M 7950N-12670E Limestone; bx, buggy, cemented and replaced by 2


sid, qtz, bar, si, and Mn-oxides.

MB-67-81 M 7950N-12670E Massive manto ore; bx and completely oxidized to 1,2


Mn-Fe oxides.

MB-68-81 M 7880N-12670E Limestone; oolitic; unalt, bx, WI trace of dia- 2,3


genetic dolomite.

MB-69-81 HPA Open Pit luff, arg alt w/ considerable dickite 2,3

MB-71-81 HPA Open Pit Jasperoid, vuggy WI massive mc in vugs 1

MB-72-81 HPA Open Pit Massive manto ore, contains py, ruby Si, gn, and 1
rhodo.

MB-74-81 HPA Open Pit Limestone, bx and replaced by rhodo, ruby si, and gn
Table Al con't

Sample Mine/Area Location Description Analyses

MB-75-81 HPA 11276N-8388E Massive manto ore; contains si, spec bin, py, and 2
clay pods; bedding plane vnits of si below tuff.

MB-77-81 HPA Massive mantoore; banded botryoidal si and sid 3


w/ coating of mc crystals.

MB-78-81 P surface Chumpe qtz porph dike; qtz-ser alt WI dissm py. 2

MB-80-81 UPA 11435N-8225E Massive manto ore; silic, vuggy w/ crystals of 2


(1 780+45, Cu 570) py, dark si, and qtz in vugs.

AND M-1 A Andaychagua Vein Vein in Catalina Volcanics; bx volcanics cemented


w/ si, gn, ruby silver; coating of native silver.

AND M-2 A Andaychagua Vein Vein in Catalina Volcanics; massive si, gn, and rhodo.

CH-1 CA surface Carahuacra qtz monz; moderate prop alt; near contact 2,4
with Excelsior Phyllite.

CH-2 CA surface Carahuacra qtz monz; moderate prop alt; 10 m from


Excelsior Phyllite.

CH-3 CA surface Carahuacra qtz monz; moderate prop alt; 300 in from 2,3,4
Excelsior Phyllite.

CH-4 CA surface Carahuacra qtz monz; moderate prop alt; near center 2,4
of intrusive.

HPA M-1 HPA 11447N-8438E Massive manto ore; contains py, qtz, si, spec hrn, sid;
(L 580, Cu 423) some vugs lined with si and sid crystals
Table Al con1t

Sample Mine/Area Location Description Analyses

HPA M-2 HPA 11379N-8451E Limestone; silic and replaced by sid; zebra texture.
(1 580, Cu 423)

HPA M-3 UPA 11106N-8631E Massive manto ore; contains py, qtz, si, and sid;
(1 580, Cu 658A) some spec hm and mag

HPA M-4 HPA 11159N-8610E Vein 199 in Pucara Group; massive ruby si and gn w/
(1 580, Cu 119A) botryoidai sid in vugs.

HPA M-5 UPA 11058N-8655E Massive manto ore; contains py, si, qtz, spec hm,
(1 630, Cu 658) and mag.

HPA M-6 UPA 11118N-8556E Massive manto ore;, contains py, si, qtz; some vugs 3
(1 630, Cu 199A) lined with qtz and si crystals.

HPA M-7 HPA 10860N-8670E Vein 755 in Pucara; massive sl w/ minor qtz, cut by
(L 630, G 799w) vnits of sid and mc, one dickite pod

HPA M-8 tWA 10774N-8386E Vein 722: silic volcanics, bx and mostly replaced by
(L 630, G 777) py, si, and qtz.

HPA M-9 HPA 11421N-8311E Massive manto ore; contains py and dark Si; vuggy;
(L 730, Cu 570) fracs coated w/ bar and gn.

HPA M-10 HPA 11420N-8312E Massive manto ore; contains alternating layers of py 3
(L 730, Cu 570) and dark si; some vugs lined with crystals

HPA M-11 HPA 11395N-8326E Massive manto ore; contains alternating layers of py 3
(1 730, Cu 570) and dark si, some vugs lined with si crystals
Table Al con't

Sample Mine/Area Location Description Analyses

HPA R-1 HPA 11475N-8400E Limestone; partially silic WI Fe-Mn alt; vnits of 2
(L 580, Cu 423) si and spec hm; somewhat porous.

UPA R-2 I-WA 11483N-8400E luff or tuffaceous limestone; arg alt; graded 2
(1 580, Cu 423) bedding.

HPA R-3 HPA 11111N-8629E Limestone; silic w/ incipient clay pods; about 2
(L 580, Cu 423) 2 m from vein 658-A.

tWA R-4 tWA 11091N-8617E luff; highly sheared, bx, w/arg alt; near vein.

HPA R-5 I-WA 11166N-8605E Limestone; completely silic; some vugs; about 2
(L 580, Cu 119A) 1 m from vein 119A.

tWA R-6 HPA 11166N-8593E luff or tuffaceous limestone; highly bx and


(1 580, Cu ll9A) and sheared; silic and replaced by Fe-Mn carb.

tWA R-7 HPA 11069P4-8619E Limestone; silic WI many incipient clay pods and 2
(L 630, Ct 658) porous areas.

I-WA R-8 FIPA 11076N-8637E luff or tuffaceous limestone, bx, arg alt, w/ vnits
(1 630, Cu 658-A) of si and sid.

I-WA R-9 HPA 11120N-8596E Limestone; completely silic and replaced by Fe-Mn 2
(L 630, Cu 119A) carb; some dissolution vugs along bedding planes.

I-WA R-l0 HPA 11120N-8590E luff or tuffaceous limestone; sheared, silic, arg
(L 630, Cu ll9A) alt; some vnits of py, si, and qtz.

UPA R-ll FIPA 10864N-8670E Limestone adjacent to vein 755, completely silic, 2
(L 630, G 799W) incipient bx wI vugs along fracs
Table Al con't

Sample Mine/Area Location Description Analyses

HPA R-12 UPA 11500N-8320E Igneous dike cross cutting limestone; unknown 2
(L 680, Cu 570) source; arg alt and vnits of py, qtz and si.

tWA R-13 HPA 11467N-8280E Igneous dike cross cutting limestone; arg alt w/ 2
(L 730, Cu 570) vnits of py, si, and qtz

HPA R-14 tWA 11384N-8315E Tuffaceous limestone; silic w/ dissm py along 2


(L 730, Cu 570) bedding laminae; vnits of py and si.

FIPA R-15 tWA 11452N-8261E Limestone; silic and frac; vnits of sid and si. 2
(L 730, Cu 570)

M-1 tWA 11118N-8642E Massive manto ore adjacent to vein 658; contains
(1 580, Cc 127) py, sl, spec tim, mag, and sid.

M2 tWA 11080N-89OOE Vein 722 in Catalina Volcanics; massive si, gn, 3


(L 630, St 777) qtz, and sid w/ vugs lined w/ sl crystals

M-3 HPA 11348N-8446E Massive manto ore; contains py, sl, spec hm, mag,
(L 630, Cu 423) sid, and qtz.

M-4 HPA 11602N-8340E Limestone; bx and vuggy; mostly replaced by py,


(L 630, Cu 570) si, qtz, and sid, many clay pods, vugs lined WI
mc and sid crystals.

M-5 HPA 11413N-8363E Massive manto ore; contains dark si w/ sid coating 3
(L 730, Cu 570) sl.

M-6 HPA 11424N-8360E Massive manto ore; contains py, qtz, spec hm; 3
(L 680, Cu 570) bedding preserved, possible sedimentary structures N)
(A)
Table Al con't

Sample Mine/Area Location Description Analyses

M-7 IWA 11454N-8289E Massive manto ore, contains alternating bands of


(1 730, Cu 570) py, sl, and qtz; some clay pods.

M-8 tWA ll599N-8247E Massive manto ore; contains coarse crystals of sl


(L 730, Cu 570) and py, many vugs

M-9 tWA (L 730, Cu 423) Limestone, bx, silic, and partially replaced by py,
sl, qtz, and sid, bx frags surrounded by ore

POLV M-1 P 10425N-ll?30E Polvarin Vein, coarse py crystals and qtz cross-
(G 958) cutting phyllite W/ qtz-ser alt.

POLV M-2 P 10615N-llO9OE Polvarin Mine, vein in Excelsior Phyllite, massive


(G 997) py and qtz

POLV R-1 P 10360N-lll7OE Excelsior Phyllite; qtz-ser + arg alt WI dissm. py.

POLV R-2 P 10634N-lll24E Chumpe qtz porph, qtz-ser ± arg alt w/dissm py 2,4

R-1 HPA 1 580 Limestone, silic, some dissolution voids along


bedding planes, zebV'a texture, carb in vugs

R-2 HPA 11107N-8627E Limestone, bx and silic, vugs around bx frags coated
(L 580, Cu 658A) WI sid, vnit of mc

R-3 HPA 11120N-8660E Tuffaceous limestone, completely silic WI some 2


(1 580, Cu 658A) preservation of bedding.

R-4 tWA lll2ON-8663E Limestone, completely siuic W/ preservation of


(L 580, Cu 6581) bedding.
Table Al con't

Sample Mine/Area Location Description Analyses

R-5 FIPA 11602N-8336E Finely laminated tuff WI conspicuous graded 2,3


(L 630, Cu 570) bedding; now mostly alt to dickite.

R-6 tWA 10880N-8700E luff, porous and friable, completely alt to clay, 2,3
(L 630, Cu 570) probably dickite

R-7 HPA 11409N-8360E Fine-grained lapilli tuff; arg alt.


(L 730, Cu 570)

R-8 HPA 11375N-8343E Jasperoid, vugs along frac and bedding planes 2
(L 730, Cu 570)

R-l0 tWA 11373N-8391E Tuffaceous limestone; partially silic, bx, and 2


(L 730, Cu 570) arg alt; si and py in fracs.

R-li HPA 11450N-8343E Igneous dike in limestone; arg. alt. w/ dissm py; 2
(L 730, Cu 570) numerous vnits of py and si.

R-12 HPA. ii454N-8297E Coarse-grained lapilii tuff-limestone; siiic and 2


(L 730, Cu 570) replaced by sid.

R-13 tWA 11591N-8247E Jasperoid; bx; frags surrounded and replaced by 2


(L 730, Cu 570) py, si, gn, and sid

R-14 HPA (L 630, Cu 424) Coarse-grained lapilli tuff; numerous graded layers; 2
partially silic; clay pods; numerous vnits of py.

SCM-i SC L 120, St 979W Vein in phyllite; coarse cp, si, and qtz; material
is highly sheared and bx.
N)
03
01
Table Al conat

Sample Mine/Area Location Description Analyses

SCM-2 Sc i 120, St 979W Vein in phyllite; bx phyilite surrounded by some- 3


what vuggy py, qtz and wolf ore

SCM-3 SC 1 120, St 820 Vein in phyllite adjacent to Chumpe qtz porph dike;
contains coarse py, qtz, wolf

SCM-4 Sc L 170, St 026 Vein in phyilite, alternating bands of cp and dark 3


si in bx phyilite WI late bar and Si in fracs

SCM-5 SC L 170, St 064 Vein in phyllite; massive banded py, qtz, wolf WI
later cp and si ore; cut by carb bands.

SCM-6 SC 1 220, St 032 Vein in bx phyllite, vnits of py, cp, qtz, and si
cutting phyllite frags, probably replacement

SCM-7 SC L 370, Ch 612 Vein in phyllite, coarse py, cp, qtz, and Si ore 3

SCM-8 Sc 1 370, G 342E Vein in phyilite adjacent to Chumpe qtz porph;


coarse py, qtz, wolf or in vx phyllite.

SCM-9 SC L 500, St 872 Vein in phyilite, coarse massive gn ore 3

SCM-10 Sc L 500, St 727 Vein in phyllite; massive ruby si WI minor cp and 3


gn cut by late bar-sid vnit

SCM-il SC 1 170, Ch 811 Vein in phyliite adjacent to Chumpe qtz porph dike;
massive py, qtz, wolf ore.

SCM-12 SC 1 270, Cc 777 Vein in phyilite, coarse py, qtz, wolf ore
N)
Table Al con't

Sample Mine/Area Location Description Analyses

SCM-13 Sc L 430, St 689E Vein in phyllite, bx and siuic phyllite w/ py,


qtz, and wolf.

SCM-14 Sc L 320, G 370 Vein in phyllite, bx black phyllite w/ qtz, cp,


and py cementing frags, some highly bx si

SCR-1 SC I Chumpe qtz porph, qtz-ser alt w/ dissm py and qtz-


py vnits.

SCR-2 SC Chumpe qtz porph, qtz-ser alt WI dissm py

SCR-3 SC L 370 Chumpe qtz porph, qtz-ser alt WI dissm py 2,4

SCR-4 SC L 170, C 987 Excelsior Phyllite; qtz-ser ± arg alt; silic WI


some dissm py

SCR-5 SC L 500 Catalina Volcanics, arg alt w/ some dissm py 4

TRI-1 SC Triunfo Mine Vein in Catalina Volcanics in easternmost part of


the San Cristobal Vein, massive tetrahedrite wI
some cp, gn, bn, si, and argentite

Co
-'3

You might also like