You are on page 1of 728

Springer Geology Field Guides

Soumyajit Mukherjee Editor

Structural
Geology and
Tectonics
Field Guidebook—
Volume 1
Springer Geology

Springer Geology Field Guides

Series Editor
Soumyajit Mukherjee, Department of Earth Sciences, Indian Institute of
Technology Bombay, Mumbai, Maharashtra, India
Springer Geology Field Guides is a book series that provides the details of both
well-known and little known transects to discover the beauty and knowledge of
Geology, worldwide. Springer Geology Field Guides aims to bring geology field
trips to professionals, students, and amateurs to find the most interesting geology
worldwide. This series includes carefully crafted guidebooks that help generations
of geologists explore the terrain with minimum or no guidance. In this series, the
audience will also find field methodologies and case studies as examples.
This book series will welcome both authored and edited field guides of all geology
disciplines, including structural geology, tectonics, sedimentology, stratigraphy,
paleontology, economic geology, among others. Photo-atlases are also welcome.

More information about this subseries at http://www.springer.com/series/16656


Soumyajit Mukherjee
Editor

Structural Geology
and Tectonics Field
Guidebook—Volume 1

123
Editor
Soumyajit Mukherjee
Department of Earth Sciences
Indian Institute of Technology Bombay
Mumbai, Maharashtra, India

ISSN 2197-9545 ISSN 2197-9553 (electronic)


Springer Geology
ISSN 2730-7344 ISSN 2730-7352 (electronic)
Springer Geology Field Guides
ISBN 978-3-030-60142-3 ISBN 978-3-030-60143-0 (eBook)
https://doi.org/10.1007/978-3-030-60143-0
© Springer Nature Switzerland AG 2021
This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part
of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission
or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt from
the relevant protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the
authors or the editors give a warranty, expressed or implied, with respect to the material contained
herein or for any errors or omissions that may have been made. The publisher remains neutral with regard
to jurisdictional claims in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Dedicated to all fieldwork-based mappers,
geoscientists and surveyors
Preface

Despite the explosive growth of structural geology and tectonics in the last few
decades, structural geological and tectonic fieldwork remain indispensable com-
ponents of geosciences education in Bachelor’s and Master’s level worldwide.
A (new) field instructor might be in search for (new) terrains close to her institute to
explore and demonstrate structures to students. This book fills up that requirement.
Through 26 main chapters, 84 authors and co-authors from 13 countries, the
book presents few well-known and several rather unknown transects where exciting
structures exist, and field-programs can be established.
Cite individual chapters in the following format:
Kaplay RD, Babar Md, Mukherjee S, Wable D, Pisal K. 2021. Granitic rocks
underlying Deccan trap along the margin of east Dharwar craton, Muntyal
(Maharashtra)—Bhaisa (Telengana), India—general description and deformation.
In: Mukherjee S. (Ed) Structural Geology and Tectonics Field Guidebook—
Volume 1. Springer Nature Switzerland AG. Cham. pp. 599-620. ISBN: 978-3-030-
60142-3.
Cite this book in the following format:
Mukherjee S. (2021) Structural Geology and Tectonics Field Guidebook—Volume
1. Springer Nature Switzerland AG. Cham. pp. 1-723. ISBN: 978-3-030-60142-3.

Mumbai, India Soumyajit Mukherjee


February 2021 soumyajitm@gmail.com
smukherjee@iitb.ac.in

vii
Acknowledgements

Dripta Dutta (IIT Kanpur), Mohamedharoon A. Shaikh (MS University Baroda),


Mohit Kumar Puniya (National Geotechnical Facility, Dehradun), and Narayan
Bose (IIT Kharagpur) assisted profusely in preparing this book. The Springer
(proofreading) team, especially Boopalan Renu, Alexis Vizcaino, Doerthe
Mennecke-Buehler, Marion Schneider, Monica Janet Michael and Annett
Buttener helped a lot in different stages of this book preparation and publication.
I thank the contributing authors and the reviewers for participation. I am espe-
cially thankful to the staff members of my Department for always cooperating. They
are Nilesh K. Borkar, Dr. Trupti V. Chandeasekhar, Shamit N. Karnekar,
Ramu K. Khandagale, Rajesh Y. Manjrakar, Dr. Shilpa V. Netravali,
P. S. Sawant, Javed M. Saikh, Srikanth Jonnala, Staphen T., Mary Thomas,
N. N. Vengulakar and Premkumar R. Verma.

ix
Introduction to Structural Geology and Tectonics
Field Guidebook

This book consists of 26 main chapters.


Chapter “Creating Geologic Maps in the Twenty-First Century: A Case Study from
Western Ireland”: Swanger and Whitmeyer (2021) discuss the modern techniques
of field mapping and the creation of geologic maps using recent software. The
authors also elaborate the same using a case study from Western Ireland.
Chapter “Strain Softening in a Continental Shear Zone: A Field Guide to the
Excursion in the Ferriere-Mollières Shear Zone (Argentera Massif, Western Alps,
Italy)”: Simonetti et al. (2021) field trip in the Western Alps shows evidences of
strain softening from ten field stops in the central portion of the Ferriere-Mollières
shear zone. The authors further constrain the shearing event between 340 and 320
Ma using in situ U-Th-Pb petrochronology on monazite.
Chapter “The Geometry and Kinematics of the Southwestern Termination of the
Pyrenees: A Field Guide to the Santo Domingo Anticline”: Field trip to the Santo
Domingo anticline in the Pyrenees by Pueyo et al. (2021) reveals complex fold
kinematics from the fold–thrust belt.
Chapter “Miocene-Quaternary Strain Partitioning and Relief Segmentation Along
the Arcuate Betic Fold-and-Thrust Belt: A Field Trip Along the Western Gibraltar
Arc Northern Branch (Southern Spain)”: Jiménez-Bonilla et al. (2021) discuss the
strain partitioning modes from the hinge to the lateral zones of the Western
Gibraltar Arc (southern Spain). Two separate itineraries are presented for the same.
Chapter “The Southern Iberian Shear Zone (SW Spain): Inclined Transpression
Related to Variscan Oblique Convergence in a HT/LP Metamorphic Belt”:
Díaz-Azpiroz and Fernández (2021) present the ductile mesostructures from the
boundary between the Ossa-Morena and South Portuguese Zones of the Iberian
Massif (Huelva Province, SW Spain). This boundary developed during the Upper
Paleozoic due to the sinistral oblique collision between Avalonia and Armorica.
Chapter “A Field Guide to the Spectacular Salt Mines of the Transylvanian Basin
and Romanian Carpathians”: Tămaș et al. (2021) describe field trips in the
Romanian Carpathians and the Transylvanian Basin to study the 3D structural

xi
xii Introduction to Structural Geology and Tectonics Field Guidebook

features of the salt domes and diapirs in the abandoned salt mines. They propose a
route with five stops to explain the link between hydrocarbon and salt tectonics.
Chapter “Spectacular Sandstone Rock Cities in the Czech Republic”: Novakova
and Novak (2021) describe sandstone rock cities from the Czech Republic. The
Cretaceous sandstones are broken into blocks during the Alpine orogenesis and
subsequently eroded to form these spectacular exposures.
Chapter “Field Guide to RODS in the Pireneus Syntaxis, Central Brazil”:
Martins-Ferreira and Rodrigues (2021) present a field guide focussing the linear
structural features from the Pireneus Range in Central Brazil. They describe the
occurrences as observed at the ten field stations from the Three Peaks area
(TrêsPicos) to the Mocó Boulders.
Chapter “Low Baric Metamorphic Belts in the Northern Tip of the Arabian–Nubian
Shield: Selected Examples from the Eastern Desert/Midyan Terranes, Egypt”:
Shallaly and Abu Sharib (2021) explore the pelitic metasediments from the LP/HT
andalusite-sillimanite-type metamorphic belts of the Arabian–Nubian shield in
Egypt. They report multiple phases of deformation in such belts.
Chapter “Review of the Geometric Model Parameters of the Main Himalayan
Thrust”: Ansari (2021) reviews geometric model parameters of the Main Himalayan
Thrust along different portions of the Himalaya. He compiles variation in the
dip-slip and strike-slip rates of this thrust along the Himalayan belt. This chapter is
not a field guide strictly speaking, but keeping this work in mind will be important
for the Himalayan field geologists.
Chapter “Traverses Through the Bagalkot Group from North Karnataka State,
India: Deformation in the Mesoproterozoic Supracrustal Kaladgi Basin”: Patil Pillai
and Kale (2021) conduct fieldwork along four different traverses within the Balakot
Group of the Kaladgi Basin. They report several mesoscale structural features,
primary sedimentary structures and bedding plane characters.
Chapter “Tectonic Framework of Northern Pakistan from Himalaya to Karakoram”:
Ali et al. (2021) explore the rocks along the Islamabad–Khunjerab transect of the
Pakistan Himalaya. They describe the lithounits encountered over a period of four
days that comprise 27 field stops.
Chapter “Structures of Lesser/Greater Himalaya in and Around an Out-of-Sequence
Thrust in the Chaura-Sarahan Area (Himachal Pradesh, India)”: Ghosh and
Mukherjee (2021) present detail field structural features near an out-of-sequence
thrust in the Western Himalaya in India. Such thrusts have been studied so far
mainly from geochemical perspectives, and field data were so far missing. This
chapter presents a good example of ductile and brittle shear sense indicators, so the
reader is referred to few recent publications”: Mukherjee (2011, 2013, 2014a, b,
2015, in press) and Mukherjee et al. (2020).
Chapter “Structural Geology Along the Nainital-Pangot Road (Kilbari Section),
Nainital Lesser Himalaya (Uttarakhand, India): Focus on Back-Structures”: Puniya
and Mukherjee (2021) study the structural geology along the Nainital–Pangot road
(Kilbari section) in the Nainital Lesser Himalaya, Uttarakhand, India. The authors
report several mesoscale back structures. Such structures have increasingly been
Introduction to Structural Geology and Tectonics Field Guidebook xiii

reported from the Himalaya (e.g., Mukherje 2013; 2019; Bose and Mukherjee
2019a, b).
Chapter “Geology, Structural, Metamorphic and Mineralization Studies Along the
Mandi-Kullu-Manali-Rohtang Section of Himachal Pradesh, NW-India”: Singh
et al. (2021) present lithounits and structures along the Mandi-Larji-Kullu-
Manali-Rohtang La transect in the NW Indian Himalaya.
Chapter “Tectonics and Channel Morpho-Hydrology—A Quantitative Discussion
Based on Secondary Data and Field Investigation”: Biswas et al. (2021) compute 30
geomorphic indices to describe the river channel morphologies and their tectonic
controls. They choose three study sites from India: the NE foreland basin of North
Bengal, the Singbhum Shear Zone (SSZ) and the Janauri–Chandigarh anticline.
Chapter “Geological Field Guide: Malvan (Maharashtra, India)”: Pundalik et al.
(2021) present detail of fieldworks from Malvan (Maharashtra, India) from the
lithologic, geomorphic and structural perspectives.
Chapter “A Field Guide to the Champaner Region, Southern Aravalli Mountain
Belt (SAMB), Gujarat, Western India”: Joshi and Limaye (2021) discuss the
structural features in the Paleoproterozoic basement gneisses to the recent sediments
of the Champaner region in Eastern Gujarat (India). They elaborate the lithounits
and structures encountered from 15 field stops along four different traverses.
Chapter “Importance of Fracturing in Uranium Mineralization in Gulcheru
Quartzite Host: A Case from Ambakapalle Area, Cuddapah Basin, Andhra
Pradesh, India”: Goswami et al. (2021a) map the fault zone in the Ambakapalle area
within the Cuddappah Basin. They focus on fractures and their influence on ura-
nium mineralization. The authors also discuss two phases of the alteration of rocks.
Chapter “Granitic Rocks Underlying Deccan Trap Along the Margin of East
Dharwar Craton, Mutnyal (Maharashtra)—Bhaisa (Telangana), India—General
Description and Deformation”: Kaplay et al. (2021) study the structural features
from the contact between the Eastern Dharwar craton and the Deccan Volcanic
Province. They detail shear tectonics along the contact.
Chapter “Structural Analyses of the Lunavada–Santrampur Area (Gujarat, India)
Using Remote Sensing Images”: Chauhan et al. (2021) analyze the folds and lin-
eaments from the Santrampur area (NE Gujarat, India) using remote sensing ima-
ges. They use Google Earth for identifying various folds geometries, viz. polyclinal
folds, second-order folds and superposed folds. This chapter will enable field
geologist to get into the detail of the terrain.
Chapter “Fundamentals of Lithostructural Mapping: Example from the SW Part
of the Proterozoic Bhima Basin, Karnataka, India: A Note on Dharwarian Crustal
Evolution”: Goswami et al. (2021b) explore the geodynamic evolution of the
Eastern Dharwar craton with the help of GPS-aided lithostructural mapping of the
SW part of the Proterozoic Bhima Basin.
Chapter “A 3D Photogrammetric Approach in Mapping Meso-Scale Folds and
Shears in Structurally Controlled Syngenetic Mn-Mineralised Zones of Shivrajpur
Region, Eastern Gujarat, India”: Joshi (2021) describes an innovative technique of
mapping mesoscale structures using 3D photogrammetry. The author maps an
xiv Introduction to Structural Geology and Tectonics Field Guidebook

outcrop scale fold from an abandoned mine from the Mn-mineralized zones of the
Shivrajpur region (Eastern Gujarat, India).
Chapter “Vein Geometry Around Bhuj (Gujarat, India)”: Omid et al. (2021) present
diverse vein geometries from Bhuj area, Kutch Basin, Gujarat, India. Detail
field-based and geochemical studies can be taken up in this hitherto unknown area
of structures.
Chapter “Oriented Rock Samples for Detailed Structural Analysis”: Gaidzik and
Żaba (2021) discuss how to collect oriented rocks from field for structural geo-
logical analyses. This chapter is particularly important to undertake kinematic
analyses of shear zone rocks under an optical microscope.

Acknowledgements The Springer (proofreading) team, especially Boopalan Renu, Alexis


Vizcaino, Doerthe Mennecke-Buehler, Marion Schneider and Annett Buttener, helped a lot in
different stages of this book preparation and publication.

Dripta Dutta
Soumyajit Mukherjee
Department of Earth Sciences
Indian Institute of Technology Bombay
Powai, Mumbai
Maharashtra, India
e-mail: soumyajitm@gmail.com
smukherjee@iitb.ac.in

References

Ali, A., Ahmad, S., Khan, M. A., Khan, M. I., Rehman, G. (2021). Tectonic framework of
Northern Pakistan from Himalaya to Karakoram. In S. Mukherjee (Ed.), Structural Geology
and Tectonics Field Guidebook—Volume 1. Springer Nature Switzerland AG. Cham.
pp. 367-412. ISBN: 978-3-030-60142-3.
Ansari, K. (2021). Review of the geometric model parameters of the Main Himalayan Thrust. In S.
Mukherjee (Ed.), Structural Geology and Tectonics Field Guidebook—Volume 1. Springer
Nature Switzerland AG. Cham. pp. 305-324. ISBN: 978-3-030-60142-3.
Bose, N., Mukherjee, S. (2019a). Field documentation and genesis of the back-structures from the
Garhwal Lesser Himalaya, Uttarakhand, India. In Sharma, I. M. Villa, S. Kumar (Eds.), Crustal
architecture and evolution of the Himalaya-Karakoram-Tibet Orogen. Geological Society of
London Special Publications 481, 111–125.
Bose, N., Mukherjee, S. (2019b). Field documentation and genesis of back-structures in ductile
and brittle regimes from the foreland part of a collisional orogen: Examples from the
Darjeeling–Sikkim Lesser Himalaya, India. International Journal of Earth Sciences, 108,
1333–1350.
Biswas, M., Pal, A., Jamal, M. (2021). Tectonics and Channel Morpho-Hydrology—A
Quantitative Discussion Based on Secondary Data and Field Investigation. In S. Mukherjee
(Ed.), Structural Geology and Tectonics Field Guidebook—Volume 1, Switzerland AG:
Springer Nature. Cham. pp. 461-494. ISBN: 978-3-030-60142-3.
Introduction to Structural Geology and Tectonics Field Guidebook xv

Chauhan, G. H., Rao, G. S., Mukherjee, S. (2021). Structural analyses of the Lunavada–
Santrampur area (Gujarat, India) using remote sensing images. In S. Mukherjee (Ed.),
Structural Geology and Tectonics Field Guidebook—Volume 1. Springer Nature
Switzerland AG. Cham. pp. 621-638. ISBN: 978-3-030-60142-3.
Díaz-Azpiroz, M., Fernández, C. (2021). The Southern Iberian Shear Zone (SW Spain): Inclined
transpression related to Variscan oblique convergence in a HT/LP metamorphic belt.
In S. Mukherjee (Ed.), Structural Geology and Tectonics Field Guidebook—Volume 1.
Springer Nature Switzerland AG. Cham. pp. 137-166. ISBN: 978-3-030-60142-3.
Ghosh, R., Mukherjee, S. (2021). Structures of Lesser/Greater Himalaya in and around an
out-of-sequence thrust in the Chaura-Sarahan area (Himachal Pradesh, India). In S. Mukherjee
(Ed.), Structural Geology and Tectonics Field Guidebook—Volume 1. Springer Nature
Switzerland AG. Cham. pp. 413-428. ISBN: 978-3-030-60142-3.
Goswami, S., Shrivastava, S., Das, S., Bhattacharjee, P. (2021a). Fundamentals of litho-structural
mapping: Example from the SW part of the Proterozoic Bhima basin, Karnataka, India: A note
on Dharwarian Crustal evolution. In S. Mukherjee (Ed.), Structural Geology and Tectonics
Field Guidebook—Volume 1. Springer Nature Switzerland AG. Cham. pp. 639-684. ISBN:
978-3-030-60142-3.
Goswami, S., Upadhyay, P. K., Natarajan, V. (2021). Importance of fracturing in uranium min-
eralization in Gulcheru Quartzite host: A case from Ambakapalle area, Cuddapah Basin,
Andhra Pradesh, India. In Mukherjee S. (Ed.), Structural Geology and Tectonics Field
Guidebook—Volume 1. Springer Nature Switzerland AG. Cham. pp. 577-598. ISBN:
978-3-030-60142-3.
Gaidzik, K., Żaba, J. (2021). Oriented Rock Samples for Detailed Structural Analysis. In: S.
Mukherjee (Ed) Structural Geology and Tectonics Field Guidebook—Volume 1. Springer
Nature Switzerland AG. pp. 715- 723.
Jiménez-Bonilla, A., Díaz-Azpiroz, M., Expósito, I., Balanyá, J. C. (2021). Miocene-Quaternary
strain partitioning and relief segmentation along the arcuateBetic fold-and-thrust belt: A field
trip along the Western Gibraltar Arc northern branch (southern Spain). In S. Mukherjee (Ed.),
Structural Geology and Tectonics Field Guidebook—Volume 1. Springer Nature
Switzerland AG. Cham. pp. 103-136. ISBN: 978-3-030-60142-3.
Joshi, A. U., Limaye, M. A. (2021). A field guide to the Champaner Region, Southern Aravalli
Mountain Belt (SAMB), Gujarat, Western-India. In S. Mukherjee (Ed.), Structural Geology
and Tectonics Field Guidebook—Volume 1. Springer Nature Switzerland AG. Cham.
pp. 529-576. ISBN: 978-3-030-60142-3.
Joshi, A. U. (2021). A 3D photogrammetric approach in mapping meso-scale folds and shears in
structurally controlled syngenetic Mn-mineralized zones of Shivrajpur region, Eastern Gujarat,
India. In S. Mukherjee (Ed.), Structural Geology and Tectonics Field Guidebook—Volume 1.
Springer Nature Switzerland AG. Cham. pp. 685-706. ISBN: 978-3-030-60142-3.
Kaplay, R. D., Babar, Md., Mukherjee, S., Wable, D., Pisal, K. (2021). Granitic rocks underlying
Deccan trap along the margin of east Dharwar craton, Muntyal (Maharashtra)—Bhaisa
(Telengana), India—general description and deformation. In S. Mukherjee (Ed.), Structural
Geology and Tectonics Field Guidebook—Volume 1. Springer Nature Switzerland AG. Cham.
pp. 599-620. ISBN: 978-3-030-60142-3.
Martins-Ferreira, M. A. C., Rodrigues, S. W. (2021). Field guide to RODS in the Pireneus
Syntaxis, central Brazil. In S. Mukherjee (Ed.), Structural Geology and Tectonics Field
Guidebook—Volume 1. Springer Nature Switzerland AG. Cham. pp. 221-264. ISBN:
978-3-030-60142-3.
Mukherjee, S. (2013). Higher Himalaya in the Bhagirathi section (NW Himalaya, India): Its
structures, backthrusts and extrusion mechanism by both channel flow and critical taper
mechanisms. International Journal of Earth Sciences, 102, 1851–1870.
Novakova, L., Novak, P. (2021). Spectacular sandstone rock cities in the Czech Republic. In S.
Mukherjee (Ed.), Structural Geology and Tectonics Field Guidebook—Volume 1. Springer
Nature Switzerland AG. Cham. pp. 189-220. ISBN: 978-3-030-60142-3.
xvi Introduction to Structural Geology and Tectonics Field Guidebook

Omid, M. W., Mukherjee, S., Dasgupta, S. (2021). Vein geometry (Bhuj, Gujarat, India). In S.
Mukherjee (Ed.), Structural Geology and Tectonics Field Guidebook—Volume 1. Springer
Nature Switzerland AG. Cham. pp. 707-714. ISBN: 978-3-030-60142-3.
Patil Pillai S., Kale, V. S. (2021). Traverses through the Bagalkot Group from north Karnataka
state, India: Deformation in the Mesoproterozoic supracrustal Kaladgi Basin. In S. Mukherjee
(Ed.), Structural Geology and Tectonics Field Guidebook—Volume 1. Springer Nature
Switzerland AG. Cham. pp. 325-366. ISBN: 978-3-030-60142-3.
Puniya, M. K., Mukherjee, S. (2021). Structural geology of the Nainital-Pangot road (Kilbari
section), Nainital Lesser Himalaya (Uttarakhand, India): Focus on back-structures. In S.
Mukherjee (Ed.), Structural Geology and Tectonics Field Guidebook—Volume 1. Springer
Nature Switzerland AG. Cham. pp. 429-436. ISBN: 978-3-030-60142-3.
Pundalik, A., Nikalje, S., Samant, A., Samant, H. (2021). Geological fieldguide: Malvan
(Maharashtra, India). In S. Mukherjee (Ed.), Structural Geology and Tectonics Field
Guidebook—Volume 1. Springer Nature Switzerland AG. Cham. pp. 495-528. ISBN:
978-3-030-60142-3.
Shallaly, N. A., Abu Sharib, A. S. A. A. (2021). Low baric metamorphic belts in the northern tip
of the Arabian Nubian Shield: selected examples from the Eastern Desert/Midyan terranes,
Egypt. In S. Mukherjee (Ed.), Structural Geology and Tectonics Field Guidebook—Volume 1.
Springer Nature Switzerland AG. Cham. pp. 265-304. ISBN: 978-3-030-60142-3.
Simonetti, M., Carosi, R., Montomoli, C. (2021). Strain softentening in a continental shear zone: A
field guide to the excursion in the Ferriere-Mollières shear zone (Argentera Massif, Western
Alps, Italy). In S. Mukherjee (Ed.), Structural Geology and Tectonics Field Guidebook—
Volume 1. Springer Nature Switzerland AG. Cham. pp. 19-48. ISBN: 978-3-030-60142-3.
Singh, P., Ao, A., Thakur, S. S., Rana, S., Sharma, R., Singh, A. K., Singhal, S. S. (2021).
Geology, structural, metamorphic and mineralization studies 2 along the
Mandi-Kullu-Manali-Rohtang section of Himachal 3 Pradesh, NW-India. In S. Mukherjee
(Ed.), Structural Geology and Tectonics Field Guidebook—Volume 1. Springer Nature
Switzerland AG. Cham. pp. 437-460. ISBN: 978-3-030-60142-3.
Swanger, W. R., Whitmeyer, S. J. Creating geologic maps in the 21st century: A case study from
Western Ireland. In S. Mukherjee (Ed.), Structural Geology and Tectonics Field Guidebook—
Volume 1. Springer Nature Switzerland AG. Cham. pp. 1-18. ISBN: 978-3-030-60142-3.
Tămaș, D. M., Tămaș, A., Jüstel, A. M., Passchier, M., Chudalla, N., Gotzen, J., Wagner, L. A. P.,
Tașcu-Stavre, T., Schléder, Z., Krézsek, C., Filipescu, S., Urai, J. L. (2021). A fieldguide to the
spectacular salt mines of the Transylvanian basin and Romanian Carpathians. In S. Mukherjee
(Ed.), Structural Geology and Tectonics Field Guidebook—Volume 1. Springer Nature
Switzerland AG. Cham. pp. 167-188. ISBN: 978-3-030-60142-3.
Contents

Creating Geologic Maps in the Twenty-First Century: A Case Study


from Western Ireland . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
W. R. Swanger and S. J. Whitmeyer
Strain Softening in a Continental Shear Zone: A Field Guide
to the Excursion in the Ferriere-Mollières Shear Zone
(Argentera Massif, Western Alps, Italy) . . . . . . . . . . . . . . . . . . . . . . . . . 19
M. Simonetti, R. Carosi, and C. Montomoli
The Geometry and Kinematics of the Southwestern Termination
of the Pyrenees: A Field Guide to the Santo Domingo Anticline . . . . . . 49
E. L. Pueyo, B. Oliva-Urcia, E. M. Sánchez-Moreno, C. Arenas,
R. Silva-Casal, P. Calvín, P. Santolaria, C. García-Lasanta, C. Oliván,
A. Gil-Imaz, F. Compaired, A. M. Casas, and A. Pocoví
Miocene-Quaternary Strain Partitioning and Relief Segmentation
Along the Arcuate Betic Fold-and-Thrust Belt: A Field Trip Along
the Western Gibraltar Arc Northern Branch (Southern Spain) . . . . . . . 103
Alejandro Jiménez-Bonilla, Manuel Díaz-Azpiroz, Inmaculada Expósito,
and Juan Carlos Balanyá
The Southern Iberian Shear Zone (SW Spain): Inclined Transpression
Related to Variscan Oblique Convergence in a HT/LP
Metamorphic Belt . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
Manuel Díaz-Azpiroz and Carlos Fernández
A Field Guide to the Spectacular Salt Mines of the Transylvanian
Basin and Romanian Carpathians . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
Dan Mircea Tămaș, Alexandra Tămaș, Alexander Magnus Jüstel,
Martijn Passchier, Nils Chudalla, Lina Gotzen,
Luis Alberto Pizano Wagner, Teodora Tașcu-Stavre, Zsolt Schléder,
Csaba Krézsek, Sorin Filipescu, and Janos L. Urai

xvii
xviii Contents

Spectacular Sandstone Rock Cities in the Czech Republic . . . . . . . . . . . 189


Lucie Novakova and Petr Novak
Field Guide to RODS in the Pireneus Syntaxis, Central Brazil . . . . . . . 221
Marco Antonio Caçador Martins-Ferreira
and Sérgio Wilians de Oliveira Rodrigues
Low Baric Metamorphic Belts in the Northern Tip
of the Arabian–Nubian Shield: Selected Examples
from the Eastern Desert/Midyan Terranes, Egypt . . . . . . . . . . . . . . . . . 265
N. A. Shallaly and A. S. A. A. Abu Sharib
Review of the Geometric Model Parameters of the Main
Himalayan Thrust . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 305
Kutubuddin Ansari
Traverses Through the Bagalkot Group from North Karnataka
State, India: Deformation in the Mesoproterozoic Supracrustal
Kaladgi Basin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 325
Shilpa Patil Pillai and Vivek S. Kale
Tectonic Framework of Northern Pakistan from Himalaya
to Karakoram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 367
Asghar Ali, Sajjad Ahmad, Sajjad Ahmad, Mohammad AsifKhan,
Muhammad Irfan Khan, and Gohar Rehman
Structures of Lesser/Greater Himalaya in and Around
an Out-of-Sequence Thrust in the Chaura-Sarahan Area
(Himachal Pradesh, India) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 413
Rajkumar Ghosh and Soumyajit Mukherjee
Structural Geology Along the Nainital–Pangot Road (Kilbari Section),
Nainital Lesser Himalaya (Uttarakhand, India): Focus
on Back-Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 429
Mohit Kumar Puniya and Soumyajit Mukherjee
Geology, Structural, Metamorphic and Mineralization Studies Along
the Mandi-Kullu-Manali-Rohtang Section of Himachal Pradesh,
NW-India . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 437
Paramjeet Singh, Aliba Ao, S. S. Thakur, Shruti Rana, Rajesh Sharma,
A. K. Singh, and Saurabh Singhal
Tectonics and Channel Morpho-Hydrology—A Quantitative
Discussion Based on Secondary Data and Field Investigation . . . . . . . . 461
Mery Biswas, Ankita Paul, and Mostafa Jamal
Geological Field Guide: Malvan (Maharashtra, India) . . . . . . . . . . . . . . 495
Ashwin Pundalik, Shiba Nikalje, Arnav Samant, and Hrishikesh Samant
Contents xix

A Field Guide to the Champaner Region, Southern Aravalli Mountain


Belt (SAMB), Gujarat, Western India . . . . . . . . . . . . . . . . . . . . . . . . . . 529
Aditya U. Joshi and Manoj A. Limaye
Importance of Fracturing in Uranium Mineralization in Gulcheru
Quartzite Host: A Case from Ambakapalle Area, Cuddapah Basin,
Andhra Pradesh, India . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 577
Sukanta Goswami, P. K. Upadhyay, and V. Natarajan
Granitic Rocks Underlying Deccan Trap Along the Margin of East
Dharwar Craton, Mutnyal (Maharashtra)—Bhaisa (Telangana),
India—General Description and Deformation . . . . . . . . . . . . . . . . . . . . 599
R. D. Kaplay, Md. Babar, Soumyajit Mukherjee, Deepak Wable,
and Kunal Pisal
Structural Analyses of the Lunavada–Santrampur Area
(Gujarat, India) Using Remote Sensing Images . . . . . . . . . . . . . . . . . . . 621
Geetika H. Chauhan, G. S. Rao, and Soumyajit Mukherjee
Fundamentals of Lithostructural Mapping: Example from the SW
Part of the Proterozoic Bhima Basin, Karnataka, India:
A Note on Dharwarian Crustal Evolution . . . . . . . . . . . . . . . . . . . . . . . 639
Sukanta Goswami, Shivam Shrivastava, Suman Das,
and Purnajit Bhattacharjee
A 3D Photogrammetric Approach in Mapping Meso-Scale Folds
and Shears in Structurally Controlled Syngenetic Mn-Mineralised
Zones of Shivrajpur Region, Eastern Gujarat, India . . . . . . . . . . . . . . . 685
Aditya U. Joshi
Vein Geometry Around Bhuj (Gujarat, India) . . . . . . . . . . . . . . . . . . . . 707
Mohammad Walid Omid, Soumyajit Mukherjee, and Sudipta Dasgupta
Oriented Rock Samples for Detailed Structural Analysis . . . . . . . . . . . . 715
Krzysztof Gaidzik and Jerzy Żaba
Creating Geologic Maps
in the Twenty-First Century: A Case
Study from Western Ireland

W. R. Swanger and S. J. Whitmeyer

Abstract Techniques of geologic mapping and geologic map creation have changed
significantly from traditional paper-based methods. Geologic mapping and data
collection in the field is now primarily facilitated by mobile devices and dedicated
geologic mapping software. Geologic map production has become a fully integrated
process, importing digital data from the field and making use of cartographic soft-
ware, such as ArcGIS and Adobe Illustrator, to create interactive geologic map prod-
ucts. Dissemination of geologic maps incorporates several types of map products
to support the variety of uses that practitioners have for geologic maps and field
data. Some of these map products include: 1. layered PDF maps, where layers can
be toggled to show different map components; 2. Google Earth KML and KMZ
files that can be viewed in the virtual 3D terrain of the geobrowser; and 3. GIS
geodatabases that include not only the geologic map interpretation of a field area but
also the primary field data. Geologic field data should also be archived in commu-
nity databases, such as StraboSpot.org, so that future field workers can access and
validate the data in their projects. This modern approach to creating geologic maps
is highlighted in a case study from the lakes region of western Ireland, where under-
graduate geoscience students have used digital mapping techniques in field exercises
for several years. A brief discussion of the history of digital field mapping and map
creation sets the stage for a discussion of modern techniques. Current best prac-
tices are highlighted for field mapping and data collection, geologic map creation,
dissemination of map-related products, and archiving of data and products.

Electronic supplementary material The online version of this chapter


(https://doi.org/10.1007/978-3-030-60143-0_1) contains supplementary material, which is
available to authorized users.

W. R. Swanger · S. J. Whitmeyer (B)


Geology and Environmental Science, James Madison University, Harrisonburg, USA
e-mail: whitmesj@jmu.edu
W. R. Swanger
Department of Earth and Environmental Sciences, University of Kentucky, Lexington, USA
S. J. Whitmeyer
Division of Earth Sciences, National Science Foundation, Alexandria, USA

© Springer Nature Switzerland AG 2021 1


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_1
2 W. R. Swanger and S. J. Whitmeyer

1 Introduction

Field data collection and the creation of geologic maps have been the cornerstone
of interpreting continental geology since the birth of the geosciences in the early
nineteenth century (e.g., Smith 1815; Cuvier and Brongniart 1822). Even with recent
dramatic advances in remote sensing (Dasgupta and Mukherjee 2017, 2019) and
laboratory-based analytical techniques, geologic maps remain one of the primary
vehicles for the interpretation and illustration of geologic and tectonic features and
processes. However, the advent of digital publication and presentation methods, as
well as mobile technologies for data collection, have resulted in modern approaches
to the creation of geologic maps that substantially differ from traditional paper-based
methods. Traditional field data collection with hardback field books and on paper
topographic maps in protective map boards was the norm throughout most of the
twentieth century, even as the production of final map products was switching to
computer-based cartography tools, such as ArcGIS and Adobe Illustrator.
Early practitioners explored the benefits of digital equipment for field mapping
early in the twenty-first century, especially after selective availability (civilian
degrading) of the Global Positioning System (GPS) was turned off in 2000. This
resulted in a proliferation of handheld GPS devices that enabled geolocation for cars
or for individuals in the field, rendering manual positioning methods, such as triangu-
lation, obsolete. Shortly thereafter, the advent of ruggedized, all-weather computer
hardware, such as Trimble handheld PCs, Panasonic Toughbooks, and other tablet
PCs (Knoop and van der Pluijm 2006), many with built-in GPS receivers, facili-
tated the digital revolution in field mapping and data collection (Pavlis et al. 2010;
Whitmeyer et al. 2010). The next big advance came with the initial release of the
iPhone in 2007, quickly followed by other enhanced cellular communication devices
(Novakova and Pavlis 2017). The development of mobile field mapping software,
coupled with the release of the iPad tablet in 2010, has ushered in the modern era
of mobile tablet-based field mapping, which has made paper-based field mapping
obsolete for today’s geoscience professionals.
A coincident evolution of software solutions accompanied the changes in field
mapping hardware. Mobile field-mapping software is initially centered on commer-
cial GIS platforms, such as ArcPad for handheld PCs and then iGIS for iPads. More
recently, geology-focused apps, such as StraboSpot, have become accepted as the
standard for both field data collection and for the archiving of field-based data in an
open, community database (Walker et al. 2019). Collectively, this steady evolution
of advances in equipment has resulted in a new digital workflow that encompasses
initial collection of geologic data in the field through the production of professional
geologic maps and other data products.
In the sections that follow, we present a modern approach to creating geologic
maps, from data collection through dissemination and archiving, recognizing that
this is still a fast-moving field and that tomorrow’s best practices will certainly differ
to some extent from what is presented here.
Creating Geologic Maps in the Twenty-First Century … 3

2 A Field Area in Western Ireland as an Example Case


Study

As an illustration of a workflow for modern field data collection and geologic map
production, we focus on a field area in western Ireland that has been the site of
digital mapping exercises for the James Madison University Geology Field Course
in Ireland for several years (De Paor and Whitmeyer 2009; Whitmeyer et al. 2019).
The field area incorporates the mountains of Bencorragh and Knock Kilbride, along
the Kilbride Peninsula, located in the lakes region north of Galway in western Ireland.
This area is along the southern margin of the South Mayo Trough, just north of the
suture with the Connemara terrane—a Dalradian terrane that was sutured to the South
Mayo Trough during the mid-Paleozoic Caledonian orogeny (Williams and Harper
1991; Williams and Rice 1989; Dewey and Ryan 2016; Fig. 1).
In the Kilbride Peninsula region the oldest lithologic units are at higher elevations
of the mountains of Bencorragh and Knock Kilbride and are predominantly composed
of Ordovician basalt pillows and flows (Williams 1990; Draut and Clift 2001) of the
Lough Nafooey Fm. (brown color, Fig. 2). On the southern and eastern shores of
Lough Nafooey (labeled on Fig. 2), north of Bencorragh, the basalts are overlain by
northwest-dipping Ordovician conglomerates, arkoses, and sandstones of the Rosroe
Fm. (gray color, Fig. 2). The southern slopes of the mountains consist of a mostly
planar, moderately to steeply southeast-dipping, Silurian transgressive sequence of
volcanic rocks (Basal Lough Mask member, purple color), terrestrial red sandstones
(Lough Mask Fm., red color), near-shore quartz arenites (Kilbride Fm., blue color),

Fig. 1 Simplified geologic map of the boundary region between the Connemara Terrane and the
South Mayo Trough. The field-mapping area is indicated by the red ellipse. After McConnell et al.
(2009)
4

Fig. 2 Authors’ crowd-sourced geologic map of the Kilbride Peninsula and the mountains of Bencorragh and Knock Kilbride in the border regions of Counties
W. R. Swanger and S. J. Whitmeyer

Mayo and Galway, western Ireland. The map area is roughly the same as in Figs. 3 and 4; Explanation for geologic map shown in below the image
Creating Geologic Maps in the Twenty-First Century … 5

siltstones (Tonalee member, green color), and shelf/slope turbidites (Lettergesh Fm.,
yellow/orange color) that overlie Ordovician basalts (Graham et al. 1989; Chew
et al. 2007; Whitmeyer et al. 2010). The Kilbride Peninsula area is broadly folded
and cross-cutby Late Devonian (Mohr 2003; Johnson et al. 2011), steeply-dipping
normal and transverse faults, with offsets that range from decimeters to decameters
(Fig. 2).
In the sections below, we make use of this field area to highlight and discuss the
three principal steps to creating geologic maps: 1. Collection of geologic data in
the field, 2. Creation of a geologic map and related products from field data, and
3. Dissemination of the new geologic map and related products to the community.
Below, we cover each of these steps in detail, starting with historical contexts and
then highlighting modern methods.

3 Modern Methods of Digital Field Data Collection

Published geologic maps that incorporate the Kilbride Peninsula field area include
the original geological map of Ireland by Griffith (1838; Fig. 3), as well as more
recent regional geologic maps, such as the South Mayo geologic map by Graham
et al. (1989) and the 1:100,000 Galway Bay map by the Geological Survey of Ireland
(Pracht et al. 2004; Fig. 4). These maps reflect a progression in the tectonic interpre-
tations of the bedrock geology of the region, but do not differ significantly in their
depictions of the areal extent of lithologic units or their offsets along faults. Impor-
tantly, none of these existing maps are detailed enough to highlight the density of
individual outcrops that occur in the field area or the abundance of faults that can be
inferred from offsets in bedding contacts given a high density of outcrop data (Whit-
meyer et al. 2019; Fig. 2). We suggest that new approaches to crowd-sourcing data
collection in the field, facilitated by modern mobile equipment, can enable better
geologic map interpretations through a significantly increased density of outcrop
data.

Fig. 3 Detail from the first geologic map of Ireland by Griffith (1838). Map is cropped to show the
same field area as Fig. 2
6 W. R. Swanger and S. J. Whitmeyer

Fig. 4 Portion of the 1:100,000 scale bedrock geology map of Galway Bay (Pracht et al. 2004).
Map is cropped to show the same extent as seen in Figs. 2 and 3

Geology-focused field investigations have traditionally been conducted by a


solitary geologist, sometimes with a field assistant. As a result, mapping strate-
gies prioritized efficient coverage of a field area with the understanding that not
all outcrops could be visited in the time available for fieldwork. Recently, crowd-
sourcing approaches to fieldwork have been advocated, where multiple field geolo-
gists collaborate on mapping a field area (House et al. 2013), or teams of (sometimes
novice) geologists map the same field area (Whitmeyer and De Paor 2014; Whitmeyer
et al. 2019). These approaches can produce a significantly greater density of field
data, which in turn, can lead to a more complete geologic map and better geologic
interpretations. For the Kilbride Peninsula field area, field data was collected by
multiple teams of James Madison University (JMU) field course students in digital
mapping exercises over several years. The inherent redundancy of several teams
mapping each section of a field area resulted in differing interpretations for some
field locations (e.g., Whitmeyer and De Paor 2014), which necessitated an expert
control dataset for validation (see the Discussion section below).
Advances in digital methods of geologic mapping and data collection in the
field help facilitate crowd-sourcing methods when field datasets are exportable in
a common format. For the Kilbride Peninsula field area, a variety of software solu-
tions were used during approximately 15 years of student mapping projects. Software
for field data collection included ArcPad and ArcMap in the early years (pre iPad),
followed by mobile apps for the iPad, such as iGIS, FieldMove (Midland Valley
2017), and most recently StraboSpot (Walker et al. 2019). Older field-mapping soft-
ware required users to upload georeferenced basemaps (topographic maps and/or
aerial photographs) and a predefined data structure before field data collection
Creating Geologic Maps in the Twenty-First Century … 7

could occur. Modern mobile apps (e.g. FieldMove, StraboSpot) automatically access
basemap imagery (generic, publicly available road maps, shaded topographic maps,
or satellite imagery), which can be cached on the mobile device for offline use.
However, users can still upload their own specialized datasets, such as high-resolution
LiDAR topography or digital elevation models (DEMs). A standard data structure
for geologic mapping and data collection is also built into the apps (Fig. 5), although
users can often customize these data structures to some extent. Notably, all of these
field-mapping software solutions allowed for the export of field data in standard GIS
formats (shapefiles). This enabled us to combine field datasets into a master database
of all field data, while retaining identifiers for novice (student) versus expert (control)
datasets.

Fig. 5 StraboSpot app on an iPad (protectively enclosed in an OtterBox case), showing the data
entry menu for a planar outcrop feature
8 W. R. Swanger and S. J. Whitmeyer

Recent initiatives in the geoscience community, as well as funding agencies like


the National Science Foundation, have championed the storage and archiving of data
in curated, often cloud-hosted, community databases. The StraboSpot data system
(strabospot.org; Walker et al. 2019) was developed in response to a community work-
shop that conceptualized a data storage system for geoscience field data (Mookerjee
et al. 2015). A mobile StraboSpot app is now available for iOS and Android devices
and functions as a field-mapping and data collection system that interfaces with the
StraboSpot.org community data system. Users can choose to keep their StraboSpot
data private or make their data public to share it with the broader community (Fig. 6).
This allows geoscientists to privately store field data during ongoing projects, prior
to releasing it to the public upon completion of the project. Student mapping projects
can be kept private and only made available to the local class community or research
group.
A significant advantage to ultimately making field data publicly available is that
published geologic maps can reference the supporting data and allow other users
to investigate geologic interpretations based on the original data (e.g., the geologic
map of Fig. 2 and the publicly available supporting data shown in Fig. 6). The ability
to provide original data as supplemental material to geologic maps allows future
geologists easy access to resources such as field notes, photographs and other data
that previously could not be included in delivered products.

Fig. 6 Screen capture of a map-viewing search page from StraboSpot.org that shows a publicly
available dataset. The expert control dataset for the Bencorragh field area (left part of Fig. 2) is
displayed. Red box shows the area of Fig. 7
Creating Geologic Maps in the Twenty-First Century … 9

4 Building a Modern Digital Geologic Map

Geologic mapping in the field is now typically undertaken with mapping apps specif-
ically designed for geologists, such as FieldMove (Midland Valley 2017) or Stra-
boSpot (Walker et al. 2019). General GIS apps, like iGIS or Collector for ArcGIS,
can also be used but take a bit more work to set up for field mapping, as the apps are not
specifically targeted at geoscientists. Other apps designed for specialized geoscience
data collection in the field include StratMobile for collecting stratigraphic data and
preparing stratigraphic columns and strip logs, and Stereonet Mobile, which builds
stereonets from orientation measurements (Allmendinger et al. 2017). All of these
apps work on iPads (some of them also work on Android tablets), which we enclose
in protective OtterBox Defender cases—necessary for field work in the frequently
wet weather of western Ireland. Some of these apps can be used on mobile phones,
including a simplified version of FieldMove called FieldMoveClino, but our experi-
ence is that the larger screen area available on mobile tablets is important for seeing
a sufficiently large expanse of the digital field area and working field map.
As almost all modern mobile devices incorporate internal sensors, such as an
accelerometer, a gyroscope, and a magnetometer (check out the Sensor Kinetics app
to see these sensors in action), several geoscience apps for the field also include a
digital geologic compass for taking orientation measurements (e.g., strike and dip
of bedding). There has been discussion in the field-mapping community regarding
whether measurements obtained from these digital compasses are accurate and
precise enough for professional field work. However, recent work by Novakova and
Pavlis (2017) and Whitmeyer et al. (2019) suggest that digital compass measure-
ments can produce measurements that are at the same level of precision as analog
geologic compasses (e.g., Brunton Pocket Transits), with an accuracy that is better
than ± 5 degrees for both azimuth (strike) and vertical (dip) angles (Whitmeyer
et al. 2019). The important caveat is that digital compasses have a sensor calibration
routine (waving the mobile device in a figure-eight pattern) that should be invoked
periodically to keep the compasses measuring surfaces correctly. Finally, in some
field situations where two-handed manipulation of an analog compass can be chal-
lenging, the ease of use of a digital compass may be preferable and more effective
(Whitmeyer et al. in review 2020).
A significant advantage of mobile-mapping devices is the capability to take a large
amount of background information into the field. This can be in the form of high-
resolution aerial or satellite imagery or a high-resolution LiDAR or photogrammetry-
based topographic DEM. The use of high-resolution shaded relief maps as basemaps
for field mapping has been demonstrated to significantly enhance the interpretation of
geologic structures (e.g., House et al. 2013). If publicly available, datasets previously
collected in a field area can be preloaded onto a mobile device for reference or
validation while in the field. We used this approach when we validated the crowd-
sourced student field data for the Kilbride Peninsula. We uploaded the student dataset
on our iPads and then traversed the field area to check the field data and sometimes
modify the students’ interpretations. In addition, maps, publications, or references
10 W. R. Swanger and S. J. Whitmeyer

that cover a region of interest can be preloaded on a mobile device as distinct files
(e.g., PDFs) or within an app such as Flyover Country, which is designed to provide
geolocated background information for a field area or field trip (Birlenbach et al.
2019).
Much of the initial drafting of a geologic map can be done on location in the field,
in the form of outcrop point data and linework (Fig. 7a), such as known contacts
and faults that can be walked out or sketched. Geologic maps drafted in the field
which include polygons and annotations (Fig. 7b) can be a very useful teaching
tool, providing a way to quickly visualize geologically complex areas. However,
the drafting of more permanent linework and polygons for the aerial expanses
of lithologic units is usually best left for a cartographic program on a laptop or
desktop computer. In general, advanced cartographic and design software, such as
ArcGIS and Adobe Illustrator, has a suite of tools that enable professional drafting
of geologic maps, and the larger screens of desktop computers make it easier to see
the broad expanse of field data and thus make more informed geologic interpreta-
tions. Modern pen-enabled touch screens, such as those made by Wacom, facilitate
easy and precise drafting of lines and polygons. Cartographic programs also provide
relative straightforward mechanisms for added auxiliary geologic map components,

Fig. 7 StraboSpot app on an iPad showing part of the Bencorragh field map a, with dashed lines
for approximate locations of contacts (black) and faults (red), and outcrop locations indicated by
black strike and dip symbols; an individual outcrop data point is highlighted in the pop-up balloon.
b shows same field map with polygons colored by lithology (pink = Lough Nafooey Fm., red =
Lough Mask Fm., blue = Kilbride Fm., green = Tonalee member)
Creating Geologic Maps in the Twenty-First Century … 11

such as map scales, north arrows, map explanations with lithologic unit descriptions
and symbology, and other map production elements (Fig. 8; see the Supplementary
File Layered_Geologic_Map.pdf for an example).

5 Geologic Map Products and Dissemination

Though today’s geologic maps include many of the same components as William
Smith’s (1815) hand-drawn and colored geologic map of England and Wales (e.g.,
surface extents of lithologic units that are represented by colored regions, vertical
interpretations of geologic structure, and orientation shown in a cross section, an
explanation that shows the color coding of the lithologic units), the modern mediums
for delivering geologic maps have come a long way from Smith’s rolled paper stock.
Paper geologic maps can still be purchased from many U.S. state geologic surveys,
but most state surveys have made their geologic and geographic map data available
via a GIS-based Web interface (e.g., the Interactive Geologic Map of Virginia;
https://www.dmme.virginia.gov/webmaps/DGMR/). Similarly, the U.S. Geological
Survey maintains the National Geologic Map Database (NGMDB; https://www.
usgs.gov/core-science-systems/national-cooperative-geologic-mapping-program/
science/national-geologic-map?qt-science_center_objects=0#qt-science_center_
objects), where an interactive Web-based GIS portal provides national- and state-
level geologic maps and information, much of which is available for download in
a digital format. Country-level international geologic maps and information can be
accessed at the OneGeology portal (http://www.onegeology.org/).
Historical maps are sometimes included on these Web portals as georectified
(geolocated) raster images, but most modern geologic maps are built in a GIS and
stored as digital data and thus are easily incorporated into GIS-based Web portals.
Most of these portals can serve the digital map data in a variety of formats to satisfy
a diverse array of end user requirements. If a user desires the raw field data and/or
interpretations of the mapping geologists, the geologic map can be made available
as a geodatabase (see the Supplementary File Bencorragh_geodatabase.gdb for an
example). Alternatively, if a user desires a professional finished map product to
either view on screen or print out, they can download a PDF file of the map. Modern
PDF files are layered and somewhat interactive, where users can turn on a selection
of available layers to display or print the desired map information (Fig. 8; see the
Supplementary File Layered_Geologic_Map.pdf as an example).
Other interactive mediums for digital display of geologic maps and information
include virtual globes or geobrowsers, such as Google Earth. Virtual globes have
the advantages of being rotatable and zoomable to any point on the globe. Most
areas in Google Earth display fairly high-resolution satellite or aerial imagery over
a moderate resolution (10-30 m) DEM, although some areas on the globe are better
resolved. Geologic maps can be draped over the Google Earth terrain to view the
maps in virtual 3D. Other location-specific information, like outcrop data and/or
outcrop photographs can also be linked to the geologic map and displayed in Google
12

Fig. 8 Authors’ multi-layered PDF file of the publishable geologic map of the Kilbride Peninsula and Bencorragh, western Ireland. Modern PDF maps can
have selectable layers to display different features of a geologic map, such as topographic and high-resolution satellite imagery basemap options. Control data
W. R. Swanger and S. J. Whitmeyer

and crowd-sourced data can also be viewed separately to highlight how increased data density enables a more detailed and accurate structural interpretation
Creating Geologic Maps in the Twenty-First Century … 13

Fig. 9 Oblique view of the crowd-sourced geologic map of Fig. 2 draped over the virtual 3D terrain
in google earth, with a pop-up balloon showing an outcrop photograph

Earth (Fig. 9). The maps and other associated data require yet another data format
for display on virtual globes, specifically a scripting language called KML (Keyhole
Markup Language; https://developers.google.com/kml). Many Web GIS servers can
export geologic maps and data in KML or KMZ (zipped packages of KML script
and associated imagery) format, although these exported files are typically less user
friendly than a KMZ map and data package that is custom developed by an expe-
rienced KML programmer (see the Supplementary File Ireland_Map.kmz for an
example).

6 Discussion

In this manuscript, we advocate for an all-digital workflow for the creation of geologic
maps, from field data collection through dissemination and archiving. However, we
recognize that there is a learning curve associated with the adoption of these methods
for both novice and expert geoscientists that are less familiar with the various aspects
of using digital technologies for geologic mapping. Indeed, our experiences with
many years of teaching digital methods of field mapping to students in undergrad-
uate field courses suggests that the use of mobile equipment in the field adds an
additional cognitive load for students that may still be grappling with the basics
of learning how to do geology in the field (De Paor and Whitmeyer 2009; Mogk
and Goodwin 2012). The JMU Geology Field Course in Ireland still starts students
14 W. R. Swanger and S. J. Whitmeyer

with analog compasses and paper field slips (topographic basemaps) for their initial
geologic mapping exercises, in order for the students to focus on basic lithologic
identification, measuring orientations, and the mechanics of making a geologic map
and interpreting the geology of a field area. Modern digital methods and equipment
are not introduced until the third or fourth mapping exercise, after the students know
how to strategize field data collection, work with multiple working hypotheses in
the field (e.g., Chamberlin 1890), and synthesize a geologic interpretation from their
field data. Even after introducing digital methods, some students still preferred to use
traditional equipment for their later culminating mapping projects (Whitmeyer and
De Paor 2014). The integration of digital methods, specifically the ability to drape
geologic maps and highlight field data in Google Earth, is crucial to the learning
process of the novice geologist. When students can rotate to oblique views of data
draped over a DEM, they can more easily grasp the vertical component of structurally
complex areas.
Throughout the several years of digital mapping exercises in the Kilbride Penin-
sula area, student field data was collected in a master GIS file of all outcrop data
points. This resulted in a dataset with over 15,000 points, some of which were redun-
dant that yielded a density of field data unprecedented for any previous geologic map
of the area. This density of field data facilitated the creation of a geologic map with a
greater resolution and better characterization of fault offsets (Fig. 2) than previously
published maps (Figs. 3 and 4) and thus arguably a more accurate depiction of the
geology of the region. However, the abundance of outcrop data from students that
were relative novices at geologic mapping in the field also highlighted local areas
where students had differing interpretations of the geology—usually highlighted as
contrasting interpretations of the lithology (Whitmeyer and De Paor 2014). We have
previously argued that the “good” data significantly dwarfs the “bad” data, but how
does one accurately characterize the “good” data?
In areas where there was a single clear outlier in a lithologic interpretation, it was
fairly easy to cull the spurious data. However, in other locations with unclear interpre-
tations, often in areas where there were gradational changes between lithologic units,
a control dataset was used to decide on the most parsimonious interpretations. For
the Bencorragh and Kilbride Peninsula field areas, we had a dataset collected by an
expert in the geology of that region (one of the field course instructors) that covered
much of the map area (e.g., Fig. 6). We used the expert dataset to constrain most
of the conflicts in our crowd-sourced dataset. However, we still found it necessary
to return to the field area with a preliminary version of the cross-sourced geologic
map loaded on our iPads in order to field check a few problematic locations. Though
not always possible in remote field areas, returning to the field with a draft geologic
map in order to constrain problematic areas or areas with little field data enables a
geologist to cover a large area in a short period of time and can greatly improve the
accuracy of the final map product.
Creating Geologic Maps in the Twenty-First Century … 15

7 Conclusion and Best Practices

We have discussed modern methods of field mapping and the creation of geologic
maps, while providing some historical context of past practices. While we understand
that technological advances to hardware and software relevant to geoscience field
work are always changing rapidly, we present below a bulleted list of some current
best practices that we have identified for efficient creation of geologic maps, ca.
2020:
Field mapping and data collection
• We suggest using the StraboSpot app for field mapping and data collection, along
with its community database storage system.
• Explore legacy geologic maps and legacy field datasets (if available) for the field
area and upload them to the mobile device as background data.
• Assemble basemaps, LiDAR images, and shaded relief maps and upload to the
mobile mapping device as desired.
• We recommend exploring crowd-sourcing methods for field data collection, if
enough personnel are available.

Geologic map creation, dissemination, and archiving


• If crowd-sourcing methods are used, then control/validation datasets by mapping
experts may be needed to constrain areas with problematic interpretations.
• High-resolution background imagery can also help constrain interpretations on
geologic maps; shaded relief maps from LiDAR and/or photogrammetry data have
proven effective.
• Produce a variety of geologic maps and map products for today’s disparate audi-
ences, including layered PDF files of maps and map data, GIS geodatabases, and
Google Earth KMZ files.
• Raw field data should be made available on publicly accessible online databases,
such as StraboSpot.org. Keep in mind that the data can be embargoed until the
project is completed.
• Finished map products can be archived on community map servers, such as those
hosted by state geologic surveys, OneGeology.org, and the National Geologic
Map Database: ngmdb.usgs.gov.

Acknowledgements The authors would like to acknowledge JMU students that have participated
in digital mapping projects over the years, as well as field course faculty: Declan De Paor, Martin
Feely, John Haynes, Steve Leslie, Beth McMillan, Eric Pyle, and Shelley Whitmeyer. This work was
supported, in part, by NSF awards 1323468 and 1841132. Any opinion, findings, and conclusions
or recommendations expressed in this material are those of the authors and do not necessarily
reflect the views of the National Science Foundation, and review comments provided by Soumyajit
Mukherjee (IIT Bombay). Thanks to Marion Schneider, Annett Buettener, Boopalan Renu, Alexis
Vizcaino, Doerthe Mennecke-Buehler, and the proofreading team (Springer). Dutta and Mukherjee
(2021) encapsulate this work.
16 W. R. Swanger and S. J. Whitmeyer

References

Allmendinger, R. W., Siron, C. R., & Scott, C. P. (2017). Structural data collection with mobile
devices: Accuracy, redundancy, and best practices. Journal of Structural Geology, 102, 98–112.
Birlenbach, D. M., Shinneman, A. C., Loeffler, S., & Myrbo, A. (2019). Flyover country: Creating
flexible field experiences using a mobile geoscience app. In The Trenches, 9(3).
Chamberlin, T. C. (1890). The method of multiple working hypotheses. Science, 15(366), 92–96.
Chew, D. M., Graham, J. R., & Whithouse, M. J. (2007). U-Pb zircon geochronology of plagiogran-
ites from the lough nafooey (= Midland Valley) arc in western Ireland: constraints on the onset
of the Grampian orogeny. Journal of the Geological Society of London, 164, 747–750. https://
doi.org/10.1144/0016-76492007-025
Cuvier, G., & Brongniart, A. (1822). Description géologique des environs de Paris (p. 428). Paris:
Chez Dufour et D’Ocagne.
De Paor, D. G., & Whitmeyer, S. J. (2009). Innovations and redundancies in geoscience field courses:
Past experiences and proposals for the future. In S. J. Whitmeyer, D. Mogk, & E. J. Pyle (Eds.),
Field geology education: Historical perspectives and modern approaches, GSA Special Paper
(Vol. 461, pp. 45–56). https://doi.org/10.1130/2009.2461(05)
Dutta, D., & Mukherjee, S. (2021). Introduction to Structural Geology and Tectonics Field
Guidebook-Volume 1. In S. Mukherjee (Ed.), Structural Geology and Tectonics Field Guide-
book—Volume 1. Switzerland AG: Springer Nature. Cham. pp. xi-xvi. ISBN: 978-3-030-60142-3.
Dasgupta, S., & Mukherjee, S. (2017). Brittle shear tectonics in a narrow continental rift:
Asymmetric non-volcanic Barmer basin (Rajasthan, India). The Journal of Geology, 125,
561–591.
Dasgupta, S., Mukherjee, S. (2019). Remote sensing in lineament identification: Examples from
western India. In: A. Billi & A. Fagereng (Eds.), Problems and solutions in structural geology
and tectonics. Developments in structural geology and tectonics book series. Series Editor: S.
Mukherjee Elsevier (Vol. 5, pp. 205–221). ISSN: 2542–9000. ISBN: 9780128140482.
Dewey, J. F., & Ryan, P. D. (2016). Connemara: Its position and role in the grampian orogeny.
Canadian Journal of Earth Science, 53, 1246–1257. https://doi.org/10.1139/cjes-2015-0125.
Draut, A. E., & Clift, P. D. (2001). Geochemical evolution of arc magmatism during arc-continent
collision. South Mayo, Ireland, Geology, 29, 543–546.
Graham, J. R., Leake, B. E., & Ryan, P. D. (1989). The geology of south mayo (p. 75). Western
Ireland, Edinburgh: Scottish Academic Press.
Griffith, R. J. (1838). In Outline of the geology of Ireland, Dublin:Railway Commissioners.
House, P. K., Clark, R., & Kopera, J. (2013). Overcoming the momentum of anachronism: american
geologic mapping in a twenty-first-century world. In V. R. Baker (Ed.), Rethinking the fabric of
geology: Geological society of America special paper (Vol. 502, pp. 103–125). https://doi.org/
10.1130/2013.2502(05)
Johnson, E. A., Sutherland, S., Logan, A. V. L., Samson, S. D., & Feely, M. (2011). Emplacement
conditions of a porphyritic felsite dyke and timing of motion along the Coolin Fault at Ben Levy,
Co galway. Irish Journal of Earth Sciences, 29, 1–13.
Knoop, P. A., & van der Pluijm, B. (2006). GeoPad: Tablet PC-enabled field science education. In
D. Berque, J. Prey & R. Reed (Eds.), The impact of pen-based technology of education: Vignettes,
evalations, and future directions (pp. 103–114). West Lafayette, Indiana: Purdue University Press.
McConnell, B. M., Riggs, N., & Crowley, Q. G. (2009). Detrital zircon provenance and Ordovician
terrane amalgamation, western Ireland. Journal of the Geological Society, 166, 473–484. https://
doi.org/10.1144/0016-76492008-081.
Valley, Midland. (2017). Field move users guide (p. 63). Edinburgh: Petroleum Experts.
Mogk, D. W., & Goodwin, C. (2012). Learning in the field: Synthesis of research on thinking and
learning in the geosciences. In K. A. Kastens & C. A. Manduca (Eds.), Earth and mind II: A
synthesis of research on thinking and learning in the geosciences. Geological society of America
special paper (Vol. 486, pp. 131–163). https://doi.org/10.1130/2012.2486(24).
Creating Geologic Maps in the Twenty-First Century … 17

Mohr, P. (2003). Late magmatism of the Galway Granite Batholith: I Dacite dykes. Irish Journal
of Earth Sciences, 21, 71–104.
Mookerjee, M., Viera, D., Chan, M., Gil, Y., Goodwin, C., Shipley, T., et al. (2015). We need
to talk: Facilitating communication between field-based geoscience and cyber infrastructure
communities. GSA Today, 25, 34–35.
Novakova, L., & Pavlis, T. L. (2017). Assessment of the precision of smart phones and tablets for
measurement of planar orientations: A case study. Journal of Structural Geology, 97, 93–103.
Pavlis, T. L., Langford, R., Hurtado, J., & Serpa, L. (2010). Computer-based data acquisition and
visualization systems in field geology: Results from 12 years of experimentation and future
potential: Geosphere, 6, 275–294.
Pracht, M., Lees, A., Leake, B. E., Feely, M., Long, B. C., Morris, J., et al. (2004). Geology of
galway bay: A geological description to accompany the bedrock geology 1:100,000 scale map
series, sheet 14, Galway Bay (p. 76). Dublin: Geological Survey of Ireland.
Smith, W. (1815). A geological map of England and Wales and part of Scotland, London: British
Geological Survey, 1 sheet.
Walker, J. D., Tikoff, B., Newman, J., Clark, R., Ash, J., & Good, J., et al. (2019). StraboSpot data
system for structural geology. Geosphere, 15. https://doi.org/10.1130/GES02039.1
Whitmeyer, S. J., & De Paor, D. G. (2014). Crowdsourcing digital maps using citizen geologists.
EOS, 95, 397–399. https://doi.org/10.1002/2014EO440001.
Whitmeyer, S. J., Nicoletti, J., & De Paor, D. G. (2010). The digital revolution in geologic mapping.
GSA Today, 20(4/5). https://doi.org/10.1130/gsatg70a.1
Whitmeyer, S. J., Pyle, E. J., Pavlis, T. L., Swanger, W., & Roberts, L. (2019). Modern approaches
to field data collection and mapping: Digital methods, crowdsourcing, and the future of statistical
analyses. Journal of Structural Geology, 125, 29–40. https://doi.org/10.1016/j.jsg.2018.06.023
Whitmeyer, S. J., Atchison, C., & Collins, T. D. (2020). Using mobile technologies to enhance
accessibility and inclusion in field-based learning, GSA Today, v. 30. https://doi.org/10.1130/
GSATG462A.1
Williams, D. M. (1990). Evolution of Ordovician terranes in western Ireland and their possible
Scottish equivalents. Transactions of the Royal Society of Edinburgh–Earth Sciences, 81, 23–29.
Williams, D. M., & Harper, D. A. T. (1991). End-Silurian modifications of Ordovician terranes in
western Ireland. Journal of the Geological Society of London, 148, 165–171. https://doi.org/10.
1144/gsjgs.148.1.0165
Williams, D. M., & Rice, A. H. N. (1989). Low-angle extensional faulting and the emplacement
of the Connemara Dalradian. Ireland: Tectonics, 8, 417–428. https://doi.org/10.1029/tc008i002
p00417
Strain Softening in a Continental Shear
Zone: A Field Guide to the Excursion
in the Ferriere-Mollières Shear Zone
(Argentera Massif, Western Alps, Italy)

M. Simonetti, R. Carosi, and C. Montomoli

Abstract Fieldwork, integrated with several techniques for microstructural and


geochronological analyses, plays a fundamental part in the study of shear zones.
In this guide, we describe a two-day field trip in the Argentera External Crystalline
Massif (Western Alps). The Massif is cross-cut by the NW-SE oriented Ferriere-
Mollières Shear Zone, a spectacular example of a nearly 25-km-long regional trans-
pressive zone with a maximum thickness of 2 km and a complex and long-lasting
evolution during the Variscan time. Recent detailed geological mapping coupled with
structural, microstructural, and petrochronological studies reveal that the Ferriere-
Mollières Shear Zone was characterized by strain softening that localized strain in its
central part during the syn-shearing exhumation and retro-metamorphism. The aim
of the field trip is to visit some key outcrops that allow to recognize such evolution
and to observe the main features of the Ferriere-Mollières Shear Zone that can be
regarded as a good example of a strain softening shear zone in the continental crust.

1 Introduction

Shear zones are portions of rock in which strain is clearly higher than in the wall
rocks that separate less strained or unstrained portions of the lithosphere. They have
a primary influence on the geometry and evolution of orogenic belts and rifts, on
crustal rheology, magma ascent and emplacement, and fluid flow.
Assessing the timing of shear zone activity is crucial to reconstruct the tectono-
metamorphic evolution of the lithosphere. Shear zones can form at all scales and
at all structural levels and can record a complex and polyphase deformation history
(Ramsay 1980; Fossen and Cavalcante 2017 and references therein). Because of

M. Simonetti (B) · R. Carosi · C. Montomoli


Dipartimento Di Scienze Della Terra, Universita‘ Di Torino, Via Valperga Caluso 35, Turin, Italy
e-mail: matteo.simonetti@unito.it
C. Montomoli
IGG-CNR PISA, Via Moruzzi 1, Pisa, Italy

© Springer Nature Switzerland AG 2021 19


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_2
20 M. Simonetti et al.

this, they vary a lot in orientation, length, thickness, displacement, strain geometry,
coaxiality, and deformation mechanisms.
Geometry, orientation, and relative movement of the walls control the deforma-
tion within the zone. Several processes may happen temporally leading to different
behavior and evolution of shear zones. For example, compaction by pressure solution
and related loss of material in the zone or strain localization where margins are left
inactive may produce a thinning of the high-strain zone. Inclusion of portions of wall
rocks and interaction between adjacent shear zones may produce widening (Fossen
2016; Fossen and Cavalcante 2017).
The lengthening and thickening of ductile shear zones are not explored to the same
extent as fault growth. Shear zones lengthen generally because they tend to connect
with other shear zones to give rise to composite systems or networks (Peacock and
Sanderson 1991; Soliva and Benedicto 2004; Ponce et al. 2013; Mukherjee 2014;
Fossen and Rotevatn 2016). Shear zone thickness is an issue that is influenced mainly
by strain and rheology of the deformed rocks.
According to Fossen (2016), four models for the evolution of the thickness of
shear zones are possible. In type I shear zones, deformation in the central part slows
down as the zone thickens as strain propagates into the walls. In type II shear zones,
the deformation localizes in the central part. Thus, the margins become inactive and
the active part gets thinner. Type III shear zones develop a fixed thickness and the
entire zone keeps deforming without internal localization. Type IV shear zones have
the same behavior as type I, but the whole shear zone remains active throughout the
deformation history.
Shear zones might behave with or without growth in thickness and with or without
localization to its central parts due to strain hardening or strain softening processes.
Strain hardening and weakening are mainly controlled by changes in grain size,
presence of fluids, metamorphic reactions, strain accumulation, dynamic recrystal-
lization process, and efficiency of recovery process(Fossen and Cavalcante 2017 and
references therein).
The evolution of shear zones over time can be studied regionally where
deformation accumulates for several millions of years and could be resolved by
combining detailed field work with structural analysis, microstructural analysis, and
geochronometers. In the last few years, this approach has led to satisfactory results
such as, for example, in the Variscan Belt (northern and central Sardinia Sardinia,
Argentera Massif and Maures Massif; Di Vincenzo et al. 2004; Carosi et al. 2012;
Simonetti et al. 2018, 2020b; Montomoli et al. 2018), in the Ross Orogen (Antarctica;
Di Vincenzo et al. 2007), in the Himalaya (Nepal; Carosi et al. 2018; Cottle et al.
2015; Iaccarino et al. 2015; Montomoli et al. 2013, 2015), in the Northern Cordillera
(Canada; Parson et al. 2018), and in the Borborema Province (Brazil; Viegas et al.
2014) were the unrevealed structural evolution of crustal-scale shear zones provided
important constraints for new tectonic models.
Examples of regional shear zones, whose evolution is well-constrained and
described in detail, are of fundamental importance and are increasingly essential in
order to test theoretical models and to verify the applicability of the process-oriented
studies that, usually, are made on small-scale structures.
Strain Softening in a Continental Shear Zone: A Field Guide … 21

The basement of the Argentera Massif, in the Western Alps, shows elegant folded
and sheared Variscan migmatites only locally affected by Alpine tectonics. This
Massif is divided in two metamorphic complexes by a 25-km-long transpressive
shear zone known as the Ferriere-Mollière shear zone (FMSZ). The FMSZ was first
recognized by Faure-Muret (1955), who refers to it as the ’Valletta Unit,’ and first
mapped by Malaroda et al. (1970). The FMSZ has been recently studied in detail
by Carosi et al. (2016) and Simonetti et al. (2017, 2018) by integrating fieldwork
with multiple techniques for structural analysis and geochronology, that constrained
a type II evolution during Variscan time. The FMSZ is therefore a good example of
a regional zone in continental crust whose evolution, controlled by strain softening
phenomena, has clear field evidences that are supported by well-constrained struc-
tural and chronological data. The aim of the field guide is to visit some key outcrops
that allow to recognize such evolution and the features of both Variscan and Alpine
deformation in the Argentera Massif.

2 The Variscan Belt in the Mediterranean Area

The Variscan Belt is the result of a continent-continent collision between Laurentia-


Baltica and Gondwana between ~ 380–280 Ma (Arthaud and Matte 1977; Burg and
Matte 1978; Tollmann 1982; Matte 1986; Di Vincenzo et al. 2004). Major expo-
sures of the Variscan units occur in Central and Western Europe (Fig. 1), parts of

Fig. 1 Distribution of the Variscan units in Europe at the present day. S = Sardinia; C = Corsica;
MTM = Maures-Tanneron Massif; MC = Massif Central; IM = Iberian Massif; AM = Armorican
Massif; RH = Rheno-Hercynian; BM = Bohemian Massif. Red circle indicate the location of the
Argentera Massif
22 M. Simonetti et al.

Morocco and Algeria and north of the West African Craton. The portion of the belt
exposed in Central and Western Europe is known in literature as the Variscides. This
region is characterized by a composite orocline showing two large-scale arcs (Matte
and Ribeiro 1975; Matte 1986): a western branch known as the Ibero-Armorican
arc (Matte and Ribeiro 1975; Brun and Burg 1982; Dias and Ribeiro 1995; Dias
et al. 2016; Fernández-Lozano et al. 2016) and a smaller eastern branch (Matte
2001; Bellot 2005; Ballevre et al. 2018; Simonetti et al. 2018, 2020a, b). The south-
eastern sector of the Variscides reworked and fragmented during the Alpine orogeny.
Because of this, reconstruction of the southern European Variscan Belt and the corre-
lation between the different fragments in the Mediterranean area is debated (Stampfli
et al. 2002; Rosenbaum et al. 2002; Advokaat et al. 2014). Key areas to under-
stand the evolution of the southern European Variscan Belt are the Maures-Tanneron
Massif in southern France, the Corsica-Sardinia Block, and the Alpine External Crys-
talline Massifs (Fig. 1). Some authors proposed that, during the Variscan time, the
future Alpine External Crystalline Massifs were in lateral continuity with southern
France and Corsica-Sardinia Block, and all these sectors were affected by a trans-
pressive regional-scale shear zone known as East Variscan Shear Zone (Matte 2001;
Corsini and Rolland 2009; Carosi et al. 2012; Simonetti et al. 2018, 2020a, b).
Evidence of transpressional deformation is well-documented in northern Sardinia
(Carosi and Palmeri 2002; Iacopini et al. 2008; Frassi et al. 2009; Carosi et al. 2005,
2012; Cruciani et al. 2015) in the Maures Massif (Schneider et al. 2014; Simonetti
et al. 2020b) and in the Alpine External Crystalline Massifs (Carosi et al. 2016;
Simonetti et al. 2018, 2020a), in places were the Alpine metamorphic overprint is
very weak.

3 Geological Setting of the Argentera Massif

The External Crystalline Massifs represent the remnants of the Variscan Orogeny
(Fig. 1) occurred between 380-280 Ma as a result of the collision between Laurussia
and Gondwana (Matte 1986; Di Vincenzo et al. 2004). They are made up of
high-to-medium grade metamorphic basement intruded by the Permo-Carboniferous
granitoids.
The Argentera Massif is located at the boundary between Italy and France (Fig. 2a)
and is the southernmost of the Alpine External Crystalline Massifs. It is composed by
two metamorphic complexes made by high-grade migmatitic gneisses (Ferrando et al.
2008; Compagnoni et al. 2010), the southwestern Tinèe Complex and the northeastern
Gesso-Stura-Vesubiè Complex (GSV), which are divided by the Ferriere-Mollières
Shear Zone (FMSZ, Fig. 2a).
The GSV complex is mainly constituted by migmatitic gneiss derived from the
Late Ordovician granitoids and migmatitic paragneiss. Field relationships between
the two lithotypes suggest an original intrusion of ortho-derived migmatites in the
para-derived lithotypes (Compagnoni et al. 2010). Migmatites show paragenesis
indicative of amphibolite facies metamorphism (Compagnoni et al. 2010). Associated
Strain Softening in a Continental Shear Zone: A Field Guide … 23

Fig. 2 a Geotectonic map of the Argentera Massif; FMSZ: Ferriere-Mollieres shear zone; FCSZ:
Fremamorta-Colle del Sabbione shear zone; BF: Bersezio fault; VLS: Valle Stura Leucogranite;
ACG: Argentera Central Granite (Compagnoni et al. 2010); 1 = geological map of the Rifugio
Migliorero area in the Ischiator valley (Fig. 3); 2 = geological map of the Prati del Vallone area
in the Pontebernardo valley (Fig. 4); b Satellite view (from Google Earth) of the Ischiator and
Pontebernardo Valley. Circles indicate the location of the stops of the field trip (red = stops of day
1; green = stops of day 2)

to the ortho- and para-derived migmatites amphibolic migmatites of the Bousset-


Valmasque Complex (Rubatto et al. 2001), bodies of metavulcanites and mafic and
ultramafic boudins are present. Ferrando et al. (2008) and Compagnoni et al. (2010),
by the study of mafic lithotypes, recognize a metamorphic evolution with four stages:
24 M. Simonetti et al.

(1) HP metamorphic peak; (2) initial decompression stage; (3) amphibolite-facies


metamorphism of HT–MP; (4) amphibolite-facies metamorphism of LT–LP.
The GSV is intruded by a large granite body at 292 ± 10 Ma (Ferrara and Malaroda
1969), known as Argentera Central Granite.
The Tinée complex is divided into three formations (Faure-Muret 1955) deformed
under amphibolite-facies conditions: (1) Valerios-Fougieret Formation constituted
by biotite and plagioclase-bearing migmatitic paragneisses with graphite and silli-
manite; (2) Anelle-Valabres Formation constituted by plagioclase, biotite-bearing
migmatitic metagreywakes sometimes with muscovite; (3) Rabuons Formation
constituted by migmatitic metapelites, with K-feldspar, biotite, muscovite and silli-
manite. In Anelle-Valabres Formation and in Rabuons Formation eclogites relicts
have been recognized (Faure-Muret 1955; Malaroda et al. 1970).The Tinée complex
underwent a minor degree of melting than the GSV complex, testified by a minor
amount of leucosomes (Compagnoni et al. 2010).
The FMSZ is a ductile shear zone cross-cutting the Paleozoic basement (Malaroda
et al. 1970; Musumeci and Colombo 2002; Compagnoni et al. 2010). It strikes NW–
SE and extends from the village of Ferriere (Valle Stura) to the northwest, to the
village of Mollières to the southeast (Fig. 2a) showing a dextral movement. Mylonites
have been recently mapped by Carosi et al. (2016) according to the percentage of
matrix and porphyroclasts both at the meso—and at the microscopic-scale following
the classification proposed by Sibson (1977) and Passchier and Trouw (2005).
Mylonites of the FMSZ result from the shearing of the migmatites belonging
to both the GSV and the Tinée complexes. A deformation gradient marked by a
transition from protomylonite, mylonite, and ultramylonite has been detected (Carosi
et al. 2016).
The main lithotypes in the shear zone are medium-grained dark-green mylonitic
schists and biotite and white mica bearing mylonitic gneisses. Embedded within
the mylonitic schists amphibole-bearing gneiss, showing variable amounts of
deformation, are present (Carosi et al. 2016).
Different interpretations have been proposed in the literature about both the
tectonic meaning and the deformation age of the FMSZ. Musumeci and Colombo
(2002) obtained a cooling age of the Rocco Verde mylonitic leucogranite of 327 ±
3 Ma by Rb/Sr analyses on the whole rock and magmatic muscovite grains. The
reduced thickness of this intrusion implies a rapid cooling in the lower grade country
rocks, allowing to interpret this age as an emplacement age of the leucogranite.
Furthermore, as it shows evidence of mylonitic deformation, these data have been
interpreted to mark the lower limit of the FMSZ activity suggesting that it is a
late Variscan shear zone. Musumeci and Colombo (2002) interpret the activity of
the FMSZ as linked to an extensional tectonic setting. On the contrary Corsini
et al. (2004) and Sanchez et al. (2011) propose a deformation age of ~22 and
~20 Ma, respectively, obtained by 40 Ar/39 Ar on phengites from FMSZ mica schists.
This suggests that the FMSZ suffered a nearly complete reactivation during Alpine
Orogeny with a strike-slip kinematics. Recently, Carosi et al. (2016) and Simon-
etti et al. (2017, 2018) demonstrated that the FMSZ is a transpressional type II
strain softening shear zone active during Variscan time. Detailed investigation of the
Strain Softening in a Continental Shear Zone: A Field Guide … 25

main mineral recrystallization mechanisms and of the syn-kinematic mineral asso-


ciations, along the gradient, revealed that the activity of the shear zone happened
under decreasing temperature conditions. Deformation in the FMSZ started at
~340 Ma under high-temperature amphibolite-facies conditions and evolved during
retro-metamorphism up to green schist-facies within at least ~20 Ma.
The Alpine greenschist-facies metamorphic overprint is generally weak in the
Argentera Massif and limited to shear zones cross-cutting the Variscan structures
(Compagnoni et al. 2010). The main alpine structures of the Argentera Massif are
the NW-SE striking Bersezio Fault Zone and the E–W striking Fremamorta-Colle
del Sabbione Shear Zone. Such brittle and brittle–ductile structures developed during
the lower Miocene as a consequence of the transpressional regime that affected the
whole Argentera Massif during this time (Baietto et al. 2009).

4 Geological Field Trip

The two-days field trip develops entirely in the Valle Stura di Demonte (CN, Italy).
The stops are located in the Ischiator (Fig. 3) and Pontebernardo (Fig. 4) valleys
(Fig. 2b). Locations of the first day are reachable following the road SS21 that
starts from Borgo San Dalmazzo (CN) village up to Bagni di Vinadio and from
here, traveling along the road for the locality of Besmorello where it is possible to
park. From the parking location, all the stops can be reached by walking along the
P26 marked trekking path. The itinerary develops in a high mountain environment
and the overall altitude difference is about 750 m. Locations of the second day are
reachable following the road SS21 up to the Pontebernardo village. The first stop is
located along the road which leads to the locality of Prati del Vallone. From here
the other stops can be reached by walking along the GTA marked trekking path and
then following directions to Colle Panieris, along a minor path. The overall altitude
difference is about 1000 m.
Day 1—stop 1: ultramylonitic schist (coord. 44°17’20.9"N; 07°01’20.6"E)
In the first stop (Fig. 3), we can observe the most sheared rocks of the FMSZ that
crop out in the lower part of the Ischiator valley (Fig. 2b). They are ultramylonites
(Fig. 5a) with > 90% matrix. They show a continuous cleavage (Fig. 5b) made by
white mica and chlorite striking NW-SE dipping steeply toward the NE (Fig. 6a). This
mineral assemblage indicates a greenschist-facies metamorphism. Mineral lineation
plunges at low angle toward the NW. In agreement with the syn-kinematic mineral
assemblage, quartz is recrystallized by subgrain rotation (Fig. 5c) (Piazolo and Pass-
chier 2002; Stipp et al. 2002). This indicates temperature of the deformation to be
~400 °C. An S–C–C’ fabric, associated to a top-to-the SW sense of shear, is well-
detectable both at the meso- and microscale (Fig. 5a and d). Flow analysis performed
by Simonetti et al. (2018) with the C’ shear band method (Kurtz and Nortrup 2008)
highlights a general shear with ~60% simple shear (Fig. 7).
26 M. Simonetti et al.

Fig. 3 Structural–geological map (1:10000 scale) of the Ischiator Valley (Fig. 2). 1: undifferentiated
debris (udb); 2: biotite-bearing mylonitic gneiss (BtMGns); 3: leucogranite (lgn); 4: ultramylonitic
leucogranite (lgnU); 5: mylonitic schist (MsBW); 6: ultramylonite (Uml); 7: mylonitic gneiss with
biotite and white mica (MgBW); 8: protomylonitic gneiss with biotite and sillimanite (PgBS); 9:
migmatitic gneiss of the Tinée complex (MgmBS)

Day 1—stop 2: mylonitic schist (coord. 44°17’14.1"N; 07°01’39.7"E)


In this stop, it is possible to observe the most classical mylonites of the FMSZ.
They are medium-grained dark-green mylonitic schist (Fig. 8a) made of quartz,
K-feldspar and plagioclase porphyroclasts in a fine-grained biotite and white mica
matrix (Fig. 8b). The amount of matrix is nearly 75%.
Foliation strikes NW-SE and dips steeply toward the NE or SW with a mineral
lineation that plunges at low angle toward the NW (Fig. 6b). The orientation of the
main foliation is perfectly concordant with the one observed in the ultramylonitic
schist at stop 1. Foliation is a penetrative disjunctive cleavage with smooth and
parallel cleavage domains (Fig. 8b), made by biotite and white mica. Both at the
meso- and microscale kinematic indicators are well recognizable. They are mainly
represented by rotated porphyroclasts, quarter mats (Fig. 8c), S–C’ fabric (Fig. 8c),
mica fish (Fig. 8d), and quartz oblique foliation representing a top-to-the SW or
top-to-the SE sense of shear, depending on the dip direction of the foliation. Quartz
is characterized by the presence of bigger grains with irregular boundaries and new
grains of smaller size. Subgrains with irregular shape also exist (Fig. 8e). These
Strain Softening in a Continental Shear Zone: A Field Guide … 27

Fig. 4 Structural–geological map (1:10000 scale) of the Pontebernardo Valley (fig. 2; modified
after Carosi et al. 2016). 1: undifferentiated debris (Udb); 2: fluvial deposit (Fdp); 3: undiffer-
entiated glacial deposit (Ugd); 4: limestone breccia (Lbc); 5: triassic quarzite(QtzTr); 6: alpine
mylonitic schist with chlorite and white mica (MsGS); 7: migmatitic gneiss of the GSV complex
(MgmB); 8: biotite-bearing mylonitic gneiss (BtMGns); 9: mylonitic leucogranite (Lgn); 10: ultra-
mylonitic leucogranite (LgnU); 11: amphibole-bearing mylonitic gneiss (AmGns); 12: ultramy-
lonitic schist (UmlS); 13: mylonitic schist (MsBW); 14: quarzite (Qtz); 15: phyllonite (Pll); 16:
ultramylonite (Uml); 17: mylonitic gneiss with biotite and white mica (MgBW); 18: mylonitic
gneiss with biotite and sillimanite (MgBS); 19: protomylonitic gneiss with biotite and sillimanite
(PgBS); 20: amphibole-bearing migmatitic gneiss (GnsAmp); 21: migmatitic gneiss of the Tinée
complex (MgmBS); 21: amphibolite (Amp). Yellow stars show the stops of the itinerary

features suggest subgrain rotation and recrystallization (Piazolo and Passchier 2002;
Stipp et al. 2002) as the main dynamic recrystallization mechanism overprinting
a previous grain boundary migration. In some domains, effects of grain boundary
migration such as window and pinning structures can still be recognized (Fig. 8f).This
indicates a deformation compatible between the transition of the two recrystallization
mechanisms at ~500 °C.
Kinematic vorticity analysis (Fig. 7) in these rocks was undertaken by Simonetti
et al. (2018) and revealed a general flow with a dominant component of pure shear
28 M. Simonetti et al.

Fig. 5 a Ultramylonite at the outcrop scale, an S-C’ fabric is indicative of a top-to-the SW sense
of shear; b continuous cleavage made by white mica and chlorite. Amount of matrix is over
90% (crossed nicols); c quartz in the ultramylonite with bimodal grainsize, subgrains and new
grains indicative of a subgrain rotation recrystallization (crossed nicols); D) S-C-C’ fabric in the
ultramylonite pointing a top-to-the SW sense of shear (crossed nicols)

between ~55–70% (C’ shear band method and stable porphyroclasts method). In situ
U-Th–Pb petrochronology on syn-kinematic monazites (Fig. 9) constrain the age of
shearing in these rocks at ~320 Ma (Simonetti et al. 2018).
Day 1—stop 3: mylonitic gneiss with biotite and white mica (coord.
44°16’58.1"N; 07°01’09.2"E)
Following the P26 path, we reach Rifugio Migliorero (Fig. 2b) in the middle of Ischi-
ator valley (Fig. 3). Here elegant outcrops of protomylonitic gneiss exist (Fig. 10a).
These rocks are made of K-feldspar, plagioclase, and quartz porphyroclasts in a fine-
grained biotite and white mica matrix (Fig. 10b). The amount of matrix is ~50%.
The foliation is oriented NW-SE and dips at high angle toward the NE (Fig. 6c),
similar to previous stops. The foliations are penetrative/disjunctive type with smooth
and parallel cleavage domains (Fig. 10b). Quartz grains, variable between micro-
metric and millimetric size, show lobate grain boundaries (Fig. 10c) indicating that
the main dynamic recrystallization mechanism is grain boundary migration. Incip-
ient subgrain rotation recrystallization is recognized locally. The mineral assemblage
indicative of amphibolite-facies conditions, and quartz microstructures reveal higher
Strain Softening in a Continental Shear Zone: A Field Guide … 29

Fig. 6 Equal angle, lower hemisphere projections: a Poles to foliation (yellow squares, 60 data) and
mineral lineation (blue triangles, 13) in the ultramylonitic schists; b Poles to foliation (green dots,
163 data) and mineral lineation (blue triangles, 27 data) in the mylonitic schists; c poles to foliation
(pink diamonds, 110 data) and mineral lineation (blue triangles, 18 data) in the protomylonitic
gneiss; d poles to foliation (blue dots, 53 data) and fold axis (red triangles, 13 data) in the Tinèe
complex; e poles to foliation (blue dots, 39 data) and fold axis (red triangles, 11 data) in the Gesso-
Stura-Vesubie complex; f poles to mylonitic foliation (gray dots, 17 data) and mineral lineation
(pink triangles, 7 data) in the Colle Panieris-Mt Peiron alpine thrust
30 M. Simonetti et al.

Fig. 7 a Polar istograms used for calculating kinematic vorticity number (Wk) with the C’ shear
band method (Kurz and Northrup 2008) for a protomylonite (Wk = 0.34), a mylonite (Wk = 0.61)
and a ultramylonite (Wk = 0.81); b results of kinematic vorticity analysis with stable porphyroclasts
method (Passchier, 1987; Wallis et al. 1993; Law et al. 2004) in a protomylonitic and in a mylonitic
sample. Data are represented with Rigid Grain Net graphs (Jessup et al. 2007). Wk values are
0.56 and 0.67 for protomylonite and mylonite, respectively. Values indicate a pure shear-dominated
deformation and point out also an increase of the simple shear component in the mylonite

deformation temperature, >~500 °C (Piazolo and Passchier 2002; Stipp et al. 2002;
Passchier and Trouw 2005), compared to the mylonitic schist at stop 2.
At the outcrop scale, asymmetric and stretched leucosomes are recognizable.
In thin section, rotated porphyroclasts and porphyroclasts with asymmetric strain
shadows (Fig. 10d), mica fish, C and C’ shear band occur. Kinematic indicators
point to a top-to-the SW shear. Vorticity estimation (Fig. 7) in these rocks revealed a
deformation with a prevalent component of pure shear (~70%; Simonetti et al. 2018).
Recent U-Th–Pb in situ petrochronology on syn-kinematic monazites, revealed a
deformation age of ~ 328 Ma (Fig. 9; Simonetti et al. 2018), older than the age of
shear in the mylonitic schist of stop 2.
Strain Softening in a Continental Shear Zone: A Field Guide … 31

Fig. 8 a Medium-grained dark-green mylonitic schist at the outcrop scale with S-C’ fabric pointing
to a top-to-the SW sense of shear is well-detectable b feldspar porphyroclasts in a finer matrix (nearly
75%), foliation is a penetrative disjunctive cleavage with smooth and parallel cleavage domain made
by biotite and white mica (crossed nicols); c S-C’ fabric and white mica quarter mats (green arrow)
around a feldpsar porphyroclast, both kinematic indicators are consistent with a top-to-the SW sense
of shear (crossed nicols); d micafish made of white mica showing a top-to-the SW sense of shear
(crossed nicols); e quartz grains with irregular boundaries and new grains of smaller size, subgrains
are also present (crossed nicols); f quartz with lobate grain boundaries in a domain where grain
boundary migration effcts are still preserved (crossed nicols)
32 M. Simonetti et al.

Fig. 9 a Age of shear deformation along the deformation gradient (U–Th–Pb on monazite; Simon-
etti et al. 2018); b Examples of textural position and zoning of monazite grains selected for in situ
dating in a protomylonite and a mylonite. Compositional map of Y shows the spot location and the
corresponding 206 Pb/238 U age (modified from Simonetti et al. 2018)

Day 1—stop 4: protomylonitic gneiss with biotite and sillimanite (coord.


44°16’46.8"N; 07°00’59.1"E)
We observe protomylonitic gneiss with biotite and sillimanite located at the boundary
of the FMSZ (Fig. 11a). At the outcrop scale, this lithotype is very similar to the
mylonitic gneiss of stop 3, the main difference is the presence of sillimanite and
the lower abundance of white mica (Carosi et al. 2016). Asymmetric and stretched
leucosomes, variable between centimetric and decimetric size, are recognizable
(Fig. 11a).The amount of matrix in this lithotype is ~40% which testifies a less intense
deformation of these rocks compared to those visited in stops 1, 2, 3 and 4. The main
foliation, oriented NW-SE and dipping at high angle toward the NE, is a penetra-
tive disjunctive cleavage (Fig. 11b) with smooth and anastomosed cleavage domains
rich in biotites and prismatic or fibrolitic sillimanite (Fig. 11b). The orientation of
the main foliation is perfectly concordant with the orientation observed in the more
sheared rocks. At the microscale kinematic indicators, such as porphyroclasts with
Strain Softening in a Continental Shear Zone: A Field Guide … 33

Fig. 10 a Protomylonitic gneiss cropping out close to the Migliorero hut, an S–C’ fabric pointing to
a top-to-the SW sense of shear; b penetrative disjunctive cleavage with smooth and parallel cleavage
domains in the protomylonitic gneiss (crossed nicols); c quartz showing lobate grain boundaries
indicative of grain boundary migration (crossed nicols); d feldspar porphyroclasts with asimmetric
strain shadows indicative of a top-to-the SW sense of shear (crossed nicols)

asymmetric strain shadows, mica fish, and C–C’ shear bands are present (Fig. 11c).
All of them point to a top-to-the SW sense of shear. Quartz is dynamically recrys-
tallized by grain boundary migration (Fig. 11d). Deformation temperature above
500 °C is in agreement with the syn-kinematic biotite + sillimanite assemblage and
quartz microstructures. The age of shear deformation, obtained with in situ U–h–Pb
petrochronology, on syn-kinematic monazites (Simonetti et al. 2018), is ~ 340 Ma
(Fig. 9). Deformation is therefore older compared to the age of shear deformation
obtained in the mylonitic schist (stop 2) and in the mylonitic gneiss (stop 3).
Day 1—stop 5: migmatitic gneiss of the Tinèe Complex (coord. 44°16’39.2"N;
07°00’29.3"E)
In the area near the Laghi dell’Ischiator (Fig. 3), migmatitic gneiss with biotite
and sillimanite (‘Rabuons Formation’ in the literature; Faure-Muret 1955; Malaroda
et al. 1970), belonging to the Tinèe Complex, crops out (Fig. 12a). Migmatites
resulted from partial melting of metasediments (Malaroda et al. 1970; Ferrando et al.
2008; Compagnoni et al. 2010) are not affected by shear of the FMSZ. Leucosomes
show variable thickness. These rocks show a disjunctive foliation (Fig. 12b) oriented
34 M. Simonetti et al.

Fig. 11 a Protomylonitic gneiss with biotite and sillimanite at the outcrop scale; b disjunctive
cleavage with smooth and anastomosed cleavage domain rich in biotites and sillimanite.Rotated
porphyroclasts show a a top-to-the SW sense of shear (parallel nicols); c S–C–C’ fabric indicative of
top-to-the SW sense of shear (parallel nicols); d quartz showing lobate grain boundaries indicative
of grain boundary migration (crossed nicols)

NW-SE and variably dipping to NE or SW (Fig. 6d). Quartz presents a big grain
size and lobated margins in accordance with grain boundary migration deformation
mechanism (Fig. 12c). Open to tight symmetric folds with subvertical axial planes,
oriented NW-SE and sub-horizontal axes, are present (Fig. 12e). Fold axial plane is
parallel to the mylonitic foliation within the FMSZ so that the folding event occurred
in a strain regime compatible with the shearing deformation in the FMSZ (Simonetti
et al. 2018). Folds contributed to accommodate the nearly horizontal shortening
resulted from the pure shear component acting during transpression.
Transpressional deformation is generally subjected to strain compatibility prob-
lems (Hudleston 1999) because it is not easy to explain how the flattening component
of the deformation is accommodated.
The presence of syn-shear folds in the unsheared migmatites testifies that the Tinée
Complex and the shear zone could have been stretched together during transpression
and therefore no strain compatibility problems are present.
Strain Softening in a Continental Shear Zone: A Field Guide … 35

Fig. 12 a Migmatitic gneiss of the Tinée complex at the outcrop scale; b disjunctive foliation of the
migmatitic gneiss (crossed nicols); c quartz showing big grainsize and lobated margins indicative
of grain boundary migration (crossed nicols); d symmetric folds with subvertical NW-SE striking
axial plane in the migmatitic gneiss

Day 2—stop 1: migmatites of the Gesso-Stura-Vesubie Complex (Coord.


44°20’00.8"N; 06°59’54.6"E)
From Pontebernardo, we follow the road toward the locality Prati del Vallone (Figs. 2b
and 4). Halfway down the road it is possible to leave the car and follow a footpath up
to a small abandoned technical building where we can reach the first outcrop of the
day. Here migmatites of the Gesso-Stura-Vesubie Complex, formed at the expense of
metasedimentary rocks, can be observed (Fig. 13a). These rocks are medium-grained
migmatitic gneiss with a quartz-plagioclase (An20–30 )-biotite-fibrolitic sillimanite-
cordierite assemblage with few K-feldspar and muscovite (Fig. 13b). Relict kyanite
and garnet rarely included in plagioclase suggest that these rocks reached P-T
conditions of the upper amphibolite facies (650–700 °C; 0.6–0.8 GPa) and subse-
quently suffered a decompression from the kyanite to the sillimanite stability field
(Compagnoni et al. 2010).
Meter-thick leucocratic portions with less biotite locally occur. About 10–20 Ma
after the Carboniferous HP metamorphism, the Argentera Massif underwent an
amphibolite-facies metamorphism with extensive development of migmatites at 323
± 12 Ma (Rubatto et al. 2001). Older ages for the anatectic event are reported from
other Variscan migmatites in different segments of the Variscan chain (Compagnoni
36 M. Simonetti et al.

Fig. 13 a Para-derived migmatites belonging to the Gesso-Stura-Vesubie Complex; b alternating


levels of red biotite and quartz + cordierite domains that materialize a gneissic foliation (parallel
nicols); c folds affecting the gneissic foliation of the GSV migmatites; d gneissic foliation folded
in thin section (parallel nicols)

et al. 2010; Oliot et al. 2015; Schneider et al. 2014), suggesting that several melting
events may have succeeded one to each other in a complex anatectic history.
Structurally, the migmatites show a disjunctive foliation that is affected by open
to tight symmetric folds (Fig. 13c, d), with fold axes gently plunging toward the
NW and subvertical NW-SE striking axial planes (Fig. 6e; Carosi et al. 2016). Axial
planes are parallel to the mylonitic foliation of the FMSZ (Fig. 14a). Similarly to the
structural framework observed in stop 5 of day 1, folding event occurred in a strain
regime compatible with shear in the FMSZ (Fig. 14b; Carosi et al. 2016; Simonetti
et al. 2018) and contributed to accommodate a nearly horizontal shortening. We come
back on the road, and we reach the locality Prati del Vallone where we can leave the
car.
Day 2—stop 2: sheared amphibole-bearing gneiss (coord. 44°19’06.7"N;
06°59’01.5"E)
In this stop, we can observe a decameter-sized lense of amphibole-bearing gneiss
(Fig. 4) within the FMSZ (“Embrechiti ed anatessiti anfiboliche” according to
Malaroda et al. 1970). These are light-colored gneiss made of amphibole (horn-
blende), quartz, biotite, and plagioclase (An 50–60%) generally with a massive
Strain Softening in a Continental Shear Zone: A Field Guide … 37

Fig. 14 Geological cross-section of the FMSZ (legend is the same as in Fig. 4; modified after
Carosi et al. 2016): a section across the Pontebernardo Valley (section A, Fig. 4); b relationship
between the mylonitic foliation in the FMSZ and the folds observed in both the GSV and the Tinée
complexes (Carosi et al. 2016); c section across the Mt Peiron-Colle di Stau-Rocco Verde (section
B, Fig. 4)

appearance (Fig. 15a) or only weakly foliated (Fig. 15b). When present, the foliation
is disjunctive(Fig. 15c). The gneiss belongs to the Tinée complex and extensively
crops out in the north-western sector of the crystalline basement of the Argentera
Massif (Fig. 4; Carosi et al. 2016). According to Compagnoni et al. (2010), the
protolith of the gneiss is an intermediate igneous rock. Several lenses of this litho-
type (Fig. 15d), showing different amounts of deformation, exist within the FMSZ
(Carosi et al. 2016). Sometimes deformation is very strong so that the original struc-
ture of the rock is nearly unrecognizable (Fig. 15e) and the percentage of matrix
in higher than 65%. Because of the position of this lense, very close to the central
part of the shear zone, it is possible to observe sheared portions with evident S-C-C’
fabric indicative of a top-to-the SE sense of shear (Fig. 15f).
Day 2—stop 3: Mt Peiron-Colle Panieris alpine thrust (coord. 44°19’31.6"N;
06°57’24.6"E)
From Prati del Vallone, we follow the ‘GTA’ path and then we go toward the Colle
Panieris on a minor path. Below the Mt Peiron, we can observe evidence of Alpine
deformation in the basement of the Argentera Massif.
In the Mt Peiron-Colle Panieris area, a mylonitic belt with metric thickness (Fig. 4)
thrusts amphibole-bearing mylonitic gneiss of the FMSZ above Permo-Triassic sedi-
mentary cover of the Argentera Massif (Fig. 16a; 14c), represented by light-colored
quartzite. Mylonitic foliation strikes nearly E–W and gently dips toward the N with
a NNW plunging mineral lineation (Fig. 6f). Mylonite are made of quartz and
feldspar porphyroclasts in a fine-grained chlorite + white mica matrix (Fig. 16b). The
38 M. Simonetti et al.

Fig. 15 a Amphibole-bearing gneiss with massive aspect; b foliated amphibole-bearing gneiss;


c amphibole-bearing gneiss in thin section, the disjunctive foliation can be recognized (parallel
nicols); d decametric lenses of amphibole-bearing gneiss (highlighted in yellow) within the FMSZ;
e low strain domain and strongly sheared domain with S–C’ fabric in the amphibole-bearing gneiss,
sense of shear is top-to-the SW; f S–C’ fabric at the outcrop scale, top-to-the SW sense of shear

matrix ranges 65–75%. The mineral assemblage indicates greenschist-facies meta-


morphism. Foliation varies from a continuos to a disjunctive cleavage with zonal and
parallel cleavage domains. An S–C fabric is well recognizable both at the outcrop
scale (Fig. 16c) and in thin section is indicative of an inverse top-to-the SSE sense
of shear. Quartz is dynamically recrystallized by subgrain rotation recrystallization,
and feldspar porphyroclasts are fractured suggesting temperature of deformation in
agreement with the syn-kinematic mineral assemblage.
Strain Softening in a Continental Shear Zone: A Field Guide … 39

Fig. 16 a Mylonitic belt with metric thickness thrusting amphibole-bearing mylonitic gneiss of
the FMSZ above Permo-Triassic sedimentary cover of the Argentera Massif in the Mt Peiron-Colle
Panieris area; b mylonite of the Mt Peiron-Colle Panieris thrust in thin section, quartz and feldspar
porphyroclasts in a fine-grained chlorite + white mica matrix can be recognized (crossed nicols);
c S-C fabric at the outcrop scale showing a top-to-the S sense of shear

The Argentera Massif was involved in the Alpine orogeny at ~ 22 Ma (Corsini


et al. 2004; Sanchez et al. 2011). Most of the deformation is concentrated in the
sedimentary covers (Malaroda et al. 1970; Barale et al. 2016; d’Atri et al. 2016)
that are detached from the metamorphic basement along evaporites and limestone
breccias levels. Alpine overprint in the basement is limited and concentrated along
ductile and brittle ductile shear zones (Compagnoni et al. 2010; Carosi et al. 2016;
Simonetti et al. 2017, 2018), the Mt Peiron-Colle Panieris thrust can be considered
as one of such shear zones.
Day 2—stop 4: undeformed Permo-Triassic quartzite (coord. 44°19’18.8"N;
06°57’12.8"E)
Not far from stop 2, following a small path toward the SW for nearly 450 m, white-
gray quartzites of Permo-Triassic age (Barale et al. 2016) crops out and lay in angular
40 M. Simonetti et al.

unconformity directly above the migmatitic basement (Fig. 17a). They are made of
quartz and very few white mica (Fig. 17b), ripple marks are often well recognizable
(Fig. 17c).
Above the quartzites tectonic carbonatic breccias known as ‘Formazione delle
Carniole’ (Malaroda et al. 1970) with both gypsum and dolomitic limestone clasts,
occur (Fig. 4). They are interpreted as weak horizons along which the detachment
of the sedimentary succession occurs.
The presence of undeformed Permo-Triassic sediments directly above the
basement is a good constrain for the age of Variscan transpression.

Fig. 17 a Permo-Triassic quarzites in angular unconformity directly above the migmatitic basement
in the Colle Panieris area; b fine-grained white mica in the quarzite (crossed nicols); c sedimentary
bedding with ripple marks
Strain Softening in a Continental Shear Zone: A Field Guide … 41

5 Discussion

The features observed in the outcrops of the field trip highlight a complex evolu-
tion of the FMSZ. Kinematic vorticity analysis (Simonetti et al. 2018) highlights a
transpression (Fig. 18a) with a variable component of simple shear both spatially
and temporally (Fig. 18b), increasing along the deformation gradient from the wall
rocks to the high-strain zone in the central part of the FMSZ (Fig. 18c). This matches
with the structural asset of the FMSZ in the field where a subvertical mylonitic foli-
ation can be recognized, with constant orientation in all the lithotypes. This is in

Fig. 18 a Diagram showing relationship between the orientation of the maximum Instantaneous
Stretching Axis (ISAmax) with respect to the shear zone boundary (angle θ) related to the kinematic
vorticity number Wk (modified after Fossen and Tikoff 1993; Fossen et al. 1994). Wk value for
which simple shear = pure shear is 0.71 according to Law et al. (2004) and Xypolias (2010). The
distribution of the samples shows a variation from a pure shear-dominated transpression to a simple
shear-dominated transpression linked to the increase of the vorticity number. Orange dashed line
represents the theoretical trend of the angle θ in relation to Wk; b Percentage of pure shear (PS)
and simple shear (SS) in relation to the calculated maximum and minimum Wk values; c variation
of the vorticity number in relation to the distance of the sample from the center of the shear zone.
The distribution of points shows increasing Wk values (i.e., increasing of simple shear component
of deformation) toward the central part of the shear zone (trend: blue-dashed line)
42 M. Simonetti et al.

agreement with the model of transpression (Fossen et al. 1994). Folds with subver-
tical axial plane parallel to the mylonitic foliation in the GSV Complex and Tinèe
Complex is another evidence of horizontal shortening attributed to the pure shear
component. Because the two walls of the FMSZ (the GSV and the Tinèe Complex)
can deform together with the shear zone, no strain compatibility problems are present
(Simonetti et al. 2018).
In situ petrochronology (combining U-Th–Pb ages with textural observations
and monazite chemistry) revealed that the Variscan transpression in the Argentera
Massif started at the metamorphic peak of the GSV complex (dated at ~340 Ma;
Compagnoni et al. 2010; Rubatto et al. 2010), or shortly after, and triggered its
exhumation (Simonetti et al. 2018). This approach revealed a long-lasting activity of
the FMSZ that activated for at least 20 Ma under changing metamorphic conditions,
from high to low temperatures.
The presence of medium- to high-grade metamorphic mylonites associated with
lower-grade ones, localized in the central part of the FMSZ, the strong deformation
gradient and the changes in the deformation regime and age of deformation along
the gradient evidence that the FMSZ evolved as a strain softening type II shear
zone (Fig. 19; Fossen 2016). In this type of shear zone, deformation progressively
concentrates in the central part because of strain softening (due to the presence
of channeled fluids, metamorphic reactions, and grain size reduction) leaving the
external parts of the structure progressively inactive (Fig. 19). Because of this the
active thickness decreases, whereas the total thickness remains constant (Fig. 19).
Strain softening can be enhanced by the presence of fluids joined with metamorphic
reactions and grain size reduction, well-detectable both at the meso—and microscale.
Evidence of the first two factors is the breakdown of amphibolite-facies minerals such
as sillimanite and biotite and consequent growth of white mica and chlorite in the
most sheared rocks. Along the deformation gradient a reduction in the amount of
quartz and feldspar in the rocks, in favor of the amount of phyllosilicates, is evident.
Because of the type II evolution, protomylonites in the external part of the FMSZ
preserve features acquired during the early stage of shear (testified by the older age of
deformation) during high-temperature conditions, whereas mylonites, in the internal
part of the shear zone, record features acquired during younger deformation stages
at lower temperature. Finally, the ultramylonites preserve evidence of the final stage
of activity of the shear zone. Despite a similar evolution, characterized by strain
localization during decreasing temperature, has been recently described in detail
in the Yukon River shear zone (Northern Cordillera, Canada; Parsons et al. 2018),
other well-described examples of regional-scale strain softening shear zones are not
so common in the literature up to now. The FMSZ should be considered as a new
example of a strain softening regional-scale shear zone, that can be useful to the
community of structural geologists for future process-oriented investigations.
Furthermore, the described data are yet another example that, as pointed out by
Fossen and Cavalcante (2017) and Oriolo et al. (2018), a multidisciplinary approach
is strictly necessary to obtain meaningful results to address unsolved problems in
modern structural geology and in the study of shear zones, were complexity arises
Strain Softening in a Continental Shear Zone: A Field Guide … 43

Fig. 19 Sketch summarizing the evolution of the Ferriere-Mollières Shear Zone according to the
type II growth model (Fossen 2016). Graphs show the theoretical variation of total and active
thickness during time and the amount of strain within the shear zone

from the coupling between long-lasting progressive and/or polyphasic deformation,


metamorphism and fluid–rock interaction.
Detailed field work and meso-structural analysis should always be combined with
microstructural observations, with modern techniques for vorticity analysis, with
deformation temperature estimation and with detailed petrochronology in order to
unreveal the fundamental steps of the tectono-metamorphic history of shear zones.

6 Conclusive Remarks

• A deformation gradient, marked by a transition from protomylonites to mylonites


and to ultramylonites, is recognizable from the marginal part of the FMSZ shear
zone toward the central part;
44 M. Simonetti et al.

• Foliation of the FMSZ shows a constant trend of orientation in all the different
types of mylonites;
• Fold axial planes the GSV complex and Tinèe complex are subparallel to the
mylonitic foliation;
• Vorticity analysis constrained a transpressional deformation with a prevalent
component of pure shear acting together with simple shear. Simple shear
component increases along the deformation gradient;
• Syn-kinematic mineral assemblage, along the deformation gradient, highlights
a syn-shear retro-metamorphism from high-temperature amphibolite-facies to
greenschist-facies. Quartz microstructures also highlight a temperature decrease
during deformation;
• Because of strain softening phenomena, deformation in the FMSZ starts to
concentrate in its central part leaving the margin inactive according to a type
II evolution;
• Across the deformation gradient it is possible to observe different stages of
formation of the FMSZ;
• In situ U-Th–Pb petrochronology on monazite revealed a long-lasting activity
of the FMSZ (between ~ 340 Ma and ~ 320 Ma) and constrained the different
strain softening evolution stages;
• The FMSZ is a very good case study of strain softening shear zone at regional
scale, for future process-oriented investigations.

Acknowledgements The staff of Rifugio Migliorero and the staff of Rifugio Prati del Vallone
are acknowledged for the hospitality during field work. Research supported by funds from Torino
University (Ricerca Locale 2017, 2018) and PRIN 2015 (resp. R. Carosi and C. Montomoli). The
Springer team (Marion Schneider, Annett Buettener, Boopalan Renu, Alexis Vizcaino, Doerthe
Mennecke-Buehler) is thanked for proofreading and other assistance. Soumyajit Mukherjee is
thanked for editorial work and constructive review. Dutta and Mukherjee (2021) encapsulate this
work.

References

Advokaat, E. L., van Hinsbergen, D. J. J., Maffione, M., Langereis, C. G., Vissers, R. L. M.,
Cherchi, A., et al. (2014). Eocene rotation of Sardinia, and the paleogeography of the western
Mediterranean region. Earth and Planetary Science Letters, 401, 183–195.
Arthaud, F., & Matte, P. (1977). Late Paleozoic strikeslip faulting in southern Europe and northern
Africa; result of a right-lateral shear zone between the Appalachians and the Urals. Geological
Society of America Bulletin, 88, 1305–1320.
Baietto, A., Perello, P., Cadoppi, P., & Martinotti, G. (2009). Alpine tectonic evolution and thermal
water circulations of the Argentera Massif (South-Western Alps). Swiss Journal of Geosciences,
102(2), 223–245.
Ballèvre, M., Manzotti, P., & Dal Piaz, G.V. (2018). Pre-Alpine (variscan) inheritance: a key for
the location of the future valaisan basin (Western Alps). Tectonics 37, 786–817. https://doi.org/
10.1002/2017TC004633
Strain Softening in a Continental Shear Zone: A Field Guide … 45

Barale, L., Bertok, C., D’atri, A., Martire, L., Piana, F., & Domini, G. (2016). Geology of the
Entracque-Colle di Tenda area (Maritime Alps, NW Italy). Journal of Maps, 12, 359–370.
Bellot, J.P. (2005). The Palaeozoic evolution of the Maures massif (France) and its potential corre-
lation with other areas of the Variscan Belt: a review. In: R. Carosi, R. Dias, D. Iacopini, G.
Rosenbaum (eds) The southern Variscan belt. Journal of the Virtual Explorer, 19, 1441–8142
Brun, J. P., & Burg, J. P. (1982). Combined thrusting and wrenching in the Ibero-Armorican arc: A
corner effect during continental collision. Earth and Planetary Science Letters 61, 319–332.
Burg, J.-P., & Matte, P. (1978). A cross section through the French Massif central and the scope of
its variscan ggeodynamic evolution. Zeitschrift der Deutschen Geologischen Gesellschaft, 129,
429–460.
Carosi, R., & Palmeri, R. (2002). Orogen-parallel tectonics transport in the Variscan belt of north-
eastern Sadinia (Italy): Implications for the exhumation of medium-pressure metamorphic rocks.
Geological Magazine, 139, 497–511.
Carosi, R., Montomoli, C., Tiepolo, M., & Frassi, C. (2012). Geochronological constraints on
post-collisional shear zones in the Variscides of Sardinia (Italy). Terra Nova, 24, 42–51.
Carosi, R., D’addario, E., Mammoliti, E., Montomoli, C., & Simonetti, M. (2016). Geological map
of the northwestern portion of the Ferriere-Mollieres shear zone, Argentera Massif, Italy. Journal
of Maps, 12, 466–475.
Carosi, R., Frassi, C., Iacopini, D., & Montomoli, C. (2005). Post collisional transpressive tectonics
in northern Sardinia (Italy). Journal Virtual Explorer, 19(3), 1–18.
Carosi, R., Montomoli, C., Iaccarino, S., & Visonà, D. (2018). Structural evolution, metamorphism
and melting in the greater Himalayan sequence in central-western Nepal. In P. J. Treloar &
M. P. Searle (Eds.), 2019 Himalayan tectonics: A modern synthesis (Vol. 483, pp. 305–323).
London:Geological Society, Special Publications.
Compagnoni, R., Ferrando, S., Lombardo, B., Radulesco, N., & Rubatto, D. (2010). Paleo-European
crust of the Italian western alps: Geological history of the argentera massif and comparison with
Mont Blanc-Aiguilles Rouges and Maures-Tanneron Massifs. In M. Beltrando, A. Peccerillo, M.
Mattei, S. Conticelli & C. Doglioni (Eds.), Journal of the Virtual Explorer, 36, 4. https://doi.org/
10.3809/jvirtex.2009.00228
Corsini, M., Ruffet, G., & Caby, R. (2004). Alpine and late Hercynian geochronological constraints
in the Argentera Massif (Western Alps). Eclogae Geologicae Helvetiae, 97, 3–15.
Corsini, M., & Rolland, Y. (2009). Late evolution of the southern European Variscan belt: Exhuma-
tion of the lower crust in a context of oblique convergence. Comptes Rendus Geosciences,
341(2–3), 214–223.
Cottle, J. M., Searle, M. P., Jessup, M. J., Crowley, J. L., & Law, R. D. (2015). Rongbuk Re-visited:
Geochronology of Leucogranites in the footwall of the south tibetan detachment system, Everest
region, Southern Tibet. Lithos, 227, 94–106.
Cruciani, G., Montomoli, C., Carosi, R., Franceschelli, M., & Puxeddu, M. (2015). Continental
collision from two perspectives: A review of Variscan metamorphism and deformation in northern
Sardinia. Periodico di mineralogia, 84(3), 657–699.
d’Atri, A., Piana, F., Barale, L., Bertok, C., & Martire, L. (2016). Geological setting of the southern
termination of Western Alps. International Journal of Earth Sciences (Geol Rundsch), 105,
1831–1858. https://doi.org/10.1007/s00531-015-1277-9.
Dias, R., & Ribeiro, A. (1995). The Ibero-Armorican Arc: A collision effect against an irregular
continent? Tectonophysics, 246, 113–128.
Dias, R., Ribeiro, A., Romão, J., Coke, C., & Moreira, N. (2016). A review of the arcuate structures
in the Iberian Variscides; constraints and genetic models. Tectonophysics, 681, 170–194. https://
doi.org/10.1016/j.tecto.2016.04.011.
Di Vincenzo, G., Carosi, R., & Palmeri, R. (2004). The relationship between tectono-metamorphic
evolution and argon isotope records in white mica: Constraints from in situ 40 Ar–39 Ar laser
analysis of the Variscan basement of Sardinia. Journal of Petrology, 45(5), 1013–1043.
Di Vincenzo, G., Talarico, F., & Kleinschmidt, G. (2007). An 40Ar–39Ar investigation of the mertz
glacier area (George V Land, Antarctica): implications for the ross orogen–east antarctic craton
46 M. Simonetti et al.

relationship and gondwana reconstructions. Precambrian Research, 152, 93–118. https://doi.org/


10.1016/j.precamres.2006.10.002
Dutta, D, Mukherjee, S. (2021). Introduction to Structural Geology and Tectonics Field Guide-
book—Volume 1. In S. Mukherjee (Ed.), Structural Geology and Tectonics Field Guidebook—
Volume 1. Switzerland AG: Springer Nature. pp. xi-xvi. ISBN: 978-3-030-60142-3.
Faure-Muret A. (1955). Etudes géologiques sur le massif de l’Argentera-Mercantour et ses
enveloppes sédimentaires. Mémoires pour servir à l’explication de la Carte géologique détaillée
de la France, Paris, Imprimerie Nationale, France, with « Esquisse Géologique du Massif de
l’Argentera-Mercantour et de sa Bordure Sédimentaire (Versant français)» at the 1/100.000 scale.,
pp 336.
Fernández-Lozano, J., Pastor-Galán, D., Gutiérrez-Alonso, G., & Franco, P. (2016). New kinematic
constraints on the Cantabrian orocline: A paleomagnetic study from the Peñalba and Truchas
synclines, NW Spain. Tectonophysics, 681, 195–208. https://doi.org/10.1016/j.tecto.2016.02.019.
Ferrando, S., Lombardo, B., & Compagnoni, R. (2008). Metamorphic history of HP mafic granulites
from the Gesso-Stura Terrain (Argentera Massif, Western Alps, Italy). European Journal of
Mineralogy, 20, 777–790.
Ferrara, G., & Malaroda, R. (1969). Radiometric age of granitic rocks from the Argentera Massif
(Maritime Alps). Bollettino della Società Geologica Italiana, 88, 311–320.
Fossen, H., & Tikoff, B. (1993). The deformation matrix for simultaneous simple shearing, pure
shearing and volume change, and its application to transpression-transtension tectonics. Journal
of Structural Geology, 15, 413–422. https://doi.org/10.1016/0191-8141(93)90137-Y.
Fossen, H., Tikoff, B., & Teyssier, C. (1994). Strain modeling of transpressional and transtensional
deformation. Norsk Geol. Tidsskr., 74, 134–145.
Fossen H. (2016). In Structural geologys (p. 510). Cambridge University Press.
Fossen, H., & Cavalcante, G. C. G. (2017). Shear zones–A review. Earth-Science Reviews, 171,
434–455.
Fossen, H., & Rotevatn, A. (2016). Fault linkage and relay structures in extensional settings—a
review. Earth-Science Reviews, 154, 14–28. https://doi.org/10.1016/j.earscirev.
Frassi, C., Carosi, R., Montomoli, C., & Law, R. D. (2009). Kinematics and vorticity of flow
associated with post-collisional oblique transpression in the Variscan Inner Zone of northern
Sardinia (Italy). Journal of Structural Geology, 31, 1458–1471.
Hudleston, P. (1999). Strain compatibility and shear zones: Is there a problem? Journal of Structural
Geology, 21, 923–932.
Iaccarino, S., Montomoli, C., Carosi, R., Massonne, H. -J., Langone, & A., Visonà, D. (2015).
Pressure-temperature-time-deformation path of kyanite-bearing migmatitic paragneiss in the Kali
Gandaki valley (central Nepal): Investigation of late Eocene–early Oligocene melting processes:
Lithos (Vol. 231, pp. 103–121). https://doi.org/10.1016/j.lithos.2015.06.005
Iacopini, D., Carosi, R., Montomoli, C., & Passchier, C. W. (2008). Strain analysis of flow in
the Northern Sardinian Varisican belt: Recognition of a partitioned oblique deformation event.
Tectonophysics, 221, 345–359.
Jessup, M. J., Law, R. D., & Frassi, C. (2007). The rigid grain net (RGN): An alternative method for
estimating mean kinematic vorticity number (Wm). Journal of Structural Geology, 29, 411–421.
Kurz, G. A., & Northrup, C. J. (2008). Structural analysis of mylonitic fault rocks in the cougar creek
complex, OregoneIdaho using the porphyroclast hyperbolic distribution method, and potential
use of SC’-type extensional shear bands as quantitative vorticity indicators. Journal of Structural
Geology, 30, 1005–1012.
Law, R. D., Searle, M. P., & Simpson, R. L. (2004). Strain, deformation temperatures and vorticity
of flow at the top of the greater Himalayan Slab, Everest Massif Tibet. Journal of the Geological
Society, London, 161, 305–320.
Malaroda, R., Carraro, F., Dal Piaz, G. V., Franceschetti, B., Sturani, C., & Zanella, E. (1970).
Carta geologica del Massiccio dell’Argentera alla scala 1:50.000 e note illustrative. Memorie
della Società Geologica Italiana, 9, 557–663.
Strain Softening in a Continental Shear Zone: A Field Guide … 47

Matte, P., & Ribeiro, A. (1975). Forme et orientation de l’ellipsoïde de déformation dans la virgation
hercynienne de Galice. Relations avec le plissement et hypothèses sur la genèse de l’arc ibéro-
armoricain. Comptes Rendus de l’Academie des Sciences de Paris, 280, 2825–2828.
Matte, P. (1986). Tectonics and plate tectonics model for the Variscan belt of Europe. Tectonophysics,
126, 329–374.
Matte P. (2001). The Variscan collage and orogeny (480–290 Ma) and the tectonic definition of the
Armorica microplate: A review. Terra nova, 13, 122–128.
Montomoli, C., Iaccarino, S., Carosi, R., Langone, A., & Visonà, D. (2013). Tectonometamorphic
discontinuities within the Greater Himalayan Sequence in western Nepal (central Himalaya):
Insights on the exhumation of crystalline rocks. Tectonophysics, 608, 1349–1370.
Montomoli, C., Carosi, R., Laccarino, S. (2015). Tectonometamorphic discontinuities in the greater
Himalayan sequence: A local or a regional feature?. In S. Mukherjee, R. Carosi, P.A. van der
Beek, B. K. Mukherjee & D. M. Robinson (Eds.), Tectonics of the Himalaya.(Vol. 412, pp. 25–41)
Geological Society of London Special Publication . https://doi.org/10.1144/sp412.3
Montomoli, C., Iaccarino, S., Simonetti, M., Lezzerini, M., & Carosi, R. (2018). Structural setting,
kinematics and metamorphism in a km-scale shear zone in the Inner Nappes of Sardinia (Italy).
Italian Journal of Geoscience, 137, 294–310.
Mukherjee, S. (2014). In Atlas of shear zone structures in Meso-scale (pp. 124). Springer
International Publishing.
Musumeci, G., & Colombo, F. (2002). Late Visean mylonitic granitoids in the Argentera Massif
(Western Alps): Age and kinematic constraints on the Ferrière-Mollières shear zone. Comptes
Rendus de l’Académie des SciencesSerie II, 334, 213–220.
Oliot, E., Melleton, J., Schneider, J., Corsini, M., Gardien, V., & Rolland, Y. (2015). Variscan crustal
thickening in the Maures-Tanneron massif (South Variscan belt, France): New in situ monazite
U-Th-Pb chemical dating of high-grade rocks. Bulletin de la Société géologique de France, 186,
145–169. https://doi.org/10.2113/gssgfbull.186.2-3.145.
Oriolo, S., Wemmer, K., Oyhantcabal, P., Fossen, H., Schulz, B., & Siegesmund, S. (2018).
Geochronology of shear zones–A review. Earth-Science Reviews, 185, 665–683.
Parson, A. J., Coleman, M. J., Ryan, J. J., Zagorevski, A., Joyce, N. L., Gibson, H. D., & Larson,
K. P. (2018). Structural evolution of a crustal-scale shear zone through a decreasing temperature
regime: The Yukon River shear zone, Yukon-Tanana terrane, Northern Cordillera. Lithosphere
(Vol. 10(6), pp. 760–72).
Passchier, C. W. (1987). Stable position of rigid objects in non-coaxial flow: A study in vorticity
analysis. Journal Structural Geology 9(5/6), 679–690.
Passchier, C. W., Trouw, R. A. J. (2005). Microtectonics (2nd edn., p 101), Berlin, Heidel-
berg:Springer.
Peacock, D. C. P., & Sanderson, D. J. (1991). Displacements, segment linkage and relay ramps in
normal fault zones. Journal of Structural Geology, 13, 721–733.
Piazolo, S., & Passchier, C. W. (2002). Experimental modelling of viscous inclusions in a circular
high-strain ring: implication for the interpretation of shape fabrics and deformed enclaves. Journal
Geophysics Reserach 107, B10, 2242 ETG, 11, 1–15.
Ponce, C., Druguet, E., & Carreras, J. (2013). Development of shear zone-related lozenges in foliated
rocks. Journal of Structural Geology, 50, 176–186.
Ramsay, J. G. (1980). Shear zone geometry: A review. Journal of Structural Geology, 2, 83–99.
Rosenbaum, G., Lister, G. S., & Duboz, C. (2002). Reconstruction of the tectonic evolution
of the western Mediterranean since the Oligocene. In G. Rosenbaum & G. S. Lister (Eds.),
Reconstruction of the Alpine-Himalayan Orogen. Journal of the Virtual Explorer, 8, 107–126.
Rubatto, D., Schaltegger, U., Lombardo, B., Colombo, F., & Compagnoni, R. (2001). Complex
Paleozoic magmatic and metamorphic evolution in the Argentera Massif (Western Alps) resolved
with U-Pb dating. Schweizerische Mineralogische und Petrographische Mitteilungen, 81, 213–
228.
48 M. Simonetti et al.

Rubatto, D., Ferrando, S., Compagnoni, R., & Lombardo, B. (2010). Carboniferous high-pressure
metamorphism of ordovician protoliths in the argentera massif (Italy), Southern European
Variscan belt. Lithos, 116, 65–76. https://doi.org/10.1016/j.lithos.2009.12.013
Sanchez, G., Rolland, Y., Schneider, J., Corsini, M., Oliot, E., Goncalves, P., et al. (2011).
Dating low-temperature deformation by 40Ar/39Ar on white mica, insights from the Argentera-
Mercantour Massif (SW Alps). Lithos, 125, 521–536.
Schneider, J., Corsini, M., Reverso-Peila, A., & Lardeaux, J. M. (2014). Thermal and mechanical
evolution of an orogenic wedge during Variscan collision: An example in the Maures-Tanneron
massif (SE France). Geological Society London, spec. publ., 405, 313–331.
Sibson, R. H. (1977). Fault rocks and fault mechanisms. Journal of the Geological Society
Conditions, 133, 191–213.
Simonetti, M., Carosi, R., & Montomoli, C. (2017). Variscan shear deformation in the Argentera
Massif: A field guide to the excursion in the Pontebernardo Valley (CN, Italy). Atti della Società
Toscana di Scienze Naturali Memorie, Serie A, 124. https://doi.org/10.2424/ASTSN.M.2017.02
Simonetti, M., Carosi, R., Montomoli, C., Langone, A., D’Addario, E., & Mammoliti, E. (2018).
kinematic and geochronological constraints on shear deformation in the Ferriere-Mollières
shear zone (Argentera-Mercantour Massif, Western Alps): Implications for the evolution of the
Southern European Variscan Belt. International Journal of Earth Sciences, 107(6), 2163–2189.
https://doi.org/10.1007/s00531-018-1593-y.
Simonetti, M., Carosi, R., Montomoli, C., Cottle, J.M., & Law, R.D. (2020a). Transpressive defor-
mation in the southern european variscan belt: new insights from the aiguilles rouges massif
(Western Alps). Tectonics, 39 (6). https://doi.org/10.1029/2020TC006153
Simonetti, M., Carosi, R., Montomoli, C., Corsini, M., Petroccia, A., Cottle, J. M., & Iaccarino,
S. (2020b). Timing and kinematics of flow in a transpressive dextral shear zone, Maures Massif
(Southern France). International Journal of Earth Science. 109, 2261–2285. https://doi.org/10.
1007/s00531-020-01898-6
Soliva, R., & Benedicto, A. (2004). A linkage criterion for segmented normal faults. Journal of
Structural Geology, 26, 2251–2267. https://doi.org/10.1016/j.jsg.2004.06.008.
Stampfli, G. M., von Raumer, L. F., & Borel, G. D. (2002). Paleozoic evolution of pre-Variscan
terranes: from Gondwana to the Variscan collision. Geol S Am S, 364, 263–280.
Stipp, M., Stunitz, H., Heilbronner, R., & Schmid, S. M. (2002). The eastern Tonale fault zone: a
“natural laboratory” for crystal plastic deformation of quartz over a temperature range from 250
to 700° C. Journal of Structural Geology, 24, 1861–1884.
Tollmann, A. (1982). Großraumiger variszischer Deckenbau im Moldanubikum und neue Gedanken
zum Variszikum Europas. Geotektonische Forschungen, 64, 1–91.
Viegas, L. G. F., Archanjo, C. J., Hollanda, M. H. B. M., & Vauchez, A. (2014). Microfabrics and
zircon U-Pb (SHRIMP) chronology of mylonites fromthe Patos shear zone (Borborema Province,
NE Brazil). Precambrian Research, 243, 1–17.
Wallis, S. R., Platt, J. P., & Knott, S. D. (1993). Recognition of syn-convergence extension in
accretionary wedges with examples from Calabrian arc and the Eastern Alps. American Journal
of Sciences, 293, 463–495.
Xypolias, P. (2010). Vorticity analysis in shear zones: A review of methods and applications. Journal
of Structural Geology, 32, 2072–2092.
The Geometry and Kinematics
of the Southwestern Termination
of the Pyrenees: A Field Guide
to the Santo Domingo Anticline

E. L. Pueyo, B. Oliva-Urcia, E. M. Sánchez-Moreno, C. Arenas,


R. Silva-Casal, P. Calvín, P. Santolaria, C. García-Lasanta, C. Oliván,
A. Gil-Imaz, F. Compaired, A. M. Casas, and A. Pocoví

Abstract We introduce a field trip to the southwestern termination of the Pyrenean


sole thrust: the Santo Domingo anticline. The field trip is articulated in three main
stops with panoramic views. We pursue to emphasize some outstanding characteris-
tics of this structure: (A) the large-scale progressive (laterally angular) unconformity
that crops out in its southern flank, fully records its folding kinematics and witnesses
for two distinct deformation periods in relation to the basement activity (Gavarnie
gentle folding and Guarga paroxism) (stop #1). (B) The remarkable conical geometry

E. L. Pueyo (B)
Instituto Geológico y Minero de España, Unidad de Zaragoza. C/Manuel, Lasala 44, 9º, 50006
Saragossa, Spain
e-mail: unaim@igme.es
E. L. Pueyo · C. Arenas · P. Santolaria · A. Gil-Imaz · A. M. Casas · A. Pocoví
Unidad Asociada en Ciencias de la Tierra, IGME-Universidad de Zaragoza, Saragossa, Spain
B. Oliva-Urcia
Dpto. de Geología y Geoquímica, Universidad Autónoma de Madrid, 28049 Madrid, Spain
E. M. Sánchez-Moreno · P. Calvín
Dpto. de Física, Universidad de Burgos, Av/Cantabria, S/N, 09006 Burgos, Spain
C. Arenas · A. Gil-Imaz · A. M. Casas · A. Pocoví
Geotransfer (IUCA), Universidad de Zaragoza, Saragossa, Spain
C. Arenas · R. Silva-Casal · A. M. Casas · A. Pocoví
Instituto Universitario de Investigación en Ciencias Ambientales, Universidad de Zaragoza,
Saragossa, Spain
P. Santolaria
Universitat de Barcelona, Geomodels Research Institute, Barcelona, Spain
C. García-Lasanta
Geology Department, Western Washington University, Bellingham, WA 98225-9080, USA
C. Oliván
Freelance Geologist. C/Duquesa de Villahermosa1, 2º, 22001 Huesca, Spain
F. Compaired
Servicio de Planificación y Gestión Forestal. Dpto. Medio Ambiente. Gobierno de Aragón,
Saragossa, Spain

© Springer Nature Switzerland AG 2021 49


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_3
50 E. L. Pueyo et al.

of the fold termination (San Marzal area) that is completely documented by structural
and palaeomagnetic analyses (stop #2). (C) The rotational kinematics of this struc-
ture recorded in its northern flank. Both the folding and the rotational kinematics can
be wholly time-bracketed by means of seven long magnetostratigraphic sections and
numerous additional palaeomagnetic sites within the syn-tectonic (syn-rotational)
sequence (stop #3). We also describe the overall structure of the Western Pyrenees
from a panoramic view from the summit of the anticline (1520 m). All these stops are
accessible by car (dirt roads). Finally, an additional field stop (#4) is devoted to the
sedimentology of the Lutetian-Bartonian carbonates following a hiking path (3 km)
from stop #3.

1 Introduction

The Southern Pyrenees (Barnolas et al. 2019; Muñoz 2019; Martín-Chivelet et al.
2019), and specifically their western part (Teixell 1992, 1996; Martinez-Peña and
Casas 2003; Casas and Pardo 2004; Millán et al. 2006; Meresse 2010; Labaume and
Teixell 2018, Labaume et al. 2016), have been the target for numerous geological field
trips for academic and industry organizations seeking to understand the sedimentary
architecture of potential reservoirs and to establish tectonic models for hydrocarbon
traps during the last decades. This is due to several reasons: the outstanding present-
day exposure conditions of the rocks favoured by the dry, Mediterranean climatic
conditions (precluding a dense forest cover). In addition to that, the exhumation level
endured by this mountain range during and after its construction adequately exposes
rocks that were buried below ~ 4000–6000 m of sediments. In addition, the endorheic
character of its foreland basin (Ebro), at least, until 11–9 Ma (Garcés et al. 2020),
considerably slowed down the erosional evolution of the landscape (low exhumation
in distal areas but relatively strong in source areas of the mountain belt) (Fitzgerald
et al. 1999).
Another additional and suggestive factor in the Pyrenees is the occurrence of
numerous syntectonic sedimentary records, especially (but not only) during the
Cenozoic (Puigdefàbregas and Souquet 1986). Classic examples of growth strata
were described in the eastern Pyrenees (so-called progressive unconformities after
Riba 1976) but they can be found all along the mountain range. Focusing on the
Southwestern Pyrenees, some classic examples are well-known; Pico del Águila
anticline (Millán et al. 1994), Balzes anticline (Barnolas and Gil-Peña 2001), Sos
del Rey Católico (Turner 1988; Teletzke 2012; Anastasio et al. 2020), etc.… (see
overview of structures in Millán et al. 2000). The opportunity given by syntec-
tonic deposits (growth strata, Suppe et al. 1992) to understand the deformation
kinematics together with the powerfulness of magnetostratigraphic studies to semi-
continuously date sedimentary piles, motivated pioneer studies of these geometries
in the 90’s (Burbank et al. 1992a, b; Bentham 1992; Hogan 1993; Holl and Anas-
tasio 1993; Hogan and Burbank 1996; Bentham and Burbank 1992, 1996; Meigs
1997, etc.). Numerous magnetostratigraphic investigations followed those works
The Geometry and Kinematics of the Southwestern Termination … 51

(Beamud et al. 2003; Oliva-Urcia et al. 2019; Anastasio et al. 2020) and were focused
on both the South Pyrenean Basin and the Ebro foreland Basin (see overviews by
Pérez-Rivarés et al. 2018; Garcés et al. 2020) producing one of the best illustrated
chronostratigraphic frames to understand the source to sink evolution of orogens and
basins.
Therefore, the collision-related geological features and the occurrence of excep-
tional and well-dated syntectonic sequences, located in easily accessible locations
preserved across the Southern Pyrenees, have facilitated increasing interest of the
scientific community to provide a detailed understanding of the geological evolution
of the mountain range. Only in the last five decades, more than a thousand scien-
tific contributions (see overview by Pocoví 2019) have targeted multiple geological
topics in the Pyrenees (e.g. Van der Voo 1969; Van Der Voo and Boessenkool 1973;
Burbank et al. 1992; Choukroune 1992; Puigdefàbregas et al. 1992; Muñoz 1992;
Coney et al. 1996; Friend et al. 1996; García-Ruiz and Valero-Garcés 1998; Vergés
et al. 1998); this scientific production is actively increasing nowadays (e.g. Chevrot
et al. 2018; Marcén et al. 2018; Núñez-Lahuerta et al. 2018; Teixell et al. 2016;
Sancho et al. 2018; Silva-Casal et al. 2019; Uzel et al. 2020).
The Southern Pyrenees are a commonly visited destination for field trips in under-
graduate and graduate academic geology programmes from institutions all around the
world. Within their academic offer, neighbouring universities, such as Universidad
de Zaragoza, Université Paul Sabatier (Toulouse), Université de Pau et des Pays
de l’Adour, Euskal Herriko Unibertsitatea, Universitat de Barcelona and Univer-
sitat Autònoma de Barcelona, include Geology and Earth Sciences programmes that
develop a wide offer of field trips (one-day or two-day visits in most cases) to multiple
Pyrenean locations, in order to instruct students in a wide range of geological topics.
But education-related geology visits to the Pyrenees are not restricted to the closest
high education institutions in Spain and France; in addition, geology field trips in the
Pyrenean Range are currently organized by numerous institutions worldwide, such
as Imperial College of London, University College of London, Universiteit Utrecht,
University of Plymouth, University of Derby, Universitetet i Bergen, University of
Missouri, South Dakota School of Mines and Technology, Karlsruher Institut für
Technologie or Nasjonal Forskerskole i Petroleumsfag (Petroleum Research School
of Norway) among many others. In the past, additional institutions organized peri-
odical Pyrenean field trips as well, including ETH Zurich, University of Leeds and
Royal Holloway University of London. In other cases, education-related geology
field trips to the Pyrenees were specific events in certain college programmes, such
as the ones organized by the University of Wisconsin-Madison in 2001 or the Univer-
sity of Michigan in 2004, as well as part of workshops and periodical training courses
developed by private companies (usually from the oil industry), like NEXT-Oil &
Gas Training and Competency Development, STATOIL, TOTAL, etc.
More than 15 years ago, the geological values of the Southwestern Pyrenees
also motivated the development of outreach activities focused on the general public
as the TransPyrenean Geological Route (www.rutatranspirenaica.com) between
Huesca and Oloron, or the very active Sobrarbe-Pirineos UNESCO Global Geopark
(https://www.geoparquepirineos.com/). Recently, other activities as the PrePyrenean
52 E. L. Pueyo et al.

Geological Train (Riglos sector) as well as several “Geolodays” organized by the


Spanish Geological Society since 2005 (https://geolodia.es/) have been devoted to
the Southern Pyrenean geology. However, most of these field trips have been focusing
on visiting some of the renowned localities, mostly, along the Central Pyrenees, such
as the Ordesa Valley (part of the Ordesa-Monte Perdido National Park), the Aínsa
and Tremp-Graus Basins and the structural features of the South Pyrenean Frontal
Thrust outcropping in Riglos and Pico del Águila areas.
The westernmost termination of the Southern Pyrenean front (west of the Riglos
pinnacles and the Gállego river transect) dominated by the large and complex
Santo Domingo anticline (Puigdefàbregas and Soler 1973; Pocoví et al. 1990)
has received very little attention and remains almost unexplored for outreach and
academic purposes except for some British and Spanish universities that developed
some research activities and academic works in the 90’s (Turner 1990; Turner and
Hancock 1990a, b; Nichols 1987 and 1989; McElroy 1990; Arenas (1993); Millán
1996; Millán Garrido 2006). During the last years, many other studies had faced
diverse research targets: structural geology (Teixell and García-Sansegundo 1995;
Hervouët et al. 2005; Oliva-Urcia et al. 1996 and 2012a; Ramón et al. 2016a), sedi-
mentology (Alegret and Aurell 1999, 2002; Silva-Casal et al. 2019), geophysics
(Calvín et al. 2018), magnetostratigraphy (Oliva-Urcia et al. 2019; Silva-Casal et al.
2019; Anastasio et al. 2020) and palaeomagnetism (Pueyo et al. 2020 and references
therein).
This article introduces a field trip in this portion of the Pyrenees emphasizing
some remarkable characteristic of its structure: (A) a large-scale progressive (laterally
angular) unconformity (growth strata) in its southern flank that fully recorded the
folding kinematics; (B) the remarkable conical closure of the fold termination (early
described by Nichols 1984 and Millán et al. 1992); (C) the rotational kinematics
of this structure (registered in its northern flank); (D) an overall description of the
Western Pyrenees structure from a panoramic view in the summit of the anticline
(1520 m). Finally, an additional field stop is focused on the sedimentology of the
Lutetian carbonates (Guara Fm) recently studied by Silva-Casal (2017).

2 General Geological Setting

We summarize the structure of the SW part of the Range. The Pyrenean Range is
an asymmetric double vergent collisional orogen (Roure et al. 1989; Muñoz 1992;
Labaume et al. 2016; Teixell et al. 2016). This fold and thrust belt developed from
Late Cretaceous to Early Miocene times in relation to the opening of the Atlantic
Ocean and the indentation of the Iberian plate pushed by the African plate under
the Euroasiatic one (Rosenbaum et al. 2002; Sibuet et al. 2004; Vissers and Meijer
2012). This collision developed five different structural zones depending upon their
characteristics (Mattauer and Séguret 1971; Séguret 1972), from north towards south
are: the Aquitanian foreland Basin, the North Pyrenean Zone, the Axial Zone, the
South Pyrenean Zone and the Ebro foreland Basin. Therefore, the Southern Pyrenean
The Geometry and Kinematics of the Southwestern Termination … 53

Zone comprises imbricated south-vergent fold and thrust system affecting Palaeozoic
to Cenozoic rocks (Séguret 1972). They developed in a foreland-breaking sequence
in this SW sector (Martínez-Peña and Casas-Sainz 2003; Teixell 1996, 1998).
The South Pyrenean Zone, which concentrates most of the shortening (between
~120 and 160 km of shortening, Teixell et al. 2016; Beaumont et al. 2000) of the Pyre-
nean Range during the alpine convergence, is located south of the Axial Zone. The
Axial Zone represents the backbone of the Range, where the Palaeozoic rocks crop
out forming an antiformal stack in the central part of the Range, which disappears
towards the west, in the area of interest (Teixell 1996). Within the South Pyrenean
Zone in this SW sector, and towards south, we come across the Internal Sierras,
the turbiditic and the molassic Jaca Basin (the piggy-back basin incorporated to the
orogen as deformation migrated towards south, forming the Jaca-Pamplona Basin)
and the External Sierras, ramping up over the Ebro foreland Basin (Fig. 1) (Labaume
et al. 2016). The Internal Sierras comprises of Late Cretaceous to Palaeogene carbon-
ates with the orogenic WNW-ESE trend. Structurally, they constitute a south-vergent
imbricated fold and thrust system branching in the basal Larra thrust that connects
towards the north with a basement thrust, the Lakora thrust (Teixell 1996). This
Larra thrust system is tilted towards the south by the Gavarnie basement thrust,
which affects, together with the Jaca basement thrust, the cover structures towards
the south in the turbiditic and the molasse basin (Labaume et al. 1985; Teixell 1996,
1998; Oliva-Urcia and Pueyo 2007; Izquierdo-Llavall et al. 2013). According to
recent interpretations of seismic sections and thermochronologic data, < 50 km to
the east of Teixell’s (1996) cross section, Broto and Fiscal basement thrust sheets
affect the cover in the Jaca-Pamplona piggy-back Basin (Labaume et al. 2016).
Finally, the Guarga basement thrust sheet produces the final shortening in the
External Sierras (the southernmost unit) and deforms the northern tip of the Ebro
foreland Basin, which is the focus of this field trip (Pocoví et al. 1990; Millán 1996;
Arenas et al. 2001). The general structure of the External Sierras can be synthetized as
the diachronous emplacement (Lutetian to lower Oligocene, Gavarnie, in age) of an
imbricate thrust system detached on evaporitic facies of Triassic age under submarine
conditions and developing in a foreland-breaking sequence that migrated towards
west (Millán et al. 2000). Later on, the incipient mountain chain was affected by
cover thrusts in a hinterland breaking sequence migrating towards west (Oligocene-
Miocene) but in a more isochronous manner (Millán et al. 2000). The distribution
of the evaporitic Triassic detachment and contractional salt remobilization might,
somehow, influence the evolution of compressional structures in the External Sierras
(Anastasio 1992; Anastasio and Holl 2001; Vidal-Royo et al. 2009; Pueyo et al.
2020).
The kinematic evolution of the South Pyrenean Zone can be determined through
the abundant syntectonic turbiditic and molasse sediments. They seem to point to the
Gavarnie Unit (comprising the basement thrusts from Gavarnie to the south except
Guarga) evolving from Eocene to Oligocene (Priabonian to Chattian, 33–24.6 Ma)
times and producing shortening in the Jaca-Pamplona Basin and the External Sierras.
The Guarga basement thrust caused the latest shortening in the External Sierras
from Chattian to Aquitanian (Oliva-Urcia et al. 2019 and references therein) times.
54

Fig. 1 a Geological map, simplified from Choukroune and Séguret’s (1973) and 1:400,000 geological map from Bureau de Recherches Géologiques et
Minières/Instituto Geológico y Minero de España (BRGM/IGME) maps. The inset map represents the structural units (simplified after 1:400,000 geological
map from BRGM/IGME) Barnolas et al. (2008), in Oliva-Urcia, 2018. b ECORS-Arzacq cross section (Teixell y García-Sansegundo, 1995, Teixell 1996, 1998).
E. L. Pueyo et al.

The southern part is modified after Millán (1996). SPZ: South Pyrenean Zone. NPZ: North Pyrenean Zone
The Geometry and Kinematics of the Southwestern Termination … 55

The deformation observed in this field trip is related to the shortening accommo-
dated in the External Sierras due to both the Gavarnie and Guarga basement thrusts
(Puigdefàbregas 1975; Oliva-Urcia et al. 2019 and references therein).
Structure of the Western External Sierras
Zooming in, the field location is placed in the westernmost sector of the Pyrenean
External Sierras, west of the Gállego river transect and the prominent conglomeratic
pinnacles of Mallos de Riglos. Back in the nineteenth century, Mallada (1878, 1881)
already mentioned the Santo Domingo anticline and range, but before the 1980s, they
draw very little attention except for the geological maps done in the 1950s by the
Geological and Mining Institute of Spain (IGME, Almela and Rios 1950a, 1951a, b).
The Santo Domingo anticline represents the topographic relief shaped by the South
Pyrenean sole thrust in its westernmost termination (Puigdefàbregas 1975; Nichols
1987; Turner 1990; McElroy 1990; Arenas 1993; Millán 1996; Pueyo 2000; Oliva-
Urcia 2000). It is the southwesternmost expression of the Pyrenean fold and thrust
front that can be tracked (around continuously) 250 km along-strike from the Oliana
anticline at the east (Sussman et al. 2004) and delineates the boundary between the
Pyrenees and the Ebro foreland Basin (Figs. 1 and 2).
The Gállego river meridian also represents a clear transition in the structural style
of the External Sierras. To the east, a thin-skinned imbricate fold and thrust system
developed during the Eocene in relation to the emplacement of the Gavarnie basement
thrust sheet, which was also responsible for many outstanding oblique structures (e.g.
Pico del Águila anticline). The system was later beheaded by a second thrusting event
(Guarga) in a break-back sequence during Oligocene-Miocene times. This second
event already affected the upper part of the very thick molasse (Campodarbe Fm)
and the Miocene alluvial fans (Uncastillo Fm), conditioned the change in structural
style and generated a remarkable thrust ramp (20–30°, dipping to the north) that can
be tracked along-strike the geological map (and in seismic sections; Labaume and
Teixell 2018) exceeding hundred km, until the Balzes anticline to the east (Rodríguez-
Pintó et al. 2013 and 2016) and beyond (Martínez-Peña and Casas 2003; Santolaria
et al. 2020).
This scenario drastically changes to the west of the Gállego river in our target
area. There, the thrust ramp laterally and progressively evolves to a tight, high-
amplitude (5 km) meso-scale fold, the Santo Domingo detachment anticline, that
finally vanishes to the west in the San Marzal periclinal closure, where the Middle
Eocene limestones disappear. The latter shows a strongly plunging termination to the
west (Nichols 1984; Oliva-Urcia et al. 1996, 2012a; Pueyo et al. 2020), which has
been related to a large-scale conical geometry (Millán 1996; Pueyo 2000; Pueyo et al.
2017a). The evaporitic nature of the detachment level (Keuper and M2 facies) has
been of key importance to delineate the transition between the ramp anticline and the
detachment folding because of a conspicuous negative gravimetric anomaly (Calvín
et al. 2018; Pueyo et al. 2020). The structural anisotropy of the Santo Domingo
anticline can be followed for many km westwards (Tafalla anticline; Puigdefábregas
1975) affecting an even higher thickness of the Campodarbe Fm in that region (Oliva-
Urcia et al. 2012a).
56

Fig. 2 Geological map of the Western External Sierras (Southern Pyrenees). UTM coordinates (30 N datum ETRS89) Source modified from the GEODE
harmonized geological map of Spain; Ebro and Pyrenees (Robador et al. 2011, 2019). The location of the cross sections (in red) in Fig. 3 is also shown as well
as the magnetostratigraphic profiles (in blue)
E. L. Pueyo et al.
The Geometry and Kinematics of the Southwestern Termination … 57

Another remarkable feature in the Western External Sierras is the occurrence of


tête plongeantes (synform anticline) structures; the Riglos, Punta Común and San
Felices imbricate thrust sheets crop out in the southern limb of the Santo Domingo
anticline. They have been linked to the westernmost oblique (and almost ignored)
structures of the External Sierras (La Peña, Fachar and Peña Ronquillo anticlines)
that crop out in its northern limb (Millán et al. 1995). These têtes plongeantes
were formerly part of the westernmost sheets of the imbricate system (coeval to
the Gavarnie stage of basement thrusting in the Axial Zone) and were affected by the
second folding event (coeval to the Guarga stage) and can be only distinguished in
the eastern sector (Salinas Range) due to the overall gentle plunging of these imbri-
cates to the east (Millán 1996; Pueyo et al. 2003a) and the current erosional level.
Along-strike shortening differences across the External Sierras front are well-known
(McElroy 1990; Millán et al. 2000: Huyghe et al. 2009); from >30 km in the Guara
Range to the east to just ~10 km in the San Marzal sector (in between, ~19 km in
the Gallego meridian) (Millán 1996; Pueyo et al. 2004). These differences imply the
accommodation of moderate vertical axis rotations revealed by numerous palaeo-
magnetic studies since the late 1990s (Hogan 1993; Pueyo et al. 1997 and 2020;
Oliva-Urcia et al. 2012a; Pueyo-Anchuela et al. 2012).
Finally, the numerous magnetostratigraphic studies (Hogan and Burbank 1996;
Oliva-Urcia et al. 2016, 2019; Teletzke 2012; Pueyo et al. 2016a; Silva-Casal et al.
2019; Anastasio et al. 2020) in the region (13 sections covering more than 14 km of
mostly syntectonic sequences) offer an outstanding chronostratigraphic frame and
help in accurately dating deformation events and unravelling their kinematics (first
and third stops of this field trip). The deformation related to the emplacement of the
Gavarnie nappe was accommodated in the Western External Sierras cover structures
(Riglos, Punta Común and San Felices imbricate thrust sheets) during Oligocene
times (Rupelian-Chattian; 31.3 and 24.55 Ma) as well as the incipient folding of
the Santo Domingo anticline (Millán et al. 1995; Oliva-Urcia et al. 2019). The San
Felices angular unconformity (Millán et al. 1995, 2000; Arenas et al. 2001) as well
as their western continuation as a progressive unconformity (growth strata) in Luesia
magnetostratigraphic section and La Peña flexure (Turner 1990 and Anastasio et al.
2020) concurs with the Oligocene ages. The onset of the emplacement of the Guarga
nappe, and its effect on the External Sierras, took place right after the emplacement of
the Gavarnie nappe (from 24.55 to 21.2 Ma; Chattian-Aquitanian times, Millán et al.
2000; Teletzke 2012; Oliva-Urcia et al. 2019; Anastasio et al. 2020). This second and
intense deformational pulse largely developed the anticline and folded the former
imbricate thrust sheets (currently visible on the southern flank) (Fig. 3).
Stratigraphy
The oldest rocks cropping out in the Southern Pyrenees and in the External Sierras
are Triassic in age: Muschelkalk facies (the upper M3-dolostomes and the lower
M2-gypsum and clays; López-Gómez et al. 2019). These rocks are found in the core
of the anticline together with the silty and gypsiferous Keuper facies. These Triassic
sequences acted as the regional detachment level. Some remnants of Jurassic rocks
(Barbed et al. 1988; Lobato and Meléndez1988) are only found in the recumbent and
58

Fig. 3 Serial cross sections in the western sector of the External Sierras (see location in Fig. 2). Cross sections by Millán (1996); San Marzal, San Felices
and Gállego. Sangüesa, Sos-Undués and Isuerre by Oliva-Urcia (2000) and Oliva-Urcia et al. (2012a). Cross sections San Marzal, San Felices (Millán 1996)
are also reinterpreted after the gravity modelling (Calvín et al. 2018) IT WILL BE POSSIBLE TO SPLIT THIS FIGURE IN TWO PAGES???, THEN THEY
SHOULD BE VERTICAL
E. L. Pueyo et al.
The Geometry and Kinematics of the Southwestern Termination … 59

southernmost unit (San Felices). A non-depositional hiatus presumably precludes the


occurrence of Lower Cretaceous rocks all along the External Sierras (Millán et al.
1994; Teixell et al. 2009; García-Sansegundo et al. 2009-MAGNA reports).
Above the Triassic and scarce Jurassic rocks very thin transitional and bioclastic
Upper Cretaceous ocher calcarenites and siltstones can be found (Adraen-Bona
Fm; ≈ 70–85 Ma, <50 m thick; Arqued et al. 1986; Alegret and Aurell 1999) as
well as the Tremp Fm (Mey et al. 1968) typically named Garumnian facies and
made of continental red sandstones and siltstones that encompass the Cretaceous-
Cenozoic boundary (ca. 50 m thick). In general, the Late Cretaceous rocks are poorly
represented but display some transitional levels enriched in marine fossil vertebrates
(Osteichthyes Chondrichthyes and Dinosauria). A tentative correlation considering
short magnetostratigraphic sections (Pueyo et al. 2016a) proposes the occurrence of
chron C32n (Campanian) at the base of the sequence, a hiatus during part of the
Maastrichtian and the recording of chrons C29r, C29n and C28r (Maastrichtian and
Danian) in the Garumn facies.
The limestones of the Guara Fm (Lutetian in age, ~40–49 Ma and <100–150 m
thick in the Western External Sierras), one of the more distinct lithologies all along
the External Sierras, are found above the Tremp Fm after another non-depositional or
erosional hiatus. These grey light limestones are mainly composed of benthic forams
(Nummulites, Alveolina, Orbitolites, acervulinids, among others) and calcareous
algae, deposited in a shallow marine carbonate ramp (Silva-Casal 2017), and are
usually found at the highest topographic ridges of the range. The abundance and
diversity of larger foraminifera biomarkers in this unit, in combination with magne-
tostratigraphic data (Rodríguez-Pintó et al. 2012 and 2013b; Silva-Casal et al. 2019),
made this area of great importance to calibrate age of Lutetian Shallow Benthic Zones
(SBZ) (Samsó et al. 1994; Serra-Kiel et al. 1998; Rodriguez-Pintó et al. 2012; Silva-
Casal et al. 2020) and develop a well-supported chronostratigraphic framework for
the Guara Fm. The base of this unit is the lower Lutetian or SBZ 13 (C21r) in the
central part of the External Sierras, and it extends into the Upper Lutetian (SBZ 16,
C19) on the Isuela section (Rodriguez-Pintó et al. 2012, Rodriguez-Pintó et al. 2019,
Silva-Casal et al. 2020). The base of this unit is younger westwards and pinches out
in the middle of the Santo Domingo sector. On its westernmost outcrops, the Guara
Fm is constrained to the Late Lutetian (SBZ 15–16, C19r; Silva-Casal et al. 2019). In
the Santo Domingo sector, above the limestones of the Guara Fm another limestone
unit, the Santo Domingo Mb (Silva-Casal et al. 2019), represents the lower part of
Arguis Fm (Mangin 1958; Puigdefàbregas 1975, see description below) in the area.
Between the Guara Fm and the Santo Domingo Mb, a very conspicuous unconfor-
mity is observed, represented by an erosive surface overlaid by glauconite-rich facies
(Silva-Casal et al. 2019). The abundance of glauconite is representative of the lower
part or the Santo Domingo Mb, and it is observed throughout the outcrops of this
unit in the southern flank of the Santo Domingo anticline. In the westernmost Santo
Domingo sector (the Campo Fenero and San Marzal sections), this unit is located
above the Tremp Fm (Silva-Casal 2017). The Santo Domingo limestone constituents
show a greater abundance of bryozoans, bivalves, echinoderms and a lower variety of
larger foraminifera, although ortofragminids (e.g. Discocyclina) are more common
60 E. L. Pueyo et al.

than in the Guara Fm. The maximum thickness of this unit is ~100 m in the Campo
Fenero section (stop #4) and it exhibits a lateral facies change to the blue marls
of the Arguis Fm, tapering to the west and pinching out around the Gállego river
(Silva-Casal 2017; Silva-Casal et al. 2019).
Right atop the limestones of both Guara Fm and Santo Domingo Mb, and often
located in hung hollows, the Bartonian-Priabonian “blue marls” of the Arguis Fm
(~36–40 Ma) crop out. This unit is mainly composed of mudstones and marls
deposited in a prodelta and outer ramp setting. In the Santo Domingo sector, the
Arguis Fm is characterized by an initial limestone member (Santo Domingo Mb,
described above), followed by a homogenous marly succession. To the east, this
monotonous marly succession is intercalated with bioclastic facies deposited in
middle to outer ramp environments (Millán et al. 1994, Morsilli et al. 2012). In
contrast to the thick sequence in the homonymous locality to the east (>1000 m;
Millán et al. 1994; Hogan and Burbank 1996; Kodama et al. 2010), the Arguis marls
hardly reach 300 m in the Santo Domingo sector. Prominent, but very discontin-
uous and laterally-changing, reef levels often with a great geomorphic expression
are observed at the top of the Arguis sequence in this sector, just below the tran-
sitional and densely laminated sandstone of the Yeste-Arrés Fm (Puigdefàbregas
1975). Equivalent coral build-ups were described in the upper part of the Arguis Fm
by Morsilli et al. (2012) to the east (close to the Rasal location). These deposits were
interpreted as mesophotic build-ups, developed in a prodelta setting. The Arguis Fm
ranges from the Upper Lutetian (C19N, Silva-Casal et al. 2019) to the Lower Priabo-
nian (C16n.2n. Oliva-Urcia et al. 2019). The rapid closing of the Atlantic gateway
led to the overall and sudden continental deposition in the entire Ebro Basin (Costa
et al. 2010; Garcés et al. 2020) during this period (C16n.2n).
The Campodarbe Fm was first defined by Puidefàbregas (1975) as a complex
unit mostly formed by alternating sandstones and siltstones of red and brownish
colours and scarce conglomerates, of alluvial origin, corresponding mainly to large
fluvial systems west-northwest flowing and alluvial fans from the north. This unit
also contains lacustrine-palustrine deposits to the west and passes laterally into (and
overlies) the deltaic Belsué-Atarés Fm (not outcropping in the Santo Domingo
area). The Campodarbe Fm shows a variable total thickness exceeding 3500 m in
the study area (and more than 5000 m to the west, Puigdefàbregas 1975). These
deposits constitute most part of the outcroping rocks in the Guarga synclinorium,
the molassic Jaca-Pamplona piggy-back basin (north of the External Sierras) and
extend further northwest (Fig. 2). Their age is Priabonian (35.6 Ma; Upper Eocene)
to Chattian (24.6 Ma; Upper Oligocene), as dated by magnetostratigraphy (Oliva-
Urcia et al. 2019). They were named Campodarbe Group by Jolley (1987) and Jolley
and Hogan (1989) and Montes (2009) to include coarse detrital units such as Santa
Orosia, Oroel, Cancias and San Juan de la Peña (the latter named Bernués Fm by
Puifdefàbregas 1975). The unit has been divided into a lower portion and an upper
portion (Puigdefàbregas 1975). The upper one contains the syntectonic unconformity
of the San Felices area. Montes (2009) differentiated 10 third-order sequences that
were grouped into 4 megasequences separated by sedimentary breaks (Fig. 4).
The Geometry and Kinematics of the Southwestern Termination …
61

Fig. 4 Synthetic stratigraphic column and pictures from all stratigraphic formations (outcrops)
62 E. L. Pueyo et al.

The Uncastillo Fm was defined by Soler and Puigdefàbregas (1970) and is equiv-
alent to the Sariñena Fm farther east (Quirantes 1969; Luzón 2005). Both units
are constituted by conglomerates, sandstones and mudstones of alluvial origin,
representing two large fluvial distributary systems (named Luna and Huesca fluvial
systems by Hirst and Nichols 1986), which coexisted with a number of small alluvial
fans (from west to east: Agüero, Murillo de Gállego, Riglos, Linás, Aniés, Nueno,
Salto de Roldán and Vadiello). In all these cases, palaeocurrent directions are directed
north–south. It ranges from Chattian to the base (24.6 Ma) to Aquitanian to the top
(aprox 21 Ma) after magnetostratigraphic dating (Oliva Urcia et al. 2019).
The Uncastillo Fm is at least 1425 m thick in the Luesia area. Its basal contact is an
unconformity over diverse units of the External Sierras strata; to the west, the basal
contact constitutes an apparent conformity with the underlying Campodarbe Fm
strata, evolving to a progressive unconformity to the east. Three genetic stratigraphic
units have been differentiated based on the sedimentary evolution and the presence
of unconformities (Arenas 1993; Arenas and Pardo 1999). The evolution of these
units was controlled by tectonics affecting their source areas, i.e. the Guarga thrust
sheet in Gavarnie thrust unit, the External Sierras formation to east (Arenas et al.
2001; Oliva-Urcia et al. 2019).

3 Field Trip to the Santo Domingo Anticline

The field trip is focused on the Santo Domingo anticline located in the homonymous
range, Sierra de Santo Domingo (in Spanish) and its continuation towards the east
in Sierra de Salinas, both ranges situated west of the Gállego river. The two ranges
(Santo Domingo and Salinas) represent the natural boundary between the Zaragoza
and Huesca provinces (Aragón Autonomous Community). The main reliefs comprise
of the maximum altitude of 1524 m, at the Peña de Santo Domingo summit (Stop
#3), and a minimum altitude of about 700 m in the southernmost sector of the map
(Fig. 2) near the Luesia village. In the central area, there are also outstanding reliefs
such as Puig Moné; 1303 m. (Stop #1). These ranges are diverse botanically and
zoologically as a result of climatic and geological (lithological-structural) contrasts.
Their location between the Mediterranean (Ebro Basin) and the Euro-Siberian and
Atlantic environments (External Sierras) together with the topographic disparities
and north and south gradients provide a wide biodiversity. Thus, the Santo Domingo
and the Salinas ranges are the first-rate natural areas protected by several preserva-
tion legal status; Site of Community Importance (SCI) and Special Protection Area
for wild birds (SPA) both managed by the European Directive of the Natura 2000
network for the biodiversity, habitats and natural heritage conservation. In addition,
it is also a Bearded-vulture nesting and Protection Area, a disappeared species in
most European regions. Lastly, this unique natural area was declared to be a Natural
Protected Landscape by the Government of Aragón in 2015 (Decree 52/2015, BOA
The Geometry and Kinematics of the Southwestern Termination … 63

72, 11767–96) and establishes the link between a series of management and protec-
tion measures (PRUG acronyms in Spanish) embracing the regulations and rules of
the natural heritage, including the points of geological interest.
Points and Locations of Geological Interest (PIG and LIG), geotrails, panoramic
views of Geological Interest, Natural Monuments as well as a water springs (Oliván
et al. 2020) are protected by national and regional laws for geological heritage and
geodiversity preservation. The recent PIGs indexing (in approval process) contains,
among others, nine locations including the western termination of the South Pyrenean
Frontal Thrust, San Marzal-Luzientes pericline, due to its outstanding conical geom-
etry. Moreover, Sánchez-Moreno (2012) and Sánchez-Moreno et al. (2015) proposed
fourteen additional locations to be considered as PIGs throughout the Santo Domingo
and Salinas ranges.

4 Logistics and Directions to the Planned Stops

The nexus city to the Santo Domingo Range is Zaragoza which has a small airport
with few international flights to main European cities (Fig. 5). Madrid and Barcelona
airports are well connected worldwide and commute Zaragoza by a dense schedule
of high-speed trains (less than two hours). Once in Zaragoza, the almost only way
to reach the field area (Luesia village) is by car taking the A-68 motorway (towards
Logroño) to the Gallur village, and then taking the A-127 to Ejea de los Caballeros (a
small city with any possible supply needed for the field trip). Alternatively, Ejea can be
reached through A124 (to Huesca) and A125 from Zuera located in the A23 (directly
north from Zaragoza). From Ejea, road A-1204 leads directly to Luesia, a very small
village with several rural guesthouses, a hostel and a camping for accommodation
as well as some restaurants and bars, a bakery and a small grocery store. Additional
accommodations and services can be found in neighbouring villages (Uncastillo,
Biel, etc.).
To get to stop #1 (Puig Moné summit 42.453709 °N; −0.968023 °W) take a
dirt road (soccer field direction) heading north, up to “la Val” valley and then cross
through a nice beech forest (Fig. 5). The main dirt road leads directly to the Puig
Moné summit (an outlet where a forest observatory is placed).
To get to San Marzal (stop #2; 42.410080 °N; −0.988345 °W), the road A-1202
is to be followed from Luesia to the west (towards Uncastillo), and ~ 500 m from
Luesia, a dirt road heading NW (directions to “Pozo de Pigalo”: a wonderful river
pond and a camping site). From there, the main dirt road heads north following the
Arba de Luesia river (it will be crossed several times) for 13 km. At this point, there
is a junction with a smaller and winding dirt road to the north, while the main road
continues to the east. Following the small road to the north, in 5 km you will find
stop #2; the San Marzal pericline. From here, the conical structure marked by the
limestone strata is surrounded by the dirt road that continues to the NE through the
centre of the anticline in the way to Longás village. In ~4 km the Hermitage of
64 E. L. Pueyo et al.

Fig. 5 Sierras de Santo Domingo y Salinas geographic map. a Logistics. Airports, highways (black
dashed line), high velocity train track (blue dashed line). b Boundaries of the Natural Park (data
from www.magrama.es and cartographic base PNOA aerial photography 2012, water points are
also shown (Oliván et al. 2020). c Geological map with topographic base detailing the directions to
the designed stops in the westernmost area of the Santo Domingo. Stops (red stars) and directions
are shown. The detail of the departure from the Luesia village is also shown. The dirt roads are
marked with red dashed line and the red circles are the key junctions on the road
The Geometry and Kinematics of the Southwestern Termination … 65

Santo Domingo (just attached to the Santo Domingo summit) is located (stop #3;
42.442949 °N; −0.916468 °W), taking the second junction to the right.
An optional visit is also planned to the Campo Fenero stratigraphic section (stop
#4; 42.431119 °N; −0.898484 °W) to the south-east of stop #3. From the Hermitage
of Santo Domingo, follow a mountain trail to the south-east, until an open field named
“Campo Fenero” is reached. The Campo Fenero section is located to the south of
that field, after another short path. The 3 km mountain path between the stop #3 and
the bonus stop is of great natural value.

5 Geological Field Stops

Stop 1: Santo Domingo anticline at the Puig Moné summit (1303 m)


42.453709 °N; −0.968023 °W
Syntectonic unconformity of the Campodarbe Fm (Panoramic view to the east)
On the southern limb of the Santo Domingo anticline, approximately to the north of
the Fuencalderas locality, the upper part of the Campodarbe Fm (Rupelian to Chattian,
31.3–24.5 Ma) is involved in a complex syntectonic unconformity (Fig. 6), which
has been linked to the emplacement of the San Felices thrust sheet (cover-related
structure of the Gavarnie basement thrust) during the External Sierras formation,
in Oligocene times (Puigdefàbregas 1975; Hogan 1993; Millán-Garrido et al. 1995;
Oliva-Urcia et al. 2016, 2019).
Close to vertical strata of sandstones and mudstones, the unconformity shows an
offlap-onlap geometry implying several successive small angular unconformities and
the overall change from overturned-to normal-polarity strata (Arenas 1993; Millán-
Garrido et al. 1995). The geometric evolution is only visible on a cartographic scale
around the San Miguel de Liso hermitage dirt road taken from Fuencalderas (to the
east of Puig Moné). This geometric wedge-out disposition opens northwestward,
thus increasing its thickness, and passing progressively to an apparent conformity,
where the upper Campodarbe unit is about 2350 m thick (Oliva-Urcia et al. 2019) and
it is laterally equivalent to La Peña flexure to the west (Turner 1990; Anastasio et al.
2020). Despite the complex sedimentary evolution, a general coarsening- followed
by a fining-upward succession has been described (Arenas et al. 2001) where the
finning sequence is related to the decrease of the San Felices thrust sheet activity,
i.e. from 28 to 24.5 Ma. The boundary and grain-size evolution can be seen along
the Luesia-San Marzal dirt road (in the way to stop #2). These deposits correspond
to fluvial systems with constant E-W palaeocurrent directions, sourced in carbonatic
and siliciclastic rocks located eastward (Puigdefàbregas 1975).
The sedimentary record involved in this unconformity has to be assigned to
tectosedimentary unit T3, according to the age provided by Oliva-Urcia et al. (2019).
Tectosedimentary units are a type of genetic units defined in continental deposits of
the Ebro Basin (see Pardo et al. 2004). The contact with the overlying Uncastillo Fm
(unit T4) in the Fuencalderas area is an apparent conformity and is characterized by
66 E. L. Pueyo et al.

Fig. 6 San Felices unconformity: a and b) Aerial photograph from the 50’s (IGN archive http://
fototeca.cnig.es/), b also includes the photointerpretation of the uncornformity, c Panoramic view
from the air looking east to the San Felices unit and recumbent thrust sheets and d geological sketch
by Millán et al. (1995)

a subtle but evident increase in grain size, with occurrence of fine-grained conglom-
erates at the base of T4. This part of the basin corresponds to the middle sector of
the Luna fluvial system, whose proximal apexes were located westward. The vertical
sedimentary evolution T3/T4 is the same as the one observed in the Arba de Luesia
valley.
Progressive unconformity of the Uncastillo Fm (Panoramic view to the west)
On the southern limb of the Santo Domingo anticline, approximately from the locality
of Biel to the west, the contact between the Campodarbe Fm (Jaca Basin) and
the Uncastillo Fm (Ebro Basin) is an apparent conformity (paraconformity). The
boundary between these two units is defined by a sharp textural and lithological
change from alternating sandstone and mudstone strata into dominantly conglomer-
ates with laterally associated sandstone and mudstone strata, near the Luesia locality.
This change coincides with the reorientation of alluvial palaeocurrent directions from
E–W into N–S (Puigdefàbregas 1975; Arenas 1993).
The Geometry and Kinematics of the Southwestern Termination … 67

Fig. 7 Panoramic view from Puig Moné summit of the Uncastillo progressive unconformity (west-
wards) and interpretation (Arenas 1993; Arenas et al. 2001 and this work). Genetic unit chronos-
tratigraphic chart (down right corner); U: Uncastillo Formation; C: Campodarbe Formation. (1)
Absolute age from Pérez-Rivarés et al. (2018) and Oliva-Urcia et al. (2019). (2) Date of the upper
boundary in the basin centre (Pérez-Rivarés et al. 2018)

The Uncastillo Fm strata form a cumulative wedge-out system with the underlying
Campodarbe strata (Fig. 7). This system is characterized by the progressive south-
ward decrease of strata dip (with southward strata-thickness increase), from vertical
to near horizontal along a distance of 6.5 km. A gentle syncline-anticline geometry
exists southward, whose axes parallel to the basin margin, close to the Luesia and
Uncastillo localities.
Three tectosedimentary units, viz. U1, U2 and U3, are differentiated within the
Uncastillo Fm, which are equivalent to units T4 (U1 + U2) and T5 (U3) (Arenas
1993). These units are ~1425 m thick in the Sierra de Luesia (measured at Bañón
section, on the western bank of the Arba de Luesia valley) but can reach ~1500 to
the east in the Fuencalderas section (Oliva-Urcia et al. 2019). Figure 7 also presents
the age equivalence among UTS and stratigraphic time scale. All of them consist
of conglomerates in proximal and proximal-middle areas that pass laterally to sand-
stones and mudstones in the middle and in the distal areas. Each unit has a cyclic,
fining-coarsening upward evolution. The boundaries between the three units are posi-
tioned at the change from coarsening to fining upward, which represents the change
from progradation to retrogradation of the alluvial systems, following the corre-
sponding change from increasing to decreasing tectonic activities in the hinterland
(Arenas et al. 2001).
68 E. L. Pueyo et al.

The Luesia fan and the Uncastillo fan are parts of what is known as the Luna fluvial
system (Late Oligocene-Miocene). This system had a radial pattern and extended
40–50 km south of the basin margin. The lithological composition of conglomerates
and sandstones indicates that the drainage area extended throughout the Jaca Basin,
Internal Sierras and part of the Axial Zone, thus being much larger than that of the
marginal fans that formed along the External Sierras, to the east of Fuencalderas.
The last panoramic view from stop #1 points to the north, towards the relief
caused by the Santo Domingo anticline (Fig. 8), which sharply ends up to the west
(San Marzal area, stop #2). To the east, the Santo Domingo Mb limestones help
to delineate the location of the southern limb of the anticline. The summit of the
range (1524 m) is also visible from Puig Moné. The prominent relief of the Internal
Sierras stands out on the background with peaks close to 3000 m (Bisaurín: 2670 m;
Collarada: 2883 m; Aspe: 2640 m).
Seven long magnetostratigraphic sections (~12,000 m investigated in total) in the
region allowed to constrain the complete kinematics of the Santo Domingo anticline;
Agüero (180 m), Ayerbe (500 m), San Felices (1300 m) and Salinas (4700 m) by
Hogan (1993) and Hogan and Burbank (1996); Luesia (4000 m) by Oliva-Urcia et al.
(2016), Fuencalderas (1200 m) by Oliva-Urcia et al. (2019) and Sos del Rey Católico
(La Peña) Sect. (3400 m) by Anastasio et al. (2020). This enviable chronostratigraphic
frame provided a very accurate timing of the deformation of the Santo Domingo
anticline (Fig. 9).
Apart from some noise in the dipping angles at the base of the profile (Luesia and
San Marzal sections), a first stage of folding (syntectonic Campodarbe) can be defined
from Rupelian to Chattian times (32.5–24 Ma). In this period, a gentle folding (just
~30° of tilting at ~3.5° Ma−1 rate) occurred in relation to the westwards progression of
the Gavarnie thrust in the External Sierras imbricate cover thrust system (blue colour
in Fig. 9). There is a second (syntectonic Uncastillo), and much more significant
folding period right after and coeval with the Guarga thrust evolution that affected
the cover units (24–21 Ma). In this period ~70° of tilting were acquired in <3 Ma
(~23° Ma−1 , red colour band on Fig. 9). These data slightly differ from those derived
from the Sos magnetostratigraphic section (Anastasio et al. 2020) across the La Peña
flexure (Turner 1990) but the variability could be associated with the diachronic
emplacement (always younger westwards) of the Gavarnie thrust. In any case, the
deformation rates are comparable; the folding pace deduced in La Peña flexure (~3°
and 35° Ma−1 ) by Anastasio et al. (2020) and those from the Luesia-Fuencaldera
sections display two distinct folding rates differing in one order of magnitude.
Stop 2: Fold termination at San Marzal

42.410080 °N; −0.988345 °W


The sharp termination of the Santo Domingo anticline represents a major landscape
feature (Figs. 8 and 10). The Santo Domingo limestones (near the core of the anti-
cline), although very thin (50–100 m) in this sector, help delineating the anticline
geometry. Abundant (>500) bedding data at all along the structure allow character-
izing the fold axis (Fig. 11). It trends N125E, slightly oblique to the main Pyrenean
The Geometry and Kinematics of the Southwestern Termination …

Fig. 8 Panoramic view of the Santo Domingo anticline and the Pyrenees ( modified from Mukherjee 2015, 2021)
69
70 E. L. Pueyo et al.

Fig. 9 Santo Domingo


anticline kinematics (dip
angle versus age [Ma]) from
Luesia and Fuencalderas
magnetostratigraphic
sections (data reworked from
Oliva-Urcia et al. 2016 and
2019, respectively) visible
from panoramic view to the
west. The chronology of
genetic units (UTS) is also
shown

trend and shows an apparent cylindrical fitting according to the eigenvalue approach
by Woodcock (1977). One significant feature is the occurrence of overturned beds in
both limbs of the anticline, an unexpected behaviour considering the general southern
vergence of the anticline. This fact has been related to salt movements at the end
of the folding event (Pueyo et al. 2020) as it was early suggested for many struc-
tures along the External Sierras front (Anastasio 1987, 1992, and Anastasio and Holl
2001). On the other hand, the San Marzal termination displays a strong plunge to the
west (305, 67) in contrasts with the pseudo-horizontal trend of the main axis (305,
05) (Fig. 11).
This plunge was described by Nichols (1984, 1987) and attracted the attention of
other researchers (Turner and Hancock 1990a; Millán et al. 1992; Pueyo 2000; Oliva-
Urcia et al. 1996, 2012a). Despite the apparent cylindrical fitting, this geometry has
been explained as a large-scale conical geometry with parallel flanks (Millán et al.
1992; Pueyo et al. 2017a). As we will show later, the main reason for the conical
geometry of the San Marzal termination is the need to accommodate vertical axis
rotations, only possible with a conical geometry but not with a cylindrical one.
First evidence comes from the tension gashes measured all around the termination
(Fig. 12), that expectedly should be perpendicular to the fold axis orientation (poles
The Geometry and Kinematics of the Southwestern Termination …

Fig. 10 Panoramic view (eastwards) of the San Marzal pericline ( modified from Mukherjee 2015, 2020)
71
72 E. L. Pueyo et al.

Fig. 11 Bedding poles in stereographic projection from the Santo Domingo anticline and its fold
termination (San Marzal). Overturned beds are plotted in the upper hemisphere and drawn as white
dots (as the palaeomagnetic convention). Cylindrical best-fitting using the eigenvectors is also
shown in both cases. Stereoplots performed using the software Stereonet 6.3.0.X (by Cardozo and
Allmendinger 2013 and Allmendinger et al. 2013)

parallel to it) but they show an apparent clockwise (CW) rotation with respect to
the fold axis. A similar rotational pattern can be observed in the joint systems found
around the pericline (additional and abundant data were provided in Turner and
Hancock 1990a). Faults (Fig. 12) show a dominant strike-slip component (in the
present coordinates) and define a conjugate system with lines intersection nearly
parallel to the fold axis. This means that they probably originally formed as normal
faults (also described by Hervouët et al. (2005) to the east in the San Felices unit)
related to local extension in the outer hinge of the fold and were passively rotated
during subsequent stages. The San Marzal area is an excellent study location to
develop academic training on structural analysis. Several outstanding outcrops occur
along the main dirt road (see, e.g., SM05 in Fig. 12) that surrounds the closure and
connects the villages of Luesia (to the south) and Longas (to the north).
However, the strongest evidence to support the conical geometry is the rotational
pattern found in the San Marzal termination after a profuse palaeomagnetic analysis
(Pueyo et al. 2020 and references therein) that evinces for more than 50° of differential
vertical axis rotation between the fold flanks, although part of them can attributed
to geometrical complexities. In addition, the processing (Fig. 13) carried out using
the VPD software (Ramón et al. 2017) in >60 unpublished sites evenly distributed
around the fold termination allows to observe the vectors, after bedding correction,
displaying a progressive CW rotation from south to north with respect to the expected
reference (Fig. 13). Similar to the palaeomagnetic information, the data derived from
the analysis of the magnetic fabrics (AMS) (Sánchez-Moreno et al. 2013) also show
an identical distribution around the fold closure. This coincidence is due to early
recording and blocking of AMS of a far-field layer parallel shortening (almost coeval
to sedimentation) and its later passive character in the southernmost portions of the
The Geometry and Kinematics of the Southwestern Termination … 73

Fig. 12 (next page): Structural data in the San Marzal pericline (bedding poles, joints, tension
gashes). An example of a specific outcrop, SM05, is also shown (see location in Fig. 13) together
with structural and AMS data in that outcrop. Finally, a schematic 3D model integrating most
observations (including simplified palaeomagnetic vectors) is presented for the entire fold closure
74 E. L. Pueyo et al.

Fig. 13 Palaeomagnetic data (yellow cones) in the San Marzal fold termination. AMS data (blue
cones) are also shown (Sánchez-Moreno et al. 2013). For every sector, the cone axis represents the
mean of the palaeomagnetic vector (or the Kmax axis in case of AMS tensors) and the semiapical
angle the confidence (α 95 by Fisher, 1953). Several sites are averaged out in every sector using the
VPD software. Stereoplots represent the palaeomagnetic means (VPD) before (in situ) and after
restoration (bedding correction) to the palaeohorizontal. Mean bedding poles (S0) are also shown
to characterize the fold axis (see also Fig. 11). The Luzientes thrust is also mapped as well as the
hanging-wall and footwall cut-offs in the Yeste-Arrés Fm (red lines)

Pyrenees and in the Ebro Basin (Larrasoaña et al. 2004; Pueyo-Anchuela et al. 2010
and 2012; Pocoví et al. 2014).
On the other hand, the San Marzal pericline also shows the westernmost evidence
of thrusting in the External Sierras. The Yeste-Arrés sandstones on top of the Arguis
marls, where the topographic depression of Luzientes is located, clearly display the
The Geometry and Kinematics of the Southwestern Termination … 75

hanging-wall and footwall cut-offs of this thrust (red lines in Fig. 13), affecting there-
fore to the entire Bartonian sequence (at least) and attesting for ~1 km of additional
shortening in the Santo Domingo anticline. The Luzientes thrust, here defined, is
the last (outcropping) imbricate of the External Sierras fold and thrust system and
was never mapped before. Apparently, the thrust seems to trend pseudo-parallel to
the main Pyrenean grain in this sector, although its original orientation (pre-conical
folding) is difficult to assess because of the complex restoration needed in these
geometries. Its relationship with the core of the anticline is obscured partially because
of a set of vertical faults in the northern part of the pericline that preclude to stablish
its continuation (or not) with the basal detachment level. In any case, the Luzientes
thrust is clearly linked to the transference of deformation to the Campodarbe unit,
where a complex intra-formational multilayer detachment system was defined and it
is responsible for the Santo Domingo hinge collapses mapped by Oliva et al. (2012).
Restoration of palaeomagnetic vectors (or any linear data) in this kind of geome-
tries is not a simple task and must be carried out considering the kinematics of
the fold. The San Marzal and Santo Domingo western termination inspired analogue
simulations (Ramón et al. 2013) and numerical modelling (Ramón et al. 2012, 2016a,
b) focused on the understanding of this complex geometry as well as on the use of
palaeomagnetic vectors as an auxiliary tool during 3D restoration of non-cylindrical
and non-coaxial structures.
Stop 3: Southwestern Pyrenees at the Santo Domingo summit

42.442949 °N; −0.916468 °W

Panoramic view of the Southwestern Pyrenees Regional Geology


Looking to the north from the Santo Domingo summit (stop #3), the Jaca piggy-
back Basin and the Internal Sierras are noted. This basin forms a part of the South
Pyrenean Zone and it is limited to the north by the Internal Sierras (although some
remains of the Axial Zone can be observed in the background), and to the south by
the External Sierras (Santo Domingo isoclinal anticline in this area).
The sediments that crop out in the Jaca piggy-back Basin are the turbiditic deposits
in the north (Hecho Group, Mutti 1984; Payros et al. 1999; Remacha and Fernández
2003), and the molasse sequence, of fluvial (mainly) and alluvial fan origin (Campo-
darbe and Bernués Fms; Puigdefàbregas 1975) (Fig. 14). The turbiditic deposits
lie on the carbonatic marine platform rocks (Upper Cretaceous-Palaeocene). These
Upper Cretaceous-Palaeocene rocks crop out in the Internal Sierras directly over
the Palaeozoic sequences. They are ~1000 m thick and are incorporated into the
Larra cover thrust system, which connects towards the north with the Lakora base-
ment thrust (Teixell 1996). The Upper Cretaceous-Palaeocene cropping out in the
External Sierras is <60–100 m thick and lay on top of the Triassic (Keuper facies).
These Triassic rocks pinch out towards the west, as registered in the Sangüesa
borehole where the Triassic Buntsandstein facies are directly underneath the Upper
Cretaceous rocks (Lanaja1987; Oliva-Urcia et al. 2012a) in the absence of Keuper
76 E. L. Pueyo et al.

Fig. 14 Panoramic view of the Western Pyrenees from Santo Domingo summit (1520 m) to the
north. Upper part; picture from Santo Domingo summit to the north (Félix Compaired). Lower
part, schematic 3D model (QGIS3) showing the location of the main structural units visible from
the summit ( modified from Arenas et al. 2001). Geological maps from Puigdefàbregas (1975) and
IGME (MAGNA maps). Digital Elevation Model by IGN (https://www.ign.es/web/ign/portal/cbg-
area-cartografia/)

facies. Following the Upper Cretaceous in the External Sierras, the Garumnian conti-
nental facies are present (Tremp Group, Cuevas, 1992). Then follows the Guara Fm
(lateral equivalent to the turbiditic Hecho Group) of Middle Eocene age (Puigde-
fàbregas 1975). The Guara Fm diminishes in thickness from east to west (Millán
1996; Silva-casal et al. 2019 and 2020). On top of the Guara limestones Fm is the
Arguis-Pamplona grey marls Fm, platform and prodelta deposits of Bartonian age,
followed by the deltaic and marsh deposits (Belsué-Atarés, Yeste-Arrés and Guen-
dulain Fms of late Eocene (Puigdefàbregas 1975)). The youngest deposits filling
up the piggy-back basin are the Campodarbe and Bernués Fms (Soler and Puigde-
fàbregas 1970; Puigdefàbregas 1975; Montes 1992). The Campodarbe Fm overlies
the previous deltaic rocks. The Campodarbe Fm (Upper Eocene– Lower Oligocene)
is a succession of sandstones and siltstones of red and brownish colours (Soler and
Puigdefàbregas 1970; Puigdefàbregas 1975) of 4500–7000 m thick (Puigdefàbregas
The Geometry and Kinematics of the Southwestern Termination … 77

1975). The studied rocks of the Campodarbe Fm belong to tectonosedimentary unit


T3 (Rupelian; Pardo et al. 2004); these continental sediments were deposited in
fluvial systems with a general WNW-flowing direction, mostly sourced on the South
Pyrenean Central Unit. Later on, the provenance of the detrital material switched to
north-coming, alluvial origin (Bernués Fm) in the Jaca piggy-back Basin. Ages of
the Campodarbe Fm in the piggy-back basin are constrained by magnetostratigraphic
studies in the Salinas section (Priabonian-Rupelian) (Hogan and Burbank 1996).
Towards the northeast of the External Sierras, the overall structure of the piggy-
back basin comprises of three main structural domains: (i) to the south, the westwards
continuation of the South Pyrenean Frontal Thrust, resulting in several tight folds
complicated by large-scale hinge collapses and intra-formational detachments; (ii)
a complex monocline that connects the south-dipping marine deposits of the Jaca-
Pamplona turbiditic Basin to the continental deposits of the Jaca-Pamplona molasse
Basin and the Ebro Basin; (iii) a structural high with outcrops of Upper Cretaceous
involved in the Leyre-Illón cover thrust system.
The southernmost zone (so-called Guarga synclinorium, Figs. 2, 3) presents broad
synclines in the northern border of the area (Bailo syncline), tight anticlines in its
central part (Botaya anticline), minor folds to the south, and the anticline separating
the Jaca-Pamplona Basin from the Ebro Basin, west of the westernmost outcrop of
the South Pyrenean Frontal Thrust: The Santo Domingo-Tafalla anticline (Tafalla
anticline in Puigdefàbregas 1975) that extends >45 km along trend. To the south of
this fold, there is the Ebro Basin. The contact between the two basins is considered to
be at places the stratigraphic contact between the Campodarbe Fm (Jaca-Pamplona
Basin) and the Uncastillo Fm (Ebro Basin) (Puigdefàbregas 1975, Arenas1993).
There is a relay of folds and an increasing number of them towards the west of the
molassic piggy-back basin, where the thickness of the Campodarbe Fm increases.
This is consistent with the behaviour of multilayer folding, which depends, among
other parameters, on the number and thickness of the layers and their mechanical
strength (Ramberg 1963; Frehner et al. 2006; Oliva-Urcia et al. 2012a).
A detailed description of the geological structure of the area can be found in Oliva-
Urcia et al. (2012a) (Figs. 3 and 4) and it is as follows: the Santo Domingo-Tafalla
anticline is interpreted to link at depth with a north-verging blind thrust rooted in
the Upper Triassic evaporites. The thrust cuts across the anticline hinge affecting
from Triassic rocks to the lower part of Campodarbe Fm. The amplitude of the
Santo Domingo-Tafalla anticline decreases towards W, in the same direction of the
plunge of the fold axis (Oliva-Urcia et al. 2012a and references therein). At surface,
it appears as a tight fold involving the Campodarbe Fm, with layer-parallel slip. As a
result, hinge collapse occurs (saddle reef structures, Ramsay 1974). Hinge collapses
are typical in multilayer systems when kink-type folds occur (Ramsay 1974). To the
north, the Longás syncline and the Botaya thrust (anticline in hanging wall) show a
southward vergence. The sedimentary wedge composed mainly by the Hecho Group
and the Arguis-Pamplona Fm thickens from the Bailo syncline towards the north.
In the northern limb of the Bailo syncline, marine platform Mesozoic-Palaeogene
rocks crop out, defining an imbricate thrust system (Leire-Illón) rooted in the Upper
Triassic detachment level. These thrust sheets have individual displacements up to ~
78 E. L. Pueyo et al.

2 km, and a general southward vergence, although bi-vergent anticlines can be found
at surface, forming pop-up structures in the hanging walls of thrusts.
The shortening calculated for the piggy-back basin is accounted by the Guarga
thrust sheet (southernmost basement thrust in Fig. 3) and probably, some of it is
related to the Gavarnie basement thrust sheet, since the Leyre-Illón system passes
laterally to folds of Late Eocene-Early Oligocene activity (Teixell 1996). However,
most of the post-Triassic cover thrust slip in the area can be accounted by the base-
ment displacement represented in the cross section of Fig. 3. The displacement of
the basement in this area, contrary to eastern or western areas, cannot be transferred
towards the south, probably due to the lower (and therefore, non-effective) thickness
of a Triassic detachment level (Keuper facies) or to the absence of evaporates in the
Eocene sequences in the Ebro Basin. Then, the folds in the Campodarbe Fm were
the pin line for the Mesozoic-Cenozoic cover, and consequently displacement was
transferred to the north. This interpretation implies that the mechanical behaviour of
the different lithologies strongly affects the structure of the sedimentary cover: struc-
tures in the cover relay along-strike, but shortening is balanced in a non-rotational
area (Oliva-Urcia et al. 2012a).
Kinematics of the vertical axis rotations in the External Sierras
Differences in shortening along-strike within fold and thrust belts are the main
driving cause to produce vertical axis rotations (VAR) (McCaig and McClelland
1992; Allerton 1998; Soto et al. 2006; Sussman et al. 2012). A truly 4D under-
standing of fold and thrust belts and orogenic systems, in a larger scale, must be
necessarily based on the study of this elusive kinematic indicator (Mukherjee 2019).
Rotational kinematics is usually tackled by characterizing the rotation magnitudes
(VAR) by means of palaeomagnetic analysis. Palaeomagnetism is a proven reliable
and accurate way to estimate VARs at different scales and, in particular, at the fold
and thrust belt scale (Norris and Black 1961; Van der Voo and Channel 1980; Pueyo
et al. 2016b; Oliva-Urcia and Pueyo 2019), with numerous data available nowadays
in almost any orogenic region.
Apart from the VAR characterization, the timing of the rotational movements
and the estimation of the velocities of rotation (both, key kinematic variables) are
very scarce (Speranza et al. 1999; Duermeijer et al. 2000; Mattei et al. 2004; see
also Mukherjee and Khonsari 2018 and Mukherjee and Tayade 2019 for review on
various rotation rates of crustal blocks). This is partially due to the difficulty in
combining the occurrence of two factors; the existence of syn-rotational sediments
and the availability of semi-continuous dating records (i.e. magnetostratigraphy). The
Southwestern Pyrenees is a unique natural laboratory to tackle these strict require-
ments and where a complete rotational kinematic record was firstly published for an
individual thrust system (the South Pyrenean sole thrust) at the Pico del Águila area
in the Central External Sierras (Pueyo et al. 2002).
Coming back to the panoramic view, one of the densest palaeomagnetic data sets
in orogenic systems can be observed from the Santo Domingo summit (Fig. 15 and
Table 1). All visible structural units have been sampled and studied for palaeo-
magnetic purposes in the frame of numerous academic (Bentham 1992; Hogan
The Geometry and Kinematics of the Southwestern Termination … 79

Fig. 15 Distribution of palaeomagnetic sites and magnetostratigraphic sections in the Western


Pyrenees (red dots) (IGME-Geode map superposed to Google Earth). Palaeomagnetic rotations
(mean values) in the southwestern Pyrenees (equal-area projection). Data obtained from the Pyre-
nean palaeomagnetic database, Pueyo et al. (2017b). Structural units with and without significant
rotations are split

1993; Meigs 1995; Pueyo 2000; Larrasoaña 2000; Oliva-Urcia 2004; Fernández
2004; Mochales 2011; Pueyo-Anchuela 2012; Teletzke 2012; Rodríguez-Pintó 2013;
Ramón 2013; Beamud 2013; Izquierdo-Llavall 2014; Pérez-Rivarés 2016; Silva-
Casal 2017) and other research works: Ebro Basin (Larrasoaña et al. 2006; Pérez-
Rivarés et al. 2002, 2004, 2016), External Sierras and the Jaca Basin (Hogan and
Burbank 1996; Pueyo et al. 1997; 2003a and b, 2020; Oms et al. 2003; Kodama
et al. 2010; Pueyo-Anchuela et al. 2012; Oliva-Urcia et al. 2012a, 2016 and 2019;
Rodríguez-Pintó et al. 2008, 2012; Silva-Casal et al. 2019; Anastasio et al. 2020),
Internal Sierras (Oliva-Urcia and Pueyo 2007a and 2007b; Oliva-Urcia et al. 2008,
80 E. L. Pueyo et al.

Table 1 Mean palaeomagnetic vectors in different structural units of the Southwestern Pyrenees.
Sites/sections: number of sites (means) and magnetostratigraphic sections considered. Polarity;
magnetic polarity (N: normal, R: reverse) P/S; Primary/secondary (remagnetized). Mean vector
data: Dec (declination) and Inc (inclination) together with Fisher’s (1953) statistical parameters
(α 95 , k and R). VAR: Robust vertical axis rotation mean estimation. A more detailed grouping of
data was performed for the External Sierras. VARs can be significantly larger in individual sites
since means tend to smooth them

2009, 2012b and 2018; Izquierdo Llavall et al. 2015 and 2018), Northwestern Pyre-
nean zone (Oliva-Urcia et al. 2010; Menant et al. 2016; Aubourg et al. 2019;
Izquierdo-Llavall et al. 2020), Pamplona Eocene Basin (Larrasoaña et al. 2003a,
b, c and 2004), as well as the westernmost locations of the Aínsa Oblique Zone
(Bentham and Burbank 1996; Mochales et al. 2010, 2012a and b, 2016; Muñoz et al.
2013; Rodríguez-Pintó et al. 2013a, b, 2016). All in all, more than 550 mean palaeo-
magnetic vectors (several thousand demagnetized rocks) are synthetized in Table 1
(Fig. 15 and Table 1).
Unrotated domains, as expected from the non-rotational convergence of Iberia
and Eurasia from Late Cretaceous to Miocene (Sibuet et al. 2004; Vissers and Meijer
2012), include all magnetostratigraphic studies from the Ebro Basin (see compila-
tions by Pérez-Rivarés et al. 2018 and Garcés et al. 2020), the unrotated Pamplona
Basin (Larrasoaña et al. 2003a) and the data located in the southern limb of the
Santo Domingo anticline (in structural continuity with the Ebro foreland Basin; see
overview by Pueyo et al. 2020). These data can be considered as the local palaeomag-
netic reference (Fig. 15 and Table 1); a vector pointing almost straight northwards
(N358E). Rotated domains comprise the data from the Jaca molasse and turbiditic
Basin (about 10–15° CW rotation), the Internal Sierras and the External Sierras,
except for the westernmost sector (≈20° CW) and the Aínsa Oblique Zone and the
westernmost sector of the External Sierras displaying significant rotations (35° CW).
Beyond the Pico del Aguila complete rotational kinematic control, additional syn-
rotational records have been investigated during the last years to the east, in the Balzes
The Geometry and Kinematics of the Southwestern Termination … 81

(Rodríguez-Pintó et al. 2016) and the Boltaña anticlines (Mochales et al. 2012a) as
well as in the Aínsa Basin (Muñoz et al. 2013). A scientific debate on the isochrony
or diachrony (Pueyo et al. 1997) of the main rotational movement in relation to
the shortening ages of the main South Pyrenean basement units is still open (see
the recent overview by Oliva-Urcia and Pueyo 2019) and the end-members of this
system will potentially shed light on the problem; the Mediano anticline (preliminary
data published by Beamud et al. 2017) to the east and the Santo Domingo anticline
to the west (here introduced).
The panoramic view in stop #3 allows us to observe the Salinas magnetostrati-
graphic section (Hogan 1993) located in the northern limb of the Santo Domingo
anticline (in the road to Sta. Bárbara pass). Furthermore, 25 previous palaeomag-
netic data drilled in the Campodarbe Fm fall along the same section and 20 additional
sites from the marine Bartonian-Priabonian rocks underneath (Pueyo 2000 Pueyo-
Anchuela et al. 2012; Pueyo et al. 2020). This set of results (Fig. 16 and Table 2) allows
us to infer the rotational kinematics of the Santo Domingo anticline, the western-
most structure of the External Sierras front (100 km away from Mediano, 75 km from
Boltaña anticlines). After the stratigraphic correlation of the individual sites (Fig. 15),
we have grouped them according to the magnetic chron age; C17&C16, C15&C13,
C12, C11&C10. These four groups present similar time spans (1–1.3 Ma). On the
other hand, the palaeomagnetic means denote an inclination shallowing (typical in
continental fluvial rocks; Garcés et al. 1996) and moderate-good qualities.
The rotation ages and magnitudes have been plotted (Fig. 17) as well as the
data from other studied structures to the east (Mediano, Boltaña, Balzes and Pico del
Águila anticlines together with the Guara thrust system). In the eastern structures, the
rotation ages, or at least the main pulse, are almost isochronous and took place during
the Bartonian-Priabonian period. However, the Santo Domingo anticline displays a
significant younger age (Upper Rupelian times) and reinforces the diachronism of the
rotational activity along-strike the Pyrenean basal thrust along the External Sierras
front. Specifically, the end of the rotational movement laterally vanishes along-strike
at a rate of ~9 km Ma−1 , considering the distance between the Balzes and Santo
Domingo anticlines. With respect to the onset of rotation, similar magnitudes are
expected if the lateral migration pattern of the deformation along the External Sierras
front is taken into account as well (Millán et al. 2000). Further data in the Mediano
anticline (in progress) may shed light on the other end-member of the system.
Interestingly, the magnitudes of the rotation velocity are very similar among the
better examples; up to 13° Ma−1 in the Balzes anticline (10° Ma−1 in average,
Rodriguez-Pintó et al. 2016), between 2.3–10° Ma−1 in Boltaña (Mochales et al.
2012a), 2–7° Ma−1 ; in the Pico del Águila anticline (Pueyo et al. 2002 and Rodríguez-
Pintó et al. 2008) and between 2–12° in the Santo Domingo anticline with two
distinct periods. This catalogue of complete folding and rotational kinematics will
help understanding the 4D architecture of the External Sierras thrust system and, in
general, the one of the Southwestern Pyrenees.
82

Fig. 16 Palaeomagnetic sites along the Sta. Bárbara road (geological map and orthophoto quad) and the stratigraphic correlation with the Salinas section (Hogan
1993)
E. L. Pueyo et al.
The Geometry and Kinematics of the Southwestern Termination … 83

Table 2 Robust palaeomagnetic vectors in different magnetic chrons along the Salinas and Santa
Bárbara Section. n/N: number of sites used/measured. Mean vector data: Dec (declination) and
Inc (inclination) together with Fisher’s (1953) statistical parameters (α 95 , k and R). Time intervals
(chron duration) also display the error
N N Dec Inc a95 (°) k R Age and error (Ma) Chron
24 27 039 37 6.3 23.2 0.9586 36.9 ± (1.3) C17 & C16
6 6 036 35 20.1 14.5 0.9310 34.4 ± (1) C15 & C13
7 7 029 30 16.9 15.9 0.9373 31.9 ± (1.3) C12
4 5 351 12 13.5 62.8 0.9841 29.6 ± (1) C11 & C10

Fig. 17 Rotation magnitudes versus age in the Southern Pyrenees. Estimating of the rotation
velocity and timing of some oblique structures; Light blue line) Mediano anticline (preliminary
data by Muñoz et al. 2013 and Beamud et al. 2017). Red line) Pico del Águila (data from Pueyo
et al. 2002 and Rodriguez et al. 2008). Blue line) Northern sector of the Guara thrust system
(Pueyo 2000; Pueyo et al. 2003b). Purple line) Balzes anticline (Rodriguez-Pintó et al. 2016, 2020).
Orange line) Boltaña anticline (Mochales et al. 2012a) and the Santo Domingo anticline (green
line) published in this work. Folding kinematics is also shown for the later anticline

Stop 4: The Campo Fenero section

42.431119 °N; −0.898484 °W

Stratigraphy and sedimentology of the middle Eocene Santo Domingo Mb


South from the Santo Domingo summit, the Upper Lutetian-Lower Bartonian lime-
stones constitute a continuous ridge along the southern flank of the Santo Domingo
anticline. This ridge can be easily tracked on aerial photography. This is because both
the underlying continental materials from the Tremp Fm and the overlying marine
84 E. L. Pueyo et al.

Fig. 18 Outcrop of the Santo Domingo Mb in the area of El Portillo (west of stop #3)

marls of the Arguis Fm are usually covered by vegetation, whereas the middle Eocene
marine limestones stand out in the Santo Domingo Range due to the differential
erosion (Fig. 18). This ridge is constituted by two litostratigraphic units, the Guara
Fm and the Santo Domigo Mb, the latter included within the Arguis Fm. The Guara
Fm pinches out towards the centre of the Santo Domingo sector and only the Santo
Domingo Mb crops out westwards (on the mentioned Campo Fenero section and on
San Marzal).
The Santo Domingo Mb can be mapped both along the southern flank of the Santo
Domingo anticline and in the San Felices and Punta Común nappes (outcropping in
the southern flank of the Santo Domingo anticline). The most representative sections
of this unit are La Osqueta (Figs. 19, 20) and Campo Fenero (Fig. 21). The uncon-
formity between the Guara Fm and the Santo Domingo Mb can be observed in La
Osqueta section (Fig. 20) as a complex surface representing the previous carbonate
ramp system (Guara Fm) sunk due to the interaction of eustasy and flexural subsi-
dence (Silva-Casal et al. 2019). Indeed, this tectonic subsidence created the neces-
sary accommodation space for the development of a new carbonate ramp system on
a former foreland area (i.e. the Santo Domingo Mb).
The Campo Fenero section
This section is easily accessible and contains impressive outcrops of the Santo
Domingo Mb that essentially have been used as a reference for the lithostratigraphic
The Geometry and Kinematics of the Southwestern Termination … 85

Fig. 19 Lithostratigraphic units outcropping in La Osqueta section in the southern flank of the
anticline. ( Modified from Silva-Casal 2017) viewed from the south

Fig. 20 Base of the glauconite layer at the lower part of Santo Domingo Mb and unconformity
in La Osqueta section. Outcrop view (a and b) and polished hand sample of the unconformity c
(Modified from Silva-Casal 2017)
86 E. L. Pueyo et al.

Fig. 21 Stratigraphic profile


as in the Campo Fenero
section ( Modified from
Silva-Casal 2017)

unit (Silva-Casal et al. 2019). The absence of Guara Fm rocks in the area precludes
observing the unconformity between both units in this section and here the Santo
Domingo Mb lies on top of the red beds of the Tremp Fm. Instead, it is a charac-
teristic glauconitic-rich interval marking the onset of the carbonate sedimentation
renewal throughout the base of the Santo Domingo Mb. This basal glauconitic layer
rests above the unconformity with Guara Fm in La Osqueta area (Fig. 20), as well
as in many western sections, such as Campo Fenero and San Marzal ones. In the
The Geometry and Kinematics of the Southwestern Termination … 87

Campo Fenero section (Fig. 21), it is constituted by skeletal facies, mainly pack-
stones, with centimetric components such as bryozoans, bivalves, equinoids and
larger foraminifera (Discocyclina). A 5 m-thick interval of Discocyclina rudstone in
a wackestone-mudstone matrix is observed after a covered area. The skeletal compo-
nents and the overall muddy texture of this basal layer on Campo Fenero evidence
oligophotic, middle ramp conditions for the deposit.
At around one-fourth of the Sect. (25 m; Fig. 21), a sharp facies change separates
the glauconitic level from the next interval, constituted by packtones with a large
amount of acervulinid foraminera and echinoid plate fragments. This level ranges
from massive facies at the base to cross-bedded towards the top and represents shallow
subtidal environments. The large concentration of acervulinid foraminifera suggests
vegetated environments, and the cross-bedding to the top points to the influence of
the wave agitation. Above, the carbonatic sedimentation is interrupted by a silici-
clastic interval (38–53 m in Fig. 21), with few bioclasts and more concentration of
organic matter. This part seems to correspond with lagoon-related environments, and
constitutes the top of an overall regressive sequence.
The following succession (53–95 m) with grainstone–packstone texture (Fig. 22)
is evenly stratified and intercalated with cross-bedded and bioturbated intervals.
Acervulinid foraminifera, echinoid plate fragments, bryozoans and occasionally
larger foraminiera (e.g. Operculina) can be observed here. To the top, the most
common components are articulated coralline algae, which constitute the uppermost

Fig. 22 Cross-bedding in shallow marine facies, upper part of Campo Fenero section
88 E. L. Pueyo et al.

grainstone almost entirely. The facies described in this interval were deposited in an
inner ramp environment with swell influence and covered at some degree by subma-
rine vegetation, as the abundance of acervulinid foraminifera and articulated coralline
algae suggests (Silva-Casal 2017). Despite the relatively homogeneous upper part
of the Campo Fenero section, the presence of tractive structures towards the top
suggests an upwards trend to shallower conditions.
On the very top, the boundary between the limestones of Santo Domingo Mb
and the blue marls of the Arguis Fm (typical crumbly aspect in outcrop) is not
observed in this section, although it is conformable in the San Marzal section, and
interpreted as a progressive deepening of the carbonate ramp in the area of Santo
Domingo. Larger foraminifera within the Santo Domingo Mb are scarce and the
only biostratigraphic data in the unit came from this specific section (a single sample
containing N. biarritzensis and N. beaumonti) allowing to interpret a Bartonian age
(SBZ 17 in Serra-Kiel et al. 1998). Photodependent larger foraminifera (i.e. Alve-
olina, Orbitolites) are characteristically scarce along the Santo Domingo Mb facies,
particularly in the shallower domains. This change in carbonate generation is prob-
ably related to a nutrient-rich environment associated to dominantly deltaic deposits
in the South Pyrenean Basin during the Bartonian (Silva-Casal 2017). The Santo
Domingo Mb represents the last stage in the development of the peripheral shallow
marine carbonate environments from the Jaca-Pamplona foreland Basin, coeval with
the beginning of the foreland basin overfill (Silva-Casal 2017).

6 Conclusions

In this book chapter, we focus on the western edge of the External Sierras (Southern
Pyrenees) as an outstanding example of complete fold kinematics in a complex fold
and thrust system termination, proposing a field trip with 4 stops. The Santo Domingo
anticline evolution is fully recorded by syntectonic sedimentation (both limb-tilting
and rotational) allowing for a fully reconstruction of its deformational history.
Folding kinematics is witnessed by two different unconformities: (1) the San
Felices complex angular unconformity (passing gradually to the west to a progressive
one; La Peña flexure) attests for the onset of folding of the Santo Domingo anticline
in Priabonian times in association to the late emergence of the Gavarnie thrust in this
sector of the foreland. (2) Just atop, the Uncastillo progressive unconformity allows
the dating of the main folding event that took place during Miocene in relation to the
emplacement of the Guarga thrust sheet in the cover units.
On the other hand, the conical geometry characterized by an abrupt fold termina-
tion of the anticline to the west (San Marzal pericline) was induced by a significant
clockwise rotational movement of the basal thrust sheet (determined by a vast palaeo-
magnetic dataset in the region). (3) The Santa Bárbara section (in the northern limb
of the anticline) has recorded the rotational kinematics of this movement during
Rupelian times, younger than other oblique structures (Boltaña, Balzes, Pico del
Aguila) in the Eastern External Sierras (Bartonian-Priabonian).
The Geometry and Kinematics of the Southwestern Termination … 89

The three syntectonic records are fully dated by long magnetostratigraphic


sections (Luesia, Fuencalderas, Salinas and Sos profiles) that allow for an accu-
rate dating of the deformation to be achieved. The field trip, centred in Luesia village
(southern limb of the anticline) proposes three stops with panoramic views of these
unique records. Besides, a fourth stop is devoted to the sedimentology of the recently
defined Santo Domingo Mb of the Arguis Fm.

Acknowledgements This research was supported by the projects CGL2006-02289,


CGL200914214 (Pmag3Drest) CGL2014-54118-C2-2-R (DR3AM), CGL2017-90632-REDT
(MABIGER Network), PRX17/00462 and PID2019-104693GB-I00/CTA (UKRIA4D) by the
Spanish Ministry of Science and Universities and the support given by the Applied Geology
and Geotransfer Research Groups (GeoAP-E0117R) by the Aragón Government. This work also
focuses on the southern sector of the target area of the GeoERA project 3DGeoEU (ERANET
Cofund action 731166 [H2020], Project code GeoE.171.005) and it is also acknowledged. We
are very grateful for the thorough corrections and improvements of Gonzalo Pardo as well as
those by Soumyajit Mukherjee who reviewed and edited this article. Dutta and Mukherjee (2021)
encapsulate this work.

References

Alegret, L., & Aurell, M. (1999). La sedimentación carbonatada en el Prepirineo Aragonés durante
el Cretácico Superior. Estudios Geol, 55, 237–246.
Alegret, L., & Aurell, M. (2002). Facies analysis and sequence stratigraphy of an Upper Cretaceous
carbonate platform (western South-Pyrenean Basin, Spain). N. Jb. Geol. Paläont. Abh, 226(1),
25–41.
Allerton, S. (1998). Geometry and kinematics of vertical-axis rotations in fold and thrust belts.
Tectonophysics, 299, 15–30.
Allmendinger, R. W., Cardozo, N. C., & Fisher, D. (2013). In Structural geology algorithms: Vectors
and tensors (pp. 289). Cambridge: Cambridge University Press.
Almela, A., & Ríos, J. M. (1950a). Mapa Geol. de España 1:50.000, serie antigua, hoja nº 247
(Ayerbe). I.G.M.E. (Ed.), (pp. 50). Madrid: 1 mapa.
Almela, A., & Ríos, J. M. (1951a). Mapa Geol. de España 1:50.000, serie antigua, hoja nº 248
(Apiés). I.G.M.E. (Ed.), (pp. 94). Madrid: 1 mapa.
Almela, A., & Ríos, J. M. (1951b). Estudio geológico de la zona subpirenaica aragonesa y sus
sierras marginales. I Congr. Int. del Pirineo. Inst. de Est. Pirenaicos. Geología, 3, 327–350.
Anastasio, D. J. (1987). Thrusting, halotectonics and sedimentation in the external Sierra, southern
Pyrenees, Spain. PhD Thesis (pp. 247). Baltimore, MD, United States: The Johns Hopkins
University.
Anastasio, D. J. (1992). Structural evolution of the External Sierra, Southern Pyrenees, Spain. In
S. Mitra & G.W. Fisher (Ed.), Structural geology of fold and thrust belts (pp. 239–251). Johns
Hopkins University Press.
Anastasio, D. J., & Holl, J. E. (2001). Transverse fold evolution in the external Sierra, southern
Pyrenees, Spain. Journal of Structural Geology, 23, 379–392.
Anastasio, D. J., Teletzke, A. L., Kodama, K. P., Parés, J. M., & Gunderson, K. L. (2020). Geologic
evolution of the Peña flexure, southwestern pyrenees mountain front Spain. Journal of Structural
Geology, 131, 103969.
Arenas, C. (1993). Sedimentología y paleogeografía del Terciario del margen pirenaico y sector
central de la Cuenca del Ebro (zona aragonesa occidental).Tesis Doctoral (pp. 858). Universidad
de Zaragoza
90 E. L. Pueyo et al.

Arenas, C., & Pardo, G. (1999). Latest Oligocene-late Miocene lacustrine systems of the north-
central part of the Ebro Basin (Spain): Sedimentary facies models and paleogeographic synthesis.
Paleogeogr. Paleoclim. Paleoecol., 151, 127–148.
Arenas, C., Millán, H., Pardo, G., & Pocoví, A. (2001). Ebro Basin continental sedimentation
associated with late compresional pyrenean tectonics (NE Iberia): Controls on margin fans and
alluvial systems. Basin Research, 13, 65–89.
Arqued, Y., Almunia, A., & Ortiga, M. (1986). Sedimentación carbonatada de plataforma durante
el Cretácico Superior en el sector oriental del Prepirineo aragonés. XI Congreso Español de
Sedimentología. GES Barcelona, Resumen de comunicaciones: 14.
Aubourg, C., Jackson, M., Ducoux, M., & Mansour, M. (2019). Magnetite-out and pyrrhotite-in
temperatures in shales and slates. Terra Nova, 31(6), 534–539.
Barbed, F., Martínez, M. B., Millán, H., Navarro, J. J., & y Pocoví, A. (1988). Observaciones
sobre la geometría de la “klippe” de San Felices (extremo occidental de las Sierras Exteriores
del Prepirineo meridional). In Symposium on the geology of the pyrennes and betics. Abstr: (Vol.
71). Barcelona.
Barnolas, A., & Gil-Peña, I. (2001). Ejemplos de relleno sedimentario multiepisódico en una cuenca
de antepaís fragmentada: La Cuenca Surpirenaica. Bol. Geol. Mine., 112(3), 17–38.
Barnolas, A., Gil-Peña, I., Alfageme, S., Ternet, Y., Baudin, T., & Laumonier, B. (2008). Mapa
geológico de los Pirineos a escala 1:400 000. IGME/BRGM ISBN: 978–2–7159-2168-9.
Barnolas, A., Larrasoaña, J. C., Pujalte, V., Schmitz, B., Sierro, F. J., Mata, M. P. et al. (2019).
Alpine foreland basins. In The geology of iberia: A geodynamic approach (pp. 7–59). Cham:
Springer
Beamud, E. (2013). Paleomagnetism and thermochronology in tertiary systectonic sediments of the
south-central pyrenees: chronostratography, kinematic and exhumation constraints. Unpublished
PhD (pp. 250). University of Barcelona
Beamud, B., Garcés, M., Cabrera, Ll, Muñoz, J. A., & Almar, Y. (2003). A new middle to late
Eocene continental chronostratigraphy from NE Spain. Earth and Planetary Science Letters,
216, 501–514.
Beamud, E., Pueyo, E.L., Muñoz, J.A., Valero, L., & Granado, P. (2017). Paleomagnetic constraints
on the kinematics of the Mediano Anticline (Aínsa Basin). Preliminary results. In X. Magiber
(Ed.), IUCA-University of Zaragoza. ISBN: 978–84-16723-40-9
Bentham, P. A. (1992). The tectono-stratigraphic development of the western oblique ramp of the
south-central Pyrenean thrust system (p. 253). Northern Spain. Ph.D: University of Southern
California.
Bentham, P., & Burbank, D. W. (1996). Chronology of Eocene foreland basin evolution along the
western oblique margin of the South-Central Pyrenees. In P. F. Friend & C. J. Dabrio (Ed.),
Tertiary basind of Spain (pp. 144–152). Cambridge University Press.
Bentham, P., Burbank, D. W., & Puigdefábregas, C. (1992). Temporal and spatial controls on the
alluvial architecture of an axial drainage system: Late Eocene Escanilla Formation, southern
Pyrenean foreland basin, Spain. Basin Research, 4, 335–352.
Beaumont, C., Muñoz, J. A., Hamilton, J., & Fullsack, P. (2000). Factors controlling the alpine
evolution of the central pyrenees inferred from a comparison of observations and geodynamical
models. Journal of Geophysical Research: Solid Earth, 105(B4), 8121–8145.
Burbank, D. W., Puigdefabregas, C., & Munoz, J. A. (1992a). The chronology of the Eocene
tectonic and stratigraphic development of the eastern Pyrenean foreland basin, Northeast Spain.
Geological Society of America Bulletin, 104(9), 1101–1120.
Burbank, D. W., Vergés, J., Muñoz, J. A., & Bentham, P. (1992b). Coeval hindward- and forward-
imbricating thrusting in the South-Central Pyrenees, Spain: Timing and rates of shortening and
deposition. Geological Society of America Bulletin, 104, 3–17.
Calvín, P., Santolaria, P., Casas, A. M., & Pueyo, E. L. (2018). Detachment fold versus ramp
anticline: A gravity survey in the southern Pyrenean front (External Sierras). Geological Journal,
53(1), 178–190.
The Geometry and Kinematics of the Southwestern Termination … 91

Cardozo, N., & Allmendinger, R. W. (2013). Spherical projections with OSXStereonet. Computers
and Geosciences, 51, 193–205.
Casas, A. M., & Pardo, G. (2004). Estructura pirenaica y evolución de las cuencas sedimentarias en
la transversal Huesca-Oloron. In F. Colombo, C. L. Liesa, G. Meléndez, A. Pocovi, C. Sancho, &
A. R. Soria (Eds.), Itinerarios geológicos por Aragón (Vol. 1, pp. 63–96). Geo Guías (Sociedad
Geológica de España). (ISBN: 84-930160-2-0).
Chevrot, S., Sylvander, M., Diaz, J., Martin, R., Mouthereau, F., Manatschal, G., et al. (2018). The
non-cylindrical crustal architecture of the Pyrenees. Scientific reports, 8(1), 1–8.
Choukroune, P. (1992). Tectonic evolution of the Pyrenees. Annual Review of Earth and Planetary
Sciences, 20(1), 143–158.
Choukroune, P., & Séguret, M. (1973). Carte structurale des Pyrénées, 1/500.000, Université de
Montpellier -ELF Aquitaine
Coney, P. J., Muñoz, J. A., McClay, K. R., & Evenchick, C. A. (1996). Syntectonic burial and post-
tectonic exhumation of the southern Pyrenees foreland fold–thrust belt. Journal of the Geological
Society, 153(1), 9–16.
Costa, E., Garces, M., Lopez-Blanco, M., Beamud, E., Gomez-Paccard, M., & Larrasoaña, J.
C. (2010). Closing and continentalization of the South Pyrenean foreland basin (NE Spain):
Magnetochronological constraints. Basin Research, 22(6), 904–917.
Cuevas, (1992). Estratigrafía del "Garumniense" de la Conca de Tremp. Prepirineo de Lérida. Acta
Geol. Hispánica, 27: 95–108.
Duermeijer, C. E., Nyst, M., Meijer, P. T., Langereis, C. G., & Spakman, W. (2000). Neogene
evolution of the Aegean arc: Paleomagnetic and geodetic evidence for a rapid and young rotation
phase. Earth and Planetary Science Letters, 176(3), 509–525.
Dutta, D., & Mukherjee, S. (2021). Introduction to Structural Geology and Tectonics Field Guide-
book - Volume 1. In S. Mukherjee (Ed.), Structural Geology and Tectonics Field Guidebook -
(Vol. 1). Springer Nature: Switzerland AG. Cham. pp. xi-xvi. ISBN: 978-3-030-60142-3.
Fernández, O. (2004). Reconstruction of geological structures in 3D. An example from the southern
Pyrenees. Unpublished Ph. D. Thesis (pp. 321). University of Barcelona
Fisher, R. A. (1953). Dispersion on a sphere. Proceedings of the Royal Society A217, 295–305.
Fitzgerald, P. G., Muñoz, J. A., Coney, P. J., & Baldwin, S. L. (1999). Asymetric exhumation across
the Pyrenean orogen: Implication for the tectonic evolution of a collisional orogen. Earth and
Planetary Science Letters, 173, 157–170.
Frehner, M., Stefan, M., & Schmalholz, S. M. (2006). Numerical simulations of parasitic folding
in multilayers. Journal of Structural Geology, 28, 1647–1657.
Friend, P, F., Lloyd, M. J., McElroy, R., Turner, J., Van Gelder, A., & Vincent, S. J. (1996). Evolution
of the central part of the northern Ebro basin margin, as indicated by its Tertiary fluvial sedimentary
infill. In P. F. Friend, & C. J. Dabrio (Ed.), Tertiary basins of Spain (pp. 166–172). Cambridge
University Press.
Garcés, M., Parés, J. M., & Cabrera, L. (1996). Further evidence for inclination shallowing in red
beds. Geophysical Research Letters, 23(16), 2065–2068.
Garcés, M., López-Blanco, M., Valero, L., Beamud, E., Muñoz, J. A., Oliva-Urcia, B., et al. (2020).
Paleogeographic and sedimentary evolution of the south-pyrenean foreland basin. Marine and
Petroleum Geology, 104105.
García Sansegundo, J., Montes, M. J., Garrido-Schneider, E., & Barnolas, A., (2009). Mapa
Geológico de España 1:50.000 (MAGNA) Hoja de Agüero (208). Mapa y Memoria- IGME
ISBN: 978-84-7840-782-8
García-Ruiz, J. M., & Valero-Garcés, B. L. (1998). Historical geomorphic processes and human
activities in the Central Spanish Pyrenees. Mountain Research and Development, 309–320.
Hervouët, Y., Espurt, N., & Dhont, D. (2005). Failles normales Paléocène à Lutétien en zone sud-
pyrénéenne (Aragon, Espagne) et flexuration de la plaque ibérique. Comptes Rendus Geoscience,
337(3), 385–392.
92 E. L. Pueyo et al.

Hirst, J. P. P.,& Nichols, G.J. (1986). Thrust tectonic controls on the Miocene alluvial distribution
patterns, Southern Pyrenees. In: P.A Allen, P. Homewood (Eds.), Foreland basins, I.A.S. Spec.
Pub. 8, 247–258.
Hogan, P. J. (1993). Geocrohonologic, tectonic and stratigraphic evolution of the Southwest Pyre-
nean foreland basin, Northern Spain. Unpublised PhD thesis (pp. 219). University of Southern
California.
Hogan, P. J., & Burbank, D. W. (1996). Evolution of the Jaca piggyback basin and emergence of
the External Sierra, southern Pyrenees. In tertiary basins of spain: the stratigraphic record of
crustal kinematics (Vol. 6, pp. 153–160). Cambridge University Press.
Holl, J. E., & Anastasio, D. J. (1993). Paleomagnetically derived folding rates, southern Pyrenees,
Spain. Geology (Boulder)., 21, 3.
Huyghe, D., Mouthereau, F., Castelltort, S., Filleaudeau, P. Y., & Emmanuel, L. (2009). Paleogene
propagation of the southern Pyrenean thrust wedge revealed by finite strain analysis in frontal
thrust sheets: Implications for mountain building. Earth and Planetary Science Letters, 288(3–4),
421–433.
Izquierdo-Llavall, E. (2014). Variaciones longitudinales en la estructura de la zona axial pire-
naica. Aaportaciones de la fábrica magnética, el paleomagnetismo, la paleotermometría y la
modelización analógica. Unpublished PhD Universidad de Zaragoza (pp. 354).
Izquierdo-Llavall, E., Aldega, L., Cantarelli, V., Corrado, S., Gil-Peña, I., Invernizzi, C., et al.
(2013). On the origin of cleavage in the Central Pyrenees: Structural and paleo-thermal study.
Tectonophysics, 608, 303–318.
Izquierdo-Llavall, E., Casas-Sainz, A. M., Oliva-Urcia, B., Burtmester, R., Pueyo, E. L., & Housen,
B. (2015). Multi-episodic remagnetization related to diachronous thrusting in the Pyrenean
Internal Sierras. Geophysical Journal International, 201, 891–914. https://doi.org/10.1093/gji/
ggv042.
Izquierdo-Llavall, E., Casas-Sainz, A. M., Oliva-Urcia, B., Villalaín, J. J., Pueyo, E. L., & Scholger,
R. (2018). Rotational kinematics of basement antiformal stacks: Paleomagnetic study of the
western Nogueras Zone (Central Pyrenees). Tectonics, 37(10), 3456–3478.
Izquierdo-Llavall, E., Menant, A., Aubourg, C., Callot, J. P., Hoareau, G., Camps, P., Péré, E.
& Lahfid, A. Pre-orogenic folds and syn-orogenic basement tilts in an inverted hyperextended
margin: the northern Pyrenees case study. Tectonics, e2019TC005719.
Jolley, E. J. (1987). Thrust Tectonics and Alluvial Architecture of the Jaca Basin. Southern Pyrenees.
Unpublished Ph. D. Thesis (pp. 365). University of Wales.
Jolley, E. J., & Hogan, P. J. (1989). The Campodarbe group of the jaca basin, pyrenean tectonic
control of Oligo-Miocene river systems. Huesca Aragon. Spain. Fourth Internatinal Fluvial
Conference Excursion Guidebook. v. Servei Geologic de Cataluña., 4, 93–120.
Kodama, K. P., Anastasio, D. J., Newton, M. L., Pares, J. M., & Hinnov, L. A. (2010). High-
resolution rock magnetic cyclostratigraphy in an Eocene flysch, Spanish Pyrenees. Geochemistry,
Geophysics, Geosystems, 11(6).
Labaume, P., & Teixell, A. (2018). 3D structure of subsurface thrusts in the eastern Jaca Basin,
southern Pyrenees. Geologica Acta, 16(4), 477–498.
Labaume, P., Séguret, M., & Seyve, C. (1985). Evolution of a turbiditic foreland basin and analogy
with an accretionary prism: Example of the Eocene south-Pyrenean basin. Tectonics, 4(7), 661–
685.
Labaume, P., Meresse, F., Jolivet, M., Teixell, A., & Lahfid, A. (2016). Tectonothermal history of
an exhumed thrust-sheet-top basin: An example from the south Pyrenean thrust belt. Tectonics,
35(5), 1280–1313.
Lanaja, J. M. (1987). Contribución de la exploración petrolífera al conocimiento de la Geología
de España. http://info.igme.es/geologiasubsuelo/GeologiaSubsuelo/Documents.aspx (pp. 465).
Instituto Geológico y Minero de España Ed., 17 mapas.
Larrasoaña, J. C. (2000). Estudio magnetotectónico de la zona de transición entre el Pirineo central y
occidental; implicaciones estructurales y geodinámicas. Tesis doctoral Universidad de Zaragoza
(p. 287 pp.).
The Geometry and Kinematics of the Southwestern Termination … 93

Larrasoaña, J. C., Parés, J. C., Millán, H., del Valle, J., & Pueyo, E. L. (2003a). Paleomagnetic,
structural and stratigraphic constraints on the role of transverse fault development during basin
inversion (Pamplona Fault, Pyrenees, N Spain). Tectonics, 22(6), 1071–1093.
Larrasoaña, J. C., Parés, J. M., del Valle, J., & Millán, H. (2003b). Triassic paleomagnetism from
the Western Pyrenees revisited: Implications for the Iberian-Eurasian Mesozoic plate boundary.
Tectonophysics, 362(1–4), 161–182.
Larrasoaña, J. C., Parés, J. C., & Pueyo, E. L. (2003c). Stable Eocene magnetization carried by
magnetite and magnetic iron sulphides in marine marls (Pamplona-Arguis Formation, southern
Pyrenees, N Spain). Studia Geophysica Geodetica, 47, 237–254.
Larrasoaña, J. C., Pueyo, E. L., & Parés, J. M. (2004). An integrated AMS, structural, palaeo-
and rock-magnetic study of the Eocene marine marls from the Jaca-Pamplona basin (Pyre-
nees, N Spain). New insights into the timing of magnetic fabric acquisition in weakly deformed
mudrocks. In F. Martín-Hernández, C. M. Lüneburg, C. Aubourg & M. Jackson (Ed.), Magnetic
fabric: Methods and applications (Vol. 238, pp. 127–144). Geological Society of London Special
Publication.
Larrasoaña, J. C., Murelaga, X., & Garcés, M. (2006). Magnetobiochronology of lower Miocene
(ramblian) continental sediments from the tudela formation (western Ebro basin, Spain). Earth
Planetary Science Letters 243, 409–423.
Lobato, A., & Meléndez, A. (1988). Análisis de las facies carbonatadas del Cretácico superior en
el sector de Arguis-Belsué (provincia de Huesca). II Congreso Geológico de España, 1, 99–102.
López-Gómez, J., Alonso-Azcárate, J., Arche, A., Arribas, J., Barrenechea, J. F., Borruel-Abadía,
V., et al. (2019). Permian-triassic rifting stage. In The geology of iberia: a geodynamic approach
(pp. 29–112). Cham: Springer.
Luzón, A. (2005). Oligocene-Miocene alluvial sedimentation in the northern Ebro Basin, NE Spain.
Tectonic control and palaeogeographical evolution. Sedimentary Geology, 177, 19–39.
Mallada, L. (1878). Geología de la provincia de Huesca (p. 559). Madrid: Mem. Com. Mapa geol.
de España.
Mallada, L. (1881). Descripción física y ge-ológica de la provincia de Huesca (p. 439). Madrid:
Mem. Com. Mapa Geol. de España.
Mangin, J. Ph. (1958). Le Nummulitique sud-pyrénéen a l’Ouest de l’Aragon. Thèse University
Dijon. Publicada en 1959–60 en Pirineos (pp. 51–58: 1-631).
Marcén, M., Casas-Sainz, A. M., Román-Berdiel, T., Oliva-Urcia, B., Soto, R., & Aldega, L. (2018).
Kinematics and strain distribution in an orogen-scale shear zone: Insights from structural analyses
and magnetic fabrics in the Gavarnie thrust, Pyrenees. Journal of Structural Geology, 117, 105–
123.
Martín-Chivelet, J., Floquet, M., García-Senz, J., Callapez, P. M., López-Mir, B., Muñoz, J. A.,
et al. (2019). Late Cretaceous Post-Rift to Convergence in Iberia. In The geology of Iberia: A
geodynamic approach (pp. 285–376). Cham: Springer.
Martínez Peña, B., & Casas, A. M. (2003). Cretaceous-Tertiary tectonic inversion of the Cotiella
Nappe (Southern Pyrenees, Spain). Int J Earth Sci (Geol Rundsch), 92, 99–113.
Mattauer, M., & Seguret, M. (1971). Les relations entre la chaine des Pyrénées et le Golfe de
Gascogne, In Histoire Structurale du Golgo de Gascogne (Vol. 1, pp. IV 4–1-IV 4–24).
Mattei, M., Petrocelli, V., Lacava, D., & Schiattarella, M. (2004). Geodynamic implications of
Pleistocene ultrarapid vertical-axis rotations in the Southern Apennines, Italy. Geology, 32, 789–
792.
Meigs, A. J. (1995). Thrust faults, thrust sheets, and thrust-belts; new insights from the Spanish
Pyrenees. PhD Thesis University of Southern California (pp. 275). Los Angeles, CA, United
States.
Menant, A., Aubourg, C., Callot, J. P., Hoareau, G., Cuyala, J. B., Lechantre, J., et al. (2016). A
post-folding thermal imprint in the Chaînons Béarnais (North Pyrenean Zone) evidenced by two
independent methods. sciencesconf.org:rst2016-caen:115312
McCaig, A. M. & McClelland, E. (1992). Palaeomagnetic techniques applied to thrust belts. In K.
R. McClay & Chapman y Hall (Ed.), Thrust Tectonics (Vol. 447, pp. 209–216). London.
94 E. L. Pueyo et al.

McElroy, R. (1990). Thrust kinematics and syntectonic sedimentation: The Pyrenean frontal ramp,
Huesca, Spain. Unpublished PhD thesis (pp. 175). University of Cambridge.
Meigs, A. J. (1997). Sequential development of selected Pyrenean thrust faults. Journal of Structural
Geology, 19(3–4), 481–502.
Meresse, F. (2010). Dynamique d’un prisme orogénique intracontinental: évolution ther-
mochronologique (traces de fission sur apatite) et tectonique de la Zone Axiale et des piémonts
des Pyrénées centro-occidentales (Doctoral dissertation, Université de Montpellier 2) (pp. 280).
Mey, P. H. W., Nagtegaal, P. J. C., Roberti, K. J., & Hartevelt, J. J. A. (1968). Lithostratigraphic
subdivi-sion of post-herynian deposits in the south-central Pyrénées, Spain. Leidse Geologische
Mededelingen, 41, 221–228.
Millán, H. (1996). Estructura y cinemática del frente de cabalgamiento surpirenaico en las Sierras
Exteriores Aragonesas. Tesis Doctoral Universidad de Zaragoza (pp. 330).
Millán Garrido, H. (2006). Estructura y cinemática del frente de cabalgamiento surpirenaico en las
Sierras Exteriores aragonesas. Colección de Estudios Altoaragoneses (Vol. 53, pp. 398). Instituto
de Estdios Altoaragoneses, Huesca. ISBN 84–8127-165-9.
Millán, H., Parés, J. M., & Pocoví, A. (1992). Modelización sencilla de la estructura del sector
occidental de las sierras marginales aragonesas (Prepirineo, provincias de Huesca y Zaragoza).
III Congreso Geol. España. Simposios, 2, 140–149.
Millán, H., Aurell, M., & Meléndez, A. (1994). Synchronous detachment folds and coeval sedimen-
tation in the Prepyrenean External Sierras (Spain). A case study for a tectonic origin of sequences
and system tracts. Sedimentology, 41, 1001–1024.
Millán Garrido, H., Pocoví Juan, A., & Casas Sainz, A. M. (1995). El frente cabalgamiento surpire-
naico en el extremo occidental de las Sierras Exteriores. Revista de la Sociedad Geológica de
España, 8(1–2), 73–90.
Millán, H., Pueyo, E. L., Aurell, M., Luzón, A., Oliva-Urcia, B., Martínez Peña, M. B., et al. (2000).
Actividad tectónica registrada en los depósitos terciarios del frente meridional del Pirineo central.
Revista de la Sociedad Geológica de España, 13(2), 117–138.
Millán, H., Oliva-Urcia, B., & Pocoví, A. (2006). La transversal de Gavarnie-Guara. Estructura y
edad de los mantos de Gavarnie, Guara-Gèdre y Guara (Pirineo centro-occidental). Geogaceta,
40, 35–38.
Mochales, T. (2011). Chronostratigraphy, vertical axis rotations and AMS in the Boltaña anticline
(Southern Pyrenees): Kinematic implications. Unpublished PhD University of Zaragoza (pp. 222).
http://zaguan.unizar.es/record/6269
Mochales, T., Pueyo, E. L., Casas, A. M., Barnolas, A., & Oliva-Urcia, B. (2010). Anisotropic
magnetic susceptibility record of the kinematics of the Boltana Anticline (Southern Pyrenees).
Geological Journal, 45(5–6), 562–581.
Mochales, T., Casas, A. M., Pueyo, E. L., & Barnolas, A. (2012a). Rotational velocity for oblique
structures (Boltaña anticline, Southern Pyrenees). Journal of Structural Geology, 35, 2–16.
Mochales, T., Barnolas, A., Pueyo, E. L., Serra-Kiel, J., Casas, A. M., Samsó, J. M., et al. (2012b).
Chronostratigraphy of the Boltaña anticline and the Aínsa Basin (southern Pyrenees). Bulletin,
124(7–8), 1229–1250.
Mochales, T., Pueyo, E. L., Casas, A. M., & Barnolas, A. (2016). Restoring paleomagnetic data in
complex superposed folding settings: the Boltaña anticline (Southern Pyrenees). Tectonophysics,
671, 281–298. https://doi.org/10.1016/j.tecto.2016.01.008.
Montes, S. (1992). Sistemas deposicionales en el Eoceno medio-Oligoceno del Sinclinorio de
Guarga (cuenca de Jaca, Pirineo central). In II Congreso geológico de España (pp. 150–160).
Salamanca. Tomo 2.
Montes, M. J. (2009). Estratigrafía del Eoceno-Oligoceno de la cuenca de Jaca (sinclinorio del
Guarga). Montes Jaca Guarga. Colección de estudios altoaragoneses, (Vol. 59, pp. 355). Instituto
de Estudios Altoaragoneses. ISBN: 8481272027
Morsilli, M., Bosellini, F. R., Pomar, L., Hallock, P., Aurell, M., & Papazzoni, C. A. (2012).
Mesophotic coral buildups in a prodelta setting (Late Eocene, southern Pyrenees, Spain): a mixed
carbonate-siliciclastic system. Sedimentology, 59, 766–794.
The Geometry and Kinematics of the Southwestern Termination … 95

Mukherjee, S. (2015). In Atlas of structural geology. Elsevier. ISBN: 978–0-12-420152-1


Mukherjee, S. (2019). In Teaching methodologies in structural geology and tectonics (pp. 1–3).
Singapore: Springer. ISBN 978-981-13-2781-0.
Mukherjee, S. (2021). Atlas of structural geology (2nd edition). Elsevier. ISBN: 9780128168028
Mukherjee, S., & Khonsari, M. (2018). Inter-book normal fault-related shear heating in brittle
bookshelf faults. Marine and Petroleum Geology, 97, 45–48.
Mukherjee, S., & Tayade, I. (2019). Kinematic analyses of brittle roto-translational planar and
listric faults based on various rotational to translational velocities of the faulted blocks. Marine
and Petroleum Geology, 107, 326–333.
Muñoz, J. A. (1992). Evolution of a continental collision belt: ECORS-Pyrenees crustal balanced
cross-section. In K. R. McClay (Ed.), Thrust tectonics (pp. 235–246). Dordrecht: Springer.
Muñoz, J. A. (2019). Alpine orogeny: deformation and structure in the northern Iberian margin
(Pyrenees sl). In The geology of iberia: A geodynamic approach (pp. 433–451). Cham: Springer.
Muñoz, J. A., Beamud, E., Fernández, O., Arbués, P., Dinarès-Turell, J., & Poblet, J. (2013). The
Aínsa Fold and thrust oblique zone of the central Pyrenees: Kinematics of a curved contractional
system from paleomagnetic and structural data. Tectonics, 32(5), 1142–1175.
Mutti, E. (1984). The hecho eocene submarine fan system, south-central Pyrenees Spain. Geo-
Marine Letters, 3(2–4), 199–202.
Nichols, G. J. (1984). Thrust tectonics and alluvial sedimentation, Aragón, Spain. Unpublished Ph.
D. Thesis (pp. 243). University of Cambridge.
Nichols, G. J. (1987). The structure and stratigraphy of the western external sierras of the Pyrenees,
northern Spain. Geological Journal, 22, 245–259.
Nichols, G. J. (1989). Structural and sedimentological evolution of part of the west central Spanish
Pyrenees in the Late Tertiary. Journal Geological Society London, 146, 851–857.
Norris, D. K., & Black, R. F. (1961). Application of palaeomagnetism to thrust mechanics. Nature
(London), 192(4806), 933–935.
Núñez-Lahuerta, C., Galán, J., Sauqué, V., Rabal-Garcés, R., & Cuenca-Bescós, G. (2018). Avian
remains from new Upper Pleistocene and Holocene sites in the Spanish Pyrenees. Quaternary
International, 481, 123–134.
Oliva-Urcia, B. (2000). Estructura y cinemática del frente surpirenaico en el sector central de la
cuenca de Jaca-Pamplona. Unpublished MSc University of Zaragoza (pp. 100).
Oliva-Urcia, B. (2004). Geometría y cinemática rotacional en las Sierras Interiores y Zona Axial
(sector de Bielsa) a partir del análisis estructural y paleomagnético. Unpublished PhD University
of Zaragoza (pp. 290).
Oliva-Urcia, B. (2018). Thirty years (1988–2018) of advances in the knowledge of the structural
evolution of the South-Central Pyrenees during the Cenozoic collision, a summary. Revista de la
Sociedad Geológica de España, 31(2), 51–68.
Oliva-Urcia, B., & Pueyo, E. L. (2007a). Rotational basement kinematics deduced from remagne-
tized cover rocks (Internal Sierras, Southwestern Pyrenees). Tectonics 26 TC4014.
Oliva-Urcia, B., & Pueyo, E. L. (2007b). Gradient of shortening and vertical-axis rotations in the
southwestern Pyrenees (Spain). Revista de la Sociedad Geológica de España, 20(2), 105–118.
Oliva-Urcia, B., & Pueyo, E. L. (2019). Paleomagnetism in structural geology and tectonics. In S.
Mukherjee (Ed.), Teaching methodologies in structural geology and tectonics (pp. 55–121, 67).
Heidelberg: Springer. ISBN: 978-981-13-2781-0. 1. https://doi.org/10.1007/978-981-13-2781-0
Oliva Urcia, B., Millán, H., Pocoví, A., & Casas, A. M. (1996). Estructura de la Cuenca de Jaca en
el sector occidental de las Sierras Exteriores Aragonesas. Geogaceta, 4(20), 800–802.
Oliva-Urcia, B., Pueyo, E. L., & Larrasoaña, J. C. (2008). Magnetic reorientation induced by pres-
sure solution: A potential mechanism for orogenic-scale remagnetizations. Earth and Planetary
Science Letters, 265(3–4), 525–534.
Oliva-Urcia, B., Larrasoaña, J. C., Pueyo, E. L., Gil, A., Mata, P., Parés, J. M., et al. (2009).
Disentangling magnetic subfabrics and their link to deformation processes in cleaved sedimentary
rocks from the internal sierras (west central Pyrenees, Spain). Journal of Structural Geology,
31(2), 163–176.
96 E. L. Pueyo et al.

Oliva-Urcia, B., Casas, A. M., Pueyo, E. L., Román-Berdiel, T., & Geissman, J. W. (2010). Pale-
omagnetic evidence for dextral strike-slip motion in the Pyrenees during alpine convergence
(Mauléon basin, France). Tectonophysics, 494(3–4), 165–179.
Oliva-Urcia, B., Casas, A. M., Pueyo, E. L., & Pocoví, A. (2012a). Structural and paleomag-
netic evidence for non-rotational kinematics in the western termination of the External Sierras
(southwestern central Pyrenees). Geologica Acta, 10(2), 125–144.
Oliva-Urcia, B., Pueyo, E. L., Larrasoaña, J. C., Casas, A. M., Román, M. T., Van der Voo, R., et al.
(2012b). New and revisited paleomagnetic data from Permian-Triassic red beds: Two kinematic
domains in the west-central Pyrenees. Tectonophysics, 522–532, 158–175. https://doi.org/10.
1016/j.tecto.2011.11.023.
Oliva-Urcia, B., Beamud, E., Garcés, M., Arenas, C., Soto, R., Pueyo, E. L., et al. (2016). New
magnetostratigraphic dating of the Palaeogene syntectonic sediments of the west-central Pyre-
nees: tectonostratigraphic implications (Vol. 425, pp. SP425–5). London:Special Publications,
Geological Society.
Oliva-Urcia, B., Gil-Peña, I., Samsó, J. M., Soto, R., & Rosales, I. (2018). A paleomagnetic inspec-
tion of the Paleocene-Eocene thermal maximum (PETM) in the Southern Pyrenees. Frontiers in
Earth Science, 6, 202.
Oliva-Urcia, B., Beamud, E., Arenas, C., Pueyo, E. L., Garcés, M., Soto, R., et al. (2019). Dating the
northern deposits of the Ebro foreland basin; implications for the kinematics of the SW Pyrenean
front. Tectonophysics, 765, 11–34.
Oliván, C., Pueyo, E. L., Garrido-Schneider, E., Azcón, A., Sánchez-Moreno, E., Larrasoaña, J.C.
et al. (2020). The Santo Domingo and Salinas ranges (protected landscape in the South-Central
Pyrenees)-1; Map of groundwater physical and chemical parameters. Journal of maps. https://
doi.org/10.1080/17445647.2020.1736192
Oms, O., Dinarès-Turell, J., & Remacha, E. (2003). Magnetic stratigraphy from deep clastic
turbidites: Aan example from the Eocene Hecho group (southern Pyrenees). Studia Geophysica
et Geodaetica, 47(2), 275–288.
Pardo, G., Arenas, C., González, A., Luzón, A., Muñoz, A., Pérez, A., et al. (2004). La cuenca del
Ebro. In Geología de España (pp. 533–543). SGE-IGME Madrid.
Payros, A., Pujalte, V., & Orue-Etxebarria, X. (1999). The South Pyrenean Eocene carbonate
megabreccias revisited: New interpretation based on evidence from the Pamplona Basin.
Sedimentary Geology, 125(3–4), 165–194.
Pérez-Rivarés, F. J. (2016). Estudio magnetoestratigráfico del Mioceno del sector central de la
Cuenca del Ebro: Cronología, correlación y análisis de la ciclicidad sedimentaria. Unpub-
lished PhD Thesis. Universidad de Zaragoza (pp. 281). https://zaguan.unizar.es/record/79504/
files/TESIS-2019-121.pdf
Pérez-Rivarés, F. J., Garcés, M., Arenas, C., & Pardo, G. (2002). Magnetocronología de la sucesión
miocena de la Sierra de Alcubierre (sector central de la Cuenca del Ebro). Revista de la Sociedad
Geológica de España, 15, 210–225.
Pérez-Rivarés, F. J.; Garcés, M., Arenas, C., & Pardo, G. (2004). Magnetostratigraphy of
the Miocene continental deposits of the Montes de Castejón (central Ebro basin, Spain):
Geochronological and paleoenvironmental implications. Geologica Acta,2(3), 221–234.
Pérez-Rivarés, F. J., Arenas, C., Pardo, G., & Garcés, M. (2018). Temporal aspects of genetic
stratigraphic units in continental sedimentary basins: Examples from the Ebro basin, Spain.
Earth-Science Reviews, 178, 136–153.
Pocoví, A. (2019). Geología pirenaica vista desde el Sur. Real Academia de Ciencias Exactas,
Físicas, Químicas y Naturales de Zaragoza. Discurso de Ingreso (pp. 122). Depósito legal: Z
592–2019: Servicio de Publicaciones. Universidad de Zaragoza.
Pocoví, A., Millán, H., Navarro, J. J., & y Martínez, M. B. (1990). Rasgos estructurales de la Sierra
de Salinas y zona de los Mallos (Sierras Exteriores, Prepirineo, provincias de Huesca y Zaragoza).
Geogaceta, 8, 36–39.
Pocovi, A., Pueyo Anchuela, Ó., Pueyo, E. L., Casas-Sainz, A. M., Román Berdiel M. T, Gil Imaz,
A. et al. (2014). Magnetic fabrics in the Central-Western Pyrenees: an overview. In B. Almqvist,
The Geometry and Kinematics of the Southwestern Termination … 97

B. Henry, M. Jackson, T. Werner & F. Lagroix (Ed.), ASM in deformed rocks a tribute to Graham
J. Borradaile. Tectonophysics, 629, 303–318.
Pueyo Anchuela, Ó. (2012) Estudio de fábricas magnéticas y su relación con la deformación en el
sector centro-occidental del Pirineo central (Aragón y Navarra). Serie Tesis de la Universidad
de Zaragoza (Vol. 68, pp. 391). https://zaguan.unizar.es/record/9908?ln=es
Pueyo Anchuela, Ó., Pocoví Juan, A., & Gil Imaz, A. (2010) Tectonic imprint in magnetic fabrics
in Foreland Basin settings. Study in the Southern Pyrenees Foreland Basin, Ebro Basin, Spain)
Tectonophysics, 492(1–4), 150–163.
Pueyo-Anchuela, O., Pueyo, E. L., Pocoví, A., & Gil-Imaz, A. (2012). Vertical axis rotations in
fold and thrust belts: Comparison of AMS and paleomagnetic data in the western external sierras
(Southern Pyrenees). Tectonophysics, 532–535, 119–133.
Pueyo, E. L. (2000). Rotaciones paleomagnéticas en sistemas de pliegues y cabalgamientos. Tipos,
causas, significado y aplicaciones (ejemplos del Pirineo Aragonés). Unpublished PhD thesis.
Universidad de Zaragoza (pp. 296).
Pueyo, E. L., Millán, H., & Pocoví, A. (1997). Rotational kinematics of the southpyrenean basal
thrust at the Sierras Exteriores Aragonesas: Magnetotectonic data. Acta Geológica Hispánica,
32(3–4), 119–138.
Pueyo, E. L., Millán, H., & Pocovı, A. (2002). Rotation velocity of a thrust: A paleomagnetic study
in the external sierras (Southern Pyrenees). Sedimentary Geology, 146(1–2), 191–208.
Pueyo, E. L., Parés, J. M., Millán, H., & Pocoví, A. (2003a). Conical folds and apparent rotations
in paleomagnetism (A case studied in the Southern Pyrenees). In C. Mac Niocaill, T. H. Torsvik
& B. A. van der Pluijm (Ed.), Paleomagnetism applied to tectonics; a tribute to Rob Van der Voo.
Tectonophysics 362(1–4), 345–366.
Pueyo, E. L., Pocoví, A., Parés, J. M., Millán, H., & Larrasoaña, J. C. (2003b). Thrust ramp
geometry and spurious rotations of paleomagnetic vectors. Studia Geophysica et Geodaetica,
47(2), 331–357.
Pueyo, E. L., Pocoví, A., Millán, H., & Sussman, A. (2004). Map-view models for correcting and
calculating shortening estimates in rotated thrust fronts using paleomagnetic data. In A. Weil &
A. Sussman (Ed.), Special publication on orogenic curvature: Integrating paleomagnetic and
structural analyses (Vol. 383, pp. 57–71). Geological Society of America Special Publication
Pueyo, E. L., Sánchez, E., Canudo, J. I., Pereda-Suberbiola, X., Puértolas-Pascual, E., Parrilla-Bel,
J., et al. (2016a). Magnetoestratigrafía del Cretácico Superior del sector Occidental de las Sierras
Exteriores (Pirineo Occidental), implicaciones bioestratigráficas. Geotemas, 16(1), 909–912.
Pueyo, E. L., Oliva-Urcia, B., Sussman, A. J., & Cifelli, F. (2016b). Palaeomagnetism in fold and
thrust belts: Use with caution. In E. L. Pueyo, F. Cifelli, A. Sussman, & B. Oliva-Urcia (Ed.),
Geological society of London special publication on palaeomagnetism in fold and thrust belts:
New perspectives (Vol. 425(1), pp. 259–276).
Pueyo, E. L., Compaired, F., & Sánchez, E. (2017a). A large-scale conical fold termination in the
southwestern Pyrenees. Journal of Structural Geology, 98. (Picture of the month)
Pueyo, E. L., García-Lasanta, C., López, M. A., Oliván, C., San Miguel, G., Gil-Garbi, H et al.
the GeoKin3DPyr working group (2017b). Metodología para el desarrollo de la BBDD pale-
omagnética de Iberia (EPOS-DDSS Iberian Paleomagnetism). In MAGIBER X. (pp. 94–99).
Instituto de Investigación en Ciencias Ambientales (IUCA). Universidad de Zaragoza. ISBN:
978-84-16723-40-9
Pueyo, E. L., Santolaria, P., Calvín, P., Casas, A. M., Sánchez-Moreno, E. M., Oliva-Urcia, B.,
et al. (2020). Local and apparent vertical-axis rotations, detachment folding and diapirism in the
southwestern Pyrenean termination. Tectonics.
Puigdefábregas, C. (1975). La sedimentación molásica en la cuenca de Jaca. Pirineos, 104, 1–188.
Puigdefàbregas, C., & Soler, M. (1973). Estructura de las Sierras Exteriores pirenaicas en el corte
del río Gállego. (prov. de Huesca). Pirineos, 109, 5–15.
Puigdefabregas, C., & Souquet, P. (1986). Tecto-sedimentary cycles and depositional sequences of
the mesozoic and tertiary from the pyrenees. In E. Banda, & S. M. Wickham (Ed.), The geological
evolution of the Pyrenees. Tectonophysics (Vol. 129(1–4) pp. 173–203).
98 E. L. Pueyo et al.

Puigdefàbregas, C., Muñoz, J. A., & Vergés, J. (1992). Thrusting and foreland basin evolution in
the southern Pyrenees. In Thrust tectonics (pp. 247–254). Dordrecht:Springer.
Quirantes, J. (1969). Estudio sedimentológico y estratigráfico del Terciario continental de Los
Monegros (Unpublished PhD thesis), Universidad de Granada.
Ramberg, H. (1963). Fluid dynamics of viscous buckling applicable to folding of layered rocks.
Bulletin of the American Association of Petroleum Geologits, 47(3), 484–505.
Ramsay, J. G. (1974). Development of chevron folds. Geological Society of America Bulletin, 85,
1741–1754.
Ramón, M. J. (2013). Flexural unfolding of complex geometries in fold and thrust belts using
paleomagnetic vectors. Unpublished PhD, University of Zaragoza, Spain. http://zaguan.unizar.
es/record/11750.
Ramón, M. J., Pueyo, E. L., Briz, J. L., Pocoví, A., & Ciria, J. C. (2012). Flexural unfolding in 3D
using paleomagnetic vectors. Journal of Structural Geology, 35, 28–39. https://doi.org/10.1016/
j.jsg.2011.11.015.
Ramón, M. J., Pueyo, E. L., Rodríguez-Pintó, A., Ros, L. H., Pocoví, A., Briz, J. L., et al. (2013). A
computed tomography approach to understanding 3D deformation patterns in complex flexural
folds. Tectonophysics, 593, 57–72.
Ramón M. J., Pueyo E. L., Caumon G., & Briz J. L. (2016a). Parametric unfolding of flexural folds
using paleomagnetic vectors. In E. L. Pueyo, F. Cifelli, A. J. Sussman, & B. Oliva-Urcia (Ed.),
Geological society of London special publication 425 on palaeomagnetism in fold and thrust
belts: New perspectives (pp. 247–258).
Ramón, M. J., Briz, J. L., Pueyo, E. L., & Fernandez, O. (2016b). Horizon restoration by best-
fitting finite element and rotation constraints: Sensitivity of the meshes geometry and pin-element
location. Mathematical Geosciences, 48(4), 419–437.
Ramón, M. J., Pueyo E. L., Oliva-Urcia, B., Larrasoaña, J. C. (2017). Virtual directions in paleo-
magnetism: A global and rapid approach to evaluate the NRM components and their stability.
Frontiers in Earth Science, 5 (8), 14
Remacha, E., & Fernández, L. P. (2003). High-resolution correlation patterns in the turbidite systems
of the hecho group (South-Central Pyrenees, Spain). Marine and Petroleum Geology, 20(6–8),
711–726.
Riba, O. (1976). Syntectonic unconformities of the Alto Cardener, Spanish Pyrenees: A genetic
interpretation. Sedimentary Geology, 15(3), 213–233.
Robador, A., Ramajo, J., Muñoz, A., Pérez, A., Luzón, A., Arenas, C. & González, A. (2011); Mapa
Geológico Digital continuo E. 1: 50.000, Zona Cuenca del Ebro (Zona-2700). In GEODE. Mapa
Geológico Digital continuo de España.
Robador, A., Samsó, J. M., Ramajo, J., Barnolas, A., Clariana, P., Martín, S. & Gil, I (2019) (2;
Mapa Geológico Digital continuo E. 1:50.000, Zona Pirineos Vasco-Cantábrica (Zona-1600). In
GEODE. Mapa Geológico Digital continuo de España.
Rodríguez Pintó, A. (2013). Magnetoestratigrafía del Eoceno inferior y medio en el frente Surpire-
naico (Sierras Exteriores): implicaciones cronoestratigráficas y cinemáticas. Unpublished PhD
University of Zaragoza (pp. 370) http://zaguan.unizar.es/record/10043
Rodríguez-Pintó, A., Pueyo, E. L., Pocoví, A., & Barnolas, A. (2008). Cronología de la actividad
rotacional en el sector central del frente de cabalgamiento de Sierras Exteriores (Pirineo
Occidental). Geotemas, 10, 1207–1210.
Rodríguez-Pintó, A., Pueyo, E. L., Serra-Kiel, J., Samsó, J. M., Barnolas, A., & Pocoví, A. (2012).
Lutetian magnetostratigraphic calibration of larger foraminifera zonation (SBZ) in the Southern
Pyrenees: The Isuela section. Palaeogeography, Palaeoclimatology, Palaeoecology, 333, 107–
120.
Rodríguez-Pintó, A., Pueyo, E. L., Pocoví, A., Ramón, M. J., & Oliva-Urcia, B. (2013a). Structural
control on overlapped paleomagnetic vectors: A case study in the Balzes anticline (Southern
Pyrenees). Physics of the Earth and Planetary Interiors.
Rodríguez-Pintó, A., Pueyo, E. L., Serra-Kiel, J., Barnolas, A., Samsó, J. M., & Pocoví, A. (2013b).
The Upper Ypresian and Lutetian in San Pelegrín section (Southwestern Pyrenean Basin):
The Geometry and Kinematics of the Southwestern Termination … 99

magnetostratigraphy and larger foraminifera correlation. Palaeogeography, Palaeoclimatology,


Palaeoecology, 370, 13–29.
Rodríguez-Pintó, A., Pueyo, E. L., Sánchez, E., Calvin, P., Ramajo, J., Ramón, M. J., et al. (2016).
Rotational kinematics of a curved fold: A structural and paleomagnetic study in the Balzes anti-
cline (Southern Pyrenees). Tectonophysics, 677–678, 171–189. https://doi.org/10.1016/j.tecto.
2016.02.049.
Rodríguez-Pintó, A., Sánchez-Moreno, E., Pueyo, E. L., Oliva-Urcia, B., Barnolas, A., & Izquierdo-
Llavall, E. (2019). Limite Luteciense/Bartoniense en la sección de Isuela (revisada), Pirineos
suroccidentales 2019. In E. Font, E. Beamud, B. Oliva-urcia, E. L. Pueyo, & F. C. Lopez (Ed.),
Magiber Xi- Paleomagnetismo Em Espanha E Portugal (Livro de Resumos)-Universidade de
Coimbra (pp. 121–124). ISBN 978-989-98914-7-0.
Rodríguez-Pintó, A., Pueyo, E. L., Serra-Kiel, J., & Barnolas, A. (2020). The chronology of the
eastern Jaca Basin revisited; implications for the kinematics of the External Sierras (Southern
Pyrenees). Geodinamica Acta
Rosenbaum, G., Lister, G. S., & Duboz, C. (2002). Relative motions of Africa, Iberia and Europe
during the Alpine orogeny. Tectonophysics, 359, 117–129.
Roure, F., Choukroune, P., Berasategui, X., Muñoz, J. A., Villien, A., Matheron, P., et al. (1989).
ECORS deep seismic data and balanced cross sections: Geometric constraints on the evolution
of the Pyrenees. Tectonics, 8, 41–45.
Samsó, J. M., Serra-Kiel, J., Tosquella, J., & Travé, A. (1994). Cronoestratigrafía de las plataformas
lutecienses de la zona central de la cuenca Surpirenaica. In A. Gonzalez Rodriguez (Ed.), II
Congreso Del Grupo Español Del Terciario (pp. 205–208). Jaca: Universidad de Zaragoza.
Sánchez-Moreno, E. M. (2012). La geología de las Sierras de Santo Domingo y Salinas como
recurso de desarrollo rural: propuesta de guía divulgativa; propuesta de gestión y conservación
(pp. 105). MSc University of Zaragoza, + 3 maps + 16 files.
Sánchez, E., Pueyo, E. L., Ramón, M. J., Oliva-Urcia, B., & Calvín, P. (2013). Vertical axis rotation
deduced from the AMS data in the Santo Domingo anticline (Western Pyrenees). In Colóquio
ASM, Tectónica e (Paleo)magnetismo de Materiais. Editado: Centro de Geofísica da Universidade
Coimbra. ISBN: 978-989-97823-2-7, pp 75-77
Sánchez-Moreno, E., Pueyo, E. L., Longares, L. A. & Compaired, F. (2015). La geología de las
Sierras de Santo Domingo y Salinas como recurso de desarrollo rural: propuesta de guía divulga-
tiva; propuesta de gestión y conservación. Instituto Geologíco y Minero de España y Consejo de
Protección de la Naturaleza de Aragón (pp. 184). ISBN 978-84-7840-965-5 NIPO: 728140149,
+ 1 Mapa.
Sancho, C., Arenas, C., Pardo, G., Peña-Monné, J. L., Rhodes, E. J., Bartolomé, M., et al. (2018).
Glaciolacustrine deposits formed in an ice-dammed tributary valley in the south-central Pyrenees:
New evidence for late Pleistocene climate. Sedimentary Geology, 366, 47–66.
Santolaria, P., Ayala, C., Pueyo, E. L., Rubio, F. M., Soto, R., Calvin, P., et al. (2020). Impact of
multidetachment units in the evolution of frontal contractional systems: Structural and geophys-
ical characterization of the western termination of the Barbastro anticline, Southern Pyrenees:
Tectonics.
Séguret, M. (1972). Etude Tectonique des nappes et series decollées de la partie centrale du versant
sud des Pyrénées. Caractére sedymentaire role de la compression et de la gravité. These Fac. Sc.
de Montpellier. Publ. de l’Univ. Des Sc. et Tec. du Languedoc (USTELA), Sér. Geol. Struct. 2,
155
Serra-Kiel, J., Hottinger, L., Caus, E., Drobne, K., Ferràndez, C., Samsó, J. M., et al. (1998). Larger
foraminiferal biostratigraphy of the tethyan Paleocene and eocene. Bull. la Soc. Geol. Fr., 169,
281–299.
Sibuet, J. C., Srivastava, S. P., & Spakman, W. (2004). Pyrenean orogeny and plate kine-
matics. Journal of Geophysical Research: Solid Earth, 109(B8).
Silva Casal, R. (2017). Las plataformas carbonatadas del Eoceno medio de la Cuenca de Jaca-
Pamplona (Formación Guara, Sierras Exteriores): análisis estratigráfico integral y evolución
sedimentaria. Unpublished PhD Universidad de Zaragoza. 345 pp.
100 E. L. Pueyo et al.

Silva-Casal, R., Aurell, M., Payros, A., Pueyo, E. L., & Serra-Kiel, J. (2019). Carbonate ramp
drowning caused by flexural subsidence: The South Pyrenean middle Eocene foreland basin.
Sedimentary Geology, 393, 105538.
Silva-Casal, R., Serra-Kiel, J., Rodríguez-Pintó, A., Pueyo, E. L., Aurell, M., & Payros, A. (2020 in
press) Systematics of Lutetian Larger Foraminifera and magneto-biostratigraphy from the South
Pyrenean Basin (Sierras Exteriores, Spain). Geologica Acta
Soler, M., & y Puigdefàbregas, C. (1970). Líneas generales de la geología del Alto Aragón
occidental. Pirineos, 96, 5–19.
Soto, R., Casas-Sainz, A. M., & Pueyo, E. L. (2006). Along-strike variation of orogenic wedges
associated with vertical axis rotations. Journal of Geophysical Research (Solid Earth), 111,
B10402–B10423.
Speranza, F., Maniscalco, R., Mattei, M., Di Stefano, A., Butler, R. W. H., & Funiciello, R. (1999).
Timing and magnitude of rotations in the frontal thrust systems of southwestern Sicily. Tectonics,
18, 1178–1197.
Suppe, J., Chou, G. T., & Hook, S. C. (1992). Rates of folding and faulting determined from growth
strata. In Thrust tectonics (pp. 105–121). Dordrecht: Springer.
Sussman, A. J., Butler, R. F., Dinarès-Turell, J., & Vergés, J. (2004). Vertical-axis rotation of a
foreland fold and implications for orogenic curvature: An example from the Southern Pyrenees
Spain. Earth and Planetary Science Letters, 218(3–4), 435–449.
Sussman, A. J., Pueyo, E. L., Chase, C. G., Mitra, G., & Weil, A. B. (2012). The impact of vertical-
axis rotations on shortening estimates. Lithosphere, 4(5), 383–394.
Teixell, A. (1992). Estructura alpina en la transversal de la terminación occidental de la zona axial
pirenaica (p. 252). de Barcelona: University
Teixell, A. (1996). The Ansó transect of the southern Pyrenees: Basement and cover thrust
geometries. Journal of the Geological Society, 153(2), 301–310.
Teixell, A. (1998). Crustal structure and orogenic material budget in the west central Pyrenees.
Tectonics, 17(3), 395–406.
Teixell, A., & García-Sansegundo, J. (1995). Estructura del sector central de la Cuenca de Jaca
(Pirineos meridionales). Revista de la Sociedad Geológica de España, 8(3), 215–228.
Teixell, A., Montes, M. J., Arenas, C., Garrido-Schneider, E., & Barnolas, A. (2009). Mapa
Geológico de España 1:50.000 (MAGNA) Hoja de Uncastillo (209). Mapa y Memoria- IGME.
ISBN: 978-84-7840-780-4
Teixell, A., Labaume, P., & Lagabrielle, Y. (2016). The crustal evolution of the west-central Pyrenees
revisited: Inferences from a new kinematic scenario. Comptes Rendus Geoscience, 348(3–4),
257–267.
Teletzke, A. L. (2012). Sedimentary and tectonic evolution of the Peñaflexure, southwest Pyrenean
mountain front, Spain.MSc Thesis Lehigh University.
Turner, J. P. (1988). Tectonic and stratigraphic evolution of the West Jaca thrust-top basin, SW
Pyrenees. Tesis Doctoral. University Bristol.
Turner, J. P. (1990). Structural and stratigrafic evolution of the West Jaca thrust-top basin, Spanish
Pyrenees. Journal of the Geological Society London, 147, 177–184.
Turner, J. P., & Hancock, P. L. (1990a). Relationships between thrusting and joint systems in the
Jaca thrust-top basin. Spanish Pyrenees. Journal of Structural Geology, 12, 217–226.
Turner, J. P., & Hancock, P. L. (1990b). Thrust systems of the southwest Pyrenees and their control
over basin subsidence. Geological Magazine, 127, 383–392.
Uzel, J., Nivière, B., & Lagabrielle, Y. (2020). Fluvial incisions in the north-western Pyrenees (Aspe
Valley): Dissection of a former planation surface and some tectonic implications. Terra Nova,
32(1), 11–22.
Van der Voo, R. (1969). Paleomagnetic evidence for the rotation of the Iberian Peninsula.
Tectonophysics, 7(1), 5–56.
Van der Voo, R., & Boessenkool, A. (1973). Permian paleomagnetic result from the western Pyrenees
delineating the plate boundary between the Iberian peninsula and stable Europe. Journal of
Geophysical Research, 78(23), 5118–5127.
The Geometry and Kinematics of the Southwestern Termination … 101

Van der Voo, R., & Channell, J. E. T. (1980). Paleomagnetism in orogenic belts. Reviews of
Geophysics and Space Physics, 18(2), 455–481.
Vergés, J., Marzo, M., Santaeulària, T., Serra-Kiel, J., Burbank, D. W., Muñoz, J. A., et al. (1998).
Quantified vertical motions and tectonic evolution of the SE Pyrenean foreland basin. Geological
Society, London, Special Publications, 134(1), 107–134.
Vidal-Royo, O., Koyi, H. A., & Muñoz, J. A. (2009). Formation of orogen-perpendicular thrusts
due to mechanical contrasts in the basal décollement in the central external sierras (Southern
Pyrenees, Spain). Journal of Structural Geology, 31, 523–539.
Vissers, R. L. M., & Meijer, P. T. (2012). Iberian plate kinematics and alpine collision in the
Pyrenees. Earth-Science Reviews, 114(1–2), 61–83.
Woodcock, N. H. (1977). Specification of fabric shapes using an eigenvalue method. Bulletin of the
Geological Society of America, 88, 1231–1236.
Miocene-Quaternary Strain Partitioning
and Relief Segmentation Along
the Arcuate Betic Fold-and-Thrust Belt:
A Field Trip Along the Western Gibraltar
Arc Northern Branch (Southern Spain)

Alejandro Jiménez-Bonilla, Manuel Díaz-Azpiroz, Inmaculada Expósito,


and Juan Carlos Balanyá

Abstract The Western Gibraltar Arc (WGA), located at the western end of the
Mediterranean Alpine orogenic belt, was built up by the westward emplacement
of the Alboran hinterland domain upon the South Iberian and Maghrebian paleo-
margin domains. This guide is focused on the strain partitioning modes developed
in the fold-and-thrust belt (FTB) of the northern branch of the WGA. The proposed
itineraries aim to show the distinctive strain partitioning modes of this arcuate FTB
from the hinge (itinerary 1) to the lateral (itinerary 2) zones of the WGA. We focus
our observations on the structural evolution from the late Miocene onwards. This
short-term (a few My) temporal frame favours that the structures accommodating
strain partitioning may greatly control a number of first-order topographic traits,
for instance, relief segmentation. Itinerary1, located in the northern WGA hinge
zone, allows one to visit a representative area of the inner FTB characterized by
a set of parallel mountain ranges enclosing the Ronda intermontane basin. In this
area, the topography is conditioned by late Miocene to Pliocene km-scale, upright
or doubly vergent, NE-SW trending folds and associated reverse faults. These struc-
tures, parallel to the orogenic grain, gave rise to a conformable relief. Additionally,
first-order along-strike relief segmentation controlled by normal faults has accom-
modated arc-parallel extension. The itinerary includes stops in the two main normal
fault zones located at the margins of the Ronda basin, the most important inter-
montane basin of the northern WGA. Itinerary2 runs along the lateral, eastern end
of the WGA, marked by a km-scale, E-W oriented brittle-ductile shear zone, the
Torcal Shear Zone (TSZ), which is formed by a set of en-echelon morphostructural
highs that define a 70 km long topographic alignment. The itinerary starts at the TSZ
western end and continues eastwards to the main structural domains of its central
part: the Valle de Abdalajís Massif and the Torcal de Antequera Massif. This permits
us to visit the main structures that characterize this highly partitioned shear zone:
NE–SW folds and reverse faults, NW–SE normal faults, and E–W dextral strike-
slip faults, especially localized along the northern and southern walls of the TSZ.

A. Jiménez-Bonilla · M. Díaz-Azpiroz (B) · I. Expósito · J. C. Balanyá


Department of Physical, Chemical and Natural Systems, Universidad Pablo de Olavide. Crtra,
Utrera km 1, 41013 Sevilla, Spain
e-mail: mdiaazp@upo.es

© Springer Nature Switzerland AG 2021 103


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_4
104 A. Jiménez-Bonilla et al.

The overall kinematics corresponds to dextral transpression with a nearly vertical


significant extrusion and associated surface uplift.

1 Introduction

Fold-and-thrust belts (FTBs) associated with arcuate orogens (e.g., Western Alps,
Carpathians, Calabrian and Gibraltar arcs) acquire a wide variety of map-scale
geometries that seem to be controlled by several factors (e.g., indenter shape, pre-
deformational thickness of the sedimentary sequence and detachment strength).
These orogens show convergence trajectories or radial tectonic transport directions
along with the arcuate structural trend (Macedo and Marshak 1999; Marshak 2004).
The degree of divergence of tectonic transport directions, which occurs by eventual
vertical axis rotation of independent blocks, is closely related to the arc forma-
tion mechanism, defining contrasting patterns when primary FTBs are compared to
secondary (oroclinal) or progressive arcuate FTBs (e.g., Hindle and Burkhard 1999;
Weil and Sussman 2004).
Due to the combination of the above-mentioned factors, arcuate FTBs display
map-view shapes with a number of linked second-order salients and recesses,
commonly connected by arc lateral transition zones. As a result, the angle of
plate convergence changes with respect to the considered FTB segment, thereby
leading to variations of the strain partitioning mode along the orogenic grain. At
arc hinges, the strain is mainly partitioned into orthogonal shortening, accom-
modated by arc-parallel folds and thrusts, and arc-parallel stretching, commonly
accommodated by arc-perpendicular normal faults and/or arc-oblique strike-slip
faults (e.g., Marshak 1988; McCaffrey 1991; Balanyá et al. 2007). By contrast, arc
lateral segments usually localize strike-slip dominated tectonics, often developing
transpressive zones, oblique to the overall arcuate orogenic trend.
In active or recent arcuate FTBs, the structures accommodating strain partitioning
in upper crustal levels may exert significant control on landscape and topography,
usually involving relief segmentation. In this respect, arc-orthogonal shortening gives
way to rugged across-strike topography, whereas arc-parallel stretching is associated
with significant along-strike relief drops.
In this guide, we visit the northern branch of the Western Gibraltar Arc (Fig. 1).
Our itinerary runs along the FTB orogenic grain and seeks to illustrate the contrasting
strain partitioning modes exhibited by arcuate FTBs from hinge to lateral zones. The
young age of the structures accommodating deformation and the occurrence of fault
zones traverse to the orogenic grain lead to a significant relief segmentation that
makes this trip a good opportunity to use landscape as a tectonic tool.
Miocene-Quaternary Strain Partitioning and Relief Segmentation … 105

Fig. 1 Geological map locating the two itineraries included in this field guide. Modified from
Jiménez-Bonilla (2017)
106 A. Jiménez-Bonilla et al.

2 Tectonic Setting

The Mediterranean-type arcs (e.g., Gibraltar, Calabria, Eastern Carpathians) are char-
acterized by high curvatures, relatively low topography and strongly attenuated upper
plate thickness due to coeval extension and collisional tectonics (Royden and Burch-
fiel 1989). The Gibraltar Arc formed during the last 25 My due to the collision
of the Alboran Domain against the South Iberian and Maghrebian paleomargins,
building up the Betic and Rif chains, respectively (Vera et al. 2004). Between the
Alboran Domain (internal zone) and both paleomargins, Flysch Trough units were
deposited and later sandwiched. Both paleomargins and Flysch units formed the
external zones and were later deformed into the Gibraltar Arc fold-and-thrust belt
(e.g. Expósito et al. 2012). The radius of curvature in the hinge zone of the arc,
measured at the internal-external zones boundary, is ~120 km, and its concave side
(hangingwall of the suture) contains a large extensional basin (the Alboran basin),
where the continental crust was thinned down to 14 km (Comas et al. 1999; Torné
et al. 2000).
The Western Gibraltar Arc (WGA; Balanyá et al. 2007, 2012) has been formally
defined as the portion of the Gibraltar Arc west of the 4° 30 meridian. Its structural
grain draws a second-order salient that ends at two transpressional zones (Fig. 1): the
Torcal Shear Zone (TSZ) to the north (Betics) and the Jebba Fault Zone to the south
(Rif). Deformation within the FTB around the WGA occurred since the early Miocene
to the Holocene. During this period, overall kinematics has been governed by the
ca. N–S Europe-Africa convergence (Mazzoli and Helman 1994) and the westward
migration of the arc/back-arc system probably related to slab retreat (Royden 1993;
Spakman and Wortel 2004) and/or asymmetric delamination (García-Dueñas et al.
1992; Duggen et al. 2004).
In the northern branch of the WGA, the FTB is made up of two types of
detached covers containing Mesozoic to Paleogene marine sedimentary sequences
(Fig. 1): (I) the Subbetic units, which correspond to the South Iberian palaeomargin,
mainly deposited onto a continental platform; and, thrust to the NW upon these,
the Flysch Trough units, mainly composed of deep-water marine sequences that
were deposited partially onto an oceanic lithosphere or an attenuated continental
lithosphere (Durand-Delga et al. 2000).
Most of the observations proposed in this guide are located within the inner
Subbetic units, namely the Penibetic, which, from bottom to top include (Martín-
Algarra 1987): (1) Middle Triassic limestones and dolostones, followed by Upper
Triassic, evaporite-rich, marls with dolostones and limestones at the top; (2) Lower
and Middle Jurassic dolostones and oolitic limestones; (3) Upper Jurassic nodular and
oolitic limestones; (4) Lower Cretaceous whitish marly limestones (“White Beds”);
and (5) Upper Cretaceous to Paleogene redish marly limestones (“Red Beds”).
The northern FTB mountain front borders the Guadalquivir foreland basin. Marine
deposits filled this basin during the middle and late Miocene, whereas a progres-
sive westward emersion took place from the Pliocene onwards (González-Delgado
et al. 2004). Several intermontane basins developed within the FTB, some of them
Miocene-Quaternary Strain Partitioning and Relief Segmentation … 107

connected with the Guadalquivir basin. Their inception took place during the late
Miocene when the current general topographic traits of the Betic Chain started to
emerge (Braga et al. 2003; Sanz de Galdeano and Alfaro 2004). Examples of these
are the Ronda and the El Chorro intermontane basins, both of them included in this
field trip.
Located in the concave side of the Gibraltar Arc, the Internal Zones (Alboran
Domain, Fig. 1) of the Betic-Rif orogen are made up, from structural bottom to
top, of the Frontal Units, the Alpujarride and the Malaguide complexes. The Frontal
Units consist of strongly deformed, Triassic to Palaeogene rocks, which crop out in
the periphery of the Alboran Domain, having been thrust upon the Flysch Trough
units. The Alpujarride and the Malaguide complexes, tectonically placed onto the
Frontal Units, contain Palaeozoic and Mesozoic sequences. The Alpujarride complex
comprise a very thick metamorphic (mostly metapelitic) sequence that contains
eclogites and carpholite-bearing HP-LT assemblages (Tubía and Gil Ibarguchi 1991)
and peridotites (Sánchez-Gómez et al. 2002). Different metamorphic stages took
place in the Alpujarride complex from the Permian to the early Miocene (Platt and
Whitehouse 1999; Sosson et al. 1998; Acosta-Vigil et al. 2014). The Malaguide
complex, on top of the Alpujarride complex, is mainly composed of a slightly
deformed and metamorphosed Ordovician-Permian sequence. La Joya Complex is
a turbiditic unit that frequently overlies the Alboran Domain (Fig. 2).
The deformation within the Betic FTB shows a typical thin-skinned style and
its evolution can be divided into two major tectonic steps. (1) A main accretionary
event occurred at early to middle Miocene time, coevally with NW-verging folds in
the Subbetic units (Expósito et al. 2012) and a thrust system affecting the Flyschs
Trough units (Luján et al. 2006). (2) From late Miocene onwards, the WGA forms
as a second-order arc of the Gibraltar Arc, and two domains with distinctive strain
partitioning modes are developed within the FTB in this sector of the Betics: the WGA
hinge zone and the Torcal Shear Zone (TSZ). The WGA hinge zone is characterized
by arc-orthogonal shortening structures (fold, thrusts and reverse faults) coeval with
arc-parallel stretching, mainly accommodated by arc-perpendicular normal faults and
conjugate strike-slip faults (Balanyá et al. 2007, 2012; Jiménez-Bonilla et al. 2015,
2017). Thrusting transport sense describes an outward divergent pattern. Conversely,
deformation in the TSZ was highly partitioned into E–W major dextral strike-slip
dominated faults and transpressive zones, NE–SW oriented shortening structures and
NW–SE strikes normal faults that accommodated subordinate extension parallel to
fold axes (Díaz-Azpiroz et al. 2014; Barcos et al. 2015).
In order to illustrate the contrasting strain partitioning modes described above,
we propose to split the field trip into two itineraries (Fig. 1): Itinerary1- devoted to
the WGA hinge zone, and Itinerary2- focused on the TSZ.
108 A. Jiménez-Bonilla et al.

Fig. 2 Geological map of Itinerary 1 with the location of its stops. Modified from Jiménez-Bonilla
et al. (2015). The red line shows the location of cross-Sect. 7b

3 Useful Information to Plan Your Trip

In addition to its geological heritage, the area described in this field guide offers a
broad range of natural and cultural values. The route goes across several protected
natural sites, such as Parque Nacional de Sierra de las Nieves, Parque Natural
de Grazalema, Paraje Natural del Torcal de Antequera, or Paraje Natural del
Desfiladero de Los Gaitanes (Caminito del Rey). Regarding its cultural interest, it is
well worth visiting, among others, the monumental towns of Antequera and Ronda,
Miocene-Quaternary Strain Partitioning and Relief Segmentation … 109

as well as the Antequera Dolmens Archaeological Site, included in the UNESCO


World Heritage Sites list.
There is a wide offer for accommodation in the area of the field trip. The touristic
towns of Ronda and Antequera are on our route, and you can find plenty of hotels
and restaurants. Málaga, the main city of the Costa del Sol, is located 50 and 100 km
away from Antequera and Ronda, respectively. Málaga also provides the closest
international airport, though you can also fly to Sevilla.
Note that the summer is quite warm in southern Spain and the temperature can
easily exceed 35-40°. Furthermore, the whole region is a common touristic hotspot,
being particularly busy in summer when some places proposed as stops in the guide
can be crowded with visitors. In this respect, if you intend to visit the “Caminito
del Rey” (see stop 2.3), you must take into account that it is a protected area with
restricted access, which needs booking in advance (tickets may be sold out months
before your planned date). You can purchase tickets through the official webpage
(http://www.caminitodelrey.info/en/) or via tourist agencies or local hotels.

4 Itineraries Description

4.1 Itinerary 1. The Northern Branch of the Western


Gibraltar Arc

This route runs along the FTB of the northern WGA hinge zone, which is an area
characterized by a set of parallel mountain ranges (up to 1918 m high) that encloses
the Ronda intermontane basin (Fig. 2).
The relief of this area was built up after the Tortonian. Its cross-strike topography
is conditioned by late Miocene to Pliocene km-scale folds displaying a conformable
relief. Therefore, mountain ranges made up of Penibetic units coincide with large-
scale NE–SW oriented late antiformal cores, whereas floor valleys coincide with
synformal cores (Jiménez-Bonilla et al. 2015, 2017). Along-strike relief segmenta-
tion also occurs, greatly controlled by arc-parallel extension (Jiménez-Bonilla et al.
2015).
One of the main relief discontinuities along the orogenic grain is the Ronda basin,
the most significant intermontane basin of the Western Betics, whose marine sedi-
mentary infill ranges from the late Tortonian to the Messinian. The basin itself is
characterized by a wavy relief with altitudes between 450 and 850 m, being limited
to the NE and SW by an abruptly rising relief, with altitudes up to 1550 m. These
sharp boundaries are controlled by post-Serravalian, NW–SE striking normal faults,
sub-orthogonal to the post-Serravallian large-scale fold axes (Jiménez-Bonilla et al.
2015). Fault surfaces dip basinward and hanging walls throw down the Penibetic and
Flysch Trough units that constitute the Ronda basin basement. Conversely, the Ronda
basin NW boundary is a dextral transpressive band. Geomorphic indices, useful to
test the recent tectonic activity in areas with low to medium deformation rates, point
110 A. Jiménez-Bonilla et al.

to a current tectonic activity of some of the structures affecting the Ronda basin
(Jiménez-Bonilla et al. 2015, 2017).
When kinematic patterns of this area are integrated within the entire WGA
(including the Western Betics and Northern Rif), the Ronda basin tectonic evolution
appears clearly controlled by the characteristic strain partitioning patterns linked to
the evolution of a progressive arc with outward divergent thrusting (Balanyá et al.
2007; Jiménez-Bonilla et al. 2015; Jiménez-Bonilla 2017). Accordingly, the aim of
this itinerary is to identify the dominant strain partitioning modes within this FTB
kinematic domain and their relationships with the development of the Ronda basin.
Stop 1.1A is devoted to introducing the tectonic organization of this FTB segment.
Stops 1.1B and 1.4 show good examples of characteristic arc-perpendicular short-
ening structures within the Subbetic units and the Ronda basin infill, respectively.
Stops 1.2, 1.3 and 1.5 present normal faults that accommodate arc-parallel stretching
at the Ronda basin margins. The relationships between the Ronda basin infill and its
basement are illustrated in stop 1.6. Finally, stop 1.7 focuses on the development of
transpressional-related structural highs in the northern boundary of the Ronda basin.

4.1.1 Stop 1.1. the Alboran Domain Mountain Front and the Innermost
Fold-and-Thrust Belt

Guarda Forestal viewpoint (36° 46 54.36 N 4° 59 14.19 W). Ronda-El Burgo
road (A-366).
Stop 1.1A. Looking to the south from the viewpoint, there is a panoramic view over
the Alboran Domain mountain front at the northern side of Sierra de las Nieves (up
to 1918 m). From bottom to top in the tectonic pile, we observe (Fig. 3): (a) the
Penibetic units, mostly represented by the Red Beds formation (Upper Cretaceous to
Paleogene); (b) the Flysch Trough units (Jurassic to Paleogene), here severely thinned
by low angle normal faults that produce the tectonic inversion of the Alboran Domain
basal thrust; and (c) the Frontal Units, mainly composed of Triassic and Jurassic

Fig. 3 General view of the Alboran Domain mountain front (left), and the Penibetic ranges (right).
Both pictures were taken nearby Stop 1.1 (see location in Fig. 2)
Miocene-Quaternary Strain Partitioning and Relief Segmentation … 111

Fig. 4 a The geologist Manuel Díaz-Azpiroz at an example of NW-verging fold affecting the Red
Beds formation, b associated stylolitic axial plane cleavage and c calcite veins perpendicular to
the fold axes. Pictures are taken nearby stop 1.1, 600 m towards Ronda along the road A-366 (see
location in Fig. 2)

carbonatic rocks. Bedding, which is especially apparent in the Penibetic, dips to the
SE all along the cross-section.
Looking to the west, we observe the Penibetic reliefs of Sierra Blanquilla (up to
1430 m), in which differences between pale grey competent rocks (Jurassic lime-
stones) and soft rocks (Cretaceous marls) are easily distinguished. Very good expo-
sures of the Jurassic sequence can be found along the road A-366 between this stop
and Puerto del Viento.
Stop 1.1B. Observations within the Penibetic can be completed walking 600 m
towards Ronda along the road A-366. Starting from the viewpoint, we observe first
the White Beds formation (white cherty marly limestones) and later the Red Beds
formation (red marly limestones). Both formations are affected by NW-verging folds,
early to middle Miocene in age (Expósito et al. 2012). Metric to decametric chevron
folds develop a pressure-solution, axial plane cleavage, which is often stylolitic
(Fig. 4). Flexural-slip-folding is evidenced by typical chevron-type morphologies and
the occurrence of calcite fibers on flexural-slip planes (bedding). Pervasive calcite
veins, mostly orthogonal to the hinge lines, accommodate fold-axis parallel extension
(Fig. 4; Expósito et al. 2012).

4.1.2 Stop 1.2. The Grazalema Fault

This stop is located on the road A-372, close to Grazalema (Fig. 5a). There is a
parking place nearby (36° 45 17.74 N 5° 21 31.10 W).
At this stop, we can observe the Grazalema fault scarp, which is marked by a
300 m topographic drop of the NE block (Fig. 5b). The SW block of the Grazalema
112 A. Jiménez-Bonilla et al.

Fig. 5 a Location of the observations and main structures of the Grazalema fault, b main scarp and
c equal area, lower hemisphere plot of fault planes and related slickenlines

fault, topographically uplifted, is composed of Jurassic Penibetic limestones and


dolostones, whilst the NE block is made up of Cretaceous to Palaeogene Penibetic
marly limestones and Miocene sandstones of the Flychs Trough units. Most fault
planes dip steeply towards the NE. In the Cretaceous marly limestones, some S–C
structures and slickenlines on C planes show dominant normal fault slip (Fig. 5c).
The Grazalema fault cuts across the Endrinal range, which corresponds with a late
Miocene to Pliocene, upright antiform (Fig. 5a, b). The vertical slip of the Grazalema
fault is higher than 1.5 km (Jiménez-Bonilla et al. 2015).
Apart from the Grazalema fault, we can also observe some features related to the
early to middle Miocene deformation event:
(1) Close to the village of Grazalema, along road CA-5311, there is a fold train
composed of metric to decametric folds, verging to the NW and affecting an
outcrop of Cretaceous marly limestones. See 1A in Fig. 5a for location (Jiménez-
Bonilla et al. 2015).
(2) Continuing to road A-372, there is another fold train made up of NW-verging,
decametric folds, but deforming in this case the Flysch Trough units. See 1B in
Fig. 5a for location (Jiménez-Bonilla et al. 2015).

4.1.3 Stop 1.3. The El Republicano Fault

Taking the road A-374 towards Villaluenga and before arriving in this town, there is
a dirty road to the left where parking is possible (36° 41 25.81 N 5° 22 19.86 W).
From this parking, we have a nice view of the El Republicano fault trace (Fig. 6).
This fault also downthrows the NE block and cuts across another NE-SW late
Miocene to Pliocene antiform. All fault planes dip steeply to the NE, where younger
Miocene-Quaternary Strain Partitioning and Relief Segmentation … 113

Fig. 6 a Picture and equal area, lower hemisphere plot (dots are slickenlines on fault planes) of
the El Republicano normal fault. b S–C structures in the normal fault zone

rocks crop out (Miocene Flyschs Trough units, see Jiménez-Bonilla et al. 2015). Both
slickenlines and kinematic criteria show that this is a dip-slip dominated, normal fault
(Fig. 6). The vertical slip of this fault surpasses 1 km.
An early to middle Miocene event deformed the Flyschs Trough units that crop
out in the fault hanging wall, giving way to a west-verging imbricate stack (Jiménez-
Bonilla et al. 2015). This stack is evidenced by the alternation of clays (without
vegetation) and sandstones (with vegetation cover). The Flysch Trough units are
currently hosted in a synform in-between the antiforms of stops 2 and 3 (Fig. 2).
This imbricate stack is cut by both the El Republicano and Grazalema normal faults,
which are the most important normal faults in the SW boundary of the Ronda basin
(Figs. 2 and 5, see also Jiménez-Bonilla et al. 2015).

4.1.4 Stop 1.4. The Salinas Fold

This stop is located at road A-374 close to the town of Ronda, in the diversion to a
petrol station (36° 45 33.20 N 5° 09 43.36 W).
Looking to the north, we observe a range within the upper Miocene infill of the
Ronda basin (Fig. 7). This NE–SW range coincides with an antiform whose fold
limbs dip 50°, as observed in resistant calcarenites eroded by deep gullies. Around
this stop, rocks of the Ronda basin infill crop out: calcarenites, conglomerates and
silts.
From this stop, we also have a wonderful panoramic view of the Ronda basin
boundaries: the SW boundary (stops 2 and 3) and the NE boundary, which will be
114 A. Jiménez-Bonilla et al.

Fig. 7 a Picture of the Salinas range and b cross-section including both boundaries of the Ronda
basin (Jiménez-Bonilla 2017; See location in Fig. 2)

visited in stops 5 and 6, looking to the west and east, respectively. Both the SW and
NE boundaries downthrow the Ronda basin and have contributed to accommodate
the WGA-parallel stretching since the late Miocene (Fig. 7). Mountain ranges at both
basin boundaries are mostly NE–SW oriented and coincide with upright late Miocene
to Pliocene antiforms cored in Jurassic limestones. The NW–SE shortening, together
with NE–SW stretching accommodated by the normal faults enlighten the post-late
Miocene strain partitioning mode in this FTB segment of the WGA (Balanyá et al.
2007; Jiménez-Bonilla et al. 2015).

4.1.5 Stop 1.5. Normal Fault Zone at the NE Boundary of the Ronda
Basin

This stop is located along road CA-414, close to the Cuevas del Becerro village
(36° 51 49.11 N 5° 03 18.63 ).
The NE boundary of the Ronda basin is partially faulted. This stop coincides
with one of the segments affected by a NW–SE normal fault zone, which have
produced SW-facing topographic scarps in the Penibetic limestones (Fig. 8). Faults
have listric geometry with decametric vertical displacements. The cumulative vertical
displacement is around 250 m. S–C type structures and slicken linespitches indicate
a dominant dip-slip, normal component of the fault (Jiménez-Bonilla et al. 2015).
Miocene-Quaternary Strain Partitioning and Relief Segmentation … 115

Fig. 8 Normal faults at the NE boundary of the Ronda basin (stop 1.5, see location in Fig. 2). The
discontinuous blue line shows the contact between Lower and Middle Jurassic limestones

4.1.6 Stop 1.6. Ronda Basin Basal Unconformity

This stop is located along the road CA-414, between the villages of Cuevas del
Becerro and Setenil. There is a dirty road to the right (36° 51 49.97 N 5° 06 0.12
W). The best place to park is at the entrance to this dirty road.
This stop shows a cross-section of a segment of the NE boundary of the Ronda
basin, which corresponds here with an unconformity. Walking 250 m to the east
from the parking place, we observe the Ronda basin basement composed of Jurassic
Penibetic limestones with gentle dips to the west. Walking westwards along the road,
we can see these limestones, but with intense karstification. Some of these karsti-
fied channels are filled with conglomerates with rounded limestone clasts, which
constitute the basal deposits of the basin. Finally, well-organized beds of similar
conglomerates are present further west (Fig. 9, see also Jiménez-Bonilla et al. 2015).

Fig. 9 Unconformity between the Penibetic limestones and the conglomerates of the Ronda basin
infill nearby Stop 1.6 (see location in Fig. 2)
116 A. Jiménez-Bonilla et al.

4.1.7 Stop 1.7. NW Boundary of the Ronda Basin

This stop is located at the Olvera castle, where we have a panoramic view (36° 56
06.56 5° 16 04.80 ).
This stop is located at the NW boundary of the Ronda basin, which corresponds to
a dextral transpressive fault zone partitioned into strike-slip faults and folds. Looking
from SW to SE many WSW–ENE-oriented rock alignments, limited by faults, can
be noticed. They define a 20 km long dextral strike-slip fault zone (Fig. 10). This
stop is located in one of these rock alignments where some dextral strike-slip fault
planes can be observed. Looking to the west, we observe a NE–SW mountain range
that coincides with a late Miocene to Pliocene antiform (Fig. 10). To the south, the
Ronda basin infill overlies a basement composed of Triassic Penibetic rocks. The
basin infill dips to the south. Hence, a structural high bounds the basin to the north,
and separates the Ronda intermontane basin from the Guadalquivir foreland basin
(Fig. 2; located 30 km to the NW; Jiménez-Bonilla et al. 2015).

4.2 Itinerary 2.—The Torcal Shear Zone

The Torcal Shear Zone (TSZ) is identified by several disconnected, en-echelon


morpho-structural highs that define an E–W, 70 km long and 5–8 km-wide alignment
(Fig. 11). To the south, the TSZ limits with the Alboran Domain, which in this area is
partially covered by lower-middle Miocene sediments, including in its upper part the

Fig. 10 Geological map of the NW boundary of the Ronda basin (see location in Fig. 2) and equal
area, lower hemisphere plot of faults bounding rock alignments
Miocene-Quaternary Strain Partitioning and Relief Segmentation … 117

Fig. 11 Geological map of the TSZ with its four principal sectors Díaz-Azpiroz et al. (2014), from
west to east: Teba-Peñarrubia ridges (TPS), Valle de Abdalajís Massif (VAM), Torcal de Antequera
Massif (TAM) and Cabras-Camorolos ridges (CCS). The inset shows the location of the TSZ within
the Betic-Rif orogen. Rectangles mark the location of Figs. 14, 16 and 19, where stops of itinerary
2 are shown

so-called La Joya olistostromic complex (Fig. 2, Suades and Crespo-Blanc 2013). To


the north, the TSZ is in contact with Triassic marls and evaporites, as well as upper
Miocene calcarenites.
The structural pattern of the TSZ includes a great variety of brittle-ductile struc-
tures that are interpreted, as a whole, as resulting from a late Miocene (or younger)
dextral transpression (Díaz-Azpiroz et al. 2014; Barcos et al. 2015). This age is based
on different criteria: (a) the interference patterns of the TSZ structures with early-
middle Miocene folds (Expósito et al. 2012, see stop 1.1B); (b) the evidence of trans-
pressive structures affecting upper Miocene calcarenites; and (c) the calculation of
focal mechanisms from crustal earthquakes and geomorphic indices, both indicating
recent strike-slip tectonics within the TSZ (Balanyá et al. 2012; Barcos et al. 2012a;
Díaz-Azpiroz et al. 2014). However, upper Miocene sediments of the El Chorro basin
locally cover some of these structures in the VAM, likely accounting for some defor-
mation diachrony. Transpression was strongly partitioned at several scales. At the
km-scale, several sectors with contrasting structural patterns and kinematics appear
along-strike (Fig. 11). The westernmost sector (Teba-Peñarrubia) represents the tran-
sition between the WGA (itinerary 1) and the TSZ (Jiménez-Bonilla et al. 2013). Two
massifs at the central sector (Valle de Abdalajís and Torcal de Antequera massifs,
VAM and TAM, respectively) accommodate the main, dextral transpressional strain,
and are the principal focus of itinerary 2. The Cabras-Camorolos sector acted as a
termination zone that links the TSZ with the central Betics (Barcos et al. 2012b).
At lower scales, late Miocene dextral transpression was partitioned into kinemat-
ically homogeneous domains that accommodated different components of the bulk
strain (Fig. 12). The TAM presents two main structural domains: an internal domain
(TAM-ID) occupying most of the massif and two narrow outer domains (TAM-OD)
118 A. Jiménez-Bonilla et al.

Fig. 12 Structural features of the TSZ central sector (Díaz-Azpiroz et al. 2014; Barcos et al. 2015).
a Geological and structural map of the TAM and VAM. Within the TAM, the inner domain (ID), and
the northern (NOD) and southern (SOD) outer domains are shown. The location of cross-sections
of the VAM (I-I’, Fig. 18b) and TAM (II-II’, Fig. 22b) is included. b and c Structural data from
the TAM b and VAM c, shown as a lower hemisphere, equal area plots of fault planes with related
slickenlines and slickenfibers (black circles), fold facing directions (asterisks) and finite strain axes
(white triangles in TAM, black squares in VAM)

that define the northern and southern limits. The TAM inner domain presents mainly
shortening structures (folds and reverse faults) oblique (NE-SW striking) to the main
TAM limits (E-W oriented). The TAM outer domains are uplifted with respect to the
neighbouring units, and dextral strike-slip dominated fault zones separate them. The
general structure of the VAM is markedly different. Its southern limit is a discrete
dextral fault zone. In turn, its inner part presents a general imbricate structure defined
by dextral-reverse fault zones, developed occasionally at antiform short-limbs. Both
structures in the inner part are subparallel to the VAM boundaries (ENE-WSW) and
homogeneously distributed within this massif. In both the TAM and the VAM, NW-
SE striking normal faults produce subordinate extension orthogonal to the principal,
shortening structures.
We analyzed this transpression in three stages: 1) description of the finite strain
geometry (orientation and, when possible, magnitude of the three main finite strain
axes) in each domain; (2) comparison of finite strain geometry with results from
the model of triclinic transpression with oblique extrusion (Fernández and Díaz-
Azpiroz 2009) to obtain the main kinematic parameters and boundary conditions
(Díaz-Azpiroz et al. 2014; Barcos et al. 2015); and (3) analogue modelling based
on the kinematic results to reproduce the main structures observed in the TAM
(Barcos et al. 2016). This analysis suggests that the different structural patterns
Miocene-Quaternary Strain Partitioning and Relief Segmentation … 119

observed, resulting from multi-scale strain partitioning, match with a single far-field
vector oriented N099°–109°E (according to the current position of the TSZ; Fig. 13).
Bulk kinematic differences between the TAM and VAM (mainly in convergence
obliquity) would likely arise from local variations in the main boundaries of each
sector (N086°E/73 N at the TAM and N064°E/82S at the VAM). In turn, distinct
strain partitioning styles could be related to the specific kinematics of each sector
and/or to lateral variations of the underlying plastic layer (Triassic evaporite-rich
materials) thickness (Barcos et al. 2015).
We integrate the TSZ kinematics in the general Betic-Rif evolution. After the
main, early-middle Miocene convergent tectonic event that structured the Betic fold-
and-thrust belt (see itinerary 1), oblique convergence with respect to the Alboran
Domain continued active from the late Miocene onwards (Balanyá et al. 2012; Barcos
et al. 2012a). The contact with the Alboran Domain represented a major rheolog-
ical boundary that localized deformation along with a narrow band that would have
resulted in the TSZ. Transpression at this shear zone uplifted the Penibetic formations
respecting the Alboran Domain to the south and the Miocene deposits to the north.
Furthermore, the dextral lateral displacement at the TSZ accommodated the progres-
sive westward migration of the WGA with respect to the rest of the Betic-Rif chain,
thus increasing the protrusion degree of this arc. Analogue models of progressive
arcs reproduce a similar kinematic evolution (Jiménez-Bonilla et al. 2020).
Itinerary 2 (Figs. 14, 16, 19) is easily accessible from either Olvera, the ending
point of itinerary 1, using road A-384 to Campillos, or directly from Ronda via road

Fig. 13 Conceptual model for the late Miocene transpressional event and the multi-scale strain
partitioning in the TSZ central sector, with the main kinematic and far-field parameters, in block
diagram a and equal area, lower hemisphere projection. b Scales in a are approximate and vertical
scale within the TSZ is double than out of this shear zone (Barcos et al. 2015)
120 A. Jiménez-Bonilla et al.

Fig. 14 Google Earth image showing the Teba-Peñarrubia ridges and the location of stops 2.1 and
2.2 (see location in Fig. 11)

A-367. Alternatively, it is possible to start this itinerary at Antequera. We present


here an overall view of the TSZ and the main structures that accommodate dextral
triclinic transpression at its central sector, in the VAM and TAM. The visitor will see
how along-strike partitioning of transpression produced different structural styles in
both massifs. The route would ideally be split into two days. The first day (stops
2.1–2.4) would be dedicated to a general view of the TSZ, the termination zone that
links the TSZ with the WGA, visited in itinerary 1, and the structures of the VAM.
The second day (stops 2.5–2.11) is focused on the TAM. A good option would be to
stay at Antequera. All field observations are based on previous works by the authors
(Díaz-Azpiroz et al. 2014, 2019; Barcos et al. 2015, 2016; Ramírez Prior et al. 2018).
The visitor is referred to these papers for further information.

4.2.1 Stop 2.1. The Teba-Peñarrubia Sector

This stop is located at the Teba castle (36°58 46.78 N 4°55 06.70 W).
Jurassic limestones crop out at the hill where the castle is located. This hill is
the western termination zone of the TSZ, which links with the WGA through the
Almargen ridge. Looking west and south, two NE–SW mountain ranges with lengths
exceeding 10 km, and corresponding to the WGA fold-and-thrust belt, are cut at their
NE end by a WNW–ESE dextral strike-slip fault. This strike-slip fault also marks
the western end of the Teba ridge. Looking east in the foreground, a change in the
structural trend from WSW–ENE in the Teba ridge to WNW–ESE in the Peñarrubia
ridge is noted. Both ridges correspond to hm-scale open antiforms. The Peñarrubia
antiform (with NNE and SSW dipping limbs) is cut at its NE limb by a dextral
Miocene-Quaternary Strain Partitioning and Relief Segmentation … 121

Fig. 15 The majestic Tajo del Molino gorge (looking southwards) crosscutting the Peñarrubia
ridge. Note the south-dipping bedding on the right cliff

Fig. 16 Google Earth image showing the VAM and the location (stop 2.3) of the starting point of
the ca. 7 km Caminito del Rey trail (see location in Fig. 11)
122 A. Jiménez-Bonilla et al.

Fig. 17 a Panoramic view from the main entrance to the Caminito del Rey and the Gaitanejos gorge
entry. From this point northwards, we can observe the uppermost part of the pre-orogenic sequence:
Middle-Upper Jurassic oolitic limestones, Lower Cretaceous White Beds and Upper Cretaceous-
Paleogene Red Beds. The upper Miocene calcarenites unconformably lie over the pre-orogenic
sequence and show spectacular taffoni (upper left). b Giants’ kettles hanged several meters above
the current riverbed. c Early-middle Miocene, north-verging chevron folds deforming the Red Beds.
d Panoramic view from the western margin of the El Hoyo valley with the interpretation of the main
structures that define the two southernmost tectonic imbrications of the VAM (Díaz-Azpiroz et al.
2019). See the text for further details

strike-slip fault, whilst the Teba antiform (with NW and SE dipping limbs) is cut by
a reverse fault at its SE limb. These faults can be appreciated by the associated scarps.
The Teba antiform and reverse fault are interpreted as a compressive bridge limited
by two WNW–ESE strike-slip faults (Jiménez-Bonilla et al. 2013). Looking ESE, in
the background, two separate topographic highs correspond to the VAM (right, stops
2.3 and 2.4) and TAM (left, stops 2.5–2.11).
Miocene-Quaternary Strain Partitioning and Relief Segmentation … 123

Fig. 18 a Panoramic view of the north-verging, imbricate structure of the VAM north of Sierra del
Valle de Abdalajís. b Interpretative cross-section of the whole massif with the location of Fig. 18a.
(Díaz-Azpiroz et al. 2019)

Fig. 19 Google Earth image showing the TAM and the location of stops 2.4–2.11 (See location in
Fig. 11)
124 A. Jiménez-Bonilla et al.

4.2.2 Stop 2.2. The Tajo Del Molino Gorge

Take road MA-465 from Teba and then C-341 to Campillos. After ~1.5 km, park to
the right (36° 58 52.4 N 4° 52 56.7 W). Walk eastwards along the road 120 m to
a bridge over Tajo del Molino.
This deep gorge (Fig. 15) is a product of an antecedent stream prior to the uplift of
the Peñarrubia ridge (Jiménez-Bonilla et al. 2013). This gorge permits us to see the
smooth SSW dip of the Jurassic limestones. Continuing ~500 m to the north along the
gorge, bedding dip progressively decreases defining the Peñarrubia antiform, which
is cut by a WNW-ESE dextral strike-slip fault. At the northern end of the gorge,
an outcrop of upper Miocene conglomerates can be observed. These sediments are
deformed by a WNW-ENE synform with beds dipping 30° to the NNE.

4.2.3 Stop 2.3. Transect Across the Valle de Abdalajís Massif

This stop consists of a ca. 7 km walk along the so-called Caminito del Rey pathway
(Fig. 16), which transects across the VAM. The trail goes through two spectacular
gorges, named Gaitanejos and Los Gaitanes. You may park at 36° 55 44.8 N 4° 48
07.6 W and walk to the entrance (36° 55 56.1 N 4° 47 21.9 W). From this point,
walking is only permitted in one direction, towards the south. After exiting the trail,
you may walk to the El Chorro railway station (36°54 24.7 N 4°45 33.2 W), where
a shuttle will take you back to the parking area. The whole route will take ~4 h. The
best roads to reach the parking lot from Teba (stop 2.2) are C-341, then A-357 to
Campillos, A-7286 to Embalses, and Parque Guadalteba, the road to Embalses del
Guadalhorce and the road to Presa Conde de Guadalhorce. This way is ~15 km in
total.
Along the Caminito del Rey it is possible to have almost continuous geological
observations of the complete pre-orogenic sedimentary sequence, from the Upper
Triassic to the Paleogene, and of the general imbricate structure of the VAM (see
stop 2.4). However, such structure is somewhat different in this sector (compare to
observations at stops 2.4 and 2.7C), which shows fewer imbrications, occasionally
affected by ENE–WSW striking normal faults. It is also possible to appreciate upper
Miocene sediments of the El Chorro basin unconformably lying over the pre-orogenic
sequence.
(A) From the parking lot to the main entrance, there is a ca. 1.4 km walk where
upper Miocene calcirudites and calcarenites extensively crop out. Calcarenites
present striking cross-bedding and spectacular taffoni (Fig. 17a).
(B) At the beginning of the trail (36°55 57.58 N 4°47 17.63 W), from the main
entrance to the entry to the first gorge (Gaitanejos), it is possible to observe,
from south to north (Fig. 17a), the uppermost part of the pre-orogenic sequence:
Upper Jurassic oolitic limestones, Lower Cretaceous whitish marly limestones
(White Beds) and Upper Cretaceous-Paleogene reddish marly limestones (Red
Miocene-Quaternary Strain Partitioning and Relief Segmentation … 125

Beds). Bedding dips moderately to the north, defining the northern limb of a hm-
scale, ENE-WSW trending antiform. Upper Miocene calcareous conglomerates
and sandstones lay unconformably over the Cretaceous-Paleogene rocks. The
rest of the Jurassic sequence (oolitic and nodular limestones) crops out along
the Gaitanejos gorge.
(C) Also at this gorge (36° 55 52.44 N 4° 47 11.96 W), giants kettles appear
hanging above the Guadalhorce current riverbed (Fig. 17b), which suggests
rapid incision.
(D) In the El Hoyo valley, which lies between the two main gorges, mainly White
and Red Beds crop out. These rocks are affected by early-middle Miocene,
north-verging chevron folds (Fig. 17c, looking WNW from 36° 55 28.64 N
4° 46 58.41 W), which are identified along the entire WGA and observed in
stop 1.1B.
(E) Looking east from 36°55 16.59 N4°46 48.15 W, it is possible to appreciate
part of the characteristic imbricate structure of the VAM (Fig. 17d). The Sierra
the Huma is in the background of the view. An open synform with Jurassic
limestones at the nucleus defines the characteristic flat summit of this range.
Triassic evaporite-rich marls and dolostones crop out at the northern limb, and
overthrust Cretaceous-Paleogene marly limestones. The southern limb dips
gently to the north and rapidly recovers to define an antiform (out of the image)
with a subvertical southern limb, which appears in the foreground, to the right,
at the other side of the river. The slope shows, to the south, dolostone beds that
overthrust the Cretaceous-Paleogene marly limestones. Further to the south, it
is possible to observe the conformable contact between the Triassic formations
and the Jurassic limestones, which are incised by the Los Gaitanes gorge.
In the central part of the view, the Cretaceous-Paleogene marly limestones lie
conformably over the Jurassic limestones, which show north-verging shortening
structures: an antiform and a reverse fault. This structure constitutes the base
of the next imbrication to the north after the Sierra de Huma.
(F) Just before entering the Los Gaitanes gorge (36°55 2.75 N4°46 27.45 W),
it can be seen, at the subvertical southern limb of the Sierra de Huma antiform
(see above), the stratigraphic contact between the Triassic and Lower-Middle
Jurassic formations. The path along this gorge crosscuts the entire Jurassic
sequence (oolitic—nodular—oolitic imestones).
(G) The opening of the gorge to the El Chorro reservoir (36° 54 56.85 N4° 46
18.66 W) coincides with the Jurassic top layers in contact with the Cretaceous-
Paleogene formations. Exiting the trail, the White Beds (with chert nodules,
36° 54 57.06 N4° 46 10.75 W) underlie the Red Beds (36° 54 55.74 N4° 46
10.63 W).
(H) In the village of El Chorro (36° 54 30.64 N4° 45 36.41 W), where the route
ends, it is possible to observe phyllites belonging to the Malaguide complex
(Alboran Domain). The contact with the Penibetic formations does not crop
out along the path, but it is the dextral transpressive fault zone that defines the
southern limit of the TSZ in the VAM.
126 A. Jiménez-Bonilla et al.

4.2.4 Stop 2.4. Panoramic View of the Northern VAM Imbricate


Structure

From the parking lot of the Caminito del Rey main entrance, make a left turn and
follow this road eastwards. After ca. 1 km, take right following signs to Antequera
and Campillos. From this point onwards, follow signs to Antequera and V. Abdalajis.
In the first part of the way (3.5 km, up to 36° 56 47.1 N 4° 47 31.5 W), the road
crosscuts upper Miocene calcareous conglomerates and sandstones with striking
cross-bedding. Between 3.5 and 9.5 km, the road goes along the northern edge of the
VAM, with either Jurassic limestones or Cretaceous-Paleogene marly limestones to
your right. At 36° 58 17.3 N 4° 44 27.8 W, looking south, there is a panoramic
view of the northern edge of the VAM. A Jurassic limestone cliff corresponds to
the hanging wall of a WSW–ESE striking, southward dipping reverse fault that puts
this formation over the Cretaceous-Paleogene marly limestones (vegetated gentle
slope below the cliff). This is the northernmost imbrication of this massif. In the
background, limestone crests mark other imbrications. Looking SSE, the Sierra de
Huma (see stop 2.3E) summit is apparent. After 16 km from the starting point, at the
crossroad (36° 57 26.1 N 4° 41 26.7 W), take road A-343 to the left. Right before
km 19 (36° 58 45.2 N 4° 39 00.4 W), there is a sign to the left to Las Lagunillas.
You may stop here or at the farmer entrance to your right.
Looking westwards, there is a panoramic view of the northernmost imbricate
structure of the VAM (see also stop 2.7C). From south to north, three of these imbri-
cations appear. They are all limited to the north by south-dipping reverse faults that
put Jurassic limestones over Cretaceous-Paleogene marly limestones (Fig. 18).

4.2.5 Stop 2.5. Northern Boundary of the Torcal de Antequera Massif

Drive southwards from Antequera along road A-343. After km 9 benchmark, at


the crossroad, continue ahead along road A-7075 towards Torcal de Antequera and
Villanueva de la Concepción. Short after the camping, at around km 47.5 benchmark
there are two dirt tracks, one at each side of the road. Stop on the left one (36° 59
06.7 N 4° 31 28.6 W).
Looking westwards (Fig. 20), the northern TAM outer domain can be observed.
In an east-facing outcrop, two reddish limestone layers can serve as markers, dipping
to the north (to the right of the view). A subvertical fault descends these layers with
respect to the rest of the TAM outer domain. Upwards, the fault zone describes a splay
structure with curved, south-dipping surfaces. Further to the north, at the lower part
of the outcrop, the reddish layers seem to define a very open, incomplete synform.
However, a thorough look shows this is an apparent effect; the structure results actu-
ally from the combined displacements of several small-scale, south-dipping faults,
which resemble that of the main fault surface. Overall, these faults define what we
could describe as a positive flower-like structure (we only see its northern half),
although it is not a typical one, as we will see in stop 2.8.
Miocene-Quaternary Strain Partitioning and Relief Segmentation … 127

Fig. 20 Northern limit of the TAM with the interpretation of the asymmetric flower-like structure
(Díaz-Azpiroz et al. 2014). Note that the structure has an important right-lateral component

4.2.6 Stop 2.6. The Boca Del Asno antiform

Continue along road A-7075 heading to the Torcal de Antequera. After km 45 bench-
mark, there are three consecutive closed curves in a place known as Boca del Asno.
Between the first two curves (36° 58 24.2 N 4° 29 42.2 W) there is an excellent
panoramic view of the eastern edge of the TAM (Fig. 21). In the central part of the
slope, grey limestone beds dip gently to the north, and define (down left) an antiform
hinge (Boca del Asno antiform, Fig. 21), which is kinematically equivalent to the
El Camorro antiform (stop 2.7B), but located further to the SE, in an en-échelon
array (Fig. 12). Short after km 42 benchmark, at the crossroad (36° 57 51.3 N
4° 30 39.5 W), make a right turn. This road finishes at the Torcal de Antequera
Interpretation Center. In the first kilometer, to the right, grey carbonatic rocks with
very thick bedding (occasionally massive) crop out. The lower part corresponds to
uppermost Triassic dolostones, which are followed by Lower-Middle Jurassic oolitic
limestones. Leave the car at the parking lot at the end of the road. Before going to
stop 2.7, it is possible to have a panoramic view to the south of the TAM from the
Mirador de las Ventanillas, 50 m away from the Interpretation Center.

4.2.7 Stop 2.7. Structural Pattern of the TAM Inner Domain

This stop consists of a 1.5 km walk from the Interpretation Center. Follow the road
from the parking lot to the NE. After ca. 300 m (36° 57 20.3 N 4° 32 29.7 W),
abandon the road and take a narrow track to the left. Walk towards the NW along the
foot of the Camorro de las Siete Mesas (an NW-SE elongated topographic high to
128 A. Jiménez-Bonilla et al.

Fig. 21 SE limit of the TAM inner domain viewed from the east and structural interpretation (Díaz-
Azpiroz et al. 2014). Rocks cropping out at the slope correspond to the lowermost layers of the
Penibetic sequence: Upper Triassic dolostones and Lower Jurassic limestones. Gently north-dipping
limestone layers define to the south an open SE-verging antiform (Boca del Asno antiform). The
overturned limb is cut by a reverse fault

your right). After ca. 1.3 km (36° 57 37.7 N 4° 32 52.5 W), there is an excellent
panoramic view to the west.
Three different observations can be made at this stop. (A) This is a good place
to observe the Upper Jurassic formations of the Penibetic stratigraphic sequence
(already presented in stop 2.3F–G). Looking eastwards, the Camorro de las Siete
Mesas presents nodular limestones at its base and oolitic limestones at its top, both
Upper Jurassic. Nodular limestones are reddish colored and rich in ammonite fossils
(it is possible to see some in this outcrop). Weathering has created a conspicuous and
contrasted layered structure that represents the most typical morphological feature
of the Torcal de Antequera (locally known as bollos). Looking to the WSW, the grey-
colored Upper Jurassic oolitic limestones produce a different weathering morphology
with dolines defining an intricate pattern locally known as callejones. The contact
between these two formations is an NW-SE striking fault with the uplifted wall
(nodular limestones) to the NE. This is one of the NW-SE striking normal faults that
produce extension parallel to the main shortening structures of the TAM inner domain
(see stop 2.9). The fault plane is not apparent here. (B) Looking to the WNW, in the
middle part, there is an excellent panoramic view of the main shortening structures
of the TAM inner domain (Fig. 22). To the right, the higher hill with limestones is the
Camorro Alto, corresponding to an hm-scale, NE–SW striking, SE-verging antiform
similar to that observed at Boca del Asno (stop 2.6A). Limestone beds mark the
Miocene-Quaternary Strain Partitioning and Relief Segmentation … 129

Fig. 22 a Shortening structures of the TAM inner domain. The hill coincides with the SE-verging
Camorro Alto antiform, whose short limb defines the SE slope. Related to this limb, the reverse El
Navazo fault puts Jurassic limestones (grey rocks) over Cretaceous-Paleogene Red Beds (cultivated
fields), which in turn define the El Navazo synform hinge zone. b Schematic cross-section showing
the main structural pattern of the TAM (see the location in Fig. 12) with the location of the panoramic
view in Fig. 22a (Díaz-Azpiroz et al. 2014)

subvertical SE limb. A NE–SW striking, SE-verging reverse fault (El Navazo fault)
is developed at this limb, and puts Jurassic limestones over Cretaceous-Paleogene
marly limestones, recognizable by the gentle relief (the El Navazo valley). These
same rocks appear at the hinge of the next synform to the SE (El Navazo synform).
(C) Looking to the west, in the background, there is a traverse panoramic view
of the neighbouring VAM imbricate structure (see Fig. 18). From south to north,
one can observe the Sierra de Huma synform (see stop 2.3E) with its characteristic
flat summit, the Sierra del Valle de Abdalajís antiform, and two smaller tectonic
imbrications limited to the north by south-dipping reverse faults (see stop 2.4).

4.2.8 Stop 2.8. Panoramic View to the North of the TAM

From the previous stop walk to the NE leaving the Camorro de las Siete Mesas to
your right. After ca. 400 m, there is an excellent view to the NNE (36° 57 55.9 N
4° 33 00.9 W).
130 A. Jiménez-Bonilla et al.

Fig. 23 NE looking panoramic view from the Camorro de las Siete Mesas. It is possible to observe
the Triassic rocks to the north of the TAM and the southern view of the flower-like structure that
defines the northern TAM outer domain, whose northern view was observed at stop 2.5

From this point, we it is possible to observe the transpression-related uplift of


the TSZ with respect to the northern formations, upper Miocene calcarenites in the
foreground to the NNW, and Triassic evaporite-rich marls to the NNE (Fig. 23). The
relative uplift of the northern TAM outer domain (see stop 2.5) respecting the rocks
located immediately to the south is also evident (Fig. 23).

4.2.9 Stop 2.9. Normal Faults of the TAM Inner Domain

Go back to the road and turn left to the east. Walk ~200 m to 36° 57 23.7 N 4° 32
19.9 W. You can also go back to the parking lot, take the car and drive to 36° 57
27.3 N 4° 32 14.5 W, where parking is also possible.
Looking to the SE, there is an almost orthogonal view of two of the NW-SE
striking normal faults of the TAM inner domain, clearly marked by the stratigraphic
top of the nodular limestones (Fig. 24). These faults dip to the NE (left), present a
somewhat listric geometry that tilted the hanging walls and contribute to extend the
TAM inner domain along a NE–SW direction. The subtle difference between the
tectonic slip and the topographic drop suggests a recent age for these structures.

4.2.10 Stop 2.10. Southern Boundary of the Torcal de Antequera Massif

Go back to the road and then drive to road A-7075. Turn right to Villanueva de la
Concepción, cross the village along Blas Infante av., and at the crossroad, continue
ahead following signs to Almogía. Drive less than 2 km along road MA-3403 and
take a right to La Joya. After 5 km along this road, where a Point of Geological
Interest (Punto de interés geologico) is marked (36° 56 01.6 N 4° 35 08.3 W),
make a right turn. Continue 2 km more to Cortijo del Robledillo (36° 56 49.6 N
4° 35 02.1 W). Leave the car, abandon the dirt track, and walk eastwards around
700 m leaving the limestone cliff to your left (36° 56 49.3 N 4° 34 38.1 W).
Miocene-Quaternary Strain Partitioning and Relief Segmentation … 131

Fig. 24 NW-SE striking, NE-dipping normal fault. Its slip is estimated from the stratigraphic top of
the reddish nodular limestones. Note the small difference between the fault throw and the resulting
topographic drop (Díaz-Azpiroz et al. 2014)

This is the southern limit of the southern TAM outer domain, which is uplifted
respecting the Alboran Domain units, including the middle Miocene La Joya olis-
tostromic complex, to the south. The limit is defined by a 20 m-thick shear zone
affecting the contact between the Jurassic limestones to the north and Cretaceous-
Paleogene Red Beds to the south. There are remarkable differences in the structural
style depending on the rock type: discrete, meter- to decimeter-spaced fault planes
in limestones, whereas marly limestones show centimeter- to decimeter-spaced S-
C like structures. This rheological partitioning is similar for all strike-slip, reverse
(see stop 2.11) and oblique faults of the TSZ, which suggests that regardless of the
kinematics, deformation occurred under similar conditions and, thus, coevally. In all
cases, slickenlines and slickenfibers are clearly seen on discrete fault and C planes.
In this outcrop, such lines present a fairly constant attitude, with low to moderate
west plunges (Fig. 25a). This, in combination with the asymmetry of S–C struc-
tures on subhorizontal observation surfaces (Fig. 25b), points to a dominant dextral
displacement with a subordinate reverse slip component.
132 A. Jiménez-Bonilla et al.

Fig. 25 Outcrop-scale structures at the southern TAM outer domain (Díaz-Azpiroz et al. 2014).
a Calcite slickenfibers on a limestone fault plane. b Asymmetric, dextral S–C like structures in
marly limestones (red beds)

4.2.11 Stop 2.11. NE–SW Reverse Fault Within the TAM Inner Domain

Go back to the vehicle and continue along the dirt track for around 3 km, until close
to Cortijo de la Fuenfría (36° 57 37.3 N 4° 35 55.9 W).
In this stop, the El Navazo reverse fault (Fig. 26), which we observed in a
panoramic view from the Camorro de las Siete Mesas (stop 2.7B), can be exam-
ined in detail. The Jurassic limestones dip to the south, defining the SE limb of the El
Camorro antiform, and overthrust the Cretaceous-Paleogene Red Beds. Some meters
to the SE, the bedding of these same marly limestones is subvertical or even dipping

Fig. 26 a The reverse El Navazo fault (see stop 2.7B) puts Jurassic limestones (grey rocks in
the background) over Cretaceous-Paleogene Red Beds (in the foreground), which show S–C like
structures. b Detailed view (location marked by red square) of S-C like structures, whose asymmetry
suggests apparent top-to-the-SE displacement (Díaz-Azpiroz et al. 2019)
Miocene-Quaternary Strain Partitioning and Relief Segmentation … 133

steeply to the north, which is coherent with their position in the short, eventually
overturned, limb. In the fault zone, S–C like structures obliterated the bedding of
the Red Beds. The high plunge of slickenfibers on C planes and the asymmetry of
structures suggest main reverse kinematics for this fault (Fig. 26b).

Acknowledgements Financial support from Project PGC2018-100914-B-100 of the Spanish


Ministry of Science, Innovation, and Universities, and Project UPO-1259543 of the Andalusian
Office of Economy and Knowledge. Edition guidance by S. Mukherjee and thorough review by
A. Azor are gratefully acknowledged. Thanks to Marion Schneider, Annett Buettener, Boopalan
Renu, Alexis Vizcaino, Doer the Mennecke-Buehler, and the proofreading team (Springer). Dutta
and Mukherjee (2021) summarize this article.

References

Acosta-Vigil, A., Rubatto, D., Bartoli, O., Cesare, B., Meli, S., Pedrera, A., et al. (2014). Age of
anatexis in the crustal footwall of the Ronda peridotites, S Spain. Lithos, 210–211, 147-167.
Balanyá, J. C., Crespo-Blanc, A., Díaz Azpiroz, M., Expósito, I., & Luján, M. (2007). Structural
trend line pattern and strain partitioning around the Gibraltar Arc accretionary wedge: Insights
as to the mode of orogenic arc building. Tectonics, 26. https://doi.org/10.1029/2005tc001932.
Balanyá, J. C., Crespo-Blanc, A., Díaz-Azpiroz, M., Expósito, I., Torcal, F., Pérez-Peña, V., et al.
(2012). Arc-parallel vs back-arc extension in the Western Gibraltar arc: Is the Gibraltar forearc
still active? Geologica Acta, 10, 249–263.
Barcos, L., Expósito, I., Balanyá, J. C., & Díaz-Azpiroz, M. (2012a). Levantamiento tectónico
asociado a transpresión en el Penibético de la Sierra del Valle de Abdalajís (Béticas): Análisis
estructural y geomorfológico. Geo-Temas, 13, 507–601.
Barcos, L., Jiménez-Bonilla, A., Expósito, I., Balanyá, J. C., & Díaz-Azpiroz, M. (2012b). Reparto
de la deformación en la terminación oriental de la Zona de Cizalla del Torcal (Béticas, S España).
Geogaceta, 56, 23–26.
Barcos, L., Balanyá, J. C., Díaz-Azpiroz, M., Expósito, I., & Jiménez-Bonilla, A. (2015). Kinematics
of the Torcal Shear Zone: Transpressional tectonics in a salient–recess transition at the northern
Gibraltar Arc. Tectonophysics, 663, 62–77. https://doi.org/10.1016/j.tecto.2015.05.002.
Barcos, L., Díaz-Azpiroz, M., Balanyá, J. C., Expósito, I., Jiménez-Bonilla, A., & Faccena, C.
(2016). Analogue modelling of inclined, brittle–ductile transpression: testing analytical models
through natural shear zones (external Betics). Tectonophysics, 682, 169–185.
Braga, J. C., Martín, J. M., & Quesada, C. (2003). Patterns and average rates of Late Neogene–Recent
uplift of the Betic Cordillera, SE Spain. Geomorphology, 50, 3–26.
Comas, M. C., Platt, J. P., Soto, J. I., & Watts, A. B. (1999). 44. The origin and tectonic history
of the Alboran Basin: insights from Leg 161 results. In R. Zahn, M. C. Comas, A. Klaus (Eds.),
Proceedings of the ocean drilling program scientific results (Vol. 161, pp. 555–580).
Díaz-Azpiroz, M., Barcos, L., Balanyá, J. C., Fernández, C., Expósito, I., & Czeck, D. M. (2014).
Applying a general triclinic transpression model to highly partitioned brittle–ductile shear zones:
a case study from the Torcal de Antequera massif, external Betics, southern Spain. Journal of
Structural Geology, 68, 316–336.
Díaz-Azpiroz, M., Balanyá, J. C., Expósito, I., Barcos, L., & Jiménez-Bonilla, A. (2019). La zona
de cizalla del Torcal (cinturón de pliegues y cabalgamientos bético). Un laboratorio natural para
analizar transpresión triclínica con alto reparto de la deformación. Geo-Guías, 11, 245–254.
Duggen, S., Hoernle, K., Van den Bogaard, P., & Harris, C. (2004). Magmatic evolution of the
Alboran region: The role of subduction in forming the Western Mediterranean and causing the
Messinian Salinity Crisis. Earth and Planetary Science Letters, 218, 91–108.
134 A. Jiménez-Bonilla et al.

Durand-Delga, M., Rossi, P., Olivier, P., & Puglisi, D. (2000). Situation structurale et nature ophi-
olitique des roches basiques jurassiques associées aux flyschs maghrébin du Rif (Maroc) et de
Sicile (Italie). Comptes Rendues de la Academie des Sciences, 331, 29–38.
Dutta, D., & Mukherjee, S. (2021). Introduction to Structural Geology and Tectonics Field Guide-
book—Volume 1. In S. Mukherjee (Ed.), Structural Geology and Tectonics Field Guidebook—
Volume 1. Springer Nature Switzerland AG. Cham. pp. xi-xvi. ISBN: 978-3-030-60142-3.
Expósito, I., Balanyá, J. C., Crespo-Blanc, A., Díaz-Azpiroz, M., & Luján, M. (2012). Overthrust
shear folding and contrasting deformation styles in a multiple decollement setting, Gibraltar Arc
external wedge. Tectonophysics, 576–577, 86–98.
Fernández, C., & Díaz-Azpiroz, M. (2009). Triclinic transpression zones with inclined extrusion.
Journal of Structural Geology, 31(10), 1255–1269.
García-Dueñas, V., Balanyá, J. C., & Martínez-Martínez, J. M. (1992). Miocene extensional detach-
ments in the outcropping basement of the Northern Alborán basin and their tectonic implications.
Geo-Marine Letters, 12, 88–95.
González-Delgado, J. A. (coord.), Civis, J., Dabrio, C. J., Goy, J. L., Ledesma, S., País, J., et al.
(2004). Cuenca del Guadalquivir. In J. A. Vera (Ed.), Geología de España (pp. 543–550). Madrid:
SGE-IGME.
Hindle, D., & Burkhard, M. (1999). Strain, displacement and rotation associated with the formation
of curvature in fold belts; the example of the Jura Arc. Journal of Structural Geology, 21, 1089–
1101.
Jiménez-Bonilla, A. (2017). Along-strike segmentation and basin evolution in curved fold-and-
thrust belts: The study case of the northern Gibraltar Arc. (Ph.D. Thesis), Pablo de Olavide
University, 276pp.
Jiménez-Bonilla, A., Barcos, L., Expósito, I., Balanyá, J. C., & Díaz-Azpiroz, M. (2013). La Zona
Transversal de Peñarrubia-Almargen (Béticas): Tectónica transpresiva tardía y segmentación del
relieve. Geogaceta, 55, 7–10.
Jiménez-Bonilla, A., Expósito, I., Balanyá, J. C., Díaz-Azpiroz, M., & Barcos, L. (2015). The role
of strain partitioning on intermontane basin inception and isolation, External Western Gibraltar.
Journal of Geodynamics, 92, 1–17. https://doi.org/10.1016/j.jog.2015.09.001.
Jiménez-Bonilla, A., Expósito, I., Balanyá, J. C., & Díaz-Azpiroz, M. (2017). Strain partitioning
and relief segmentation in arcuate fold-and-thrust belts: A case study from the Western Betics.
Journal of Iberian Geology, 43, 497–518.
Jiménez-Bonilla, A., Crespo-Blanc, A., Balanyá, J. C., Expósito, I., & Díaz-Azpiroz, M. (2020).
Analog models of fold-and-thrust wedges in progressive arcs: A comparison with the Gibraltar
Arc external wedge. Frontiers in Earth Science, 8. https://doi.org/10.3389/feart.2020.00072.
Luján, M., Crespo-Blanc, A., & Balanyá, J. C. (2006). The Flysch Trough thrust imbricate (Betic
Cordillera): A key element of the Gibraltar Arc orogenic wedge. Tectonics, 25, 1–17.
Macedo, J., & Marshak, S. (1999). Controls on the geometry of fold-thrust belt salient. Geological
Society of America Bulletin, 111, 1808–1822.
Marshak, S. (1988). Kinematics of orocline and arc formation in thin-skinned orogens, Tectonics,
7, 73–86.
Marshak, S. (2004). Salients, recesses, arcs, oroclines, and syntaxes—A review of ideas concerning
the formation of map-view curves in fold-thrust belts. In K.R. McClay (Ed.), Thrust tectonics
and hydrocarbon systems (Vol. 82, pp. 131–156). AAPG Memoir.
Martín-Algarra, A. (1987). Evolución geológica alpina del contacto entre las zonas internas y
externas de la Cordillera Bética. (Tesis Doctoral), Universidad de Granada, 1171pp.
Mazzoli, S., & Helman, M. (1994). Neogene patterns of relative plate motion for Africa-Europe:
Some implications for recent central Mediterranean tectonics. Geologische Rundschau, 83, 464–
68.
McCaffrey, R. (1991). Slip vectors and stretching of the Sumatra fore arc. Geology, 19, 881–884.
Platt, J. P., & Whitehouse, M. J. (1999). Early Miocene high-temperature metamorphism and rapid
exhumation in the Betic Cordillera (Spain): Evidence from U-Pb zircon ages. Earth and Planetary
Science Letters, 171, 591–605.
Miocene-Quaternary Strain Partitioning and Relief Segmentation … 135

Ramírez Prior, A., Díaz-Azpiroz, M., & Balanyá, J. C. (2018). Caracterización geológica del
cinturón de pliegues y cabalgamientos bético en el transecto del “Caminito del Rey” (Málaga).
Bases para la interpretación de su patrimonio geológico. Geogaceta, 63, 75–78.
Royden, L. H. (1993). Evolution of retreating subduction boundaries formed during continental
collision. Tectonics, 12, 629–638.
Royden, L. H., & Burchfiel, B. C. (1989). Are systematic variations in thrust belt styles related to
plate boundary processes? (The western Alps versus the Carpatians). Tectonics, 8, 51–61.
Sánchez-Gómez, M., Balanyá, J. C., García-Dueñas, V., & Azañon, J. M. (2002). Intracrustal
tectonic evolution of large lithosphere mantle slabs in the western end of the Mediterranean orogen
(Gibraltar Arc). In G. Rosenbaum, G. S. Lister (Eds.), Reconstruction of the Alpine-Himalayan
Orogen. Journal of Virtual Explorer, 8, 23–34.
Sanz de Galdeano, C., & Alfaro, P. (2004). Tectonic significance of the present relief of the
BeticCordillera. Geomorphology, 63, 175–190.
Sosson, M., Morillon, A. C., Bourgois, J., Feraud, G., Poupeau, G., & Saint-Marc, P. (1998).
Late exhumation stages of the Alpujarride Complex (western Betic Cordillera, Spain): New
thermochronological and structural data on Los Reales and Ojen nappes. Tectonophysics, 285,
253–273.
Spakman, W., & Wortel, M. J. R. (2004). A tomographic view on western Mediterranean
geodynamics. In W. Cavazza, F. Roure, W. Spakman, G.M. Stampfli, & P. Ziegler (Eds.),
The TRANSMED Atlas-The Mediterranean region from crust to mantle (pp. 31–52). Berlin,
Heidelberg: Springer.
Suades, E., & Crespo-Blanc, A. (2013). Gravitational dismantling of the Miocene mountain front of
the Gibraltar Arc system deduced from the analysis of an olistostromic complex (western Betics).
Geologica Acta, 11, 215–229.
Torné, M., Fernández, M., Comas, M. C., & Soto, J. I. (2000). Lithospheric structure beneath
Alboran Basin: Results from 3D gravity modelling and tectonic relevance. Journal of Geophysical
Research, 105, 3209–3228.
Tubía, J. M., & Gil Ibarguchi, J. I. (1991). Eclogites of the Ojén nappe: A record of subduction in
the Alpujarride Complex (Betic Cordilleras, southern Spain). Journal of the Geological Society,
London, 148, 801–804.
Weil, A. B., & Sussman, A. J. (2004). Classifying curved orogens based on timing relationships
between structural development and vertical-axis rotations. Geological Society of America Special
Paper, 383, 1–15.
Vera, J. A., Arias, C., García-Hernández, M., López-Garrido, A. C., Martín-Algarra, A., Martín-
Chivelet, J., et al. (2004). Las zonas externas Béticas y el paleomargen Sudibérico. In J. A.
Vera (Ed.), Geología de España (pp. 354–361). Madrid: Sociedad Geológica de España-Instituto
Geológico y Minero de España.
The Southern Iberian Shear Zone (SW
Spain): Inclined Transpression Related
to Variscan Oblique Convergence
in a HT/LP Metamorphic Belt

Manuel Díaz-Azpiroz and Carlos Fernández

Abstract In this field guide, we present the boundary between the Ossa-Morena
and South Portuguese Zones, two of the main domains of the Iberian Massif, in the
northern central Huelva province (SW Spain). This boundary forms part of the Rheic
Ocean suture, one of the major sutures of the Variscan belt in Europe, and was formed
by the Paleozoic sinistral oblique convergence and collision between Avalonia and
Armorica in SW Iberia. Strain was partitioned between mainly shortening structures
and left-lateral, eventually transpressional, shear zones. The former suture portrayed
in this guide presents an oceanic domain with metasediments and MORB-derived
metabasites, and a former continental margin with HT/LP metamorphic rocks. The
oceanic-derived rocks are partially deformed by the Southern Iberian Shear Zone,
an excellent example of ductile triclinic transpression. The field guide is organized
in four transects across the Ossa-Morena–South Portuguese boundary, particularly
focused on the oceanic metabasic unit and the Southern Iberian Shear Zone, to show:
(1) the old, deep structure of this orogenic suture; (2) ductile mesostructures related to
complex 3D transpression; and (3) interesting HT/LP rocks. Transects are achievable
on foot and are close to each other, thus the whole trip could be completed in 2–3 days.
The field area is located in the Sierra de Aracena and Picos de Aroche Natural Park.
It is easy to reach and has varied accommodation facilities and additional geological,
natural, cultural and gastronomic attractions. This field trip could be combined with
that to the Western Gibraltar Arc northern branch, just 220 km away, and presented
also in this issue.

M. Díaz-Azpiroz (B)
Department of Physical, Chemical and Natural Systems, Universidad Pablo de Olavide, Crtra.
Utrera Km 1, 41013 Seville, Spain
e-mail: mdiaazp@upo.es
C. Fernández
Department of Earth Sciences, Universidad de Huelva, Campus de El Carmen, 21071 Huelva,
Spain
e-mail: fcarlos@uhu.es

© Springer Nature Switzerland AG 2021 137


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_5
138 M. Díaz-Azpiroz and C. Fernández

1 Introduction

The displacement of tectonic plates on the Earth’s spherical surface is defined


by small circles around specific rotation poles, thus it is normally oblique to
tectonic boundaries (e.g., Philippon and Corti 2016). Equivalent oblique situations
are produced extensively at smaller-scale structures, such as major orogenic bound-
aries and shear zones. In oblique convergent cases, deformation can be partitioned
between mainly shortening and strike-slip zones or accommodated at transpressional
zones where shortening normal to the zone is combined with zone-parallel simple
shearing. At the usual scale of these structures, we can assume that the displace-
ment vector between deforming blocks related by the transpressional zone (i.e.,
far-field vector, Jones et al. 2004) is presumably subhorizontal (subparallel to the
overall Earth surface at a specific place). Consequently, the orientation of the simple
shearing direction on the shear zone plane is directly related to the inclination of
that shear zone. As such, simple shear is horizontal in vertical transpressional zones
(Sanderson and Marchini 1984; Fossen and Tikoff 1993) and oblique when these
are inclined (e.g., Lin et al. 1998). In the latter case, the resulting deformation is
typically triclinic, that is, none of the three finite deformation axes coincides with
the external reference frame defined by the deformed zone and the vorticity vector.
This situation can be observed in several natural ductile shear zones where foliation
attitude remains roughly constant whereas the orientation of tectonic lineations is
strongly variable. After the definition and modeling of such structures, the reported
number of natural cases has greatly increased in the last years (Díaz Azpiroz et al.
2019). However, well-studied examples are still needed.
The Paleozoic convergence and collision between Avalonia and Armorica at SW
Iberia were mostly oblique and produced strain partitioning between mainly short-
ening structures and left-lateral shear zones, and also to more complex transpres-
sional deformation zones (e.g., Crespo-Blanc and Orozco 1988; Díaz-Azpiroz and
Fernández 2005; Díaz-Azpiroz et al. 2006; Simancas et al. 2003, 2005), thus repre-
senting an excellent place to analyze triclinic transpression. Moreover, lithologically,
uncommon HT/LP rocks characterize this former suture (Bard 1969; Crespo-Blanc
and Orozco 1991; Díaz-Azpiroz et al. 2006) and make it even more interesting.
Accordingly, in this field guide, we present a view of this old, eroded boundary
(loosely identified as the Rheic Ocean suture, although its exact former location is
still controversial, see for example, Nance et al. 2010; Kroner and Romer 2013;
Pérez-Cáceres et al. 2015), with particular attention to the Southern Iberian Shear
Zone (Crespo-Blanc and Orozco 1988).
The Southern Iberian Shear Zone (SW Spain): Inclined … 139

2 Geological Setting

The Variscan belt in Western Europe (Fig. 1) resulted from the Paleozoic convergence
and subsequent collision between Gondwana, along with some related microconti-
nents (e.g., Armorica), and Avalonia (e.g., Matte 2001; Simancas et al. 2005; Nance
et al. 2010; Kroner and Romer 2013). The Iberian Massif represents the largest
and best exposed portion of the European Variscan orogen. Based on the seminal
works of Lotze (1945) and Julivert et al. (1972), the Iberian Massif was divided
into several domains with distinctive stratigraphic, paleontological, petrological,
and structural features. The two major domains appearing at the SW-most sector
of the Iberian Massif are the Ossa-Morena Zone (OMZ) and the South Portuguese
Zone (SPZ), which represent two former continental blocks (likely Armorica and
Avalonia, respectively) once separated by the Rheic Ocean, that collided during the
Variscan orogenic cycle in Devono–Carboniferous times. Therefore, the OMZ, the
SPZ, and their boundary constitute an excellent example of an old orogenic belt with
an oceanic suture, now profoundly eroded. The IBERSEIS seismic profile (Simancas
et al. 2003) shows a clear image of the deep crust of these two zones and their contact.
Three main units appear at the major tectonic boundary between the OMZ and
the SPZ (Fig. 2): (1) the southernmost edge of the OMZ, whose metasedimentary
sequence would represent the former continental margin of Armorica; (2) the Beja-
Acebuches metabasites (BAM), with oceanic petrological and chemical affinities;
and (3) the Pulo do Lobo unit (Bard 1969; Ribeiro and Silva 1983), a cryptic unit
interpreted either as a former accretionary prism (e.g., Eden 1991; Silva et al. 1990;
Onézime et al. 2003) or as the northernmost edge of the SPZ (Oliveira 1990; Pérez-
Cáceres et al. 2015). The first two units constitute a high temperature/low pressure
(HT/LP) belt defined by Bard (1969) as the Aracena metamorphic belt (AMB); they
were defined, respectively, as the continental and oceanic domains of this belt (Castro
et al. 1996a, b). Further to the north, the Cubito-Moura unit, with high-pressure

Fig. 1 Possible plate


configuration of the Variscan
belt at late Carboniferous
times (simplified from Matte
2001; Simancas et al. 2005
and Pérez-Cáceres et al.
2015). The orange star marks
the approximate location of
the field area
140 M. Díaz-Azpiroz and C. Fernández

Fig. 2 Geological map of the boundary between Ossa-Morena and South Portuguese Zones at
SW Iberian Peninsula (from Castro et al. 1996a), with the main tectonic domains and shear zones
(SISZ: Southern Iberian Shear Zone; CSZ: Calabazares Shear Zone; CASZ: Cortegana-Aguafría
Shear Zone; BVFZ: Beja-Valdelarco fault zone). The four proposed transects (T1 to T4) and the
approximate location of Fig. 3 (dark blue square) are also shown

metamorphism and intercalations of MORB-derived metabasites, is interpreted as


an allochtonous domain that could be derived from the emplacement of the Rheic
Ocean (Pérez-Cáceres et al. 2015 and references therein). This field guide does not
include visits to this latter unit.
The HT/LP metamorphic rocks of the southernmost edge of the OMZ are inter-
preted as the high-grade equivalents of the Precambrian continental shelf broken
by a bimodal volcano-sedimentary sequence produced during the lower–middle
Cambrian rifting stage (Expósito et al. 2003; Chichorro et al. 2008). The main
lithostratigraphy is defined from correlations with similar rock units described in
the central, low-grade OMZ. Accordingly, the visited sector includes from bottom
to top: mainly pelitic gneisses, migmatites, nebulites, and charnockites with black
quartzites and minor kinzigites (corresponding to the Ediacaran Serie Negra, Bard
1969; Crespo-Blanc and Orozco 1991; Giese et al. 1994; Chichorro et al. 2008);
lower Cambrian marbles and an overlying complex unit comprising feldspathic and
throndhjemitic leucogneisses, calc-silicate rocks, amphibolitic migmatites, amphi-
bolites, and marbles, dated between 526 and 505 Ma (Chichorro et al. 2008). Syn-
to late-metamorphic noritic bodies (Castro et al. 1996b) intrude these metasedimen-
tary rocks. Geothermobarometric estimations have yielded maximum temperatures
of 850–1000 °C at pressures of 4–6 kbar (Díaz-Azpiroz et al. 2006 and references
therein). According to structure superposition criteria (Díaz-Azpiroz et al. 2006),
the HT/LP metamorphism, dated as Mississippian (351–323 Ma, Castro et al. 1999;
350–335 Ma, Pereira et al. 2009), was coeval with an extensional deformation stage
(OMZ-D2 ), described elsewhere in the OMZ and interpreted as an intraorogenic
mainly transtensional phase in SW Iberia (e.g., Pereira et al. 2009; Pérez-Cáceres
et al. 2015). Folds at microlithons account for a contractional deformation (OMZ-D1 )
The Southern Iberian Shear Zone (SW Spain): Inclined … 141

phyllites sheared
quartzites banded
Pulo do Lobo metabasiteamphibolite
domain

Beja-Acebuches
metabasites
Aracena
metamorphic
belt Continental domain CD-D
(high-grade zone)
3 fo
CD-D4 folds Southern Iberian

lds
shear zone
Calabazares
Cortegana-Aguafría shear zone
shear zone leucogneisses
marbles
pelitic series (gneisses,
migmatites, granulites)
1 km calc-silicate rocks

N
0 1 Cortegana-Aguafría
2 km shear zone

Fig. 3 Block diagram: Main structural features of the different domains at the OMZ-SPZ boundary
(from Díaz-Azpiroz et al. 2006). CD-D3 and CD-D4 correspond to the composite final deformation
stage affecting the HT/LP metamorphic rocks of the southernmost edge of the OMZ

previous to this extension (Díaz-Azpiroz et al. 2006). OMZ-D1 has been related to
the upper Devonian subduction/closure of the Rheic Ocean at SW Iberia (cf. Díaz-
Azpiroz et al. 2006; Pérez-Cáceres et al. 2015). The main HT/LP foliation (OMZ-S2 )
was affected by a composite left-lateral transpressional deformation (OMZ-D3−4 )
that produced complex fold interference patterns and ductile shear zones (Fig. 3).
This stage likely resulted from the final, Upper Carboniferous, oblique continental
collision.
The BAM is a narrow (up to 2 km wide) band that extends discontinuously from
Beja (Portugal) to Almadén de la Plata (Spain) along ca. 150 km. It strikes WNW-ESE
(N100-110°E in average) and dips moderately (average dip angle of 60–70°) to the NE
(Fig. 4a). There are two sectors, at Beja (Portugal) and the northern Huelva province
(Spain), where the BAM is better represented. However, specific features described in
the former, such as serpentinized mantle rocks, older N-directed thrusting structures
and a large-scale fold affecting the unit (Fonseca and Ribeiro 1993; Quesada et al.
1994; Pérez-Cáceres et al. 2015), are not observed in the Spanish sector.
The BAM are mainly amphibolites, with mafic schists near the southern contact
with the Pulo do Lobo unit, and were formed from the metamorphism of basic
magmatic rocks likely formed at a mid-ocean ridge, as suggested by geochemical
studies (Bard and Moine 1979; Dupuy et al. 1979; Quesada et al. 1994; Castro et al.
1996a; Azor et al. 2009). Zircon ages, interpreted as those of the magmatic event
(Azor et al. 2008), and Ar/Ar ages supposedly dating the metamorphism (Castro et al.
1999) yielded similar results (ca. 340 Ma), which has led to some controversy on their
tectonic interpretation (Azor et al. 2008, 2009; Pin and Rodríguez-Aller 2009). The
northern half of the BAM unit is represented by medium to coarse-grained banded
Hb-Pl amphibolites that include also diopside (Di) in its uppermost levels (upper
amphibolites–granulites facies transition), near the contact with the OMZ. Pressures
are estimated to be around 4–5 kbar (Castro et al. 1996a) and temperatures range
from near 800 °C at the uppermost levels, consisting of hornblende–plagioclase–
diopside (Hb–Pl–Di) amphibolites, to ca. 700 °C at the Hb–Pl amphibolites (Castro
142 M. Díaz-Azpiroz and C. Fernández

Fig. 4 a Schematic map of the Beja-Acebuches metabasites and the four proposed transects. Equal
area, lower-hemisphere projections of the mylonitic foliation (BAM-S2 ) and tectonic lineation
(BAM-L2 ) from three along-strike domains (see Fernández et al. 2013 for plotting details). Note
that BAM-L2 plunges to the E-SE at western and eastern domains and to the NW at the central
domain. b Detailed map of the Beja-Acebuches metabasites and the Southern Iberian Shear Zone
(location in Fig. 4a), corresponding to transects 1–3 (modified from Fernández et al. 2013). c Inverted
field metamorphic gradient of the BAM as illustrated from Pl-Hb temperatures (Díaz-Azpiroz et al.
2006). Different strain parameters indicate fabric strength decrease away from the SISZ base. These
are: Concentration parameter of the distribution of long axes of amphibole crystals (k Amp), shear
folds interlimb angle, axial ratio of amphibole crystals (R Amp), crystal size of plagioclase crystals
( Pl), and slope angle of CSD diagrams of plagioclase crystals (β CSD Pl). See Fernández et al.
(2013) for details
The Southern Iberian Shear Zone (SW Spain): Inclined … 143

et al. 1996a, Díaz-Azpiroz et al. 2006). Therefore, these rocks present a field inverted
metamorphic gradient (Figs. 4b, c and 5), rendering tectonic interpretation problem-
atic (cf., Castro et al. 1996a; Díaz-Azpiroz and Fernández 2005; Pérez-Cáceres et al.
2015). A syn-metamorphic deformation (BAM-D1 ) produced the main mesostruc-
ture of these amphibolites, a grain-size layering, often remarked by the preferred
orientation of amphibole blasts (BAM-S1 ), subparallel to the main boundaries of the
unit, WNW-ESE striking with moderate dips to the N. Occasionally, in the upper-
most levels, a strongly NE-plunging, faint mineral lineation (BAM-L1 ) is apparent.
Shear-sense criteria suggest SW-verging thrusting for BAM-D1 (Díaz-Azpiroz and
Fernández 2005). The uppermost BAM contact is affected by a meter-thick late
thrust, viz., the Calabazares Shear Zone (CSZ), which trends ca. WNW-ESE, dips
to the north and extend discontinuously along at least 10 km (Díaz-Azpiroz and
Fernández 2005). The southernmost half of the BAM unit is affected by the Southern
Iberian Shear Zone (SISZ, Crespo-Blanc and Orozco 1988) and by a related retro-
grade metamorphism that reached the lower amphibolite–greenschist facies transi-
tion at the lowermost levels (mafic schists with Pl–Hb–Ac–Ep–Chl), thus enhancing
the inverted field metamorphic gradient. The deformation associated with the SISZ
(BAM-D2 ) is described extensively by Díaz-Azpiroz and Fernández (2005) and is
summarized here. BAM-D2 produced shear folding of the BAM-S1 observable at
the upper structural levels of the SISZ. Toward the structural base, intense stretching
at fold limbs and strong grain-size reduction produced by BAM-D2 resulted in a
new, penetrative mylonitic foliation (BAM-S2 ), remarked in the mafic schists by Ep-
Chl rich layers, which obliterated BAM-S1 . Progressive strain also increases fabric
strength (Díaz-Azpiroz and Fernández 2003; Díaz-Azpiroz et al. 2007). Therefore,
both the deformation and the retrometamorphism produced at the SISZ were more
intense at the structural base of the shear zone and progressively decreased upward
(Figs. 4 and 5). However, meter-scale heterogeneous distribution of this activity
produced lozenges of higher grade, less deformed rocks surrounded by lower grade,
strongly foliated bands. A conspicuous lineation (BAM-L2 ) defined by plagioclase
ribbons and the preferred orientation of amphibole blasts is often observed. BAM-S2
presents a rather constant attitude (WNW–ESE striking and dipping moderately to
the NE), whereas BAM-L2 shows gentle to moderate plunges either to the E–SE or
to the NW, depending on the analyzed sector, covering a wide range >90° (Fig. 4).
These structural features fit well with a triclinic, simple shear-dominated transpres-
sion with along-strike variations in the extrusion direction of the coaxial component
(Fernández et al. 2013). Numerous shear-sense criteria observed on gently S-dipping
surfaces (i.e., close to the vorticity normal section, VNS) indicate a main left-lateral
slip for the simple shear component of transpression. Both the SISZ and the CSZ are
kinematically compatible with the late, left-lateral collisional stage registered in the
OMZ (OMZ-D3−4 ).
The SISZ marks the contact between the BAM and the Pulo do Lobo unit, affecting
also a decameter-thick band of the latter. Here the Pulo do Lobo unit is composed of
phyllites to fine-grained quartz-potassium feldspar-muscovite (Qtz-Kfs-Ms) schists,
with intercalated, cm- to m-thick quartzite layers. A meter-thick band of the Pulo do
Lobo uppermost levels occasionally presents also garnet (Grt), andalusite (And) and,
144 M. Díaz-Azpiroz and C. Fernández

Fig. 5 Schematic log of the Beja-Acebuches metabasites, including microstructural features,


constructed with combined information from transects 1–3 (Fernández et al. 2012)
The Southern Iberian Shear Zone (SW Spain): Inclined … 145

more sporadically, cordierite (Crd). This assemblage suggests the BAM may have
induced some sort of contact metamorphism to the underlying Pulo do Lobo unit
(Crespo-Blanc and Orozco 1991; Díaz-Azpiroz et al. 2006). The phyllites-quartzites
of the Pulo do Lobo unit affected by the SISZ show a penetrative S-C structure,
with strongly N-dipping S-planes and gently N-dipping C-planes, compatible with
SW-verging thrusting.
This field guide presents four different transects across the OMZ-SPZ boundary,
particularly focused on the oceanic metabasitic unit (BAM), including the Southern
Iberian Shear Zone, which also affects the oceanic metasediments of the Pulo do
Lobo unit and with some outcrops of the HT/LP metamorphic rocks from the former
Armorica continental margin (OMZ). The main goals are to present: (1) the old, deep
structure of a Paleozoic orogenic suture; (2) mesostructures related to complex 3D
deformation produced at a transpressional, mainly ductile shear zone; and (3) a suite
of rocks that underwent a somewhat particular HT/LP metamorphic event

3 Location, Accessibility, and Useful Information

This field guide is organized in four transects (Figs. 2, 4 and 6) located at the
northern Huelva province, in the Sierra de Aracena and Picos de Aroche Natural
Park (SW Spain). The closest international airport is Sevilla (100 km), with flights to
several destinations in Europe and domestic connections with Madrid, Barcelona, and
Palma. Sevilla has also high-velocity railway connections with Madrid, Barcelona,
and Málaga, with airports including transoceanic flights. The closest towns to the
proposed routes are Almonaster la Real (transects 1 and 2), Cortegana (transect 3)
and Aracena (transect 4). They all have facilities for long stays, such as hotels and
touristic apartments. Localities are communicated by local roads, with a maximum

Fig. 6 Satellite image of the field area showing the four proposed transects (T1 to T4)
146 M. Díaz-Azpiroz and C. Fernández

distance (between Stops 3.5 and 4.4) of 35 km and 50 min. Each transect is achievable
on foot, although in transects 3 and 4, a car may be preferred. Transects 2 to 4 are
mainly located along roads, where wearing reflective safety vests is compulsory. The
perfect season to accomplish this field trip is March–May or September–November,
although some rain could be expected in both periods. Southern Spain is particularly
hot from late May–early June to August–early September (maximum temperature
>35 °C at this time of year), so doing this field trip is not recommended in this
season (however, you may check weather conditions in Aracena, the main town in
the area, at the Spanish meteorological agency Web page, http://www.aemet.es/es/
eltiempo/prediccion/municipios/aracena-id21007). Winter is just cool, so it is also
an alternative moment of the year for this fieldtrip. There are no water springs along
the proposed routes, so carrying your own water is recommended.
This field guide is focused on few particular geological issues, but the area presents
some other interesting geological localities, such as the Cumbres Mayores, lower
Cambrian pillow-lavas, the Peña de Arias Montano spring and travertine formation,
the Aroche wollastonite skarns and stripped marbles or the many mines of the Iberian
Pyrite belt. The visitor is referred to Olías Álvarez et al. (2008) and Sáez et al. (2012)
for detailed information about these and other geological sites. Almadén de la Plata,
60 km eastward of Aracena, is the easternmost outcrop of the OMZ-SPZ boundary
portrayed in this field guide and one of the entry points to the Sierra Norte de
Sevilla UNESCO Geopark, which offers a wide variety of interesting geological
sites. Besides its great geology, the Sierra de Aracena y Picos de Aroche Natural
Park possesses also relevant natural and cultural offers. The Gruta de las Maravillas
(Grotto of the Wonders), which is included here as part of transect 4 (see Stop 4.1),
is a world famous cave on Cambrian marbles (http://www.aracena.es/es/municipio/
gruta/). Furthermore, several small, charming villages, with white houses and stoned
streets, are worth a visit. Our favorites are Aracena, Almonaster La Real, Linares
de la Sierra, Alájar, Los Marines, and Valdelarco, but all of them have nice things
to offer. This region is one of the most important in Spain for the raising of Iberian
pork and includes the village of Jabugo, the world’s capital of the celebrated Iberian
ham (Jamón Ibérico). It is also a well-known region among mushrooms collectors,
and it is possible to find several succulent species from fall to spring.

4 Routes Description

4.1 Transect 1: Almonaster La Real

Along this route (Fig. 7), it is possible to observe several examples of the main
lithologies from the HT/LP metamorphic belt located at the southernmost edge of
the OMZ; a transect to the Beja-Acebuches metabasites, from the HT, Di-bearing
amphibolites to the mafic schists; and also rocks of the Pulo do Lobo unit affected
by the SISZ.
The Southern Iberian Shear Zone (SW Spain): Inclined … 147

Fig. 7 Transect 1 with proposed stops on satellite image (a) and geological map (b). c Longitudinal
profile of this route
148 M. Díaz-Azpiroz and C. Fernández

This route mostly goes on dirt tracks and trekking trails. However, the last part from
Stops 1.11–1.15 goes mostly along roads, where wearing visible vest is mandatory.
Three stops (1.3, 1.8, and 1.10) are located in private land; the owners normally
accept crossing these fences, but they must be kept closed at all times.

Practical information
Route type: circular
Length: 4400 m (with additional 800 m to return back to the starting point)
Total upward walk: 1750 m
Average upward slope: 12%
Maximum upward slope: 30%
Difficulty: medium (some walks on trekking trails, steep slopes)
Estimated time: 5.5 h (including final return from Stop 1.15 to the starting point)
Start walking NNE from the village of Almonaster La Real, near the restaurant
“Las Palmeras” (37° 52 23.93 N 6° 47 2.08 W), where a signpost indicates a marked
track.
Stop 1.1: After a short walk of ca. 200 m, go up a zigzag path from 37° 52 26.3 N
6° 46 55.8 W.
On the way up, observe the medium to coarse-grained banded amphibolites of the
BAM. Looking carefully with lenses, green diopside crystals can be detected.
Stop 1.2: At the top of the zigzag path, at the junction, take left and walk around
50 m to 37° 52 29.7 N 6° 46 57.5 W.
There is a small outcrop of amphibolites with diopside from the uppermost levels
of the BAM affected by the Calabazares Shear Zone (Fig. 8a). This is a meter-thick
ductile shear zone, with reverse kinematics, which could be related to the latest
collisional stage between the SPZ and the OMZ. Heterogeneous deformation at
this shear zone produced lozenges of folded amphibolites surrounded by mylonitic
bands where intense dynamic recrystallization (Díaz-Azpiroz and Fernández 2005)
produced strong grain-size reduction (compare with Stop 1.1).
Stop 1.3: Go back and walk 250 m to the NE. At 37° 52 35.22 N 6° 46 50.27 W,
there is an outcrop on the left side, across a fence.
Cambrian marbles show a HT foliation (OMZ-S2 ) defined by the preferred orien-
tation of amphiboles (black) and pyroxenes (green) and remarked by differential
solution produced by weathering. This foliation is affected by SW-verging recum-
bent folds of the OMZ-D3-4 , which is related to the final continental collision stage
(Fig. 8b).
Stop 1.4: Go back to the south around 50 m and reach the junction with a small
track to the left at 37° 52 34.1 N 6° 46 50.9 W. About 20 m to the south of the track
junction, at 37° 52 33.3 N 6° 46 51.5 W, it is possible to observe calc-silicates (with
Pl–Hb–Di), with cm-thick intercalations of marbles. Continue 20 m to the north and
take the small track to the right. At 37° 52 34.3 N 6° 46 50.3 W, <10 m from the
junction toward the NE, on the left, a small outcrop of graphitic gneisses can be
The Southern Iberian Shear Zone (SW Spain): Inclined … 149

Fig. 8 Examples of different HT/LP lithologies observed between Stops 1.1–1.10 a the Calabazares
Shear Zone affecting the uppermost structural levels of the BAM (note the undeformed lozenge
surrounded by strongly foliated mylonite) in Stop 1.2. b Wollastonite marbles affected by OMZ-
D3-4 , SW-verging folds (Stop 1.3). c Calc-silicate boudins in wollastonite marbles (Stop 1.5).
d Very coarse-grained amphibolites from the OMZ (Stop 1.8). e The same rocks mylonitized
by the deformation of an OMZ-D3-4 shear zone (Stop 1.8). f Typical aspect of norites (note the
characteristic holes) intruding calc-silicate rocks (Stop 1.9)

observed at the base of a small trench. The presence of organic matter in the lower
Cambrian sediments is typical of rift settings (e.g., Crespo et al. 2004). Moreover,
this organic matter recrystallized during HT/LP to form a highly crystalline form of
graphite, compatible with metamorphic temperatures around 900 °C (Crespo et al.
2004). Graphite appears discontinuously along a narrow band at the southernmost
150 M. Díaz-Azpiroz and C. Fernández

edge of the OMZ and was mined in the surroundings of Almonaster La Real in
the first third of the twentieth century (Fernández-Caliani and Fernández-Rodríguez
2008).
Stop 1.5: Follow the track from 37° 52 34.3 N 6° 46 50.3 W to 37° 52 35.6 N
6° 46 38.2 W.
This is an almost continuous outcrop of marbles. The N-dipping main foliation
(OMZ-S2 ) is apparent from differentially weathered surfaces and more competent
calc-silicate layers, which appear strongly boudinaged (Fig. 8c). Elsewhere in the HT
belt, chocolate-tablet boudinage accounts for syn-metamorphic axial flattening defor-
mation (Díaz-Azpiroz et al. 2006). This foliation is affected by minor folds attributed
to the OMZ-D3-4 . These marbles show wollastonite (bright-white to yellowish fibrous
mineral observable on fresh surfaces), which has been associated with HT skarns
produced by basic intrusions in the marbles (Díaz-Azpiroz et al. 2004 and references
therein).
Stop 1.6: Continue along this track until reaching a small road at 37° 52 39.1 N
6° 46 31.2 W. Going up the road around 150 m, to the left, a low-quality outcrop
on the ground permits to observe nebulites with charnockitic affinity (with Kfs, Crd,
and orthopyroxene–Opx-, and no Grt). These rocks are the metamorphic equivalent
of the Ediacaran Serie Negra and register peak temperatures ~900 °C (Díaz-Azpiroz
et al. 2006).
Stop 1.7: Continue upward along this road around 350 m to 37° 52 49.07 N
6° 46 28.12 W.
On the right, there is an outcrop of leucocratic gneisses. This is one of the most
common lithotypes of the southernmost edge of the OMZ, and its main mineral
assemblage is dominated by Qtz and Kfs, although Pl is also common. Different
minor phases include Amp (amphibole), Cpx (clinopyroxene), Bt (biotite), Grt
(garnet), or Gr (graphite). In this outcrop, tiny garnet grains occur. Geochemical
features suggest these gneisses could be derived from felsic volcanic rocks related
to the lower–middle Cambrian rifting stage (Chichorro et al. 2008).
Looking SSE from this same point there is a good panoramic view of the main
part of this transect (Fig. 9). From NE (left) to SW (right), there appear coarse-
grain amphibolites (Stop 1.8), marbles, and calc-silicate rocks intruded by norites
(Stop 1.9), kinzigites (Stop 1.10), and the BAM unit (Stops 1.11–1.12). The valley
corresponds to the SISZ affecting amphibolites and mafic schists of the BAM (Stops
1.13–1.14) whereas the high reliefs at the back are produced by quartzite layers of
the Pulo do Lobo unit. Looking south from this point, we also have a beautiful view
of Almonaster la Real and its antique mosque–church–castle at the top of one of the
previously mentioned Pulo do Lobo hills.
Stop 1.8: Go down the road 700 m to 37° 52 34.4 N 6° 46 28.5 W. On your left,
across the fence (be careful to keep it closed), there are two large boulders.
On the left boulder, amphibolites with highly variable grain sizes occur (Fig. 8d).
Black hornblende crystals are up to several cm long. Such coarse grain is interpreted
as the effect of significant fluid circulation related to the HT skarn metamorphism
produced by basic intrusions into a carbonate-rich sequence (Díaz-Azpiroz et al.
2004 and references therein). These rocks register peak temperatures around 900 °C
The Southern Iberian Shear Zone (SW Spain): Inclined … 151

Fig. 9 NE-SW panoramic view from Stop 1.7 with geological cross section (simplified from Castro
et al. 1996b) and the approximate location of Stops 1.8–1.11

(Díaz-Azpiroz et al. 2006, 2007). On the lower part of the right boulder (with difficult
access), these amphibolites are strongly affected by a mylonitic shear zone, assigned
to the OMZ-D3–4 phase (Fig. 8e). This deformation was coeval with retrograde meta-
morphism and produced strong grain-size reduction via Pl and Hb crystal plasticity
(Díaz-Azpiroz et al. 2007).
Stop 1.9: Go back to the road and continue downward 300 m to 37° 52 28.0 N
6° 46 38.2 W.
On the left side, banded Di-rich calc-silicate rocks crop out. Several meters down-
ward, these rocks are cut by metanorites (mafic granulites, according to Castro et al.
1996b), a metaigneous rock with texture characterized by multiple round holes
(Fig. 8f). These are formed by differential weathering of biotite aggregates that
substituted hornblende crystals, which, in turn, grew on Opx crystals (El-Hmidi
2000). These rocks appear along the southernmost edge of the OMZ, very close to
the contact with the BAM, in form of small plutons usually with subparallel contacts
with the main foliation (OMZ-S2 ) of the country rocks (Bard 1969; Castro et al.
1996a, b; El-Hmidi 2000). Metanorites geochemistry is that of high-Mg andesites
(i.e., boninitic affinity, Castro et al. 1996a). Therefore, they are interpreted as near-
trench, syn-metamorphic intrusions produced in an atypical subduction process (see
Castro et al. 1996a, Díaz-Azpiroz et al. 2006 and references therein).
Stop 1.10: Continue down the road 150 m to 37° 52 25.37 N 6° 46 42.14 W.
There is a plantation of chestnut trees. Go up the hill around 100 m to a large boulder
(37° 52 21.95 N 6° 46 40.35 W).
152 M. Díaz-Azpiroz and C. Fernández

This rock presents black crystals (hercynite) and large, dark red, round crystals
(Grt) in a gray matrix (mostly Crd). In other outcrops at the top of the hill, Kfs also
appears. Microscopic analyses show fibrous sillimanite in Crd cores, whereas very
scarce Qtz and Bt appear as tiny, isolated inclusions within Crd and Grt, respectively.
This rock is a felsic granulite (Castro et al. 1996b), interpreted as a kinzigite, a solid
rest from a partial melting event that could have reached 1000 °C (Díaz-Azpiroz
et al. 2006).
Stop 1.11: Go back to the road and continue downward 100 m to 37° 52 25.20 N
6° 46 46.07 W. There is a small outcrop on the right side.
Going down the road, we find again an outcrop of the BAM, where de Calabazares
Shear Zone (see Stop 1.2) is not clearly observed. These rocks represent the levels of
the BAM with highest metamorphic grade. Their mineralogy is dominated by Pl–Hb–
Di (Fig. 10a), suggesting they reached the upper amphibolites–granulite facies tran-
sition. Metamorphic temperatures estimated via the Pl-Hb thermometer (Holland and
Blundy 1994) for P = 4 kbar in samples from this transect reach 800 °C (Castro et al.
1996a). These rocks present a grain-size layering (BAM-S1 ), defined by medium- to
coarse-grained layers, that dips moderately to the NNE.

Fig. 10 Examples of different lithologies observed between Stops 1.11–1.15. a Banded, Di-bearing
(circled green crystal) amphibolites of the BAM (Stop 1.11). b Amphibolites of the BAM with
a penetrative mylonitic foliation (BAM-S2 ) remarked by Ep-Chl layers (Stop 1.13). c Tectonic
lineation (BAM-L2 ) defined by plagioclase ribbons and elongated amphibole grains (Stop 1.14).
d Pulo do Lobo phyllites and quartzite intercalations with a S–C structure suggesting top-to-the-
south shear (Stop 1.15)
The Southern Iberian Shear Zone (SW Spain): Inclined … 153

Stop 1.12: Continue down the road ~275 m to a zone with three consecutive close
curves (37° 52 21.10 N 6° 46 54.22 W). In the second one, there is an outcrop on
the left side.
Here, the amphibolites still present a grain-size layering, although grain size is
finer than in Stop 1.11. The mineral assemblage does not include Di (amphibolite
facies) and the registered metamorphic temperature (Pl-Hb thermometer, 4 kbar) is
around 750 °C (Castro et al. 1996a).
Stop 1.13: Continue down the road 250 m to a junction with local road HU-
8105. Take this road to the left (to the east) and walk around 100 m to a junction
(37° 52 13.3 N 6° 46 55.8 W) with signs for “La Escalada”. Just before this junction,
there are outcrops on both sides of the road (be aware, despite its narrowness, this is
a rather busy road, it is mandatory to use reflective safety vest).
The rocks here are amphibolites (mineral assemblage dominated by Pl + Hb).
In this case, the grain-size layering is no longer apparent. Instead, a very penetra-
tive mylonitic foliation (BAM-S2 ) represents the activity of the SISZ (Fig. 10b).
This foliation strikes WNW–ESE and dips moderately to the NNE. On the foliation
planes, up to two lineations can be observed. The most prominent one is defined
by elongated amphibole blasts and/or plagioclase ribbons and presents a low pitch,
with gentle plunges to the E. Occasionally (partly depending on the light), a high
pitch intersection lineation is present. This is interpreted as related to late crenulation
affecting BAM-S2 (Díaz-Azpiroz and Fernández 2005).
Stop 1.14: Take the road to La Escalada and walk 200 m to 37° 52 12.6 N
6° 47 04.2 W. Along this walk, there are different outcrops on the left.
These outcrops show mafic schists of the BAM with dominantly Pl-Hb-Ac
(actinolite)-Ep (epidote)-Chl (chlorite). The assemblage corresponds to the lower
amphibolites–greenschists facies transition. The mylonitic foliation (BAM-S2 ),
related to the SISZ, is defined by the preferred orientation of amphibole blasts and
remarked by Ep-Chl-rich layers. This suggests that the retrometamorphism and the
deformation affecting the BAM produced by the SISZ are strongly heterogeneous. In
this outcrop, the tectonic lineation defined by the preferred orientation of elongated
amphibole blasts and plagioclase ribbons is particularly evident (Fig. 10c).
This transect across the BAM from the uppermost structural levels (Stop 1.11) to
the lowermost level (Stop 1.14) depicts the inverted metamorphic field gradient that
characterizes this unit.
Stop 1.15: Walk down the road around 125 m to 37° 52 10.20 N 6° 47 6.97 W.
Halfway (37° 52 11.7 N 6° 47 06.2 W), the contact between the BAM and the Pulo
do Lobo unit crops out behind a ruined hut.
This is the best outcrop in the area to observe the phyllites and quartzites of the
Pulo do Lobo unit affected by the SISZ (Fig. 10d). The phyllites present a phyllonitic
foliation subparallel to the original bedding, which is evident from the phyllites
and quartzites layers. This foliation dips moderately to the N. Northward gently
dipping C-planes also exist. The geometrical relationship between S-plane and C-
plane suggests thrusting kinematics for the deformation affecting these rocks. This
reverse displacement contrasts with the main left-lateral motion deduced for the
SISZ when affecting the BAM. We interpret that SISZ transpression was plausibly
154 M. Díaz-Azpiroz and C. Fernández

partitioned into two main domains defined by distinct lithologies. As such, strike-
slip dominated deformation was mostly accommodated at the BAM whereas mainly
reverse shear took place at the Pulo do Lobo unit.

4.2 Transect 2: Calabazares

This route (Fig. 11) goes along road HU-8105 (wearing safety vets is compulsory at
all times), from 37° 51 57.36 N 6° 45 28.12 W to 37° 51 53.74 N 6° 45 45.33 W
(Fig. 12). It is possible to park the car at 37° 51 56.18 N 6° 45 29.43 W. The route
shows a continuous transect across the BAM unit, from the NNE boundary with the
HT belt of the OMZ to the first levels of the mafic schists of the BAM. Two additional
stops permit to observe, respectively, the mafic schists located at the lowermost levels
of the BAM and the rocks of the Pulo do Lobo unit located at the contact with the
former. All in all, this is the best exposure of the entire BAM unit in Spain and
permits to analyze the inverted field gradient of this unit, as well as some of the most
characteristic structures of both the banded amphibolites and of the SISZ.

Practical information
Route type: linear
Length: 500 m (one way) + 400 m to reach Stops 2.7 and 2.8
Average upward slope: 0%
Difficulty: easy

Fig. 11 Satellite image with transect 2 and proposed stops


The Southern Iberian Shear Zone (SW Spain): Inclined … 155

Fig. 12 Different structures shown by the BAM along transect 2. a BAM-L1 on BAM-S1 foliation
plane (note that this photograph is not from this transect). b Grain-size layering (BAM-S1 ) in
banded BAM amphibolites (Stop 2.2). c Extensional shears interpreted as the C/ -planes affecting
the contact between layers of different grain sizes, and suggesting apparent left-lateral displacement.
d Open folds affecting BAM-S1 (Stop 2.3). e Disharmonic folding in BAM-S1 tight folds (Stop
2.3). f Isoclinal fold affecting BAM-S1 , which is now completely obliterated by BAM-S2 . g and
h show respectively, a sigmoid plagioclase porphyroclast/fish and S–C/ structures, both indicating
left-lateral shear-sense
156 M. Díaz-Azpiroz and C. Fernández

Estimated time: 3 h (including returning to the starting point)


At the beginning of the route, calc-silicate rocks of the OMZ crop out. The first
outcrops of the BAM unit (Stop 2.1, 37° 51 56.87 N 6° 45 28.95 W) correspond to
banded amphibolites with Di (small light green crystals, often within black horn-
blende grains). Depending on the light, it is sometimes possible to observe on foli-
ation planes (BAM-S1 ) a high pitch, NE-plunging, faint mineral lineation defined
by amphibole grains (BAM-L1 ). Elsewhere in the BAM unit, this lineation is more
evident (Fig. 12a). Further to the SSW there is the best outcrop of banded amphibo-
lites in the area (Stop 2.2, 37° 51 55.60 N 6° 45 31.31 W). It is possible to observe
several layers with medium to coarse grain sizes (Fig. 12b). Grain-size differences,
instead of modal proportion or chemical composition, account for distinct colors in
these bands (Castro et al. 1996a). A foliation defined by the preferred orientation of
amphibole elongated grains appears subparallel to grain-size layering, both defining
BAM-S1 . Planes located at low angles to BAM-S1 produce mm-scale extensional
displacements (Fig. 12c) and are thus interpreted as C/ -planes. According to the
main orientation of BAM-L1 , these and other similar microscopic shear-sense criteria
suggest a main top-to-the-SSW thrusting (Díaz-Azpiroz and Fernández 2005).
Further to the south, the banded amphibolites are deformed by the SISZ, which
produced shear folds, a mylonitic foliation (BAM-S2 ) and a related tectonic lineation
(BAM-L2 ). These structures become progressively more strongly developed along
this continuous transect (see, for example, folds geometrical variations in Fig. 13),
suggesting that deformation at the SISZ gradually increased from the uppermost
to the lowermost levels (Díaz-Azpiroz and Fernández 2005). At the uppermost
levels (Stop 2.3, 37° 51 54.77 N 6° 45 33.20 W), folds range from gentle to tight
(interlimb angles = 40–140°), but open prevail, have low amplitude (types 1–3 of
Hudleston 1973), present no associated axial-planar foliation (Fig. 12d), and their
axes plunge consistently to the NNE (Fig. 13b). These levels are considered as the
uppermost, gradual boundary of the SISZ. Southward (Stop 2.4, 37° 51 53.88 N
6° 45 35.28 W), folds are mostly tight (interlimb angle = 25–85), with interme-
diate amplitude (types 3–4 of Hudleston 1973) and present incipient limb stretching
and axial planar foliation, although BAM-S1 is still dominant. Original rheolog-
ical contrast between different layers provokes disharmonic folding (Fig. 12e). Even
further to the south (e.g., Stop 2.5, 37° 51 53.14 N 6° 45 42.84 W, but also in other
sites along the transect) folds become isoclinal (interlimb angle = 0–25°), usually
show strongly stretched, eventually boudinaged limbs (types 4–5 of Hudleston 1973)
and isolated intrafolial hinges (Fig. 12f, see also Mukherjee et al. 2015 for a review).
Such stretching likely produced strong grain-size reduction and a mylonitic folia-
tion (BAM-S2 ) that obliterated BAM-S1 . Fold axes trend spans more than 90° (from
NW to ESE) along the average BAM-S2 orientation (Fig. 13c). A tectonic lineation
(BAM-L2 ) defined by plagioclase ribbons and the preferred orientation of elongated
amphibole prismatic grains is incipient here and becomes more penetrative toward
the south (Stop 2.6, 37° 51 53.74 N 6° 45 45.33 W, see also Fig. 10c). BAM-S2
strikes WNW–ESE and dips moderately to the NNE, and BAM-L2 shows gentle to
moderate plunges to the ESE (Figs. 4a, 13b, c). All measured folds along this transect
show S-shaped asymmetry and their axes plot to the NW of the average orientation
The Southern Iberian Shear Zone (SW Spain): Inclined … 157

of BAM-L2 (Fig. 13b, c). In analogous folds from other transects not represented
in this field guide (e.g., Alájar, to the east), axes located to the SE of BAM-L2
show Z-shaped asymmetry (Díaz-Azpiroz and Fernández 2005). These features are
compatible with sheath folds produced by deformation at a mainly sinistral, reverse
shear zone (Díaz-Azpiroz and Fernández 2005). Other asymmetric shear-sense indi-
cators compatible with this kinematics include sigmoid porphyroclasts or mineral
fish (Fig. 12g) and S-C/ structures (Fig. 12h), which can be observed at Stops 2.5,
2.6, and 2.7 (37° 51 51.74 N 6° 45 47.04 W).
We could go down the road around 250 m to the junction to the village of
Calabazares. After 100 m along this road, take a dirt track to your right and walk
near the river along 200 m to Stop 2.7. 200 m further, at Stop 2.8 (37° 51 48.36 N
6° 45 38.33 W), Pulo do Lobo phyllites crop out (see Stops 1.14–1.15 for a detailed
description of these rocks).

Fig. 13 SISZ shear fold geometry evolution along transects 2 and 3 (Díaz-Azpiroz and Fernández
2005). a Variation of interlimb angles versus distance to the base of the shear zone, measured at
folds in transects 2 and 3. b and c are equal area, lower hemisphere plots of BAM-D2 fold axes
(BAM-B2 ), BAM-S2 (great circles) and BAM-L2 (asterisks) from the upper structural part b and
the lower structural part c of transect 2. The gray pattern marks the average loci of BAM-L2 .
Anticlockwise arrows on fold axes indicate the S asymmetry of these folds
158 M. Díaz-Azpiroz and C. Fernández

4.3 Transect 3: Acebuches–Veredas

This route (Fig. 14) goes along a local road that communicates two small localities,
Acebuches and Veredas. Interestingly, J.P. Bard used the former name to define
for the first time, in 1969, the oceanic metabasitic unit (BAM) as the Acebuches
amphibolites. This route is an almost continuous transect from the NNE boundary
with the HT belt of the OMZ, across the entire BAM unit to the mafic schists of
the lowermost levels. There is an additional stop (Stop 3.5) further to the west. We
recommend parking the car near Stop 3.1, return from Stop 3.4 and drive to 3.5.
Walking around 2 km along the same road from Stop 3.4 to Stop 3.5 is also possible.

Practical information
Route type: linear
Length: 1500 m (one way) + Stop 3.5
Total upward walk: 400 m
Average upward slope: 6.5%
Maximum upward slope: 10%
Difficulty level: easy
Estimated time: 3 h (including returning to the starting point and driving 3.5 km to
Stop 3.5)
The rocks of the OMZ at the beginning of this route (37° 53 20.34 N
6° 48 47.92 W) are mainly calc-silicate rocks. The first outcrops of the BAM unit
(Stop 3.1, 37° 53 18.69 N 6° 48 50.68 W) correspond to banded amphibolites with

Fig. 14 Satellite image with transect 3 and proposed stops


The Southern Iberian Shear Zone (SW Spain): Inclined … 159

Di. A particularity of these outcrops is the abundance of trondhjemitic-tonalitic veins


(Fig. 15a). In accordance with experimental results (López and Castro 2001), these
veins are interpreted as leucosomes produced by anhydrous partial melting of the
uppermost amphibolite levels, at temperatures ~800 °C (Díaz-Azpiroz et al. 2004).
Southward along the road, note several outcrops of the banded amphibolites (without
Di) with their characteristic grain-size layering (BAM-S1 ). We set the uppermost
boundary of the SISZ at Stop 3.2 (37° 53 10.05 N 6° 49 8.22 W), where the first
folds affecting BAM-S1 appear. As in transect 2, the geometry of these folds varies
toward the lowermost structural levels of the unit accounting for progressive strain
accumulation, with decreasing interlimb angles (Fig. 13a) along with increasing
stretching at the limbs and axial planar foliation intensity. Progressive strain also
reduces grain size and strengthens the mineral fabric (Díaz-Azpiroz and Fernández
2003; Díaz-Azpiroz et al. 2007). From Stop 3.3 (37° 53 4.64 N 6° 49 32.99 W), the
new mylonitic foliation (BAM-S2 ) completely obliterates the previous one (BAM-
S1 ) and intrafolial hinges are their only witnesses (Fig. 15b). Sporadic Ep + Chl-rich
layers enhance this second foliation. Further southward (Stop 3.4, 37° 52 57.71 N
6° 49 34.08 W), mafic schists crop out and Ep–Chl-rich layers become more abun-
dant. In rocks with penetrative BAM-S2 (Stops 3.3 and 3.4), on foliation planes the
BAM-L2 presents orientations that vary from gentle ESE plunges to subhorizontal.

Fig. 15 Features of the BAM along transect 3. a Trondhjemitic veins cross-cut the banded
amphibolites (Stop 3.1). b Isoclinal folds affecting BAM-S1 and resulting in BAM-S2 (Stop 3.3).
c NW-plunging tectonic lineation in mafic schists at Stop 3.5
160 M. Díaz-Azpiroz and C. Fernández

In Stop 3.5 (37° 53 14.67 N 6° 50 39.30 W), mafic schists show penetrative
BAM-S2 and BAM-L2 . This outcrop is essential to understand the overall kinematics
of the SISZ. In contrast with the other outcrops shown in this guide, here BAM-L2
plunges gently to the NW (Fig. 15c). This orientation is maintained along a sector
located between Veredas, 1.2 km to the east, and El Hurón (Aroche), 5.5 km to the
west (Fig. 4a). By contrast, further east and west, BAM-L2 plunges to the ESE (as
in transects 1, 2 and 4). This orientation distribution is a typical feature of inclined
transpression shear zones with triclinic symmetry (e.g., Fernández and Díaz-Azpiroz
2009; Fernández et al. 2013).

4.4 Transect 4: Aracena

This transect offers a poorer exposure of the BAM unit than in the previous routes.
Nevertheless it is interesting because the SISZ shows here an extremely heteroge-
neous character, which contrasts with that described before. The route (Fig. 16)
departs from the town of Aracena, the largest locality in the Sierra de Aracena y
Picos de Aroche Natural Park, and continues along the road A-479 that connects
Aracena with Campofrío. Although it is always possible to complete this transect
on foot, the absence of outcrops in much of the route makes it advisable to visit the
proposed outcrops (Stops 4.1 to 4.4) separately. If walking is preferred, it is possible
to leave the car in any of the places allowed to park in the city of Aracena. Otherwise,

Fig. 16 Satellite image with transect 4 and proposed stops


The Southern Iberian Shear Zone (SW Spain): Inclined … 161

the most appropriate place to park will be indicated here below, where each stop is
described. Remember that the reflective vest is mandatory when walking on roads.

Practical information (for the walking option)


Route type: linear
Length: 4000 m (one way; most of the route from Stop 4.1 to Stop 4.4 is downward)
Total upward walk: 150 m
Average upward slope: 3.8%
Maximum upward slope: 35%
Difficulty: easy
Estimated time: 4 h (including returning to Aracena)
Stop 4.1: The first stop is located in the town of Aracena itself. From the parking
place go up to the castle hill (37° 53 23.84 N 6° 33 46.23 W). There is a church
located at the top of the hill (Prioral Church; with small cars it is possible to cross
the entrance door to the top of the hill, which is used as a bell tower, and park next to
the church). Take the branch of the path that starts from the main door of the church
(37° 53 21.73 N 6° 33 43.35 W), surrounding the ruined walls of the castle on the
side facing SSW. To the right of the path, in the trench under the wall, in a walk
of ~150 m, one finds a good outcrop of the high-grade marbles of the AMB, which
show a high crystallinity (grain size of the carbonate crystals) and intercalations of
cm to dm thick calc-silicate rocks. The large competence contrast between marble
and calc-silicate rock conditioned the boudinage of the latter (Fig. 17a). Some good
examples of calc-silicate boudins can be observed at 37° 53 23.62 N 6° 33 48.74 W
(cf. Fernández and Díaz-Azpiroz 2008). Depending on the exact composition of
the involved calc-silicate boudinaged layer, boudins range from rounded to tabular,
with up to 10:1 aspect ratios in the latter case. It is interesting to note that the
structures that are observed in this stop are the same that can be seen in the so-called
Grotto of the Marvels (“Gruta de las Maravillas”), a system of natural caves with
beautiful karst formations, which extends precisely below the hill of the castle. It is
strongly recommended to visit these caves, not only for their intrinsic beauty, but
because, together with the surface outcrop, they offer a 3D-view of the structure
of this marbles massif of the AMB. The schedules and prices for the visit can be
consulted at the tourist office of Aracena or at the entrance of the cave (37° 53 27.77 N
6° 33 57.12 W).
Stop 4.2: From Grotte of the Marvels square (Plaza de la Gruta de las Maravillas)
walk SW ~250 m along Tenerías street. When reaching a roundabout, head toward the
S, still along the same street. At ~300 m a second roundabout will be reached, at the
exit of the village (there is a parking lot to the right, close to the roundabout). Continue
toward SSE, now on the A-479 road. About 150 m from the second roundabout, the
road turns to the right. In the curve, the road is limited by trenches. The best outcrop
is in the trench on the right (37° 53 05.37 N 6° 34 05.67 W). The rocks belong to the
162 M. Díaz-Azpiroz and C. Fernández

Fig. 17 Observations along transect 4. a Calc-silicate boudins within marbles (Stop 4.1).
b Lozenges of slightly deformed metabasites embraced by metabasites with a penetrative mylonitic
foliation (BAM-S2 ). The asymmetry of the lozenge suggests apparent left-lateral displacement
(Stop 4.3). See reviews on lozenges by Ponce et al. (2013) and Mukherjee (2017)

upper part of the BAM. Banded amphibolites (Pl–Hb), with a prominent BAM-S1
foliation striking WNW–ESE, and steeply dipping to the NNE, occasionally show
small Di crystals. This outcrop, therefore, belongs to the upper part of the BAM
and is very similar to those in transects 1–3. Unfortunately, it is not possible to see
the contact of these amphibolites with the high-grade metamorphic materials of the
AMB (such as the marbles of Aracena, Stop 4.1).
Stop 4.3: Continue toward the S ~750 m on the road A-479 to 37° 52 43.00 N
6° 34 13.96 W, just before the road turns left. On the right, there is the entrance to
some farms and private houses (this is a good place to park those who have chosen to
follow the transect by car). From here, continue southward along a small path limited
by stone fences leaving the road on your left. On the ground of the path, and along
~ 100 m, note a good exposure of the lower half of the BAM, affected by the SISZ.
Here, differently from the characteristics of the SISZ described in Transects 1–3, the
The Southern Iberian Shear Zone (SW Spain): Inclined … 163

deformation is strongly heterogeneous. Instead of appearing folds of the foliation


BAM-S1 , lozenges of metric size composed of apparently undeformed fine-grained,
dark metabasites exist. Such lozenges are embraced by shear bands of highly foli-
ated (BAM-S2 ) green metabasites with a subhorizontal lineation (BAM-L2 ) on the
foliation planes. Kinematic criteria (S-C structures, asymmetric porphyroclasts, and
the geometry of the lozenge blocks) consistently indicate a sinistral displacement, in
accordance with the large-scale kinematics of the SISZ.
Stop 4.4: To get to this stop from the previous one, it is advisable to do it walking.
Return to the A-479 road and walk toward the SE and S about 1.2 km. A curve to the
left will be passed first, and then a wider one to the right. Finally, at 37° 52 11.00 N
6° 34 05.18 W, located in a curve to the left, a trench several meters high on the left
side of the road shows a good outcrop of the basal part of the BAM. Here, the SISZ
shows similar characteristics to those described at Stop 4.3, but is possible to observe
cm- to m-scale, asymmetric sinistral lozenges of undeformed, very fine-grained,
black amphibolite (Fig. 17b). This rock shows very small idiomorphic crystals of
pyroxene and plagioclase, interpreted as igneous in origin. Therefore, it corresponds
to the type of rocks closer to the protolith of the BAM unit that can be found in the
region, scarcely affected by any of the main deformation phases. This stop can be
completed, ~500 m south, on the same road, with another outcrop visible at point
37° 51 55.57 N 6° 33 57.52 W, in both trenches (right and left). Here, close to the
lower boundary of the BAM, the foliated mafic schists resulting from the activity of
the SISZ are predominant. Again, the southern contact of the BAM is not exposed in
this transect, and also the Pulo do Lobo unit is absent due to the presence of plutonic
rocks belonging to the Castilblanco de los Arroyos batholith (e.g., at 37° 51 26.67 N
6° 33 41.74 W), that intruded after the activity of the SISZ.

Acknowledgements Financial support from Projects PGC2018-100914-B-I00 and PGC2018-


096534-B-100 of the Spanish Ministry of Science, Innovation and Universities is gratefully acknowl-
edged. Soumyajit Mukherjee (IIT Bombay) reviewed and edited this chapter. I. Pérez Cáceres (IES
Jaume Almera) also reviewed the manuscript. Their suggestions improved the final result and
are gratefully thanked. Thanks to Marion Schneider, Annett Buettener, Boopalan Renu, Alexis
Vizcaino, Doerthe Mennecke-Buehler and the proofreading team (Springer). Dutta and Mukherjee
(2021) summarize this article.

References

Azor, A., Rubatto, D., Simancas, J. F., González Lodeiro, F., Martínez Poyatos, D., Martín Parra,
L. M., & Matas, J. (2008). Rheic Ocean ophiolitic remnants in Southern Iberia questioned by
SHRIMP U-Pb zircon ages on the Beja-Acebuches amphibolites. Tectonics, 27, TC5014. https://
doi.org/10.1029/2009tc002527.
Azor, A., Rubatto, D., Marchesi, C., Simancas, J. F., González Lodeiro, F., Martínez Poyatos, D.,
et al. (2009). Reply to comment by C. Pin and J. Rodríguez on “Rheic Ocean ophiolitic remnants in
southern Iberia questioned by SHRIMP U-Pb zircon ages on the Beja-Acebuches amphibolites".
Tectonics, 28, TC5014. https://doi.org/10.1029/2009tc002527.
164 M. Díaz-Azpiroz and C. Fernández

Bard, J. P. (1969). Le Metamorphisme régional progressiff des Sierras d’Aracena en Andalousie


Occidentale (Espagne): Sa place dans le segment hercynien sud-Ibérique. Thése Université de
Montpellier, 397pp.
Bard, J. P., & Moine, B. (1979). Acebuches amphibolites in the Aracena hercynian metamorphic
belt (southwest Spain): Geochemical variations and basaltic affinities. Lithos, 12, 271–282.
Castro, A., Fernández, C., De la Rosa, J. D., Moreno-Ventas, I., & Rogers, G. (1996a). Significance
of MORB-derived amphiboles from the Aracena metamorphic belt, southwest Spain. Journal of
Petrology, 37, 235–260.
Castro, A., Fernández, C., De la Rosa, J. D., Moreno-Ventas, I., El-Hmidi, H., El-Biad, M., et al.
(1996b). Triple-junction migration during Paleozoic plate convergence: the Aracena metamorphic
belt, Hercynian massif, Spain. Geologische Rundschau, 85, 180–185.
Castro, A., Fernández, C., El-Hmidi, H., El-Biad, M., Díaz, M., De La Rosa, J. D., et al. (1999).
Age constraints to the relationships between magmatism, metamorphism and tectonism in the
Aracena metamorphic belt, southern Spain. International Journal of Earth Sciences, 88, 26–37.
Chichorro, M., Pereira, M. F., Díaz-Azpiroz, M., Williams, I. S., Fernández, C., Pin, C., et al. (2008).
Cambrian ensialic rift-related magmatism in the Ossa-Morena Zone (Évora-Aracena metamor-
phic belt, SW Iberian Massif)—Sm-Nd isotopes and SHRIMP zircon U-Th-Pb geochronology.
Tectonophysics, 461, 91–113. https://doi.org/10.1016/j.tecto.2008.01.008.
Crespo, E., Luque, J., Fernández-Rodríguez, C., Rodas, M., Díaz Azpiroz, M., Fernández-Caliani,
J. C., et al. (2004). Significance of graphite occurrences in the Aracena Metamorphic Belt, Iberian
Massif. Geological Magazine, 141, 687–697.
Crespo-Blanc, A., & Orozco, M. (1988). The Southern Iberian Shear Zone: A major boundary in
the Hercynian folded belt. Tectonophysics, 148, 221–227.
Crespo-Blanc, A., & Orozco, M. (1991). The boundary between the Ossa-Morena and Southpor-
tuguese Zones (Southern iberian Massif): A major suture in the European Hercynian Chain.
Geologische Rundschau, 80, 691–702.
Díaz-Azpiroz, M., & Fernández, C. (2003). Characterization of tectono-metamorphic events using
crystal size distribution (CSD) diagrams. A case study from the Acebuches metabasites (SW
Spain). Journal of Structural Geology, 25, 935–947.
Díaz-Azpiroz, M., & Fernández, C. (2005). Kinematic analysis of the Southern Iberian Shear Zone
and tectonic evolution of the Acebuches metabasites (SW Variscan Iberian Massif). Tectonics,
24, TC3010. http://dx.doi.org/10.1029/2004TC001682.
Díaz-Azpiroz, M., Castro, A., Fernández, C., López, S., Fernández-Caliani, J. C., & Moreno-Ventas,
I. (2004). The contact between the Ossa-Morena and the South-Portuguese zones. Characteristics
and significance of the Aracena metamorphic belt in its central sector between Aroche and Aracena
(Huelva). Journal of Iberian Geology, 30, 23–51.
Díaz-Azpiroz, M., Fernández, C., Castro, A., & El-Biad, M. (2006). Tectono-metamorphic evolution
of the Aracena metamorphic belt (SW Spain) resulting from ridge-trench interaction during
Variscan plate convergence. Tectonics, 25, TC1001. http://dx.doi.org/10.1029/2004TC001742.
Díaz-Azpiroz, M., Lloyd, G. E., & Fernández, C. (2007). Development of lattice preferred orien-
tation in clinoamphiboles deformed under low-pressure metamorphic conditions. A SEM/EBSD
study of metabasites from the Aracena metamorphic belt (SW Spain). Journal of Structural
Geology, 29, 629–645.
Díaz Azpiroz, M., Fernández, C., & Czeck, D. M. (2019). Are we studying deformed rocks in
the right sections? Best practices in the kinematic analysis of 3D deformation zones. Journal of
Structural Geology, 125, 218–225.
Dupuy, C., Dostal, J., & Bard, J. P. (1979). Trace element geochemistry of paleozoic amphibolites
from SW Spain. Tscherm Mineral Petrograph Mitt, 26, 87–93.
Dutta, D., & Mukherjee S. (2021). Structural Geology and Tectonics Field Guidebook—Volume 1.
In S. Mukherjee (Ed.), Structural Geology and Tectonics Field Guidebook—Volume 1. Springer
Nature Switzerland AG. Cham. pp. xi-xvi. ISBN: 978-3-030-60142-3.
Eden, C. P. (1991). Tectonostratigraphic analysis of the northern extent of the oceanic exotic terrane,
Northwestern Huelva Province, Spain. (Ph.D. thesis), Univ. of Southhampton, 214pp.
The Southern Iberian Shear Zone (SW Spain): Inclined … 165

El-Hmidi, H. (2000). Petrología y geoquímica de los sistemas andesíticos ricos en Mg: estudio
petrológico y experimental de las noritas de la Banda Metamórfica de Aracena, SO de España.
Tesis Doctoral, Universidad de Huelva, 239pp.
Expósito, I., Simancas, J. F., González Lodeiro, F., Bea, F., Montero, P., & Salman, K. (2003).
Metamorphic and deformational imprint of Cambrian-Lower Ordovician rifting in the Ossa-
Morena zone (Iberian Massif, Spain). Journal of Structural Geology, 25, 2077–2087.
Fernández, C., & Díaz-Azpiroz, M. (2008). Estructuras de deformación en mármoles: el ejemplo del
cerro del Castillo de Aracena. In M. Olías Álvarez, T. Donaire Romero, C. Fernández Rodríguez,
E. Mayoral Alfaro, J. M. Morales González, F. M. Alonso Chaves, G. Ruiz de Almodóvar
(Coords.), Geología de Huelva. Facultad de Ciencias Experimentales, Universidad de Huelva,
90–91.
Fernández, C., & Díaz-Azpiroz, M. (2009). Triclinic transpression zones with inclined extrusion.
Journal of Structural Geology, 31, 1255–1269.
Fernández, C., Díaz-Azpiroz, M., & Castro, A. (2012) Geología de la sierra de Huelva: una antigua
sutura entre continentes. In R. Sáez, C. Moreno, Ruiz de G. Almodóvar (Eds.), Geología de la
Provincia de Huelva (pp. 51–66). Universidad de Huelva.
Fernández, C., Czeck, D. M., & Díaz-Azpiroz, M. (2013). Testing the model of oblique transpres-
sion with oblique extrusion in two natural cases: Steps and consequences. Journal of Structural
Geology, 54, 85–102. https://doi.org/10.1016/j.jsg.2013.07.001.
Fernández-Caliani, J. C., & Fernández-Rodríguez, C. (2008). Las mineralizaciones de grafito de la
Sierra de Aracena. In M. Olías Álvarez, T. Donaire Romero, C. Fernández Rodríguez, E. Mayoral
Alfaro, J. M. Morales González, F. M. Alonso Chaves, & G. Ruiz de Almodóvar (Coords.),
Geología de Huelva. Facultad de Ciencias Experimentales, Universidad de Huelva, 76–77.
Fonseca, P., & Ribeiro, A. (1993). Tectonics of the Beja-Acebuches ophiolite: A major suture in
the Iberian Variscan Foldbelt. Geologische Rundschau, 82, 440–447.
Fossen, H., & Tikoff, B. (1993). The deformation matrix for simultaneous simple shearing, pure
shearing and volume change, and its application to transpression/transtension tectonics. Journal
of Structural Geology, 15, 413–422.
Giese, U., Walter, R., & von Winterfeld, C. (1994). Geology of the southern Iberian Meseta II. The
Aracena Metamorphic Belt between Almonaster La Real and Valdelarco, Huelva province (SW
Spain). Neues Jahrbuch für Geologie und Paläontologie, Abh, 192, 333–360.
Holland, T. J. B., & Blundy, J. D. (1994). Non-ideal interactions in calcic amphiboles and their
bearing on amphibole-plagioclase thermometry. Contributions to Mineralogy and Petrology, 116,
433–447.
Hudleston, P. J. (1973). Fold morphology and some geometrical implications of theories of fold
development. Tectonophysics, 16, 1–46.
Jones, R. R., Holdsworth, R. E., Clegg, P., McCaffrey, & K., Tavarnelli, E. (2004). Inclined
transpression. Journal of Structural Geology, 26, 1531–1548.
Julivert, M., Fontboté, J. M., Ribeiro, A., & Conde, L. E. (1972). Mapa Tectónico de la Península
Ibérica y Baleares. E. 1:1 000 000. IGME, Madrid.
Kroner, U., & Romer, R. L. (2013). Two plates—Many subduction zones: The Variscan orogeny
reconsidered. Gondwana Research, 24, 298–329.
Lin, S., Jiang, D., & Williams, P. F. (1998). Transpression (or transtension) zones of triclinic
symmetry: natural example and theoretical modeling. In R.E. Holdsworth, R.A. Strachan, J.F.
Dewey (Eds.), Continental transpressional and transtensional tectonics (Vol. 135, pp. 41–57),
Geological Society of London, Special Publications.
López, S., & Castro, A. (2001). Determination of the fluid-absent solidus and supersolidus phase
relationship of MORB-derived amphibolites in the range 4–14 kbar. American Mineralogist, 86,
1396–1403.
Lotze, F. (1945). Zur Gliederung der Varisziden der Iberischen Meseta. Geoteckt Forsch, 6, 78–92.
Matte, P. (2001). The Variscan collage and orogeny (480–290 Ma) and the tectonic definition of the
Armorica microplate: A review. Terra Nova, 13, 122–128.
166 M. Díaz-Azpiroz and C. Fernández

Mukherjee, S. (2017). Review on symmetric structures in ductile shear zones. International Journal
of Earth Sciences, 106, 1453–1468.
Mukherjee, S. Punekar, J., Mahadani, T., & Mukherjee R. (2015). A review on intrafolial folds and
their morphologies from the detachments of the western Indian Higher Himalaya. In S. Mukherjee,
& K.F. Mulchrone (Eds.), Ductile shear zones: from micro- to macro-scales (pp. 182–205). New
York: Wiley Blackwell.
Nance, R. D., Gutiérrez-Alonso, G., Keppie, J. D., Linnemann, U., Murphy, B. J., Quesada, C.,
et al. (2010). Evolution of the Rheic Ocean. Gondwana Research, 17, 194–222.
Olías Álvarez, M., Donaire Romero, T., Fernández Rodríguez, C., Mayoral Alfaro, E., Morales
González, J. M., Alonso Chaves, F. M., et al. (2008). Geología de Huelva. Facultad de Ciencias
Experimentales: Universidad de Huelva, 176pp.
Oliveira, J. T. (1990). Part VI: South Portuguese Zone, stratigraphy and synsedimentary tectonism.
In R.D. Dallmeyer, E. Martínez García (Eds.), Pre-Mesozoic Geology of Iberia (pp. 334–347).
New York: Springer.
Onézime, J., Charvet, J., Faure, M., Bourdier, J. L., & Chauvet, A. (2003). A new geodynamic
interpretation for the South Portuguese Zone (SW Iberia) and the Iberian Pyrite Belt genesis.
Tectonics, 22(4), 1027. https://doi.org/10.1029/2002tc001387.
Pereira, M. F., Chichorro, M., Williams, I. S., Silva, J. B., Fernández, C., & Díaz-Azpiroz, M., et al.
(2009) Variscan intra-orogenic extensional tectonics in the Ossa-Morena Zone (Évora-Aracena-
Lora del Río metamorphic belt, SW Iberian Massif): SHRIMP zircon U-Th-Pb geochronology.
In J.B. Murphy, J.D. Keppie, & A.J. Hynes (Eds.), Ancient Orogens and Modern Analogues (Vol.
327, pp. 215–237). Geological Society of London, Special Publications. https://doi.org/10.1144/
SP327.11.
Pérez-Cáceres, I., Martínez-Poyatos, D., Simancas, J. F., & Azor, A. (2015). The elusive nature of
the Rheic Ocean suture in SW Iberia. Tectonics, 34, 2429–2450. https://doi.org/10.1002/2015TC
003947.
Philippon, M., & Corti, G. (2016). Obliquity along plate boundaries. Tectonophysics, 712, 171–182.
Pin, C., & Rodríguez-Aller, J. (2009). Comment on: “Rheic Ocean ophiolitic remnants in southern
Iberia questioned by SHRIMP U-Pb zircon ages on the Beja-Acebuches amphibolites” by Azor
et al. Tectonics, 28, TC5013. https://doi.org/10.1029/2009TC002495.
Ponce, C., Druguet, E., & Carreras, J. (2013). Development of shear zone-related lozenges in foliated
rocks. Journal of Structural Geology, 50, 176–186.
Quesada, C., Fonseca, P. E., Munha, J., Oliveira, J. T., & Ribeiro, A. (1994). The Beja-Acebuches
Ophiolite (Southern Iberia Variscan fold belt): Geological characterization and geodynamic
significance. Boletín Geológico y Minero, 105, 3–49.
Ribeiro, A., & Silva, J. B. (1983). Structure of the South Portuguese Zone. In A. Sousa, & J.T.
Oliveira (Eds.), The Carboniferous of Portugal (Vol. 29, pp. 91–97). Lisboa: Memória Servicio
Geológico de Portugal.
Sáez, R., Moreno, C., & Ruiz de Almodóvar, G. (Eds.) (2012). Geología de la Provincia de Huelva.
Universidad de Huelva, 197 pp.
Sanderson, D., & Marchini, R. D. (1984). Transpression. Journal of Structural Geology, 6, 449–458.
Silva, J. B., Oliveira, J. T., & Ribeiro, A. (1990). Part VI: South Portuguese zone. Structural outline.
In R. D. Dallmeyer & E. Martínez García (Eds.), Pre-Mesozoic geology of Iberia (pp. 348–362).
Berlin: Springer.
Simancas, F., Carbonell, R., González-Lodeiro, F., Pérez Estaún, A., Juhlin, C., Ayarza, P., et al.
(2003). Crustal structure of the transpressional Variscan orogen of SW Iberia: SW Iberian deep
seismic reflection profile (IBERSEIS). Tectonics, 22. http://dx.doi.org/10.1029/2002TC001479.
Simancas, J. F., Tahiri, A., Azor, A., González Lodeiro, F., Martínez Poyatos, D. J., & El Hadi, H.
(2005). The tectonic frame of the Variscan-Alleghanian orogen in southern Europe and northern
Africa. Tectonophysics, 398(3–4), 181–198.
A Field Guide to the Spectacular Salt
Mines of the Transylvanian Basin
and Romanian Carpathians

Dan Mircea Tămas, , Alexandra Tămas, , Alexander Magnus Jüstel,


Martijn Passchier, Nils Chudalla, Lina Gotzen,
Luis Alberto Pizano Wagner, Teodora Tas, cu-Stavre, Zsolt Schléder,
Csaba Krézsek, Sorin Filipescu, and Janos L. Urai

Abstract Salt diapirs/bodies are remarkable structures often presenting a rich


internal structure, with features like isoclinal folds, sheath folds, curtain folds, boud-
inage, etc. Salt structures are themselves valuable as a resource but are also often
associated with significant hydrocarbon accumulations, as well as valuable as storage
sites. While significant progress has been made to imagine the external geometry of
salt bodies, especially using 3D seismic, the complex internal structure of the salt
diapirs is less known. Such structures can mainly be investigated using salt mines,
outcrops, physical models and rarely seismic for larger structures. There are just a
few places in the world where salt mines or outcrops can be studied, and Romania
is such an example. The Romanian Carpathians and the Transylvanian Basin offer
access to multiple salt mines (and outcrops) which are easily accessible especially
because most of the salt mines have been converted into public treatment facilities.
The walls of the galleries have been cleaned, polished and well lighted, offering
unique 3D exposures which can be used to observe and map in detail the defor-
mation of these salt domes. In this field-trip guide, we propose a route with five
stops, visiting four salt mines and one important location from the link between salt
tectonics and hydrocarbons. Each of the chosen locations is unique as it records a
different regional kinematic evolution and offers the opportunity to compare and
contrast the specific internal structures.

D. M. Tămas, (B) · T. Tas, cu-Stavre · S. Filipescu


Department of Geology and Center for Integrated Geological Studies, Babes, -Bolyai University,
Cluj-Napoca, Romania
e-mail: danmircea.tamas@ubbcluj.ro
A. Tămas,
Department of Earth Sciences, Durham University, Durham, UK
A. M. Jüstel · M. Passchier · N. Chudalla · L. Gotzen · L. A. P. Wagner · J. L. Urai
Structural Geology, Tectonics and Geomechanics, RWTH Aachen University, Aachen, Germany
Z. Schléder
OMV Exploration and Production GmbH, Vienna, Austria
C. Krézsek
OMV Petrom S.A., Exploration B.U., Bucharest, Romania

© Springer Nature Switzerland AG 2021 167


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_6
168 D. M. Tămas, et al.

1 Introduction

Aspects of salt tectonics have been studied extensively due to their economic and
scientific importance (i.e. Mukherjee et al. 2010; Hudec and Jackson 2012; Warren
2016; Jackson and Hudec 2017). One key topic, making salt tectonics a high focus
for research is its relation to hydrocarbon systems globally. Our understanding of salt
tectonics evolved over the past 160 years. The first models explaining the discordant
salt bodies invoked ideas ranging from chemistry to tectonics (see Jackson 1995;
Tămas, et al. 2018 for historical reviews). Later, models of salt tectonics were based on
the concepts of density inversion in a layered system of two viscous fluids (Nettleton
1934; Trusheim 1960; Kukla et al. 2019). New insights after 1990 have shown that the
overburden is not viscous, but brittle-frictional (Vendeville and Jackson 1992a, b).
Experimental determination of the mechanical properties of salt, and extrapolation
of these data to conditions of slow geological deformation (Wawersik and Hannum
1980; Urai et al. 1986; Carter et al. 1993; Li and Urai 2016; Rowan et al. 2019) has
laid the basis for a quantitative understanding of the mechanics of salt deformation.
Nowadays, the salt is viewed as a weak, thermally conductive fluid, which acts as a
mechanically decoupling agent and a seal for different kinds of fluids (Schultz-Ela
et al. 1993; van Keken et al. 1993; Urai et al. 2008; Barabasch et al. 2019; Raith et al.
2016, 2017).
Romania has historical importance when it comes to salt tectonics (Tămas, et al.
2018 and reference therein). One of the first mentions of discordant salt structures in
the world is the Praid diapir, in the Transylvanian Basin (Posepny 1871). In 1900, at
the International Exhibition in Paris, Mrazec presented the map of the salt structures
in the Romanian Carpathians (Mrazec and Teisseyre 1902). Mrazec (1905, 1907).
Mrazec also noted the discordance between the strata adjacent to the salt body, and
the fact that it was mostly found in anticlinal structures (results mainly derived from
surface exposures). These structures were named in different ways, like “anticlines
with salt core” or “folds with piercing core”, and this phenomenon of piercing he
considered to be controlled by the orogenic forces.
The Third International Petroleum Congress was held in Bucharest, 1907. This
clearly demonstrates the importance of the Romanian oil industry at that time. It
was here where Ludovic Mrazec had a chapter where he defined and described the
“diapir folds” (Fig. 16) in a discussion “on the formation of the Romanian petroleum
deposits” (Mrazec 1910). The term diapir comes from the Greek word “διαπ είρειν”,
which means to pierce. For more details on the history of salt in the Romanian
Carpathians, see Tămas, et al. (2018).
The Romanian Carpathians and Transylvanian Basin (Figs. 1 and 2) host many
salt mines, salt exploitations and outcrops (i.e. Tămas, et al. 2020a). Four salt mines
and one historically important location were chosen to feature in this guide. The
salt mines, namely Turda, Praid, Slănic Prahova and Ocnele Mari (Fig. 2), offer a
world-class opportunity to study the internal structure of different salt bodies due
to their unique 3D underground exposure. Most of these mines can be visited in
abandoned galleries that have now been converted to touristic and health facilities.
A Field Guide to the Spectacular Salt Mines … 169

Fig. 1 Regional map of the Alpine—Carpathian—Pannonian region, showing the Transylvanian


Basin (TB) and the Romanian Southern and Eastern Carpathians (Schléder et al. 2019)

For this reason, the galleries are well lighted, and the pillars are cleaned, making
these fascinating salt structures visible.

2 Field-Trip Itinerary

The field-trip guide presents two different salt regions with different salt tectonics
styles, the Transylvanian Basin (Stops 1 and 2; Fig. 2) and the Romanian Carpathians
(Stops 3–5; Fig. 2). The trip will start and end in Cluj-Napoca, Transylvania, which
is one of the largest cities in Romania and hosts the Babes, -Bolyai University.
The trip starts at the western margin of the Transylvanian Basin, by visiting the
Turda salt mine and then crossing the hilly Transylvanian Basin (towards the E)
170 D. M. Tămas, et al.

Fig. 2 Simplified map of the distribution of the “lower salt formation” (lower Burdigalian) and
the “upper salt formation” (mid-Badenian) in Romania (Popescu et al. 1973; Tămas, et al. 2018).
The field-trip is marked as red lines. The start-end locality is marked with a black filled dot and the
stops with red dots

to reach the Praid salt mine (Fig. 2). Then we continue SSE, along the Eastern
Carpathians with a view to the Neogene volcanic mountains until reaching the histor-
ical city of Bras, ov. From here, we cross the Eastern Carpathian Bend Zone (heading
SE), to reach the Slănic salt mine, located in the Slănic syncline (Fig. 2).
Leaving Slănic, we continue (heading WSW) in the most prolific onshore hydro-
carbon area in Romania, the Diapir Fold Zone. Here, one optional stop is on top of
the Băicoi salt diapir, which offers a view on the Moreni–Băicoi salt diapir linea-
ment (~30 km long, with a roughly NE–SW orientation) and the effects of the latest
deformation in the area, the Wallachian phase (Fig. 2). This is a place of historical
and economic importance as it hosts the largest onshore oil fields in Romania and is
the type locality for the term diapir (Mrazec 1910; Tămas, et al. 2018, 2019; Schléder
et al. 2019).
The trip continues along the Carpathians (heading W) to reach the Ocnele Mari
salt mine, which is located at the southern slope of the Southern Carpathians. From
here, the road crosses the Southern Carpathians (heading NNW) back into the Tran-
sylvanian Basin and to Cluj-Napoca. The entire trip is ~900 km of driving and can
be done in four days to a week (Fig. 2). Many other outcrops can be visited along
the route.
A Field Guide to the Spectacular Salt Mines … 171

3 The Geology of Salt in the Transylvanian Basin

The Transylvanian Basin (Figs. 1, 2 and 3) is interpreted as a post-Cenomanian sag


basin (Krézsek and Filipescu 2005; Krézsek and Bally 2006; Tilit, ă et al. 2015). On
top of the basement nappes that were stacked by mid-Cretaceous, the basin locally
contains >5 km of sedimentary rocks (Krézsek and Bally 2006).
The mid-Miocene evaporites in the Transylvanian Basin (Fig. 2) are related to
the Badenian salinity crisis that took place at Carpathian—Central Paratethys scale
(Peryt 2006; de Leeuw et al. 2010). The salt was deposited in a deep marine desiccated
setting, with initial thicknesses ~300 m (Krézsek and Bally 2006; Tilit, ă et al. 2015).
The first salt movements are presumably linked to differential loading by local
sedimentation depocentres (Tilit, ă et al. 2015) that developed salt pillows within
the central part of the basin (Fig. 3; Krézsek and Bally 2006). There are two main
diapir lineaments present in the Transylvanian Basin (Fig. 3). While the eastern
diapir lineament developed initially as reactive diapirs due to extension caused by
gravitational gliding, the western lineament is characterized by toe thrusting (Krézsek
and Bally 2006). Later, the eastern diapir lineament was compressed plausibly due
to the additional loading produced by the Pliocene volcanism (Fig. 3; Krézsek and
Bally 2006).
The presence of salt near the surface is the result of a different regional kinematic
evolution, from a gravitational toe-thrust diapir (Turda) to a reactive and then passive
diapir, which was later shortened (Praid).

4 Stop 1: Turda Salt Mine

4.1 Road, Access and History

The Turda salt mine is located in the town Turda, a municipality in Cluj county along
the Aries, River within the Transylvanian Basin. The drive from Cluj-Napoca to Turda
salt mine entrance (46° 35 15.7 N 23° 47 14.8 E) follows the E81 road (towards
the south) for ~35 km (Fig. 2). Tickets for visiting the mine can be acquired from the
entrance. Along the hallway leading to the mine, beautiful hand-drawn geological
sections and maps from the mine archive are available.
Salt exploitation activities in Turda were first documented in 1075 and lasted until
1932 (Salina Turda 2019). During World War II, the mine was used as an anti-aircraft
shelter. In 1992, the mine was opened for tourism and modernized in 2008 with a
grand reopening in 2010 and a newly built visitor centre (Fig. 4) (Salina Turda 2019).
The temperature in the mine ranges 10–12 °C with a relative humidity of 75–80%.
The main attractions are the Joseph mine (Fig. 4), a 112 m deep and up to 67 m
wide opening, the Crivac room, the Theresa mine, the Rudolph mine (Fig. 4) and the
Ghizela mine (Salina Turda 2019) (Fig. 5).
172

Fig. 3 Seismic profile, oriented E–W across the Transylvanian Basin (Krézsek and Bally 2006). Note the locations of the western diapir alignment (i.e. Stop
1—Turda salt mine) and the eastern diapir alignment (i.e. Stop 2—Praid salt mine)
D. M. Tămas, et al.
A Field Guide to the Spectacular Salt Mines … 173

Fig. 4 Turda salt mine—photograph looking at the Rudolph mine and the access staircase

Fig. 5 Turda salt mine—photograph looking vertically upwards showing complex fold structures
with locally horizontal fold axes
174 D. M. Tămas, et al.

A gallery makes it possible to walk along the top of the mine. On either side of
the Rudolph mine, a staircase with around 170 steps connects the amusement park
area with the remaining parts of the Turda salt mine complex. On the staircase, the
year at which the salt exploitation reached the level is graved in the salt wall. The
Rudolph mine is 42 m deep, up to 50 m wide and 80 m long (Salina Turda 2019).

4.2 Description of Geology

Salina Turda is part of the western diapir lineament (Fig. 3) within the Transylvanian
Basin (Figs. 1, 2 and 3). The evolution of this diapir was actively controlled by several
factors: deformation rates of the toe-thrust bringing the salt towards the surface,
differential unloading (erosion rates) and differential loading (sedimentation rates in
the basin) (Krézsek and Bally 2006).
The spectacular exposures in Turda salt mine with near-polished salt walls (Figs. 4,
5 and 6) show predominantly sub-vertical bedding of rock salt with darker and lighter
salt layers (10 cm scale). These are marked by different content of impurities. Locally,
some shale layers are disrupted and sheared into a salt mylonite with tectonic clasts
(near the edge of the diapir). The bedding is N-S striking with tight isoclinal folds
with variable fold axial planes, ranging from horizontal to upright.

Fig. 6 Turda salt mine—photograph showing stalactites formed by seepage of brines


A Field Guide to the Spectacular Salt Mines … 175

The ceiling of the Rudolph Mine shows fold structures which are either flat laying
salt layers cut parallel to the bedding or sheath folds (Fig. 5). In the Turda mine, the
ceiling presents salt stalactites, pointing to the seepage of groundwater through more
impure layers of rock salt (Figs. 4 and 6). Regularly spaced white rock salt veins
characterize the boundary of the salt diapir.
The salt composition is >99% halite with impurities such as clay minerals, iron
oxides or gypsum not surpassing 0.7% (Salina Turda 2019). Whereas the salt and
gypsum are authigenic minerals, the remaining minerals have a sedimentary origin
and were transported into the basin. The alternating brighter and darker salt layers of
coarse crystalline halite are due to impurities (Figs. 4, 5 and 6) and define the bedding.
No visible differences regarding the grain sizes for both layers can be observed. Pure,
clean salt is transparent to translucent, recrystallized with a grain size up to several
centimetres at polished sections. Dirty salt, containing impurities, has a brownish
colour, caused most likely by clay minerals between the salt grains.

5 Stop 2: Praid Salt Mine

5.1 Road, Access and History

The road from Turda salt mine to Praid crosses the Transylvanian Basin (Fig. 2),
taking the E60 road (~E) towards Târgu Mures, city and later other national roads that
lead to Praid–Sovata region. The drive distance between the two mines is ~150 km.
Praid and Sovata are balneary resorts due to their thermal and salty water springs
and lakes.
The Praid salt mine is located in the town of Praid, Harghita County (Fig. 2). Salt
outcrops can be visited in both Sovata and Praid. Praid also hosts a beautiful trail
through the salt canyons located above the mine. The first mining activity was during
Roman times (~100 AD). Commercial mining started in 1762 with the opening
of the Joseph Mine (Salina Praid 2019). The temperature in the mine is around
16 °C(~120 m below the surface). The accessible area is on level 50, around 120 m
below the surface (Salina Praid 2019).
The access to the Praid salt mine is made by bus. Tickets can be acquired at the
mine bus station (46° 33 03.6 N 25° 07 09.8 E).

5.2 Description of Geology

Praid has an essential role in the evolution of salt tectonics ideas in Romania and
worldwide. The first mention of discordant salt structures and the mechanism of salt
emplacement in Romania, and among the first worldwide had been described from
Praid (Fig. 7; Posepny 1871). Many details illustrated by Posepny (1871) regarding
176 D. M. Tămas, et al.

Fig. 7 Cross section through the Praid diapir, as first drawn by Posepny (1871). The image illustrates
the discordant nature of the salt structure, as well as the internal folding. The dashed areas represent
the eroded part of the diapir (Tămas, et al. 2018)

the internal deformation of the Praid diapir and the surrounding rocks are ahead of
his time. Some of the structures depicted by Posepny (1871) are still valid today and
nowadays termed rim synclines and upturned flaps (Fig. 7; Tămas, et al. 2018).
The Praid diapir is part of the eastern diapir lineament (Fig. 3). The sub-elliptic
diapir is 1.4 km long (NE–SW), 1.2 km wide and up to 2.7 km deep (Salina
Praid 2019), striking NW–SE to NE–SW. The evolution is interpreted as a reac-
tive extensional diapir, which progressed to a passive diapir and was later shortened
by more recent contractional events. The diapirs in the region show discordant struc-
tures. This could be the result of reverse faults crosscutting the flanks during the
Wallachian Event or due to reactive and passive diapir growth before shortening
occurred (Jackson and Hudec 2017; Krézsek and Bally 2006).
The Praid salt mine is characterized by isoclinally folded grey to white salt layers
(Figs. 8 and 9). The bedding dominantly strikes NE–SW, but many other orientations
are present pointing to curtain folding (Figs. 8 and 9). Bedding-parallel spalling of
the anhydrite-rich layers show well-developed lineation on these layers.
The rock salt in this mine also consists of multiple different coloured layers
with variable thickness and consists of 94–98% halite (Salina Praid 2019). The
composition can be separated into authigenic (halite, anhydrite, gypsum, calcite
and dolomite) and of allogenic origin (quartz, feldspars, clay minerals) (Har et al.
2010). The lighter-coloured layers are cm to dm-scale while the darker layers are
much thicker here (dm-m) than in other mines.
Near the entrance to the salt mine, a brown boudinaged mudstone slab can be found
in a brownish salt layer of the wall. Other boudinaged layers can be found on the
northern wall in the centre of the mine (clay-containing mud layer cut perpendicular
to the c-axis) and near the exit of the mine.
A Field Guide to the Spectacular Salt Mines … 177

Fig. 8 Praid salt mine-view towards the chapel, showing isoclinal folds exposed on the polished
floor

Fig. 9 Praid salt mine—orthophotograph of part of the mine ceiling, as viewed from above. The
figure highlights the intricate folding patterns
178 D. M. Tămas, et al.

Fig. 10 Praid salt


mine—slicken sides on an
anhydrite-rich layer within
the salt mine

At least one fault (steeply dipping (70°) towards NW) can be located in the centre
of the Praid salt mine. The hanging wall has broken off, after mining activities,
leaving the fault plane exposed (Fig. 10).

6 The Geology of Salt in the Romanian Carpathians

The Romanian Carpathians are a highly arcuate Alpine orogen (Fig. 1) which
records the evolution of the Alpine Tethys during latest Jurassic—middle Miocene
(Săndulescu 1984, 1988; Mat, enco and Bertotti 2000; Csontos and Vörös 2004;
Schmid et al. 2008). One younger deformation event (late Miocene to recent), the
Wallachian phase (Hippolyte and Sandulescu 1996), had an essential impact on salt
tectonics in the Romanian Carpathians (Schléder et al. 2019).
One of the particularities of the Carpathians is the presence of two different salt
horizons (Figs. 2 and 11): the “lower salt”, considered early Burdigalian (lower
Miocene) and the “upper salt”, middle Badenian (mid-Miocene) (i.e. Murgoci 1905;
Popescu et al. 1973). These evaporites initially deposited in the foreland area of
the Carpathians were later incorporated within the Carpathian nappes. While the
A Field Guide to the Spectacular Salt Mines …

Fig. 11 Composite seismic profile (a), interpretation (b) and resaturation (c) illustrating the interpreted structural style of the Eastern Carpathian Bend Zone
(Schléder et al. 2019)
179
180 D. M. Tămas, et al.

Burdigalian salt is restricted to the Carpathians foreland basin, the Badenian salt
also extends into the, e.g. Transylvanian Basin, East Slovakian Basin (e.g. Popescu
et al. 1973; Babel 2004; Krézsek and Bally 2006; de Leeuw et al. 2010).
Two of the following stops, the Slănic Prahova and Ocnele Mari salt mines expose
mid-Miocene salt (Fig. 2). Their evolution was controlled by both the middle Miocene
and upper Miocene to recent compressional events.
The Moreni–Băicoi diapir lineament and other similar structures in the Diapir
Fold Zone are composed of lower Miocene salt and have a different evolution than
the middle Miocene ones (Figs. 2 and 11; Schléder et al. 2019; Tămas, et al. 2019).

7 Stop 3: Slănic Prahova Salt Mine

7.1 Road, Access and History

The road from Praid to the Slănic Prahova salt mine crosses the Eastern Carpathian
Bend Zone, taking the E60 road (south) towards Bras, ov and later national road 1A
(SSE) that leads to Slănic salt mine entrance (45° 14 08.4 N 25° 56 29.9 E). The
distance of this drive is ~240 km (Fig. 2). The entrance is made by bus, and the tickets
for visiting the mine can be bought in the bus station.
Salt extraction in Slănic started more than three centuries ago. The first written
evidence of the salt mining is the contract signed by Prince Mihail Cantacuzino in
1685, who bought land in Slănic to open a salt mine. This document also notes that
some other salt mines near Slănic (Teişani) have been already active by that time.
Early (seventeenth to eighteenth century) salt mining applied the bell extraction
technique, which consists of descendent vertical mining of salt coeval with lateral
enlargement of the mine cavity (Schléder et al. 2016; Salina Slănic 2019).
Since the nineteenth century, the room-and-pillar extraction method was employed
on several levels and in different mines. Current mining activity is confined to the
New Mine, in the south-eastern part of the diapir (Schléder et al. 2016 and reference
therein).
The publicly accessible Old Mine was active from 1938 to 1970. One of the upper
extraction levels was opened to the public in 1970. The temperature in the mine is
constant 13 °C, humidity around 60% (at 208 m depth). The trapezoidal rooms vary
in height from 55 to 70 m. The width of the chambers is 10 m at the ceiling and 32 m
at the base (Salina Slănic 2019).

7.2 Description of Geology

The Slănic salt diapir is located in the axis of the Slănic syncline (ENE–WSW),
which is part of the internal Tarcău Nappe (Fig. 12). In this part of the nappe, both
A Field Guide to the Spectacular Salt Mines … 181

Fig. 12 Slănic salt mine—image highlighting the large isoclinal fords in 3D

evaporitic events are recorded (Fig. 2). The middle Miocene halite is mined in Slănic.
The salt appears as a lens-like accumulation and is thicker along the synclinal axis
(Schléder et al. 2016).
The Slănic Prahova salt mine is characterized by 10-15 m scale isoclinally folded
layers, with axial planes dipping mainly ~45° to the SE (Figs. 13 and 14). Most of the
fold axial lines plunge gently to the NE, but close to the mine entrance, they gently
rotate to E–W. Some of the isoclinal folds can be correlated across the trapezoidal
gallery, which is 32 m wide at its base, and 54 m tall (Fig. 13). This shows that
some folds are quite cylindrical. Rare anhydrite-rich layers show strong boudinage
structures, related to the large strains in this salt body. Rare open fractures are filled
with large halite crystals pointing to brittle deformation close to the surface.
The mid-Miocene salt in this mine was deposited in a shallow water environ-
ment. Detailed microstructural investigation reveals the widespread presence of
typical shallow water syn-sedimentary features such as fluid-inclusion-rich bands
and dissolution pipes filled with clear halite (Schléder et al. 2016).
The deformation in this salt mine is interpreted to relate to the most recent
(Wallachian) deformation phase. The dominant shortening direction characteristic for
this tectonic event in the Bend Zone is roughly NW–SE. This gave rise to widespread
ENE–WSW to NNE–SSW trending folds and thrusts in the area, and the NE–SW
trending folds observed in the salt mine fit in this general trend.
182 D. M. Tămas, et al.

Fig. 13 Slănic salt mine -photograph looking vertically upwards at the folding pattern

Fig. 14 Ocnele Mari mine—photograph of the corner of one of the room pillars showing salt
folding
A Field Guide to the Spectacular Salt Mines … 183

8 Stop 4: On Top of the Băicoi Diapir (Optional)

8.1 Road and Access

The road leading from the Slănic Prahova to this optional stop is about 30 km long
(Fig. 2). There is no exact visiting point; the option is to use the small roads used for
access to the hydrocarbon wells to get a good, high viewpoint (e.g. location 45°02
31.6 N 25°54 13.3 E). This area is of high historical importance, for both salt
tectonics and hydrocarbon production worldwide.

8.2 Description of Geology

This stop is located on top of the Băicoi salt diapir and provides an overview of the
regional topography of the GuraOcnit, ei—Moreni—Băicoi diapir lineament (looking
west; Fig. 11). On both the north and the south of these diapirs lie the largest onshore
oil fields in Romania.
The main production of these fields comes from the supra-salt upper Miocene
stratigraphy, mainly from the areas flanking the salt diapirs. Less production is asso-
ciated with the supra-salt lower and mid-Miocene stratigraphy (Schléder et al. 2019).
Due to their mature status, recent work has focused on understanding the structural
style and salt deformation history, with a target to understand the sub-salt geome-
tries (Schléder et al. 2019; Tămas, et al. 2019). North of this area, where the sub-salt
reservoirs are exposed, they exhibit further complications, specifically the presence
of sand injectites (Tămas, et al. 2020b).

9 Stop 5: Ocnele Mari Salt Mine

9.1 Road, Access and History

The road from the Băicoi salt diapir or the Slănic Prahova salt mine goes first to
Bucharest, the capital of Romania and then take the A1 highway towards Pitesti and
then the road E81 to Ocnele Mari salt mine (45° 05 06.9 N 24° 18 33.7 E). The
drive from stop 4 to stop 5 is ~250 km (Fig. 2). The entrance in the salt mine is made
by bus, and the tickets can be acquired from the bus stop, near the salt mine parking
lot.
The Ocnele Mari salt mine is located in the town Ocnele Mari, Vâlcea County
(Fig. 2). Mining activities were ongoing during Roman times (eleventh to thirteenth
century) (Salina Ocnele Mari 2019). The mining is still ongoing with blasting in
184 D. M. Tămas, et al.

Fig. 15 Ocnele Mari mine—image highlighting the folding pattern on a side-view of one salt pillar

progress after the visitor part has closed down. The temperature in the mine ranges
between 13 and 15 °C and the humidity of ~50%(at ~80 m depth).
The available (public) part of the mine is called horizon +226, indicating the
height above sea level. It consists of around 50–60 squared pillars (Figs. 15 and
16) with an edge length and spacing between the pillars of ~15 m and a height of
~8 m. Active mining is still being carried out in this horizon and Horizon 210 (Salina
Ocnele Mari 2019).

9.2 Description of Geology

The salt deposit is located at the Ocnele Mari-Govora antiform. Miocene uplift
exposes pre-and syn-kinematic sediments in this area. With the oldest strata cropping
out in the west. The fold axis of the anticline plunges towards east. To the north and
south of the crest, post-tectonic rocks cover the anticline (Răbăgia et al. 2011). It is
assumed that structural inversion in the area resulted from oblique thrusting between
the Southern Carpathians and the Moesian Platform (Răbăgia et al. 2011).
The salt deposit at the Ocnele Mari-Govora is 7.5–8 km long (striking E–W) and
~3.5 km wide showing a stratified lenticular shape (Zamfirescu et al. 2007). The
salt body dips 15–20° to the north. In the central part, the salt body is >400 m thick
(Zamfirescu et al. 2007).
The Ocnele Mari salt structure consists of tightly folded package with north or
east-dipping layers (Figs. 15 and 16). The layering/bedding is more pronounced
compared to the other mines due to ~ 10 cm thick layers of darker salt. The isoclinal
A Field Guide to the Spectacular Salt Mines … 185

folds are dm to m scale, with local deviations from similar folding, pointing to small
variations in lithological contrast. In addition, thin, cm-scale anhydrite-rich layers
show more concentric folding with angular hinges, pointing to a larger rheological
contrast. The fold axes trend ~ E–W, consistent with the south-verging thrusting of
the nappes.

Acknowledgements The authors thank Soumyajit Mukherjee (IIT Bombay) for his review and
editorial comments which helped improve this manuscript. D. M. Tămas, acknowleges financial
support from Babes, -Bolyai University through the young researcher grant no. 35275/18.11.2020
and the project: Entrepreneurship for innovation through doctoral and postdoctoral research,
POCU/380/6/13/123886 co-financed by the European Social Fund, through the Operational
Program for Human Capital 2014–2020. Thanks to Marion Schneider, Annett Buettener, Boopalan
Renu, Alexis Vizcaino, Doerthe Mennecke-Buehler and the proofreading team (Springer). Dutta
and Mukherjee (2021) encapsulate this work.

References

Babel, M. (2004). Badenian evaporite basin of the northern Carpathian Foredeep as a drawdown
salina basin. Acta Geologica Polonica, 54(3), 313–337.
Barabasch, J., Urai, J. L., Raith, A. F., & de Jager, J. (2019). The early life of a salt giant: 3D
seismic study on syntectonic Zechstein salt and stringer deposition on the Friesland Platform,
Netherlands. Zeitschrift Der Deutschen Gesellschaft FürGeowissenschaften, 170, 273–288.
Carter, N. L., Horseman, S. T., Russell, J. E., & Handin, J. (1993). Rheology of rocksalt. Journal
of Structural Geology, 15(9–10), 1257–1271.
Csontos, L., & Vörös, A. (2004). Mesozoic plate tectonic reconstruction of the Carpathian region.
Palaeogeography, Palaeoclimatology, Palaeoecology, 210, 1–56.
Dutta, D., & Mukherjee, S. (2021). Introduction to Structural Geology and Tectonics Field Guide-
book—Volume 1. In S. Mukherjee (Ed.), Structural Geology and Tectonics Field Guidebook—
Volume 1. Switzerland AG: Springer Nature. Cham. pp. xi-xvi. ISBN: 978-3-030-60142-3.
De Leeuw, A., Bukowski, K., Krijgsman, W., & Kuiper, K. F. (2010). Age of the Badenian salinity
crisis; impact of Miocene climate variability on the circum-Mediterranean region. Geology, 38(8),
715–718.
Har, N., Rusz, O., Codrea, V., & Barbu, O. (2010). New data on the mineralogy of the salt deposit
from Sovata (Mures, county-Romania). Carpathian Journal of Earth and Environment Sciences,
5(2), 127–135.
Hippolyte, J. C., & Săndulescu, M. (1996). Paleostress characterization of the ‘Wallachian’ phase
in its type area. Southeastern Carpathians, Romania: Tectonophysics, 263, 235–269.
Hudec, M. R., & Jackson, M. P. A. (2012). De Re Salica: Fundamental principles of salt tectonics.
In D. G. Roberts & A. W. Bally (Eds.), Regional geology and tectonics: Phanerozoic passive
margins, cratonic basins and global tectonic maps (pp. 19–41). Amsterdam, The Netherlands:
Elsevier.
Jackson, M. P. A., & Hudec, M. R. (2017). Salt tectonics: Principles and practice. Cambridge:
Cambridge University Press.
Jackson, M. P. A. (1995). Retrospective salt tectonics. In M. P. A. Jackson, D. G. Roberts & S.
Snelson, (Eds.), Salt tectonics: A global perspective: AAPG Memoir, (Vol. 65, pp. 1–28).
Krézsek, C., & Filipescu, S. (2005). Middle to Late Miocene sequence stratigraphy of the
Transylvanian Basin (Romania). Tectonophysics, 410(1–4), 437–463.
186 D. M. Tămas, et al.

Krézsek, C., & Bally, A. W. (2006). The Transylvanian Basin (Romania) and its relation to the
Carpathian fold and thrust belt: Insights in gravitational salt tectonics. Marine and Petroleum
Geology, 23, 405–442.
Kukla, P. A., Urai, J. L., Raith, A., Li, S., Barabasch, J., & Strozyk, F. (2019). The European Zechstein
Salt Giant—Trusheim and Beyond. In Presented at the 37th Annual GCSSEPM Foundation
Perkins-Rosen Research Conference.
Li, S., & Urai, J. L. (2016). Rheology of rock salt for salt tectonics modeling. Petroleum Scince,
13(4), 712–724.
Mat, enco, L., & Bertotti, G. (2000). Tertiary tectonic evolution of the external East Carpathians
(Romania). Tectonophysics, 316, 255–286.
Mrazec, L. (1905). Contribuţiune la geologiaregiuneiGuraOcnit, ei—Moreni. MonitorulPetrolului,
28, 785–788.
Mrazec, L. (1907). Despre cute cu sîmbure de străpungere. BuletinulSocietăt, ii de S, tiint, e, 16, 6–8.
Mrazec, L. (1910). Über die Bildung der rumänischenPetroleumlagerstätten. Third International
Petroleum Congress (1907), Bucharest, Comte Rendu, 2, 80–134.
Mrazec, L., & Teisseyre, W. (1902). Priviregeologicăasupraformaţiunilorsalifereşizăcămintelor de
sare din România. MonitorulPetroluluiRomân, 3, 1–55.
Murgoci, G. M. (1905). Tertiary formations of the Oltenia with regard to salt, petroleum and mineral
springs. The Journal of Geology, 13(8), 670–712.
Mukherjee, S., Talbot, C. J., & Koyi, H. A. (2010). Viscosity estimates of salt in the Hormuz and
Namakdan salt diapirs, Persian Gulf. Geological Magazine, 147, 497–507.
Nettleton, L. L. (1934). Fluid mechanics of salt domes. AAPG Bulletin, 18, 1175–1204.
Peryt, T. M. (2006). The beginning, development and termination of the Middle Miocene
Badeniansalinity crisis in Central Paratethys. SedGeol, 188–189, 379–396.
Popescu, G., Ciupagea, D., Georgescu, C., Balteş, N., & Motaş, C. (1973). Privire de
ansambluasuprageologieiformaţiunilorsalifere din România. Petrol şi Gaze, 24(9), 525–532.
Posepny, F. (1871). StudienausdemSalinargebieteSiebenbürgens. Kaiserlich-
KöniglichenGeologischenReichsanstaltJahrbuch, 21, 123–186.
Răbăgia, T., Mat, enco, L., & Cloetingh, S. (2011). The interplay between eustacy, tectonics and
surface processes during the growth of a fault-related structure as derived from sequence
stratigraphy: The Govora—Ocnele Mari antiform, South Carpathians. Tectonophysics, 502,
196–220.
Raith, A., Strozyk, F., Visser, K., & Urai, J. L. (2016). Evolution of rheologically heterogeneous
salt structures: a case study from the NE Netherlands. Solid Earth, 7(1), 67–82.
Raith, A., Urai, J. L., & Visser, K. (2017). Structural and microstructural analysis of K-Mg salt
layers in the Zechstein 3 of the Veendam Pillow, NE Netherlands: development of a tectonic
mélange during salt flow. Netherlands Journal of Geosciences, 96(4), 331–351.
Rowan, M. G., Urai, J. L., Fiduk, J. C., & Kukla, P. A. (2019). Deformation of intrasalt competent
layers in different modes of salt tectonics. Solid Earth, 10, 987–1013.
Săndulescu, M. (1984). Geotectonica României, Editura Tehnică Bucureşti, 335 p.
Săndulescu, M. (1988). Cenozoic tectonic history of the Carpathians. In L. Royden, & F. Horváth
(Eds.), The Pannonian Basin: A study in basin evolution (Vol. 45, pp. 17–25). AAPG Memoir.
Salina Turda (2019). Retrieved from https://www.salinaturda.eu/.
Salina Praid (2019). Retrieved from http://www.salrom.ro/praid.php.
Salina Slănic (2019). Retrieved from http://www.salrom.ro/slanic-prahova.php.
Salina Ocnele Mari (2019). Retrieved from http://www.salrom.ro/ocnele-mari.php.
Schléder, Z., Man, S., & Tămas, , D. M. (2016). History of salt tectonics in the Carpathian Bend
Zone, Romania. In AAPG European Regional Conference and Exhibition, Bucharest, Field-trip
guide, May 18, 2016, 22p.
Schléder, Z., Tămas, , D. M., Arnberger, K., Krézsek, C., & Tulucan, A. (2019). Salt tectonics style
in the Bend Zone sector of the Carpathian fold and thrust belt, Romania. International Journal
of Earth Sciences, 108(5), 1595–1614.
A Field Guide to the Spectacular Salt Mines … 187

Schmid, S. M., Bernoulli, D., Fügenschuh, B., Mat, enco, L., Schefer, S., Schuster, R., Tischler, M.,
& Ustaszewski, K. (2008). The Alpine–Carpathian–Dinaridic orogenic system: Correlation and
evolution of tectonic units. Swiss Journal of Geoscience, 101(1), 139–183.
Schultz-Ela, D. D., Jackson, M. P. A., & Vendeville, B. C. (1993). Mechanics of active salt diapirism.
Tectonophysics, 228(3–4), 275–312.
Tămas, , D. M., Schléder, Z., Tămas, , A., Krézsek, C., Copot, B., & Filipescu, S. (2019), Middle
Miocene evolution and structural style of the Diapir Fold Zone, Eastern Carpathian Bend,
Romania: Insights from scaled analogue modelling, In J. Hammerstein, R. Di Cuia, P. Griffiths,
M. Cottam, G. Zamora, R. Butler (Eds.), Fold and thrust belts; fold and thrust belts: Struc-
tural style, evolution and exploration (Vol. 490). London: Geological Society of London, Special
Publications. https://doi.org/10.1144/sp490-2019-091.
Tămas, , D. M., Tămas, , A., Barabasch, J., Rowan, M. G., Schléder, Z., Krézsek, C., & Urai, J.L.
(2020a). Low-angle shear within the exposed Manzalesti salt diapir, Romania: incipient decap-
itation in the Eastern Carpathians fold-and-thrust belt. https://doi.org/10.1002/essoar.105044
78.1.
Tămas, , A., Tămas, , D. M., Krezsek, C., Schleder, Z., Palladino, G., & Bercea, R. (2020b). The nature
and significance of sand intrusions in a hydrocarbon-rich fold and thrust belt: Eastern Carpathians
Bend Zone, Romania, Journal of the Geological Society, 177(2), 343–356.
Tămas, , D. M., Schléder, Z., Krézsek, C., Man, S., & Filipescu, S. (2018). Understanding salt in
orogenic settings: the evolution of ideas in the Romanian Carpathians. AAPG Bulletin, 102(6),
941–958.
Tilit, ă, M., Scheck-Wenderoth, M., Mat, enco, L., & Cloetingh, S. (2015). Modelling the coupling
between salt kinematics and subsidence evolution: Inferences for the Miocene evolution of the
Transylvanian Basin. Tectonophysics, 658, 169–185.
Trusheim, F. (1960). Mechanism of salt migration in northern Germany. Bulletin of the American
Association of Petroleum Geologists, 44(9), 1519–1540.
Urai, J., Means, W. D. & Lister, G. (1986). Dynamic recrystallization of minerals. In B. E. Hobbs,
& H. C. Heard (Eds.), Mineral and rock deformation: Laboratory studies: The Paterson volume
(Vol. 36, pp. 161–199). American Geophysical Union Geophysical Monograph.
Urai, J. L., Schléder, Z., Spiers, C., & Kukla, P. (2008). Flow and transport properties of salt rocks. In
R. Littke, U. Bayer, D. Gajewski, & S. Nelskamp (Eds.), Dynamics of Complex Intracontinental
Basins: The Central European Basin System (pp. 291–304). Berlin Heidelberg: Springer.
van Keken, P. E., Spiers, C. J., van den Berg, A. P., & Muyzert, E. J. (1993). The effective viscosity
of rocksalt; implementation of steady-state creep laws in numerical models of salt diapirism.
Tectonophysics, 225(4), 457–476.
Vendeville, B. C., & Jackson, M. P. A. (1992a). The rise of diapirs during thin-skinned extension.
Marine and Petroleum Geology, 9, 331–354.
Vendeville, B. C., & Jackson, M. P. A. (1992b). The fall of diapirs during thin-skinned extension.
Marine and Petroleum Geology, 9, 354–371.
Warren, J. K. (2016). Evaporites, a geological compendium (2nd edn., 1813pp). Switzerland:
Springer International Publishing.
Wawersik, W. R., & Hannum, D. W. (1980). Mechanical behavior of New Mexico rock salt in
triaxial compression up to 200 Degrees C. Journal of Geophysical Research, 85(B2), 891–900.
Zamfirescu, F., Mocut, ă, M., Constantinescu, T., Nita, C. & Danchiv, A. (2007). The main causes
and processes of instability at Field II of Ocnele Mari, Romania. In Solution Mining Research
Institute, Technical Conference, pp. 1–7.
Spectacular Sandstone Rock Cities
in the Czech Republic

Lucie Novakova and Petr Novak

Abstract We describe five spectacular geological and geomorphological sandstone


rock cities in the Czech Republic. Each region is famous and exceptional for its
reliefs and geological interestingness. These five localities are: (1) Hrubá Skála, (2)
Suché skály, (3) Kokořín area, (4) Dutý Kámen and (5) Pravčická brána. It is worth
to spend a day in each locality to admire the architecture of nature and the physical
and chemical processes like selective weathering. Rocks described in this guide
were formed during Cretaceous when the seawater attended the Bohemian Massif
~100 Ma years ago. In the Cretaceous, the sand was subsided into the Bohemian
Cretaceous Basin and compacted in the shallow water at ~200 m depth. During the
Alpine orogenesis, the Bohemian Massif was cut by NW–SE- and NE–SW-oriented
main faults. Blocks of compact sandstones were broken and eroded. Spectacular
sandstone rock cities formed eventually. There are substantial sandstone labyrinths,
tiny sandstone towers, windows, caves, stone plates, mushrooms, balanced boulders
and other features.

1 Introduction

By definition, a sandstone rock city would be a landscape where an originally compact


sandstone rock massif eroded into a complex maze of deep narrows and furrows.
Upright walls of blocks and towers rise several tens of metres (Sitenský 1984). Such
definition unveils both origin of the rock city and its basic morphology. Rock cities
are similar and yet different. Individual forms, however, provide insight into the
prerequisites needed to form a sandstone rock city. Sitenský (1984) suggested the

L. Novakova (B)
Institute of Rock Structure and Mechanics, Czech Academy of Sciences, V Holesovickach 41,
182 09 Prague, Czech Republic
e-mail: lucie.novakova@irsm.cas.cz
P. Novak
Watrad Ltd, Osadni 26, 170 00 Prague, Czech Republic
e-mail: pnovak@watrad.cz

© Springer Nature Switzerland AG 2021 189


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_7
190 L. Novakova and P. Novak

following conditions and characteristics necessary for evolvement of a sandstone


rock city.
(1) The area should be petrologically uniform (e.g. made up of sandstone). This
condition is not strict since in several rock cities veins and local intrusions are
known (Jetřichovické skály, Dubské skály).
(2) The sandstone must be massive and consistent enough to form vertical walls,
towers and pillars.
(3) The rock, on the other hand, must also be hydraulically permeable allowing
infiltration of a significant quantity of rainwater. Water erosion should not only
work on the surface but also enlarge deeper fractures.
(4) The massif must be fractured. Numerous long vertical or sub-vertical fractures
present an almost ideal environment for a rock city not only in sandstones but
also in granites.
(5) The massif should be exposed or otherwise situated to allow effective water
erosion forming deep valleys. Typical are table mountains (Ostaš, Adršpach-
Teplice area), valley upper edge positions (Hřensko, Drábské světničky) or in
some cases tectonic rearrangement (Suché skály, Pantheon). The massif exposi-
tion also means no or little soil cover is present to limit the rainwater infiltration.
The same condition applies to vegetation.
(6) The massif must have been exposed to erosion for an extended period (up to tens
to hundreds of thousands or even millions of years). The exposition, however,
must not be too long to allow erosion to evolve into complete denudation of the
original massif.
The optimal environment is therefore the area where erosion is effective but not
too much. Intensive and rapid weathering, as in tropics, does not provide an optimal
environment. Considering the listed limitations, Sitenský (1984) pointed out that the
temperate climate provides the optimal conditions for a rock city to emerge and last
(Fig. 1).

2 Geological Settings

The Bohemian Massif is a part of the Variscan orogenic belt formed 380–300 Ma
ago. The old part of continental crust is composed of Precambrian and Palaeozoic
rocks. Afterwards in Mesozoic, the old rocks were covered by younger sediments.
The most significant area with the sedimentary cover in the Bohemian Massif is
called the Bohemian Cretaceous Basin. The Bohemian Cretaceous Basin ranges
from Germany and comprises most of the Czech Republic with the area of 14,600
km2 . Limnic sediments deposited during the Early Cenomanian (ca. 100 Ma). After
the Cenomanian transgression, the basin was flooded by a shallow sea. Adamovič
et al. (2011) indicated the youngest well-maintained sediments of Santonian age (ca.
85 Ma). Up to 1100-m-thick Upper Cretaceous sedimentation was caused by the
seawater (Chlupáč 2002).
Spectacular Sandstone Rock Cities in the Czech Republic 191

Fig. 1 Origin of the


sandstone rock city. Drawing
by the author

During the Alpine orogenesis, the sedimentary cover was fractured. The Bohemian
Cretaceous Basin was segmented into small descending parts connected to the
transtensional blocks (Ulicný et al. 2003). Major younger faults striking NW–SE were
reactivated—Lusatian Fault (Coubal et al. 2015), bounding the Bohemian Cretaceous
Basin on the north and far away to the east the Sudetic Marginal Fault (Novakova
2015). Both fault zones belong to the fault set activated during the NE–SW compres-
sion in the latest Cretaceous and earliest Tertiary (much like the Franconian Fault in
Bavaria) and now characterize the boundaries of the Bohemian Massif (Adamovič
et al. 2011). The intraplate volcanism during the Late Oligocene and Early Miocene
initiated the development of extensional basins associated with the faulting (Peterek
et al. 1997).

3 Geomorphologic Settings

The tectonic deformations are monitored by the growth of large-scale landforms,


such as tilting of blocks and jointing. Bohemian rock cities were generally formed
by orthogonal jointing during fracturing the big blocks of sandstone. The lower parts
of the basins are mostly built of fine-grained sandstone. Adamovič et al. (2011)
described the differences in sedimentary textures and structures maintained by the
transformations of small-scale weathering forms. The sediments that lay in the layers
horizontally at the bottom of the basins are termed beds or strata, separating each
other, and form sedimentary layers and bedding planes. Bedding planes with undu-
lations expose paleo-current trend. The inclined bedding, on the other hand, proves
192 L. Novakova and P. Novak

the tilting of the block. The vertically sorted bedding may suggest turbidity flows
subsequent to a continental gradient.

3.1 Petrography of the Sandstones

Sedimentary rocks compose only 5% of the volume of the continental crust. Never-
theless, they cover most of the ocean floors and 75% of continents (Lambert D &
the Diagram Group 2007). Most sedimentary rocks are clastic from particles eroded
from rocks on land. These fragments range in size from microscopic grains (clays)
up to boulders. Besides the clastic sedimentary rocks, others, especially carbonates,
come from chemical precipitates or the remains of living organisms. The material
forming the Bohemian rock cities is mostly medium- to coarse-grained sandstones.
Sometimes, layers of fine-grained conglomerates also can be found. Quartz grains,
white, grey or milky white, predominate though white feldspar grains, dark grey
quartzite grains, pale mica flakes, clay particles and other accessories might be
present (Fig. 2). Later on, quartz, calcite, clay minerals and even iron oxides had
precipitated in pores of the yet unconsolidated sediment. Statistically speaking, most
of the rock cities’ sandstones, ~80%, are quartzose. The rests are feldspathic sand-
stones and subgreywackes, and even arkoses (Sitenský 1984). The vast majority
of the Bohemian sandstone rock cities have evolved in Turonian and Coniacian
block sandstones. Block disintegration so typical for most rock cities originates
from characteristic orthogonal fracturing of these sandstones.

Fig. 2 Different colours of quartz grains of up to 19 mm in size. Kokořín area (Kokorin lids),
GPS50.4561072°N, 14.5899383°E. Coin 1 Euro for scale
Spectacular Sandstone Rock Cities in the Czech Republic 193

3.2 Physical and Chemical Weathering

Variety of the rock has been extended even more due to secondary processes like
weathering and cementation. Weathering and erosion of sandstones produce land-
forms such as arches, alcoves, pedestal rocks and pillars (Bruthans et al. 2014) and
honeycombs (Bruthans et al. 2018). Honeycomb weathering typically develops in
coarse-grained sediments like sandstones (Fig. 3). Honeycombs are conserved for a
long time in an exact shape and size. The horizontal bases of honeycombs incline to
sustain in a horizontal position. However, the honeycombs itself follow the landform
shapes.
Fresh rock colour varies from the most common light yellow and ochre to clean
white and grey or vivid brown-reddish tones (Sitenský 1984). The combination of
intensive impregnation of sandstones with limonite (a mixture of iron hydroxides) and
selective weathering results in a variety of, often almost fantastic, shapes (Fig. 4a).
On the surface, the rock towers and block in the rock cities are usually greys due to
weathering salt crust (Fig. 4b). Typically, 2–3-cm thick iron layers are sometimes
formed by nearly pure limonite. In some cases, the formations of iron sandstones,
various concretional, partly rosy, head-like and tubular reddish-brown to black, are
of bowl forms. The sandstone rocks are divided by cracks into several blocks, the
widening of which, together with the foot-shaped boulder heaps, testifies to intense
Pleistocene selective frost weathering (Fig. 5).
The blowing wind, water erosion and the frost are responsible for fracturing
and breaking the sandstone to form the windows or pillars (Fig. 6a) and gates or
arches (Fig. 6b). The rock windows might be the originators for arches. The primary
mediator of the weathering is water, while frost weakens the rock. Sand grains that
are adjourned by the wind cause the rounded surface. The combined forces push to

Fig. 3 Honeycomb weathering with inclined bedding (Hrubá Skála, the Podmokelská Rock Tower,
GPS 84 50.5519639°N, 15.1772506°E). Coin 1 Euro for scale
194 L. Novakova and P. Novak

Fig. 4 a Undulating layers rich of iron (Hrubá Skála, a roadside outcrop, GPS84 50.5461719°N,
15.1875586°E), b White clean sand eroded by the water is covered by grey salt crust. Hrubá Skála,
the Větrník Rock Tower (The Windmill Rock Tower), north side, (GPS84 50.5480025°N,
15.1944989°E). Hammer Estwing E3-14P, overall length 279 mm

Fig. 5 Selective weathering of the sandstone rocks. During the late Tertiary and Quaternary, the
weathering and erosion caused the origin of separated blocks. The acid rainwater started the process
of salt weathering. Together with frost, the weather is responsible for relief modelling. The top cover
is rich in iron and more stable than the lower easily eroded sandstone layers. Drawing by the author
Spectacular Sandstone Rock Cities in the Czech Republic 195

Fig. 6 a Rock pillar with windows and pillars supports a freestanding arch with mass layers on the
top. b The weak parts are eroded and gradually extended. Hrubá Skála, the Větrník Rock Tower
(The Windmill Rock Tower), west side. GPS84 50.5478836°N, 15.1945269°E. Hammer Estwing
E3-14P, overall length 279 mm

the weak rounded grains and form a window or even in an extreme force give rise to
arches (Cruikshank and Aydin 1994).

4 Localities

Five localities described in this field guide offer an opportunity to explore amazing
geological sandstone features. These are: (1) Hrubá Skála, (2) Suché skály, (3)
Kokořín area, (4) Dutý kámen and (5) Pravčická brána (Fig. 7). The sandstone rock
cities are located approximately in one and half hour driving distance from Prague.
We recommend to spend at least one day at each locality, but it is also possible to
visit all localities in one or two days. The selection is significant, and one does not
need to be a geologist to enjoy the scenarios of the sandstone rock cities. All places
can be visited accessible on foot. The walking distances differ within the localities
from one to seventeen kilometres.
196 L. Novakova and P. Novak

Fig. 7 Schematic map of the Czech Republic with excursion localities. (1) Hrubá Skála, (2)
Suché skály, (3) Kokořín area, (4) Dutý kámen and (5) Pravčická brána

4.1 Hrubá Skála (Course Rock Sandstone City)

The Hrubá Skála sandstone rock city is one of the best-known rock areas in the
Czech Republic. It belongs to the Bohemian Paradise area. It is characterized by
extraordinary sandstone towers and also by steep canyons. There are >400 rock
towers reaching up to 55 m. The Hrubá Skála sandstone rock city is divided into
three main areas: the Dračí Skály (The Dragon Rocks) and Zámecká rokle (The
Chateau Valley), Kapelník (The Bandmaster), Maják (The Lighthouse) and Údolíčko
(The Small Valley). The famous trail leading along the top parts of the rock city is
called golden trail of the Bohemian Paradise. This trail offers picturesque lookout
points, such as the Mariánská Vyhlídka (Marian Lookout), Vyhlídka Na Kapelu
(Lookout band) or Vyhlídka U Lvíčka (Lion’s Lookout). The sandstone is relatively
low resistant, so there are many honeycombs and joints. For this reason, it is a popular
destination for climbers (Fig. 8).
Spectacular Sandstone Rock Cities in the Czech Republic 197

Fig. 8 Schematic plan of the trail with the stops

GOLDEN TRAIL—8 km.


STOP P: Parking lot Sedmihorky spa.
50.5541714°N, 15.1959050°E.
From the parking lot, follow the blue trail to Janova Vyhlídka (Jan Lookout). Within
the road, it is possible to observe the sandstone towers with vertical joints. On the
trail, there is a narrow, 50-m-high tower called Taktovka (The Baton).
STOP 1: Janova Vyhlídka (Jan Lookout)
50.5527642°N, 15.1757992°E.
The viewpoint offers spectacular views of the Kozákov ridge and the area of the
Maják (The Lighthouse) and Čertova Ruka (The Devil’s Hand) rocks. Huge solely
tower with horizontal and vertical jointing forms orthogonal system of joints (Fig. 9).
STOP 2: Vyhlídka U Lvíčka (Lion’s Lookout).
50.5527642°N, 15.1757992°E.
If we follow the red tourist trail along the ridge from Hrubá Skála to Valdštejn Castle,
beautiful views of the rock town below us will begin to open up among the trees—
one of the viewpoints that show the Hrubá Skála rock town in all its beauty is the
Vyhlídka U Lvíčka (Lion’s Lookout). The original viewpoint was located only on
a short rocky promontory with limited views. Vertical joint on the tower top splits
198 L. Novakova and P. Novak

Fig. 9 Rock towers from Jan Lookout. On the right Lighthouse and Devil’s Hand Rock Towers

the “head” of the tower. Under the “head”, very nice honeycombs are developed
(Fig. 10).
STOP 3: Vyhlídka Na Kapelu (Band Lookout).
50.5512672°N, 15.1790111°E.
The Band Lookout can be found by the red-marked golden trail of the Bohemian
Paradise, about halfway between Valdštejn Castle and Hrubá Skála. As the name of
the Lookout suggests, it offers a view of a group of rocks called Band and the lonely
rock tower Bandmaster (Fig. 11).
STOP 4: Zámecká vyhlídka (Chateau Lookout).
50.5464708°N, 15.1931806°E.
We can get to the Chateau Lookout by following the yellow sign from the crossroads
at Adam’s bed. The viewpoint offers a beautiful view of the HrubáSkála chateau
and the surrounding rock formations. The rocks have variable colours from grey to
orange, meaning richness of iron. In the distance, we can observe the castle Trosky,
which is of the Tertiary volcanics (Fig. 12).
Spectacular Sandstone Rock Cities in the Czech Republic 199

Fig. 10 Vertical joint from


the Lion’s Lookout

STOP 5: Mariánská Vyhlídka (Marian Lookout).


50.5480192°N, 15.1935925°E.
Mariánská Vyhlídka is one of the most famous and most visited in the Bohemian
Paradise. It offers the famous postcard panorama of Hrubá Skála, Trosky Castle and
Kozákov Hill. Besides, there is a beautiful view of the group of Sahara rocks, the
Twelve Apostles, the Pinwheel, the Dragon Tower and other rocks. It lies at an altitude
of 360 m.n.m. on a rock block in the area of Dračské rocks, about 1.5 km from the
Hrubá Skála Chateau. Continue by the yellow tourist sign from the Adamovo lože
crossroads. Below Mariánská Vyhlídka is a symbolic cemetery of climbers. From
this Lookout, follow the yellow and blue trail back to the parking lot (Fig. 13).
200 L. Novakova and P. Novak

Fig. 11 Bandmaster tower.


Solely tower of 55 m high,
with well-developed
honeycombs and alcoves

4.2 Suché skály (Dry Rocks)

Formerly Suché skály were called Kantor’s organ; today, they are also called Czech
Dolomites. The stone backdrop was the result of the shifting of the Earth’s shrubs
when the lower layer of sandstone lifted and rotated perpendicular to the Earth
surface. Cenomanian rock cities (Suché skály, Pantheon) are the exceptional geolog-
ical locality. Dry rocks are a great climbing terrain. It is divided into about 20 towers—
Main, Falken, Middle and others. In 1965, Suché skály was declared a national natural
monument (Fig. 14).
SMALL DOLOMITE TRAIL—4 km.
STOP P: Parking lot Suchéskály, Koberovy.
50.6352097°N, 15.2107997°E.
There are no marked touristic trails. Follow the individual beaten roads. From the
parking lot, make a circle around the rock city and return.
Spectacular Sandstone Rock Cities in the Czech Republic 201

Fig. 12 View from Chateau Lookout. Far on the left Trosky castle built on the tertiary volcanics
chambers

STOP 1: Baba (The Old Woman).


50.6337506°N, 15.2154961°E.
Lonely woman (Baba) is a sandstone tower in the easternmost part of the Suché skály.
On the walls of Baba rock, the conjugate system of deformation bands is well devel-
oped. The bands can be used as kinematic indicators to determine the principal stress
directions (Fig. 15).
STOP 2: Hlavní věž (Main Tower).
50.6367225°N, 15.2120472°E.
In this region, regional tectonics (the Lusatian Fault) fetch out and turn up buried
sandstone layers exposing them right enough. Tectonic predisposition gave these
rock cities their distinctive ridge-like appearance. This is a unique geomorphologic
locality, which has no analogues in our country (Fig. 16).
STOP 3: Sokolívěž (Falken Tower).
50.6369564°N, 15.2114397°E.
Following the trail from the Main Tower, we can get to Falken Tower. Many
fault planes can be observed here. Well-developed slickenlines and tectonic mirrors
are used for paleo-stress analysis to determine the Lusatian Fault tectonic history
(Fig. 17).
202 L. Novakova and P. Novak

Fig. 13 Well-developed
rock window in the weak
layer of the rock tower

STOP 4: Střední věž (Middle Tower).


50.6368525°N, 15.2117200°E.
The Dry Rocks sandstone is more resistant than sandstones in other rock cities.
Conjugate deformation bands with the apparent offset in massive sandstone form the
resistant ledge on the walls. Grains sliding along shear plane have fewer points of
contact. Stress concentrations lead to the fracturing of the grains (Crider and Peacock
2004). The sandstone is heavily and irregularly jointed and does not show regular
horizontal and vertical segmentation as can be seen elsewhere in rock cities (Fig. 18).
STOP 5: Vyhlídka Suché skály (Dry Rocks Lookout).
50.6328836°N, 15.2064653°E.
The Lookout provides spectacular view or the rock formation. It is an articulated ridge
formed of Upper Cenomanian sandstone resembling the Italian Dolomites. The steep
rock with vertical layers aligned by the Lusatian Fault is a Czech geomorphologic
Spectacular Sandstone Rock Cities in the Czech Republic 203

Fig. 14 Schematic plan of the trail with the stops

Fig. 15 Well-developed honeycombs and deformation bands with apparent thrust offset. Hammer
Estwing E3-14P, overall length 279 mm
204 L. Novakova and P. Novak

Fig. 16 Rotated sandstone layers along the Lusatian fault due to the Alpine orogenesis

phenomenon. From this view, the Suché skály seem really like a small dolomites
(Fig. 19).
STOP 6: Sokolík (Little Falken).
50.6323225°N, 15.2046167°E.
The Little Falken is a small rock tower on the opposite side of the road to Malá Skála.
The trip is optional but recommended. On the tower, there is undulating bedding
within the almost whole tower, except the top part where horizontal bedding occurs
(Fig. 20).
Spectacular Sandstone Rock Cities in the Czech Republic 205

Fig. 17 Slickenlines on the fault plane of Falken Tower showing thrust movements. Coin 1 Euro
for scale

Fig. 18 Photograph of conjugate deformation bands on the Middle tower wall. Coin 1 Euro for
scale

4.3 Kokořínsko (Kokořín Area)

Kokořín area is a Cretaceous chalk plate area of Bohemia with unique geomorpho-
logical relief made of block sandstone. The central relief features are determined
by the relationship between the two main groups of surface forms: (1) plateaus and
often deep, terraced valleys, on the edges of which rock-column formations have
arisen rock towers, and (2) numerous medium-sized forms and micro-forms have
arisen by selective weathering in forms and to an extent not seen elsewhere in the
206 L. Novakova and P. Novak

Fig. 19 Main ridge of Suché skály

Czech Republic. Thanks to the large forested area and the sparse structure of housing
estates, this countryside is unusually conserved, and its unique beauty is also under-
lined by numerous examples of original log, half-timbered or brick-wall architecture
(Figs. 21 and 25).
HANDLES TRAIL—0.8 km.
STOP P: Parking lot Kokořín lids.
50.4547411°N, 14.5863650°E.
From the parking lot, follow the blue trail with signs “Pokličky”. Follow the path
and climb the stairs.
STOP 1: Kokořínské Pokličky (Kokořín lids).
50.4553989°N, 14.5889156°E.
The turbid towers in the medium-grained to coarse-grained sandstones of the Jizera
Formation are covered with a resistant layer of ferrous sandstones to conglomerates,
which form the so-called lids. These rock objects are the result of selective weathering
of sandstones. The coarsest sediments in the upper parts of sandstone towers with
rich incrustations of iron compounds are more resistant to erosion than the sandstones
below. The source of ironing of coarse-grained sandstones and conglomerates is the
penetrations of Tertiary volcanics further east. During Tertiary volcanism, iron-rich
groundwater actively circulated in these coarse-grained sediments. The source of
iron was the basic volcanics themselves, in which the magnetite and biotite reacted
(Fig. 22).
Spectacular Sandstone Rock Cities in the Czech Republic 207

Fig. 20 Sokolík (Little Falken) pillar

Fig. 21 Schematic plan of the trail with the stops


208 L. Novakova and P. Novak

Fig. 22 Kokořín lids. GPS 50.4553989°N, 14.5889156°E

STOP 2: Rock windows.


50.4553158°N, 14.5910619°E.
About 200 m east of “Pokličky” along the hiking trail, we pass sandstones, which
are located in the overburden. Numerous cavities and rock windows have formed
in these sandstones as a result of selective erosion of sandstone. The layer with a
more prosperous calcareous component was more easily cracked during diagenesis
than the surrounding, relatively more plastic quartz sandstone. The calcareous putty
eventually wholly dissolved in the perimeter of today’s cavities and the incoherent
sand was removed. From this stop, go back to the parking lot (Fig. 23).
STOP 3: Honeycomb tower.
50.4547917°N, 14.5879153°E.
The tower is situated in the trail with no signs. Follow the beaten trail along the
blue trail, and after ~200 m there is an “iron” tower showing many typical honey-
comb weathering features. The honeycombs are developed in the horizontal layers
following the sedimentation. Relations between rock crust and honeycomb pits can
be demonstrated; pits are missing at places with fresh exfoliations (Fig. 24).
Spectacular Sandstone Rock Cities in the Czech Republic 209

Fig. 23 Rock window with alcoves and honeycombs. Hammer Estwing E3-14P, overall length
279 mm

Fig. 24 Rock tower with


honeycomb wall. Hammer
Estwing E3-14P, overall
length 279 mm
210 L. Novakova and P. Novak

Fig. 25 Schematic plan of the trail with the stops

IRON ROSE TRAIL–0.7 km.


STOP P: Parking lot Křenov.
50.5231211°N, 14.5748908°E.
From the parking lot, the fastest way how to manage the hill is climbing up. Several
hundred metres up the rocks can be seen.
STOP 1: Stone hill Křenov.
50.5218364°N, 14.5738022°E.
According to Adamovič (1994), a short rocky ridge of the ENE–WSW direction
was formed by quartz sandstones of the central part of the Jizera Formation (Middle
Turonian), in places permeated with iron putty. In the western part, 28 m long, 7 m
wide and 12 m high, meso- and micro-forms of relief are developed, conditioned,
among other things, by the occurrence of iron inlays (cornices, slabs, stone roses).
The top of the wooded hill is protected as a geological peculiarity. The altitude is of
370–382 m above sea level. These rock shapes—incrustations created by selective
weathering of rocks—are protected. They look like iron roses (Fig. 26).
STOP 2: Balanced boulder.
50.5218797°N, 14.5729725°E.
The boulderstone is balanced of about 5 cm. Its dimensions are 1.8 × 1.7 × 1.5 m.
The collapsed sandstone blocks and boulders show the effects of frost in the cold
periods of the Pleistocene. Predominantly medium- to coarse-grained or coarse-
grained, occasionally to gravelly quartz block sandstones in the top part of the hill
belong to the middle part of the Jizera Formation (Malkovčin and Sedláček 2009).
In the lower part of the slopes, the sandstones of the lower part of the formation are
also represented (Fig. 27).
Spectacular Sandstone Rock Cities in the Czech Republic 211

Fig. 26 Iron roses on the top of Křenov hill. The cover from Canon D250 camera, diameter 67 mm,
for scale

Fig. 27 Boulderstone on the foot of the Křenov hill. Hammer Estwing E3-14P, overall length
279 mm

4.4 Dutý kámen (Hollow Stone)

The unique occurrence of sandstone formations with pentagonal and hexagonal


columns in contact with the primary igneous rocks has been protected as a natural
monument since 1963. It is a sandstone ridge elevated 30 m above the surrounding
terrain. In the past, sandstone was mined here for construction use (Fig. 28).
212 L. Novakova and P. Novak

Fig. 28 Schematic plan of


the trail with the stops

Columnar Jointing Trail—2 km.


STOP P: Parking lot Motel Dutý kámen.
50.7746972°N, 14.6562244°E.
From the parking lot, follow the trail along the main road in the meadow. Then,
climb small hill to the ridge of the rocks. The trail is then ridge-marked with green
triangles.
STOP 1: Beautiful wall.
50.7706111°N, 14.6567500°E.
Walking on the ridge, we will meet a rock tower called Beautiful Wall. There is a
spacious cave on the base of the tower (Fig. 29).
STOP 2: Columnar tower.
50.7703036°N, 14.6560656°E.
The contact of sandstone and basalt vein on Hollow Rock exhibits an interesting
columnar jointing of sandstone. Joints have a diameter of 2.5–5 cm, up to 2.5 m
long. In some places, they are arranged parallel and steeply inclined (80–60° to the
NW); in some places, they are arranged in a fan shape. A polzenite vein caused the
formation of sandstone columns on the Hollow Stone. The hot gases and vapours that
accompanied its penetration rose the cracks and heated the surrounding sandstone
to a high temperature. Although this temperature was not enough to melt the rock,
it caused it to solidify by crystallization. During the subsequent cooling, together
with the reduction of the rock volume, contraction cracks formed and the sandstone
Spectacular Sandstone Rock Cities in the Czech Republic 213

Fig. 29 Beautiful Wall with


the cave 4 m high

cracked into thin vertical slabs. Near the cracks, where the heating was most intense,
these slabs produced transverse cracks. Where there was substantial compaction of
transverse cracks, small quadrilateral to hexagonal sandstone columns formed, they
can be referred to as contact columns. Therefore, the columns developed only close to
cracks and at greater distances, they quickly pass through plate-shaped sections into
solid sandstone (Balatka and Sládek 1972; Havránek 1982). On Hollow Stone, the
columnar separation of sandstone is visible in several places, but it is best developed
on a 2.5-m-high rock gorge, which stands about in the middle of the ridge. The
sandstones forming the columns are strongly silicified. There are small abandoned
pits in the sandstone ridge (Fig. 30).
STOP 3: Vyhlídka Dutý kámen (Hollow Stone Lookout).
50.7702011°N, 14.6555956°E.
On one of the sandstone rocks called Dutý kámen, there is a Lookout, from which
there is a nice view of the landscape. You will see Zelený vrch (The Green Hill) with
214 L. Novakova and P. Novak

Fig. 30 Columnar jointing


of the sandstones. The cover
from Canon D250, diameter
67 mm, for scale

the settlements of Drnovec, Cvikov, Klíč (The Key Hill) and the hill Ortel. On the
circular rock plateau, there is a stone table with a bench. On the east side of the table
is carved the date of manufacture 7/19/1914 (Fig. 31).

4.5 Pravčická brána (Pravčice Gate)

The Pravčice Gate is the largest natural rock gate in Europe. It is considered to be the
most beautiful natural formation in Bohemian Switzerland and forms a symbol of
the whole area. In the forests near Hřensko town, surrounded by a mysterious calm,
stands the Pravčice Gate. For centuries, it has attracted lovers of natural beauty with
its charm. From the entrance to the Pravčice Gate, there are groomed paths and stairs
to the individual viewpoints rising high above the treetops. In beautiful weather, there
is the possibility to see to Děčínský Sněžník. Nowadays, to protect the gate, it is not
allowed to visit the top of the gate (Fig. 32).
Spectacular Sandstone Rock Cities in the Czech Republic 215

Fig. 31 View from the Hollow Rock Lookout

Fig. 32 Schematic plan of the trail with the stops

GATE TRAIL—17 km.


STOP P: Parking lot in Hřensko town “Restaurant Klepáč”.
50.8750267°N, 14.2538756°E.
Park the car on the parking lot in Hřensko town. Then, follow the yellow trail to the
Edmund Gorge. The route is not accessible by foot, and it is necessary to use a boat.
There are two boats on the trail. The boats are operational from April to October. If
216 L. Novakova and P. Novak

you plan to come another time, or not willing to use the boats, then use an optional
trail by biking road n. 21 to Mezná village.
STOP 1: Edmundova soutěska (Edmund Gorge).
50.8695386°N, 14.2759125°E.
Edmund Gorge, formerly called Silent, is a rocky canyon of the Kamenice River, the
last of three gorges on the lower reaches of the river, east of Hřensko town. Its walls
form steep, often perpendicular rock walls, rising above the water level to a height
of 50–150 m, with a visible blocky decay of sandstones, and some boulders slid to
the bottom of the canyon. In the section where the rocks fall directly into the water,
boating on boats replaces the path. Access from the upper edge of the Hřensko area
is on foot along the left and then along the right bank of the Kamenice under rock
overhangs with three tunnels to the raft, above which is the lower boat dock. Above
the Gorge rise several rocks with exciting formations, such as the Rock Family or
Strážce (Guardian) (Fig. 33).
STOP 2: Divoká soutěska (Wild Gorge).
50.8640464°N, 14.3057319°E.
The Wild Gorge is a deep rocky canyon of the Kamenice River, adjacent to the
Edmund Gorge at the Mezní Bridge. From the Limit Bridge, the trail leads to the
lower dock, another 450 m long section accessible only by boat. From the Wild
Gorge, follow the yellow sign, and then change to the blue sign Mezní louka. Where
the blue trail ends in Mezní louka village, change the trail to red to Pravčická brána
(Fig. 34).
STOP 3: Homole (Cone Tower).

Fig. 33 Rocks in the Edmund Gorge


Spectacular Sandstone Rock Cities in the Czech Republic 217

Fig. 34 Scenery of the Wild Gorge

50.8826661°N, 14.2943889°E.
The Cone Tower is a lonely tower of 40 m height with horizontal jointing. On the
upper part, there are well-developed honeycombs oriented horizontally according to
the jointing (Fig. 35).
STOP 4 Vyhlídka Pravčická brána (Pravčice Gate Lookout).
50.8837122°N, 14.2810267°E.

Fig. 35 Cone tower


218 L. Novakova and P. Novak

Fig. 36 Pravčice Gate. The view from far Lookout

The Pravčice Gate is the biggest and most spectacular natural sandstone arch in the
Europe. The dimensions of the Pravčice Gate are respectable, but when viewed at
close range, they seem even more significant. The span of the arch at the bottom is
26.5 m, the height of the opening is 16 m, the width is 7–8 m, the minimum thickness
is 3 m, and the top platform of the gate is 21 m above its bottom (Fig. 36).
From Pravčice Gate, follow the red trail to Three Streams and continue to Hřensko
town.

5 Summary

The sandstone rock cities in the Czech Republic provide spectacular views of all
types of features. There are many more trails and routes than described in this guide.
It is a paradise sensulato—the Bohemian and geological paradise. We choose only
a few of the trails and Lookouts to attract readers. Depending on the time reserved
for the field trips, we recommend to plan and visit all suggested localities. They are
worth to visit.

Acknowledgements We thank our friends and colleagues for their support. This work was carried
out thanks to the support of the long-term conceptual development research organization RVO:
67985891. Soumyajit Mukherjee reviewed three times and made minor comments. Thanks to
Marion Schneider, Annett Buettener, Boopalan Renu, Alexis Vizcaino, Doerthe Mennecke-Buehler
and the proofreading team (Springer). Dutta and Mukherje (2021) summarize this article.
Spectacular Sandstone Rock Cities in the Czech Republic 219

References

Adamovič, J. (1994). Paleogeography of the Jizera formation (Late Cretaceous sandstones), Kokořín
area, central Bohemia. Sbornik Geologickych Ved: Geologie, 46, 103–123.
Adamovič, J., Mikuláš, R., Schweigstillová, J., & Bohmová, V. (2011). Porosity changes induced by
salt weathering of sandstones, Bohemian Cretaceous Basin, Czech Republic. Acta Geodynamicaet
Geomaterialia, 8(1), 29–45.
Balatka, B., & Sládek, J. (1972). Sloupkovitýrozpadpískovců V Ralsképahorkaťině. Ochranapřírody
R., 10, 235–243.
Bruthans, J., Soukup, J., Vaculikova, J., Filippi, M., Schweigstillova, J., Mayo, A. L., et al. (2014).
Sandstone landforms shaped by negative feedback between stress and erosion. Nature Geoscience.
https://doi.org/10.1038/ngeo2209.
Bruthans, J., Filippi, M., Slavík, M., & Svobodová, E. (2018). Origin of honeycombs: Testing the
hydraulic and case hardening hypotheses. Geomorphology, 303, 68–83.
Chlupáč, I. (2002). Geological history of the Czech Republic (in Czech).
Coubal, M., Málek, J., Adamovič, J., & Štěpančíková, P. (2015). Late Cretaceous and cenozoic
dynamics of the Bohemian Massifinferred from the paleostress history of the Lusatian fault belt.
Journal of Geodynamics, 87, 26–49. https://doi.org/10.1016/j.jog.2015.02.006.
Crider, J. G., & Peacock, D. (2004). Initiation of brittle faults in the upper crust: A review of field
observations. Journal of Structural Geology, 26(4), 691–707.
Cruikshank, K. M., & Aydin, A. (1994). Role of fracture localization in arch formation, Arches
National Park, Utah. Geological Society of America Bulletin, 106(7), 879–891.
Dutta, D., & Mukherjee, S. (2021). Structural Geology and Tectonics Field Guidebook—Volume
1. In Mukherjee S. (Ed) Structural Geology and Tectonics Field Guidebook—Volume 1. Springer
Nature Switzerland AG. Cham. pp. xi-xvi. ISBN: 978-3-030-60142-3.
Havránek, P. (1982). Hollow Rock. Pam.a Příroda, 1, 59–60. Praha (in Czech).
Lambert D., and the Diagram Group. (2007). The field guide to geology (New, p. 296). New York:
Facts on File.
Malkovčin, P., & Sedláček, M. (2009). Protected landscapes of the Czech Republic. Middle
Bohemia, XIII. Agenturaochranypřírody a krajiny ČR aEkoCentrum Brno, Praha, 904 pp. (in
Czech).
Novakova, L. (2015). Tectonic phase separation applied to the sudetic marginal fault zone (NE part
of the Czech Republic). Journal of Mountain Science, 12(2), 251–267. https://doi.org/10.1007/
s11629-014-3297-5.
Peterek, A., Rauche, A., Schröder, B., Franzke, H.-J., Bankwitz, P., & Bankwitz, E. (1997). The
late- and post-Variscan tectonic evolution of the western border fault zone of the Bohemian massif
(WBZ). Geologische Rundschau, 86, 191–202.
Sitenský, J. (1984). Rock cities in Bohemia. Dita. pp. 219 (in Czech with English resume).
Ulicný, D., Cech, S., & Grygar, R. (2003). Tectonics and depositional systems of a shallow-marine,
intra-continental strike-slip basin: Exposures of the Cesky Raj region, Bohemian Cretaceous
Basin, in excursion guide, First meeting of the central European tectonics group and the eighth
meeting of the Czech tectonic studies group. Geolines, 16, 133–148.
Field Guide to RODS in the Pireneus
Syntaxis, Central Brazil

Marco Antonio Caçador Martins-Ferreira


and Sérgio Wilians de Oliveira Rodrigues

Abstract Linear structures are extremely important in structural mapping, they can
be used to separate deformation phases and to determine the kinematics of defor-
mation. Quartz rods are one of the most eye-catching linear structures in deformed
rocks. Despite relatively rare, rods are described in several places worldwide. Wher-
ever rods occur, they are promptly noticed. Rods form a conspicuous coarse lineation,
frequently highly contrasting with the surrounding rock in regions that was under high
strain. The term rod or rodding, in geology, broadly refers to a mass of rock, which has
assumed a cylindrical shape while accommodating strain; however, different defini-
tions are found in the literature. The mechanisms of rod formation can be constrained
from field observations. They are frequently parallel to fold axes and lie at right angles
to the direction of maximum compression. Rods are commonly confused to mullions,
boudins and pencil structures, since these are also elongated and rounded structures.
But striking differences in their geometry and genesis can be used to distinguish
between them. In this guide, the various definitions of rods and how they distin-
guish from other structures are discussed. An outstanding field occurrence of rods
is presented in a location where prominent rods were formed in the context of a
long and narrow corridor of high-strained rocks, the central region of the Pireneus
Orogenic Syntaxis, in central Brazil. There, nearly all veins, conglomerate clasts and
the rock mass itself have turned into quartz cylinders. This guide brings instructive
comparisons, photos and illustrations that will aid on distinguishing between these
structures in the field. The ten best locations to observe rods and the structures related
to their development are presented with maps and plenty of pictures and interpreta-
tions. For each observation station, a stereoplot is provided with the attitudes of the
main planes and lines measured at that station as well as their mean attitudes. Nearly
every object (clasts, veins, layers, etc.) has been caught in the process of rodding. A
direct link of the process of rodding to the development of fold-related foliations is

M. A. C. Martins-Ferreira (B) · S. W. de O. Rodrigues


Faculdade de Ciências e Tecnologia, Universidade Federal de Goiás (UFG), RuaMucuri, Setor
Conde dos Arcos, Aparecida de Goiânia, GO 74968755, Brazil
e-mail: martinsmarco@gmail.com
S. W. de O. Rodrigues
e-mail: sergiowilians@gmail.com

© Springer Nature Switzerland AG 2021 221


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_8
222 M. A. C. Martins-Ferreira and S. W. de O. Rodrigues

established. It is possible to observe that the variations in constriction and flattening


deformation are highly dependent on the fold geometry.

1 Introduction

One of the most eye-catching deformational structures are rods. Rods are relatively
rare structures, but where they occur are promptly noticed because they form a
conspicuous coarse lineation, frequently highly contrasting with the surrounding
rock in regions submitted to high strain.
Linear structures are extremely important in structural mapping as they can be used
to separate deformation phases and to determine the kinematics of deformation. In
poly-phase deformed terranes, the consistency of lineation orientations is a key factor
in the subdivision of a map into structurally homogeneous sub-areas. Therefore,
lineations must be systematically measured and recorded while mapping (McClay
1987).
In this guide, an outstanding field occurrence of rods is presented. These prominent
rods have formed in the geological context of a long and narrow corridor of high-
strained rocks in the central region of the Pireneus Orogenic Syntaxis (PSX), in
central Brazil. Here, nearly all veins, meta-conglomerate clasts and also the rock
mass itself have turned into quartz cylinders (Fig. 1).
The term rod or rodding, in geology, broadly refers to a mass of rock which has
assumed a cylindrical shape while accommodating strain. However, the definitions
of rods in literature are a bit loose, varying from author to author. While defining
rods, some authors consider only their geometrical characteristics, and others also
impose genetic conditions, such as the material they were formed from, or the process
involved in their formation. A review of these definitions is presented below.

1.1 Geometrical Definitions of Rods

Rodding is a process that occurs in metamorphic rocks, where the stronger parts,
such as vein quartz or quartz pebbles, have been shaped into parallel rods, forming a
linear structure. Whether the structure is formed parallel to the direction of transport
or parallel to the fold axes has been debated (Bates and Jackson 1987).
Rods are stretched and elongated fragments of competent material in a ductile
matrix. Pebbles, quartz veins or dyke fragments commonly form rods that are parallel
to either “a” or “b” tectonic axis. Rodding produces L-tectonites (McClay 1987).
Rod is a morphologic term for elongate, cylindrical and usually monomineralic
bodies of segregated mineral (quartz, calcite, pyrite, etc.) in metamorphic rocks of all
grades. Rods may have any profile outline, from elliptical to irregular, dismembered
rounded structures. Rods are generally parallel to local fold axes and often are isolated
fold hinges detached from their limbs (Burg 2019).
Field Guide to RODS in the Pireneus Syntaxis, Central Brazil 223

Fig. 1 Rodded clasts in clast-supported meta-conglomerate in station 9. Picture taken from


horizontal view. Hammer is ~40 cm long

1.2 Geometric-Genetic Definitions of Rods

Wilson (1982) is undoubtedly the most comprehensive study of rods in literature and
consequently the main reference on the subject. According to the author, “Rods are
cylindrical bodies of quartz or other minerals which have segregated in, or have been
introduced into the country-rock during deformation” (Wilson 1982, p. 91). However,
on p. 93 he further states: “Elongated pebbles do not conform to the definition of rods
given on p. 91, which applies to those derived from segregation quartz, nevertheless,
the final products of the two may be so difficult to separate in the field, that until
the original parentage of the structure can be established, the term rod, which is
descriptive and has no genetic significance, may with reason be applied to both.
Later it can be qualified if necessary”.
224 M. A. C. Martins-Ferreira and S. W. de O. Rodrigues

The idea of rods being formed exclusively from segregated or introduced material
in the country rock has been later abandoned when several other examples of rods
were observed worldwide, giving space to more broad definitions.
According to Davis and Reynolds (1996), quartz rods can be formed in a number
of ways, some may be boudined, necked expressions of once-continuous layers of
quartz whereas some may be interpreted as the linear equivalent of veins, produced
by open-space filling and/or replacement not along fractures but rather along the
hinge zones of penetrative folds. In the latter case, quartz may represent mineral
re-precipitation of silica made available by local pressure solution in the same host
rock in which the rods were formed.
The mechanisms of rod formation can be constrained from field observations.
They are frequently parallel to fold axes of different scales and lie at right angles
to the direction of maximum compression. The much longer x-axis of the strain
ellipsoid in rods suggests constrictional deformation, but this is not always the case;
rods can assume an elliptical shape in the yz basal plane, where y >> z and fall into
the flattening or oblate deformation fields.
Rods are commonly confused to mullions, few varieties of boudins and pencil
structures, since these are all elongated and rounded structures. However, striking
differences in their geometry and genesis can be used to distinguish between them.
In the next sections, this guide brings an instructive comparison that will aid on
distinguishing between these structures in the field.

1.3 Rods and Mullions

The term mullion stems from the old French “moinel”, designing the vertical columns
in tall windows of Gothic architecture. Mullions and rodding structures are both forms
of coarse lineation developed in strongly deformed rocks. However, rods differ from
mullions in that they are round all over their volume and not only at one side, see
Fig. 2. Also, their geneses differ. Mullions are coarse corrugation of the bedding
surface between a competent and an incompetent layer. Ultimately, mullions are
the result of the lack of strain continuity in rock layers with contrasting mechanical
properties.

Fig. 2 Cartoons showing the main morphological differences between rods, mullions and boudins
Field Guide to RODS in the Pireneus Syntaxis, Central Brazil 225

Mullions form at any size in the original rock material and are never formed by
segregated or introduced material, as rods can sometimes be. Their ribbed or grooved
appearance is often due to broad, smoothly curved convex surfaces of the competent
layer rather regularly separated by narrow, sharp, inward-closing hinges (Burg 2019).
Mullions of cuspate, lobate or bulged forms occur at later interfaces (Fig. 2). Rods,
on the other hand, have little or no direct relationship to layer interfaces and are thus
not restricted to them, being frequently found amid more ductile matrices.

1.4 Rods and Boudins

Like rods, boudins tend to be long rock bodies, but the two can be easily distinguished
in the field because like mullions, boudins are usually restricted to certain layers. This
happens because boudins and mullions both form by (i) heterogeneous layer-parallel
strain and (ii) contrast of layer competence. Boudins are long and parallel segments
of a competent layer sandwiched in a less competent rock (Fig. 2). Each segment is
separated by an extensional gap. The segmentation or thinning of the more competent
layer is induced by the flow of the less competent layer. Thus, boudins, unlike rods,
are most commonly aligned to each other in their intermediate axis (Fig. 2).
Boudins form by layer-parallel extension and their shape is highly dependent on
the ductility contrast between the strong and the weak layers. Boudins can assume
various shapes, viz. rectangle, rhomb and lozenge shape. When competent layers
do not reach the breaking point, narrow and thinned necks separate boudins. In
this case, a pinch-and-swell structure is formed, with do not much resemble with
rods. Boudin profiles can assume lenticular shape or have convex (sausage boudins),
straight (blocky boudins) and concave to extremely concave (fish-mouth) extremities
(Burg 2019). Sausage boudins are the ones that most resemble rods; however, they
can be diagnosed as boudins by the observation of aligned repetition and filling of
spaces between the gaps by the rock matrix.

1.5 Rods and Pencil Structures

While rods might occasionally resemble pencil structures, they are completely
different structures in terms of genesis and tectonic significance. The formation
of pencil structures is related to the interference between cleavages, frequently the
product of a compaction cleavage cut by a subsequent tectonic cleavage, or by the
interference between two tectonic cleavages developed at right angles to each other.
Pencil structures tend to be much less rounded than rods. They form a lineation
parallel to the intersection lineation of the two cleavages and differently from rods,
which are typical of non-metamorphic or very low-grade metamorphic rocks.
226 M. A. C. Martins-Ferreira and S. W. de O. Rodrigues

1.6 Classification Schemes

Alone, the terms rod, mullion and boudin have no genetic significance and do
not comprise geometric variations. However, further classification can be applied
to differentiate between genetic processes and geometric variations. This type of
classification has not yet been proposed for rods, but they exist for mullions and
boudins.
Wilson (1953, 1961, 1982) classifies mullions into three varieties that suggest
differences in the processes involved in their formation: (a) fold-mullions (cylindrical
structures forming undulations of folded bedding plane) or bedding mullions (same
as fold-mullions but not related to folding); (b) cleavage-mullions; and (c) irregular
mullions (most common variety, cylindrical structures very irregular in cross section).
The classification of boudins is more frequently related to their geometry in the
cross-sectional profile. Davis and Reynolds (1996) classify boudins as: (a) rectan-
gular or rhombic (separated by tension fractures on layers with very high ductility
contrast); (b) barrel-shaped (suffer some ductile necking prior to tension fracture,
forming boudins that resemble barrels); (c) fish-head (after separation by fracturing
the boudins deform by ductile flow, forming c-shaped walls between them, similar
to fish mouths); (d) lenticular (separated by ductile necking on low ductility contrast
forming pinch-and-swell structures).
Two classification schemes for rods are presented in this guide to be used in
the field: according to their geometry and to their origin, the previous structure or
composition from which they have been formed. Sometimes it is impossible to iden-
tify the prior structure or the parental material of rods, in which case the geometric
classification can be used alone. If the protolith, or material of origin (previous struc-
ture and/or composition), can be identified, a combination of the two classification
schemes is possible.
As per genesis, (a) vein rod or rodded vein (formed from tabular vein by the
rotation of vein segments cut by foliation planes during flexural slip and flexural
flow between layers of folds); (b) mega-rod (formed from sin-tectonic silicification
in hinge zones of mesoscopic scale); (c) bulk-rod (formed from bedding layers or
any portion of the bulk rock that was cut and rotated by foliation related to flexural
slip and flexural flow of folds, can be larger than mega-rods with diameters larger
than 0.3 m in cross section usually occurring in regional fold hinge zones); (c)
clast-rod or rodded clast (formed from conglomerate clasts embedded or not in a
more ductile matrix in folded layers, usually accompanied by a well-developed slip
foliation responsible for rotation and clast constriction) (Table 1).
As per geometry, (a) circular rod: cylinder-shaped rods with circular or nearly
circular cross sections, where relations of the strain ellipsoid axes are Z = Y < X; (b)
elliptical rod: cylinder-shaped rods with lenticular or elliptical cross sections, where
relations of the strain ellipsoid axes are Z < Y < X; (c) sigmoidal rods: rods with
sigmoid-shaped cross sections that have not been submitted to enough simple shear
rotation to form circular basal sections, where relations of the strain ellipsoid axes
are usually Z << Y < X; (d) tabular rod: clapboard or belt-shaped, usually formed
Field Guide to RODS in the Pireneus Syntaxis, Central Brazil 227

Table 1 Classification of rods according to the protolith (previous structure or composition),


proposed in this work
Name Protolith Mechanism of rod Related figures
formation
Vein rod or Rodded Planar veins Cutting and rotation of Figure 10b, d
vein vein segments cut by Figure 22b, c, d
foliation during
deformation via
flexural flow
Mega rod Hinge zone/Silica Siyn-tectonic Figure 17b, c, d, f, g
silicification in
mesoscopic hinge
zones
Bulk rod Thick quartzite layer Interlayer cutting and Figure 14a, d
(metric scale) rotation of thick Figure 29d
quartzite layers by
foliation near fold
hinge zones deformed
by flexural flow
Layer rod or rodded Rock layer of higher Intralayer cutting and Figure 22a, e, f
layer viscosity (sub-metric rotation by an S-C-like
scale) foliation during
deformation via
flexural flow
Clast rod or rodded Conglomerate clasts Rotation and clast Figure 10a, e, f, e
clast constriction during Figure 14a, b
folding by flexural flow Figure 15a, b, c, d, e

from originally tabular veins or from bulk rock and clasts that have been intensely
flattened after being rolled, relations of the strain ellipsoid axes are Z <<< Y < X
(Table 2 and Fig. 3).

1.7 Global Occurrences of Rods

Rods are described in the international literature in several places worldwide. An


exclusive selection of references is presented in Fig. 4.

2 Geological Settings of the Pireneus Syntaxis

The Pireneus Syntaxis is a pronounced concave-to-the-foreland curve in the other-


wise north–south trending structural grain of the Brasilia orogenic belt (AraújoFilho
1999). The region has other names such as Pireneus Lineament, Pireneus Megaflexure
228 M. A. C. Martins-Ferreira and S. W. de O. Rodrigues

Table 2 Rod classification according to their geometry in cross-section


Name Geometry Strain ellipsoid axes Related figures
Circular rod Cylinder-shaped rods with Z =Y<X Figure 9b, Figure 10g,
circular or nearly circular Figure 21b, Figure 27a
cross-sections
Elliptical rod Cylinder-shaped rods with Z <Y<X Figure 10a, Figure 11a,
lenticular or elliptical Figure 12a
cross-sections
Sigmoidal rods Rods with sigmoid-shaped Z << Y < X Figure 26a, b, c, d
cross-sections
Tabular rod Clapboard or belt-shaped, Z <<< Y < X Figure 10b, Figure 22a,
usually formed from b, d
tabular veins or from bulk
rock and clasts that have
been intensely flattened
after rolled

Fig. 3 Cross-sectional view


of rods and their geometrical
classification end-members
Field Guide to RODS in the Pireneus Syntaxis, Central Brazil 229

Fig. 4 Selected world localities of rod occurrence: (1) Ben Hutig, North Sutherland, Scotland
(Peach and Horne 1907; Wilson 1953; Holdsworth 1987, 1990); (2) Arnisdale, Scotland (Ramsay
1960); (3) Tortolita Mountains, Southern Arizona (Davis and Reynolds 1996); (4) Devonian Excel-
sior Phyllite, Central Andes, Peru (Davis and Reynolds 1996); (5) Berkshire Massif, Western
Massachusetts (Hatch 1975); (6) Southeast of Walvis Bay, Namibia (Sawyer 1981); (7) Hyde-
Macraes Shear Zone, Eastern Otago, New Zealand (Winsor 1991); (8) Ketchikan and Bradfield
Canal, Southeastern Alaska (Smith 1977); (9) Debari Formation, Rajasthan, India (Roy and Jakhar
2002). Black star: Pireneus Range, Goiás State, Brazil (this guide)

(Fonseca et al. 1995), Pireneus Corridor or the Pireneus Zone of High Strain (D’el-
Rey Silva et al. 2008, 2011). In this guide, the name Pireneus Syntaxis (PSX) is
adopted because it is non-genetic and only descriptive of the geometry of an orogenic
feature. The geographical location and geological settings of the PSX are thus here-
after presented. It is located in the state of Goiás, central Brazil, ~130 km from the
national capital, Brasília (Fig. 5a). The place of major occurrence of rods in the
PSX is the Pireneus Range, a mountain range in the Pireneus Range State Park,
located between the towns of Pirenópolis and Cocalzinho de Goiás (Fig. 5b). The
area is accessible from both cities via an unpaved roadway (BR-070) that passes
through the observation stations presented in this guide (Fig. 5c). Secondary road-
ways provide access to the observation stations, except station 7, which must be
accessed by foot trail in ~20-min walk from station 2 towards the south.

2.1 Tectonic Evolution

In the Neoproterozoic, during the amalgamation of West Gondwanaland, an orogen


named Tocantins Province formed as a result of the collision between the São Fran-
cisco craton in the east, the Amazonas craton in the north-west and the Paranapanema
block in the south-west, the last one hidden beneath the Paraná Basin (Fig. 6a). This
orogen is divided into three belts: the Brasília, the Araguaia and the Paraguay belts.
The Pireneus Syntaxis (PSX) divides the Brasília Belt in its north and south segments
(Fig. 6b). These segments present distinct geology and deformation trends, being the
230 M. A. C. Martins-Ferreira and S. W. de O. Rodrigues

Fig. 5 Maps: a the location covered by Fig. 3b in Central Brazil; b the regional access roads and
towns around the area covered by Fig. c; c the local roads and towns in the Pireneus Range zone

PSX the zone of interference of these two segments. Two main tectonic interpre-
tations are provided for the PSX. (1) Two distinct non-coeval orogens would have
interfered to produce the PSX (AraújoFilho and Marshak 1997; AraújoFilho 1999,
2000). (2) The PSX would be the axial zone of rotation of an orocline with compres-
sion in the inner part and extension in the outer part of the orocline bending hinge
(D’el-Rey Silva et al. 2011). It is not in the aims of this guide to discuss these inter-
pretations in depth, but it is quite obvious that they are conflicting and history shows
that in these cases the two interpretations might have assertive aspects that must be
considered when putting together a unified working hypothesis.
The geologic map of the PSX (Fig. 6c) shows segments of Paleo-, Meso- and
Neoproterozoic meta-volcano-sedimentary rocks (Rio do Peixe, Paranoá, Canastra
and Araxá groups) that are intruded by sin-tectonic granites (Itapuranga Suite) and
are cut by thrusts and strike-slip faults. These rocks are surrounded by a window
of Paleoproterozoic crystalline basement and the mafic granulite of the Barro Alto
Field Guide to RODS in the Pireneus Syntaxis, Central Brazil 231

Fig. 6 Regional geological maps showing: a the location of the Tocantins Structural Province in
South America in the context of cratons and basins of Brazil; b the location of the area in the
Pireneus Syntaxis, central Brasília Belt and; c the geological map of the PSX, outlining Fig. 7 area
(modified from LacerdaFilho et al. (1999) and AraújoFilho (2000))
232 M. A. C. Martins-Ferreira and S. W. de O. Rodrigues

complex to the NW and the mafic granulite of the Anápolis-Itauçú complex to the SW.
These granulites are interpreted as the orogenic metamorphic nucleus respectively
of the northern and southern Brasília Belt. Between these segments, the PSX occurs
as a corridor of high ductile strain. The P–T conditions for the region have not yet
been constrained although a few good geothermobarometers are present, e.g., garnet,
abundant in the schist, elongated quartz crystals and sin-kinematic kyanite, abundant
where rods are most prominent. Further detailed investigation is being conducted by
the authors regarding P–T conditions and tectonic mechanisms of rod formation in
the PSX.

2.2 Pireneus Range Geology

The geological map of the Pireneus Range (Fig. 7), where abundant rods occur,
shows meta-sedimentary rocks and meta-volcanic rocks of the Araxá and Rio do
Peixe groups, deformed ductilely, forming fold trains with ~E–W fold axes parallel
to the rods as well as elongated clasts and mineral lineation. Late strike-slip fault
zones cut these previous structures, including rods, and form transpressive corridors.
The main anticlinal fold axes have been preferentially eroded and now represent
valleys that deepen towards the west; the lithology plays a keyhole in controlling the

Fig. 7 Local lithostructural map of the area where rods are most prominent in the Pireneus range
Field Guide to RODS in the Pireneus Syntaxis, Central Brazil 233

Fig. 8 Satellite image of the Pireneus range, same area as Fig. 7. Source Microsoft Bing

local geomorphology, where quartzite and conglomerate hold the elevated areas and
the schists and meta-volcanic rocks are located in the lower relief areas (Figs. 7 and
8).
The area of prominent occurrence of rods is marked in light blue colour in the
map (Fig. 7). It is presumably associated with folds that are upright and tight to
isoclinal. Both the fold axes and rods are nearly horizontal and dip to both E and
W, except where they have been deformed by the late transpression. The rodded
structures occur predominantly in the kyanite-muscovite-quartzite facies and in the
meta-conglomerate facies as elongated clasts. Rods also occur occasionally in the
meta-volcanic/plutonic rocks and in the garnet-muscovite schist, with the same E–W
orientation.
In the Pireneus Range, two distinct deformation stages can be easily distinguished:
D1 and D2 . Stage D1 is responsible for the N–S shortening and generates nearly
E–W-oriented fold axes, rods and mineral lineation. Stage D2 is only identified in
restrict areas, where transpressive right-lateral shear zones occur oriented approxi-
mately in the NW–SE direction. Elsewhere in the PSX, the structural record allows
to distinguish different phases within these two stages, resulting in different stage
configurations (AraújoFilho 2000; D’el-Rey Silva et al. 2011).

3 Rods in the Pireneus Syntaxis

Different geneses of rods can be deciphered in the PSX. All these processes are,
however, associated with the primary process of folding. Here, isoclinal folding has
produced imposed flux parallel to the fold b-axis.
234 M. A. C. Martins-Ferreira and S. W. de O. Rodrigues

Rods occur in different lengths and thicknesses in both the hinge zones and the
limbs of regional and local folds. In hinges of folds formed by flexural flow, pure shear
prevails and the flux of matter along the b-axis (or the fold axis) is plausibly most
efficient. As zones of relatively low pressure, fold hinges form very long rods with
ellipsoidal shape in their cross section and constant shape and thickness along their
lengths. Rods developed on fold limbs display a greater diversity of morphology.
This happens because in the limbs different deformation mechanisms may occur
during buckling, such as flexural slip or flexural flow (Ramsay 1967; Davis 1984;
Twiss and Moores 1992; Hudleston et al. 1996; AraújoFilho 2000; Fossen 2010).
Fold limbs with heterogeneous beds with high rheological contrast between each
other tend to develop S–C-like fabrics in order to accommodate folding via flexural
flow mechanism (i.e. layers of meta-conglomerate and micaceous quartzite). Limbs
with homogeneous beds and low rheological contrast tend to accommodate folding
via flexural slip mechanism (Hudleston et al. 1996), and the less is mica content,
greater is the tendency of flexural slip instead of flexural flow (i.e. repeated layers
of pure quartzite). Also, close to fold hinges, rods appear less flattened than in away
from hinge zones, where rods are flattened or do not occur. In some limbs that
are distal from hinge zones, instead of rods, pancake-shaped quartz aggregates or
conglomerate clasts are formed (Fig. 9a). Rods are frequently striated parallel to their

Fig. 9 a Pancake-shaped conglomerate clasts formed in fold-limb zones that are distant from
hinges; b cylindrical rod with partition planes perpendicular to its length; c striated rod. Pictures
taken from inclined view. Hammer is ~40 cm long
Field Guide to RODS in the Pireneus Syntaxis, Central Brazil 235

lengths and present fracture planes perpendicular to their lengths, along which they
are commonly broken apart (Fig. 9b and c, respectively).

3.1 Parental Material and Rodding Mechanisms

A great variety of rod shapes and sizes occur in the PSX (Fig. 10). The material
or object involved in the process of rodding seems to be a key factor to define the
resulting rod morphology. In the PSX region, these original materials involved in
rodding may vary. Quartz veins, sedimentary bedding and meta-conglomerate clasts
can be involved in rodding. In some cases, it seems that silica from the country rock
filled pressure shadows created by the mylonite-like shearing produced by the S1 –S0
pair indicating movement similar to a S–C mylonite foliation (Fig. 22b, c, e and f).
Commonly, matrix-supported meta-conglomerates form the LS-tectonites
(Fig. 10a and b) and clast-supported meta-conglomerates the L-tectonites (Fig. 10e
and g). Rods formed from meta-conglomerate clasts that were elongated are easily
distinguished in the field, and they get narrow towards both ends (Fig. 10e and f).
The narrowing is usually easily noticed, unless the rods are very long and have been
broken (Fig. 10a). Planar quartz veins can be cut and rotated by S1 and become
rod-shaped (Fig. 10d) or form clapboard-shaped rods (Fig. 10b). Evidence that a
boudinaged or disrupted and rotated layer can evolve into rods can be seen on the
field, as demonstrated in Figs. 20b, 22a and 26e. A well-marked striation parallel to
the rod length is quite common and occurs in most rods formed from bedding layers;
these rods usually have basal diameters comparable to the thickness of the bed they
have been formed from (Fig. 10c). The cross-sectional shape of rods can vary from a
perfect circle (Fig. 10g), a flattened ellipse (Fig. 10a), tabular (Fig. 10b) or sigmoid
(see classification scheme in Table 2). Rods in the PSX sometimes exceed 5 m in
length.

3.2 Mineral Lineation and Rods

The rods observed in the PSX form a conspicuous coarse lineation, impossible to pass
unnoticed. But the matrix in which rods develop also shows a pronounced mineral
lineation parallel to the rods. When rods occur in clast-supported conglomerates
(Fig. 11a), this mineral lineation is either absent or faint. However, in the matrix-
supported meta-conglomerates, a strong mineral lineation forms in quartz from the
matrix (Fig. 11b, c, e). Also, in pure quartzite, quartz elongation produces a quite
impressive mineral-oriented fabric (Fig. 11d).
236 M. A. C. Martins-Ferreira and S. W. de O. Rodrigues

Fig. 10 Photographs showing the great diversity of rod geometries and protoliths. a Highly rodded
conglomerate pebbles not showing thinning towards the tips; b tabular rod, usually formed from
planar veins or layers; c highly striated rod formed from country rock bedding (bulk-rod); d rods
formed from planar veins; e rodded clasts showing thinning towards the tips in clast-supported
meta-conglomerate; and f in matrix-supported meta-conglomerate; g rodded conglomerate forming
a nucleus of circular rods surrounded by a cover of elliptical rods. All pictures are inclined views,
except f (vertical). Hammer is ~40 cm long
Field Guide to RODS in the Pireneus Syntaxis, Central Brazil 237

Fig. 11 a Cross section of clast rods in clast-supported meta-conglomerate forming elliptical to


sigmoidal rods (horizontal view); b weakly rodded and flattened conglomerate clast, immerse in
matrix of micaceous quartzite exhibiting prominent mineral lineation (horizontal view); c zoom
over mineral lineation marked by quartz needles (horizontal view); d quartz rods immerse in matrix
marked by penetrative foliation and mineral lineation (vertical view). Pencil and ruler ~15 cm long

3.3 Foliation and Rodding

In the PSX, observations suggest that rod formation can be a result of constriction
related to the rotation provided by the interaction of two main slip planes, the S1
tectonic foliation and the S0 bedding. The S0 imposes a great variation of flow
capacity between layers of different competences such as the clast-supported meta-
conglomerate and micaceous quartzite. The S1 foliation involves the rods, which
develop their long x-axis parallel to the strike of S1 (Fig. 12). The long x-axis coincides
with the axis of rotation. The amount of rotation seems to affect rod morphology
(i.e. more rotation results in well-formed and longer rods and little rotation will form
underdeveloped and shorter rods). Rod morphology also can be observed to vary
238 M. A. C. Martins-Ferreira and S. W. de O. Rodrigues

Fig. 12 Interpreted photographs showing the relationship between foliation and rods. a Elliptical
rods mimicking bamboo canes or flutes, surrounded by penetrative tectonic S1 foliation. b Inter-
preted picture showing the traces of S1 foliation and its relation to the axes of the strain ellipsoid
(inset). c Rodded conglomerate clasts surrounded by penetrative foliation S1 . d Zoom at Fig. c area
where the foliation is clearly seen to go around the clast and not through the clast. e Interpreted
picture showing the traces of S1 foliation and its relation to the axes of the strain ellipsoid (inset) in
a rodded conglomerate clast ellipsoid. Coloured lines represent the principal axis of the finite strain
ellipsoid: red = x; green = y; blue = z. All pictures taken from inclined view. Hammer is ~40 cm
long

greatly depending on their location in relation to the fold’s geometry (i.e. rods in
hinge zones are prone to have a circular to elliptical cross section).
Field Guide to RODS in the Pireneus Syntaxis, Central Brazil 239

3.4 Kyanite and Rodding

Kyanite is commonly associated with the rodded structures (Fig. 13a), especially in
the area coincident to the main regional fold axis, where generalized silicification
is also stronger. The kyanite crystals are elongated along rod surfaces with their
c-crystallographic axis usually aligned to the rods x-elongation axis (Fig. 13b), but
rotated crystals can also be seen (Fig. 13c and d) as well as non-oriented kyanite
clusters are also observed. In thin sections, beautiful kyanite fish can be seen.

Fig. 13 Kyanite crystals associated with rods; a in a rod surface; b and c in a mega-rod surface
and d in detail showing preferential orientation of kyanite crystals parallel to the rods’ long axis
and in some cases rotated. All pictures taken from inclined view. Pencil is ~15 cm long
240 M. A. C. Martins-Ferreira and S. W. de O. Rodrigues

Fig. 14 Mega-rods and bulk-rods formed from the country rock itself (a and e) or from silicifi-
cation in fold hinges (b, c and f). The cross section of mega-rods formed from volume affected
by silicification is most commonly cylindrical with elliptical cross sections and the mega-rods are
frequently surrounded by smaller rods with the y-axis oriented parallel to the mega-rod surface (b,
d and g). Rods formed from the country rock (bulk-rods) are not surrounded by smaller rods (a and
e). All pictures taken from horizontal view. Hammer is ~40 cm long

3.5 Mega-Rods and Bulk-Rods

Mega-rods and bulk-rods are terms suggested, in this guide, for metric scale wide
and long projecting ribs, usually found in hinge zones of regional folds. They are
not formed from objects such as veins or clasts. Detailed field observations on their
morphology and related structures suggest that bulk-rods are product of rotation
of large segments of rock layers that are rolled up at right angles to their lengths
into metric-sized rods (Fig. 14a and e). In some cases, they represent isolated fold
Field Guide to RODS in the Pireneus Syntaxis, Central Brazil 241

hinges (Fig. 17a and b). Mega-rods are most probably formed by silicification during
foliation slip in tight to isoclinal fold hinge zones (Fig. 14b, d, f and g). In the PSX,
bulk-rods vary from 1 to 3 m in diameter and have been observed to be as much as
7 m long. Mega-rods are a bit smaller and vary from 0.3 to 1.5 m in diameter and
have been observed as long as 5 m. The cross sections of silicified mega-rods are
usually elliptical. Interestingly, smaller-scale rods occur surrounding the mega-rods
with their y-axes oriented parallel to the mega-rod surfaces, suggesting that these
smaller rods are associated with mega-rod formation (Fig. 14b, d and g).
The cross-sectional dimensions of the bulk-rods are usually controlled by layer
thickness. For mega-rods, however, cross-sectional dimensions are controlled mainly
by the scale of the fold hinge involved in the rodding process.

4 Observation Stations

The 10 best locations to observe rods and the structures related to their development
were selected. For each station, a stereoplot (lower hemisphere) is provided with the
attitudes of the main planes and lines measured at that station as well as their mean
attitudes. The structures presented in the stereoplots are briefly described below:
S0 : The original sedimentary bedding plane. Bedding (S0 ) is folded and can be
mistaken by a tectonic foliation S1 but with the help of this guide it can be
recognized in most observation stations, especially in outcrops where layers of
meta-conglomerate are present.
S1 : A tectonic foliation plane S1 is related to the development of tight to isoclinal
folding by flexural flow in an early deformational phase D1 . It is more penetrative
in fold limbs and less frequent in hinge zones, except in hinges affecting finer rock,
where it occurs as an axial plane foliation. The S1 foliation is easily identified in
micaceous quartzite and is interpreted as responsible for the development of rods. S1
presents an S–C pair relationship with S0 , becoming nearly parallel to S0 when near
to it, resembling the foliation pair developed in mylonites, but being S0 coincident
with the bedding planes, as shown in the interpreted photos of station 9.
S2 : A late tectonic foliation plane generated by a late D2 phase of transpressive char-
acter dominated by pure shear rather than simple shear. This deformation phase folded
all previous structures such as rods, fold axes, S0 and S1 planes and is frequently seen
in the field as a crenulation cleavage. The S2 is a post-rod foliation and hence has
no genetic relationship to rods. The intersection of this foliation to the other planes
results in Li2 , an intersection lineation described below. In the Pireneus Range area,
S2 occurs restricted to a NW–SE corridor ~1 km wide; its boundaries are indicated
in the geological map by dextral shear faults (Fig. 7). At stations 5 and 6, the effects
of D2 are most prominent and S2 foliation usually dips 30–50° to the SW quadrant.
242 M. A. C. Martins-Ferreira and S. W. de O. Rodrigues

Li1 : The intersection lineation formed by S0 and S1 . This lineation is commonly


parallel to sub-parallel to the fold axes and also to the rods and can be easily measured
in limbs of folded micaceous quartzite, where S1 is more penetrative.
Li2 : The intersection lineation formed by the intersection of planes S0 /S1 and S2 .
The Li2 is a post-rod lineation and hence has no genetic nor geometric relationship
to rods.
Lrod : Rods form a coarse lineation, represented as Lrod in the stereoplots. Rods are
sub-horizontal and sub-parallel to fold hinges and to Li except where the late phase
D2 has changed their attitude. Their area of major rod occurrence is indicated in the
local geological map by the light blue colour.
Lclast : At some stations, it is clear that the coarse lineation formed by rodded structures
is formed by conglomerate clasts that have been rolled into rods, named Lclast . Rodded
clasts are wider at the centre and get thinner towards both ends. In this guide, Lclast
was measured separately from rods of different origins for means of comparison. The
results show no difference in orientation from the other types of rodded structures.
Lclast is parallel to Lrod and B1 .
Lm : The mineral lineation, very common structure in the area. It is marked by the
oriented and/or stretched minerals quartz, muscovite and kyanite. The Lm lineation
is also parallel to the other D1 structures such as Lrod , Lclast , Li and B1 .
B1 : The b tectonic axis of folds in outcrop scale. Rods and Li lineation are both
sub-parallel to B1 except where S2 occurs changing their attitude.

4.1 Station 1. Three Peaks Area (Três Picos), Near Little


Chapel (Capelinha)

Coordinates UTM 22L WGS-84: 0731550N/8252979E

Rods are not abundant in station 1; however, the local folding pattern is well exposed
here offering an opportunity to understand how folds have developed in the area,
an important issue, since rods are intimately related to the development of folds.
Besides, this is a place of astonishing landscape views, where hectometric folds and
Field Guide to RODS in the Pireneus Syntaxis, Central Brazil 243

their folding pattern and geometry can be inferred in a broader scale on the landscape.
Figure 15 shows south-verging cylindrical folds with nearly horizontal axis (Fig. 15a,
c) and their relationships to the geometry of greater folds (Fig. 15a). Figure 15a is
snapped from a northern limb of a regional fold. Figure 15b and c are snapped in the
hinge zone of the second fold from N to S in Fig. 15a (marked in red) on the way to
station 2.

Fig. 15 a Landscape view showing km-scale fold train (East view). Hinge zones at this scale have
been eroded and are now valleys (see satellite image on Fig. 4). b Outcrop-scale cylindrical fold
with horizontal axis and gentle vergence to the south (marked in red at the centre of Fig. a, West
view); c continuation to E of fold from Fig. b (outlined by dotted square in the upper left corner).
Hammer is ~40 cm long
244 M. A. C. Martins-Ferreira and S. W. de O. Rodrigues

4.2 Station 2. Park Rangers’ Office Area

Coordinates UTM 22L WGS-84: 0732320N/8252414E

Spectacular folds of textbook quality can be seen in this station. Folds are upright,
tight to nearly isoclinal and clearly present a thickened hinge and thinned limbs,
configuring Ramsay’s type-3 folds (Fig. 16a; Ramsay 1967). The mean fold axis

Fig. 16 a Upright tight fold, nearly isoclinal, with thickened hinge and thinned limbs (type-threefold
of Ramsay); b fold hinge of regional anticline showing “M” pattern of parasitic folds (hammer for
scale); c detail view of “M” pattern seen on fold hinges. All pictures are horizontal views. Hammer
is ~40 cm long. Pencil is ~15 cm long
Field Guide to RODS in the Pireneus Syntaxis, Central Brazil 245

dips 5° to the 102 azimuth direction. The “M” pattern of parasitic folds, indicative
of fold hinge zones, is well exposed here and can be observed at different scales
(Fig. 16b and c).

4.3 Station 3. East Antenna

Coordinates UTM 22L WGS-84: 0734516N/8250880E

The East Antenna station 3 cannot be missed! A great diversity of rods occurs here.
In a merely 200-m walk, one comes across rodded clasts, folds with rods in the
hinge zone, several mega-rods of all sizes and quartz pancakes. Kyanite crystals are
abundant and occur aligned with the striae that run along the trend of the rods. In a N–
S profile, one crosses a fold from the north-dipping limb to the south-dipping limb
some 300 m further. In the northern limb, still far from the hinge zone, pancake-
shaped conglomerate clasts occur (Fig. 9a) where deformation by flattening was
predominant and the S1 foliation is under-developed. Walking to the south, towards
the hinge zone, the bedding becomes vertical although the S1 tectonic foliation dips
to the north. This foliation may be confused with the bedding, but in places where
conglomerates occur, the bedding plane is obvious (Fig. 17c and d). Still in the
northern limb but closer to the hinge zone, first ellipsoidal, nearly tabular, rods occur
(Fig. 17e) and those gradate for oval-shaped to circular cylindrical rods near and in
the hinge zone. Mega-rods occur in the central portion or nucleus of the hinge zone,
where the bedding is vertical and kyanite is more abundant. The mega-rods here are
quite impressive even from the distance (Fig. 17a and b). They are oval-shaped in
their cross section and their intermediate y-axis is usually vertical or plunging to the
same direction of S1 foliation.
Local folds offer a good opportunity to observe the relationship between folding
and the different products of rodding. In Fig. 18a, one of these local folds is presented,
in which it is possible to see that rods within the inner fold hinge zone tend to be
symmetrical and mainly elliptical, besides larger in cross-sectional area (Fig. 18b,
outlined in red) and rods occurring in the limb and near the outer fold hinge zone
tend to assume an asymmetric sigmoid shape and are smaller in cross-sectional view
(Fig. 18b, outlined in green). The observer should always look at the structures
from different angles to avoid misinterpretation of geometric and kinematic orders.
246 M. A. C. Martins-Ferreira and S. W. de O. Rodrigues

Fig. 17 a Mega-rods in hinge zone of regional fold. Bedding here is vertical and kyanite is abundant
and oriented along the striae (hammer for scale); b detail from figure A showing oval-shaped, highly
striated, cylindrical mega-rod; c example of outcrop where the vertical bedding can be seen in the
contact between meta-conglomerate and micaceous quartzite; d interpretation of Fig. c showing
vertical contact (black line) from meta-conglomerate and micaceous quartzite and the S1 tectonic
foliation associated with rodding; e ellipsoidal, nearly tabular rods formed from conglomerate clasts.
All pictures are inclined views. Hammer is ~40 cm long
Field Guide to RODS in the Pireneus Syntaxis, Central Brazil 247

Fig. 18 a Inclined tight fold containing fold-related rods; fold axis B1 = 100/07; rods L rod =
107/05; b interpreted version of B showing the morphological differences between rods formed in
the inner (red) and outer (green) parts of a local fold. c and d Folded rods crop out and its two possible
geometric and kinematic interpretations depending on the angle of analysis: c in cross-sectional
view, where the dashed white line shows the folded bedding forming an “S” pattern of parasitic
folds, indicating here a short limb parasite fold train. d in plane-view, where the dashed white line
and arrows show how the inappropriate observation angle can lead to erroneous interpretation of a
top-to-the-west displacement. Horizontal views. Hammer is ~40 cm long

Figure 18c and d show a beautiful outcrop with folded rods that can have two possible
geometric and kinematic interpretations (one of them erroneous) depending on the
angle of analysis. In Fig. 18c, favouring a cross-sectional view, the dashed white line
shows the folded bedding forming an “S” pattern of parasitic folds, indicating that
this outcrop represents the short limb of a larger south-verging fold, and in Fig. 18d,
favouring a plane-view of the bedding, the dashed white line and arrows show how
the inappropriate observation angle can lead to erroneous interpretation of a top-to-
the-west displacement, movement not observed anywhere in the region. See Dutta
and Mukherjee (2019) review article on cautions required in interpreting other shear
senses.
248 M. A. C. Martins-Ferreira and S. W. de O. Rodrigues

4.4 Station 4. South of West Antenna Road

Coordinates UTM 22L WGS-84: 0733011N/8251489E

Station 4 is a special place with plenty of wonderful structures, it is also very easy to
access and find the structures presented in this field guide, which are around 10–30 m
from the road, near the road truncation that leads to the West Antenna.
One observes the vertical bedding of pure quartzite layers resisting to the formation
of foliation S1 , whereas the micaceous quartzite (<1 m to the right) is fully allowing
the development of S1 and thus the formation of rods by disrupting and rotating
thin layers of pure quartzite and quartz veins (Fig. 19a and c). Also, as in station 3,
the local folds show, in a very instructive way, how folding leads to the rodding of
competent layers by disrupting and rolling them along a foliation plane resulting in
a variety of rod morphologies depending on the location, in the fold, that the rod is
developed (Fig. 19b and d).
Walking further to the south to ~20 m, a great variety of rods can be observed;
those formed from competent beds and quartz veins like the ones shown in Fig. 19
are sometimes side by side with those formed from conglomerate clasts. However,
the untrained eye might have a hard time identifying them. A few tips might be useful
for that purpose, as presented below.
The fact of rods being aligned parallel to the bedding (Fig. 20a) is not a good
criterion for distinction because quartz veins here are also mostly parallel to the
bedding in the region, rods formed from competent layers are obviously also parallel
to the bedding, and conglomerate layers were also deposited respecting the bedding
plane.
However, conglomerate clasts are usually not well sorted in the region, so rods
formed from clasts usually vary considerably in size, both in width and in length
(Fig. 20d). Additionally, rods formed from clasts usually thin towards their ends. If
rods align parallel to bedding, do not get narrower to both ends and present nearly
constant x- and z-axes lengths but vary in their y-axis length, they are most probably
formed from competent layers (Fig. 20b, c). Quartz rods formed from tabular veins
also occur aligned and frequently are parallel to bedding, but they can be easily
distinguished. The most obvious criterion is the composition, observable with a 10×
hand lens. Rods of pure quartz with no signs of meta-sedimentary quartz grains or
muscovite means one is looking at rods formed from veins. Also, rods from veins
Field Guide to RODS in the Pireneus Syntaxis, Central Brazil 249

Fig. 19 a Vertical bedding of pure quartzite layers (left and centre) and micaceous quartzite to the
right where the more competent layers (pure quartzite) are being rodded while the less competent
(micaceous quartzite) the S1 foliation; b folds on hinge zone of local asymmetric fold (see inset
on Fig. d) forming “M” pattern with quartzite layers in the process being split and rotated in the
process of rodding; c geometrical and kinematic interpretation of Fig. a showing evidence that the
outcrop is located in the short limb of an asymmetrical fold being the upper part of the photograph the
partially eroded fold hinge zone (see inset), note how smaller rods abound in the micaceous quartzite
and bigger rods (bulk-rods) form from thick quartzite layers in the hinge zone; d geometrical and
kinematic interpretation of Fig. b showing the relationship of fold geometry, S1 foliation and rod
morphology. The S0 bedding planes are marked by black lines, the S1 foliation by dashed lines and
rodded quartz is outlined in white. Horizontal views. Hammer is ~40 cm long
250 M. A. C. Martins-Ferreira and S. W. de O. Rodrigues

Fig. 20 a Outcrop of micaceous quartzite with sparse clast lines (horizontal view); b detail area
of Fig. a and c its geometric and kinematic interpretation showing the vertical bedding marked by
a line of rods and their relationship to the S1 foliation (dashed lines) that seem to be parallel to
the bedding planes at both ends, like an S–C mylonitic foliation pair, connoting slip and clockwise
rotation of clasts; d striated rodded conglomerate clasts (inclined view). Hammer is ~40 cm long.
Pencil is ~15 cm long

tend to be sigmoidal or tabular in their cross sections and are usually not wider than
~20 cm in their y-axis of strain.
Field Guide to RODS in the Pireneus Syntaxis, Central Brazil 251

4.5 Station 5. West Antenna

Coordinates UTM 22L WGS-84: 0733179N/8251901E

Walking ~50 m to the north from the cell phone antenna, excellent outcrops can be
seen. Here, the bedding is nearly vertical and its relationship to the S1 foliation can
be observed generating rotation and the consequent formation of rods even in pure
quartzite (Fig. 21a and b). In the micaceous quartzite, rods appear as isolated quartz,
which can be rounded (Fig. 21c) or flattened (Fig. 21d). In several places. rods and
the Lm mineral lineation are folded by a late S2 foliation (Fig. 18e and f), which also
crenulates the S0 bedding, forming a pronounced L2 lineation at the intersection of
bedding and the crenulation planes (Fig. 21g).
Further towards west, mega-rods occur (Fig. 14a and e) and meta-conglomerate
clasts are also involved in the process of rodding.

4.6 Station 6. Cora Pathway

Coordinates UTM 22L WGS-84: 0732663N/8252008E

This is one of the most complete stations in terms of rod variety. Here, it is possible
to see quartz veins, meta-conglomerate clasts as well as quartzite layers, in different
degrees of the process of rodding (Fig. 22). Also, this is a great place to observe rods
deformed by a late S2 foliation. Quartzite layers that have been cut and rolled by S1
form a massive set of malformed tabular rods (Fig. 22a). Quartz veins, usually parallel
252 M. A. C. Martins-Ferreira and S. W. de O. Rodrigues

Fig. 21 a Quartzite outcrop shows vertical bedding and north-dipping S1 foliation and associated
rods formed by the precipitation of silica; b interpreted version of Fig. a outlining the trace of
bedding in black line, the S1 foliation in dashed lines and the cross sections of the rods filling the
spaces opened by the S1 –S0 pair; c detail of cylindrical rods within micaceous quartzite; d detail of
folded rods; e and f photograph and interpretation, respectively, of folded and flattened rod showing
also the folded Lm mineral lineation and the trace of S2 foliation coincident with the axis of the
folds on the rods. Horizontal views. Hammer is ~40 cm long. Pencil is ~15 cm long

to S0, have also been cut and rolled by S1 and can be seen here as aligned sigmoid
rods (Fig. 22b, c, d, e and f). These sigmoidal rods indicate similar shear sense as
the S–C-like foliation pair formed by S0 and S1 . Folded rods are evident (Fig. 22g).
At this station, some rods are twisted obliquely to their lengths, indicating that two
phases of movement D1 and D2 affected the rods with different axes of rotation
(Fig. 22h).
Field Guide to RODS in the Pireneus Syntaxis, Central Brazil 253

Fig. 22 a Rodded quartzite layer forming massive set of rods; b quartz vein in the process of being
rodded; c geometric and kinematic interpretation of Fig. b showing the S1 foliation and the sense of
movement; d rods formed from tabular veins parallel to bedding; e rods formed from quartz veins in
quartzite; f geometric and kinematic interpretation of Fig. g showing the S1 foliation and the sense
of movement; g folded rods (vertical view); h twisted rod. All horizontal views except g. Hammer
is ~40 cm long. Pencil is ~15 cm long
254 M. A. C. Martins-Ferreira and S. W. de O. Rodrigues

4.7 Station 7. Cabeludo Hill

Coordinates UTM 22L WGS-84: 0731819N/8251023E

The Cabeludo Hill is an impressive geomorphologic feature in the area (Fig. 23). It
is formed by E–W-oriented vertical quartzite and meta-conglomerate layers and is
stroke by a N–S-oriented spaced cleavage. The result from the intersection of these
two planar structures is the high quartzite columns easily noted from the distance.
At the foothills of the Cabeludo, isoclinal folds up to 3 m high can be seen with
plenty of rodded structures, as shown in the photographs of Fig. 24. The bedding
plane is well marked by conglomerate clasts (Fig. 24b), and its relationship to the S1
foliation indicates the outcrop location in the folds related to their development (figure
insets). Intra-bedding quartz veins are cut by the S1 foliation and some segments have
been rodded, forming sigmoidal rods (Fig. 24c and d).

Fig. 23 Cabeludo Hill in the distance. The vertical structures looking like columnar joints are
merely the result of the intersection of vertical bedding planes and a perpendicular vertical spaced
cleavage related to the folding of these quartzites
Field Guide to RODS in the Pireneus Syntaxis, Central Brazil 255

Fig. 24 a Vertical limb of upright isoclinal local parasitic fold b = 098/05; b northern limb of
local upright synformal fold where meta-conglomerate layers are interbedded with quartzite layers.
Note strongly rodded meta-conglomerate clasts. The bedding plane is nearly vertical, but dips to the
south, whereas the S1 foliation dips to the north; c detail of intra-bedding quartz vein from the lower
centre area of Fig. a; d interpretation of Fig. c evidencing the S1 foliation affecting the intra-bedding
quartz veins, in an initial process of rodding, producing malformed rods with sigmoid shape and
the x-axis parallel to the local and regional fold axis. Horizontal views. Hammer is ~40 cm long

4.8 Station 8. Road to Pirenópolis

Coordinates UTM 22L WGS-84: 0729262N/8251717E


256 M. A. C. Martins-Ferreira and S. W. de O. Rodrigues

This is one of the few places where nearly horizontal bedding can be seen with the
intercalation of beds composed of pure quartzite, meta-conglomerate and micaceous
quartzite (Fig. 25a). In this station, the S1 tectonic foliation exerts a determinant
control over the morphology and orientation of rods, resulting in a majority of sigmoid
rods. Depending on the localization of the rod in relation to the S0 –S1 pair (S–C-
like), the y-axis can vary from sub-horizontal (Fig. 25b, lower left) to sub-vertical
(Fig. 25b, upper centre). The geometric and the kinematic analyses of these outcrops
position them in the long limb of a south-verging anticline (Fig. 25c).
In this station, the process of rodding of bedding layers is quite obvious (Fig. 26a).
A layer with under-developed rods shows the S1 foliation cutting the layer and
promoting rotation of the resulting fragments (Fig. 26b). Impressive sigmoid flutes
are seen (Fig. 26c and d).

Fig. 25 a Rods in nearly horizontal bedding outcrop, in the long limb of a regional asymmetric
fold. The y-axis of these rods varies from nearly horizontal (lower left) to nearly vertical (upper
centre), the upper bed is composed of pure quartzite and meta-conglomerate, whereas the lower bed
is composed of micaceous quartzite; b quartzite layer in the process of becoming a rod. c Geometric
and kinematic interpretation of figure B, where the trace of the sigmoid S1 foliation is marked by
dashed lines, the nearly horizontal bedding plane is marked by the solid line and the shear sense
top-to-the-south is indicated by black half arrows. Horizontal views. Hammer is ~40 cm long. Pencil
is ~15 cm long
Field Guide to RODS in the Pireneus Syntaxis, Central Brazil 257

Fig. 26 a Basal section of rods showing quartzite layers formed by the development of S1 foliation;
b geometric and kinematic interpretation of Fig. a showing S1 foliation marked by dashed line and
the simple shear sense; c rods cropping out in 3D view; d geometric and kinematic interpretation
of Fig. c showing S1 foliation marked by dashed line and the direction of simple shear. Horizontal
views. Pencil is ~15 cm long

4.9 Station 9. Road to Sonrisal Waterfalls

Coordinates UTM 22L WGS-84: 0730505N/8251223E

In this station, thick layers of clast-supported conglomerate as well matrix-supported


conglomerate have been deformed so that all clasts have been rolled into rodded bars,
forming impressive outcrops that resemble stacks of cut wood piled for the winter
fires (Figs. 1 and 27). Some of these rodded clasts are up to 5 m long and 20 cm wide.
These structures are easily distinguished from the rods formed by quartz segregation
since all of them are wider in the middle and get thinner towards both ends (Fig. 27d,
e).
258 M. A. C. Martins-Ferreira and S. W. de O. Rodrigues

The short limbs of slightly asymmetric south-verging folds can be seen and
the bedding can be easily distinguished by the metre-scaled interlayering of meta-
conglomerate and micaceous quartzite (shown in detail in Fig. 17c). Foliation S1 is
also easily identified, as shown in the interpreted Fig. 28b and d. The S1 foliation
refracts when crossing between layers of different mechanical behaviour (see Bose
et al. 2020). Also, in the conglomerate, the S1 foliation is less penetrative. A careful
analysis of the outcrops as in Fig. 28 is the key to understand the relationship between
the mechanisms of folding and rodding.

4.10 Station 10. Mocó Boulders

Coordinates UTM 22L WGS-84: 0729990N/8250836E

This station is located in the nucleus of the rod-occurring area, aligned with the
Cabeludo Hill station. Here, silicification is the highest, propitiating one of the best
crags for boulder rock climbing in the region. Higher silicification is accompanied
by the higher P/T conditions, as evidenced by abundant kyanite. At some places, the
quartzite itself is bluish due to numerous kyanite crystals (Fig. 29a and b). In other
places, kyanite grains are as large as 5 cm, forming clusters or lying oriented (Lm )
parallel to and between rods and fold axes (Fig. 29c). In this station, bulk-rods occur
in major fold hinges (Fig. 29d).
A visit to the Sonrisal waterfalls is highly recommended after visiting station 10
by getting the walking trail to the south-west direction for ~10–15 min. There, the
vertical layering is evident and a narrow rod-like quartzite bridge can be crossed
over the clear turquoise-coloured water. From the Sonrisal stream, a few hundreds of
metres further to the west, finger-sized rods can be found in quartz-muscovite-garnet
schist.
Field Guide to RODS in the Pireneus Syntaxis, Central Brazil 259

Fig. 27 Conglomerate clasts rodded in station 9, ~50 m to the west of the fold on the next
figure: a rodded clasts of clast-supported meta-conglomerate exhibiting circular to sigmoid cross
sections seen on a fracture plane; b vertical contact of original sedimentary bedding between clast-
supported meta-conglomerate and micaceous quartzite; c massive layer of rodded clast-supported
meta-conglomerate, some of these rods are nearly 4 m long and 25 cm wide at the centre; d matrix-
supported meta-conglomerate showing clasts with different degrees of rodding; e detail of rodded
conglomerate clasts thinning at both ends. Horizontal views except for b and d (inclined). Hammer
is ~40 cm long
260 M. A. C. Martins-Ferreira and S. W. de O. Rodrigues

Fig. 28 a Outcrop of micaceous quartzite in vertical contact with meta-conglomerate in the short
limb of a flexural flow fold; b interpreted version of figure A showing the bedding planes (solid lines),
the S1 foliation (dashed lines), the kinematic analysis (arrows) and the geometrical location of the
outcrop in the local fold (inset); c fold affecting interlayered clast-supported meta-conglomerate and
micaceous quartzite. Note how well-developed rods occur in the meta-conglomerate; d interpreted
version of Fig. c showing the bedding planes (solid lines), the S1 foliation (dashed lines), the
kinematic analysis (arrows) and the geometrical location of the outcrop in the local fold (inset).
Horizontal views. Hammer is ~40 cm long

5 Conclusions

Rods are extraordinary features that deserve our attention and study. Despite being a
relatively rare structure on Earth, rods are a very common structure in the Pireneus
Range. This field guide presents to the reader a great diversity of rods, with distinct
geometries, scales and origins, all in the same place, the high strain corridor of the
Pireneus Syntaxis. This is, therefore, a great natural laboratory to study rods. The
geologist leaves this place with the impression that nearly every object (clasts, veins,
layers, etc.) has been caught in the process of rodding. A direct link of this process
to the development of fold-related foliations can be established. The variations in
constriction and flattening deformation are highly dependent on the locus relative
the fold geometry. Near hinge zones constriction prevails; at the limbs, flattening is
predominant.
Field Guide to RODS in the Pireneus Syntaxis, Central Brazil 261

Fig. 29 a Detail of Fig. b showing bluish kyanite-quartzite matrix amid the cross sections of
yellowish rods; b highly silicified kyanite-quartzite boulder with sparse agglomerates of clasts.
White spots are MgCO3 powder, used by local rock climbers; c longitudinal cross section showing
kyanite crystals formed between rods and not inside them; d bulk-rod formed near hinge zones of
major folds. Horizontal views. Hammer is ~40 cm long. Pencil is ~15 cm long

The authors invite geologists and rock sympathizers from all over the world
to witness these outstanding structures, which are accompanied by nice landscape
views, crystal clear waterfalls and a lot of fun in the field. Those prone to collection
will certainly relish on swords and cigars (Fig. 30) and the inquisitive minded will
certainly leave with more questions than answers.
262 M. A. C. Martins-Ferreira and S. W. de O. Rodrigues

Fig. 30 Geologists Marco A. C. Martins-Ferreira (left photo, savouring a rod cigar) and Sérgio W.
O. Rodrigues (right photo, wielding a medieval rod sword) are currently professors at the Federal
University of Goiás, Brazil, and are passionate for Structural Geology and Tectonics

Acknowledgements We acknowledge the Pireneus State Park rangers on their receptivity. The
field campaigns were funded by the authors. We are thankful to Lívia Helena Carrera Silveira,
who has taken care of the children while the authors went hunting rods on the weekends. This
chapter is a result of a wider research project being developed in the area by the Tectonic Analysis
of Terranes and Basins research group, hosted by the Federal University of Goiás (UFG) and the
Brazilian National Research Council (CNPq). Further details on the research group at: https://dgp.
cnpq.br/dgp/espelhogrupo/9122454495366464. This chapter was edited and reviewed by Soumyajit
Mukherjee (IIT Bombay). Thanks to Marion Schneider, Annett Buettener, Boopalan Renu, Alexis
Vizcaino, Doerthe Mennecke-Buehler and the proofreading team (Springer). Dutta and Mukherjee
(2021) summarize this article.

References

Araújo Filho, J. O. (2000). The Pireneus Syntaxis: An example of the intersection of two Brasiliano
fold-thrust belts in Central Brazil and its implications for the tectonic evolution of Western
Gondwana. Revista Brasileira De Geociências, 30(1), 144–148.
Bates, R. L., & Jackson, J. A. (1987). Glossary of geology (p. 788). Alexandria, Virginia: American
Geological Institute.
Bose, N., Dutta, D., & Mukherjee, S. (2020). Refraction of micro-fractures due to shear-induced
mechanical stratigraphy in a low-grade meta-sedimentary rock. Journal of Structural Geology,
133, 103995.
Burg, J. P. (2019). Script to structural geology. In Lectures 651-3422-00L and 651-3422-00V. ETH
Zurich.
Davis, G. H. (1984). Structural geology of rocks and regions (p. 492). New York: Wiley.
Davis, G. H., & Reynolds, S. J. (1996). Structural geology of rocks and regions (p. 776). New
Jersey: Wiley.
Araújo Filho, J. O. (1999). Structural characteristics and tectonic evolution of the Pireneus Syntaxis,
Central Brazil. Ph.D. Thesis (p. 433). Urbana-Champaign, USA: University of Illinois.
Field Guide to RODS in the Pireneus Syntaxis, Central Brazil 263

Araújo Filho, J. O. & Marshak, S. (1997). Formation of the Pireneus Syntaxis: Evidence for two
episodes of Brasiliano (Pan African) deformation in the Brasília Orogenic Belt, Central Brazil.
GSA abstract with programs A-228, BTH 44. Salt Lake City.
D’el-Rey Silva, L. J., de Oliveira, Í. L., Pohren, C. B., Tanizaki, M. L. N., Carneiro, R. C., Fernandes,
G. L. D. F., & Aragão, P. E. (2011). Coeval perpendicular shortenings in the Brasília belt: Collision
of irregular plate margins leading to oroclinal bending in the Neoproterozoic of central Brazil.
Journal of South American Earth Sciences, 32(1), 1–13.
D’el-Rey Silva, L. J., De Vasconcelos, M. A., & Silva, D. V. (2008). Timing and role of the Maranhão
River Thrust in the evolution of the Neoproterozoic Brasília Belt and Tocantins Province, central
Brazil. Gondwana Research, 13(3), 352–374.
Dutta, D., & Mukherjee, S. (2019). Opposite shear senses: Geneses, global occurrences, numerical
simulations and a case study from the Indian Western Himalaya. Journal of Structural Geology,
126, 357–392.
Dutta, D., & Mukherjee S. (2021). Introduction to Structural Geology and Tectonics Field Guide-
book - Volume 1. In: S. Mukherjee (Ed.) Structural Geology and Tectonics Field Guidebook -
Volume 1. Springer Nature Switzerland AG. Cham. pp. xi-xvi. ISBN: 978-3-030-60142-3.
Fonseca, M. A., Dardenne, M. A., & Uhlein, A. A. (1995). Faixa Brasília, Setor Setentrional: Estilos
Estruturais e Arcabouço Tectônico. Revista Brasileira De Geociências, 25(4), 267–278.
Fossen, H. (2010). Structural geology. Cambridge: Cambridge University Press. https://doi.org/10.
1017/CBO9780511777806.
Hatch, N. L. (1975). Tectonic metamorphic, and intrusive history of part of the east side of the Berk-
shire Massif, Massachusetts. In Tectonic studies of the Berkshire Massif, Western Massachusetts,
Connecticut and Vermont. United States Geological Survey Professional Paper, 51.
Holdsworth, R. E. (1987). Basement/cover relationships, reworking and Caledonian ductile thrust
tectonics of the Northern Moine, NW Scotland. Unpublished Ph.D. thesis, University of Leeds.
Holdsworth, R. E. (1990). Progressive deformation structures associated with ductile thrusts in the
Moine Nappe, Sutherland N. Scotland. Journal of Structural Geology, 12(4), 443–452.
Hudleston, P. J., Treagus, S. H., & Lan, L. (1996). Flexural flow folding: Does it occur in nature?
Geology, 24(3), 203–206.
LacerdaFilho J. V., Rezende, A., & Silva, A. (1999). Programa de levantamentos geológicos básicos
do Brasil—Geologia e recursosminerais do estado de Goiás e Distrito Federal (p. 184). Goiânia:
CPRM, METAGO S.A., UnB, 2 ed.
McClay, K. R. (1987). The mapping of geological structures (p. 164). Wiley.
Peach, B. N., & Horne, J. (1907). The geological structure of the North-West highlands of Scotland.
Scotland: Memoirs of Geological Survey.
Ramsay, J. G. (1960). The deformation of early linear structures in areas of repeated folding. The
Journal of Geology, 68(1), 75–93.
Ramsay, J. G. (1967). Folding and fracturing of rocks (p. 568). New York: McGraw-Hill.
Roy, A. B., & Jakhar, S. R. (2002). Geology of Rajasthan: Precambrian to recent (p. 421). Jodhpur,
India: Scientific Publishers.
Sawyer, E. W. (1981). Damaran structural and metamorphic geology of an area southeast of Walvis
Bay, South West Africa/Namibia. In: South West Africa (p. 94), Department of Economic Affairs,
Geological Survey, 1981–Science.
Smith, J. G. (1977). Geology of the Ketchikan D-1 and Bradfield Canal A-1 quadrangles, south
eastern Alaska (Vol. 1425). Department of the Interior, Geological Survey.
Twiss, R. J., & Moores, E. M. (1992). Structural geology (p. 532). New York: W. H. Freeman.
Wilson, G. (1953). Mullion and rodding structures in the Moine Series of Scotland. Proceedings of
the Geologists’ Association, 64(2), 118–151.
Wilson, G. (1961). The tectonic significance of small-scale structures, and their importance to the
geologist in the field. Annales de la Société Géologique de Belgique, 84, 510–517.
Wilson, G. (1982). Mullion and rodding structures. Introduction to small-scale geological structures
(pp. 86–93). Dordrecht: Springer.
264 M. A. C. Martins-Ferreira and S. W. de O. Rodrigues

Winsor, C. N. (1991). The relationship between the Hyde Macraes Shear Zone, deformation
episodes, and gold mineralisation potential in eastern Otago, New Zealand. New Zealand Journal
of Geology and Geophysics, 34(2), 237–245.
Low Baric Metamorphic Belts
in the Northern Tip
of the Arabian–Nubian Shield: Selected
Examples from the Eastern
Desert/Midyan Terranes, Egypt

N. A. Shallaly and A. S. A. A. Abu Sharib

Abstract Low-pressure metamorphic assemblages are very common in many


orogenic belts. Characteristically, they are related to the thermal metamorphism
caused by post-orogenic rift-related plutonism. However, it has proven correct that
magmatic activity at any stage of the orogen can provide the required temperature
to initiate and drive the metamorphism. The Arabian–Nubian Shield (ANS), the
northern part of the Neoproterozoic East African Orogen (EAO), is characterized by
widespread exposures of the pelites-dominated metasedimentary terranes originated
in different tectonic settings. Low-pressure metamorphic belts constitute important
components, controversial in places, of some of these terranes. The current chapter
documents the field characterization of four multiply deformed and metamorphosed
low-pressure/medium- to high-temperature metamorphic belts located in the Eastern
Desert and Midyan terranes in the extreme north-eastern tip of the ANS.

1 Introduction

Low-pressure–high-temperature (LP–HT) metamorphism or Abukuma gradient


defines metamorphic terranes having high heat flow (25–50 °C km−1 ) and formed
in different tectonic settings (Kornprobst 2003; Kearey et al. 2009). Extension-
related LP–HT metamorphism is the most common type that is well documented
in modern and recent rifts (e.g. Wickham and Oxburgh 1987; Sandiford and Powell,
1986). Examples include the Late Paleozoic Oslo Rift (Jamtveit et al. 1992a, b,
1997; Jamtveit and Andersen 1993; Svensen and Jamtveit 1998), the Rhine graben,
and the Cenozoic East African Rift. On the other hand, accretionary-related LP–
HT terranes have been recorded in many orogenic belts that are characterized by
crustal thickening and widespread shallow magmatic intrusions. Examples include

N. A. Shallaly (B)
Geology Department, Faculty of Science, Cairo University, Giza 12613, Egypt
e-mail: shallalynahla@yahoo.de
A. S. A. A. Abu Sharib
Geology Department, Faculty of Science, Beni-Suef University, Beni-Suef 62521, Egypt

© Springer Nature Switzerland AG 2021 265


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_9
266 N. A. Shallaly and A. S. A. A. Abu Sharib

the Buchan belt of the Dalradian Supergroup (Cambrian), the Central Western Belt
of western Main, USA (Devonian), and the Ryoke-Abukuma belt in Japan (Creta-
ceous). In all settings, the heat expelled from the shallow extension-related or slightly
deeper convergent-related magmatic intrusions is the main source for temperature
required for the metamorphism. The voluminous magmatic intrusions together with
the extensive correlation between the pluton boundaries and metamorphic isograd led
some authors to give the term regional thermal metamorphism to describe these LP–
HT metamorphisms (e.g. Solar and Brown 1999; Holdaway 1982). Porphyroblastic
andalusite and/or cordierite are the main metamorphic phases to grow in pelitic and
calc–pelitic rocks metamorphosed at LP–HT metamorphic conditions (Bucher and
Frey 1994). Porphyroblasts are attributed to intrusion-related post-tectonic growth
(e.g. De Yoreo et al. 1989). However, deformation-related syn-tectonic growths have
been recorded in some orogenic belts, e.g. the Acadian orogeny (Brown and Solar
1998a, b, 1999; Solar and Brown 2001; Sanislav 2010; Bell and Sanislav 2011;
Sanislav and Bell 2011).
The Late Cryogenian–Ediacaran East African Orogen (EAO) is a N–S-trending
collisional belt (Fig. 1) comprising metamorphic terranes exhibiting high geothermal
gradients. It is differentiated into the amphibolite-to-granulite facies-dominated
Mozambique belt, to the south, and the greenschist facies-dominated juvenile
Arabian–Nubian Shield (ANS), to the north (Fig. 1) (Stern 1994; Meert 2001
and references therein). The ANS is composed of a number of ophiolite-decorated
accreted arc terranes (Johnson et al. 2003, 2011), the northern most of which, the
Midyan-Eastern Desert, exposes the entire Precambrian basement rocks of Egypt
(Fig. 2). The basement rocks of Egypt comprise three main divisions (Fig. 3) (Ries
et al. 1983; Sturchio et al. 1983a, b; El Gaby et al. 1984; El-Ramly et al. 1984;
Greiling et al. 1984, 1988; Habib et al. 1985; El Gaby et al. 1988; Greiling et al.
1988; Kamal El-Din et al. 1992; Rice et al. 1993; Hamimi et al. 1994; Kröner et al.
1994; Abdeen et al. 1996; Bregar 1996; Fritz et al. 1996; Kamal El-Din et al. 1996;
Messner 1996; Naim et al. 1996; Neumayr et al. 1996a, b; Blasband et al. 2000;
Fowler and Osman 2001): (1) a number of structurally lower tectonic windows of
amphibolite facies gneisses, and migmatites (Hafafit, Meatiq, El-Shalul, El-Sibai, and
Feiran–Solaf belt), (2) a structurally upper greenschist facies-dominated eugeoclinal
succession of dominantly ophiolitic mélange–island arc, and (3) post-amalgamation
molasse-type sediments (Hammamat sediments). All units are intruded by a number
of syn- to late and post-tectonic intrusions (Fig. 3) (Andersen et al. 2009; Sturchio
et al. 1983a, b; Stern and Hedge 1985; Loizenbauer et al. 2001).
The Precambrian terranes exposed in Sinai and the Eastern Desert are separated
by the Oligo–Miocene Gulf of Suez Rift (Fig. 3) and bear common features as well as
some differences. Both are metamorphosed at greenschist to amphibolite facies meta-
morphic conditions, whereas they differ dramatically in the significant proportion of
post-collision granitoids (mainly alkaline pink granites) and the widespread expo-
sures of ophiolitic and/or ophiolitic mélange rocks. More than 30% of the exposed
Precambrian rocks in Sinai belong to the post-orogenic granites. On the other hand,
ophiolitic rocks are restricted to the Eastern Desert, specifically the Central and
Southern Eastern Desert, CED and SED, respectively. Low baric metamorphic belts
Low Baric Metamorphic Belts in the Northern Tip … 267

Fig. 1 A general map shows the distribution of the Neoproterozoic belts in Africa. The East
African Orogen (EAO) is differentiated into the Arabian–Nubian Shield (ANS) to the north and the
Mozambique belt (MB) to the south

are exposed in a few occurrences in Sinai Peninsula and the Eastern Desert. This
research aims to shed the light on the tectono-metamorphic evolution of the LP–
HT metamorphic assemblages/rocks in four occurrences: Wadi Kid, W. Samra, and
W. Um Had and W. El Miyah in Sinai Peninsula and the Central Eastern Desert,
respectively (Fig. 3).

2 Geologic Setting

2.1 Sinai Belts

The Kid Group is the youngest of three Proterozoic arc complexes (Sa’al ~ 1.1 Ga;
Elat ~ 870 Ma; and Kid ~ 640 Ma: Jarrar et al. 1983; Eyal et al. 2014) cropping out in
the Sinai Peninsula (Fig. 4a) and occupies the extreme north-eastern tip of the ANS.
268 N. A. Shallaly and A. S. A. A. Abu Sharib

Fig. 2 A general map shows the different tectonic terranes constituting the Arabian–Nubian Shield.
The Eastern Desert (ED) Terranes, the Nubian equivalent to the Midyan terrane in the Arabian Shield,
are divided into three sub-terranes that are arranged, geographically from north to south, into: the
Northern, Central, and Southern Eastern Desert, NED, CED, and SED, respectively

Located in SE Sinai, it comprises metamorphosed volcano-sedimentary successions


that are differentiated into four E–W-trending tectono-stratigraphic units, arranged
geographically from north to south: the Melhaq, Um Zariq and Heib formations,
and the Tarr complex. The four units are juxtaposed along major tectonic contacts
that have been interpreted to be NE–SW- and E–W-trending low-angle megashears,
mylonites, and thrusts (Shimron 1980, 1983, 1984a, b, 1987; Abu Sharib 2003;
El-Bayoumi et al. 2006; Hafez et al. 2006; Shallaly 2007; Abdel Wahed and Shal-
laly 2007; Abdel Wahed et al. 2008; Fowler et al. 2010; Bogoch 1990). However,
normal stratigraphic and gradational contacts have been interpreted (Reymer and
Yogev 1983; Navon and Reymer 1984; Abdallah 1989). Lithologically, the Melhaq
and Heib are volcanic- and volcaniclastic-dominated formations composed mainly
of metamorphosed intermediate to felsic volcanic flows and their associated metapy-
roclastics intercalated with volcaniclastic sediments that are metamorphosed at the
greenschist facies condition. On the other hand, Um Zariq Formation and the Tarr
complex are mainly metasedimentary rocks consisting of alternating successions of
psammitic, pelitic, psammo-pelitic, calc–pelitic, and calcareous assemblages with
a northward increase in metamorphic grade from greenschist to amphibolite facies
conditions for the Tarr and Um Zariq formations, respectively.
Low Baric Metamorphic Belts in the Northern Tip … 269

Fig. 3 A generalized geologic map shows the distribution of the Precambrian rock units exposed
in the Eastern Desert and Sinai terranes
270 N. A. Shallaly and A. S. A. A. Abu Sharib

Fig. 4 A geologic map of the Kid–Um Zeriq area, Southern Sinai, Egypt (modified after Shallaly
2007). UZM: Um Zeriq Formation MMV: Melhaq metavolcanics TM: tectonic melange HMV:
Heib metavolcanics

2.2 Kid–Um Zeriq Belt

This belt is located in the south-eastern part of the Sinai Peninsula and constitutes
one of the most conspicuous LP/HT andalusite–sillimanite metamorphic belts in the
Precambrian rocks of Egypt (Hafez et al. 2006; Shallaly 2007). This structurally
lowest Um Zeriq metasedimentary (UZM) unit is overthrust by: a lower basic to
intermediate Melhaq metavolcanic (MMV) unit to the north (Fig. 5a), and an upper
intermediate to acidic Heib metavolcanic (HMV) unit to the east along two major
N- to NE- and E- to SE-trending thrust faults, respectively (Fig. 4). The MMV is
in turn overthrusted by a tectonic mélange (TM), a uniquely non-coherent unit of a
Low Baric Metamorphic Belts in the Northern Tip … 271

Fig. 5 a The Um Zeriq Formation (UZM) in contact with the Melhaq metavolcanics (MMV) along a
NNW-dipping thrust fault, Wadi Um Zariq, looking W. Reproduced from Abdel Wahed and Shallaly
2007. b A highly sheared thrust contact between the tectonic mélange (TM) and the overlying Heib
metavolcanics (HMV). Wadi Madsus, looking NE. Reproduced from Abdel Wahed and Shallaly
2007. c Huge elliptical blocks of MMV that are embedded in foliated metapelitic melange matrix,
TM, Wadi Madsus, looking W. d Tabular cross-bedding in a thinly laminated garnet quartzite, UZM.
Wadi Kid, looking N. e Rhythmically layered chiastolite schist with darker porphyroblast-rich and
paler porphyroblast-poor layers exhibiting bedding parallel foliation, S1 = S0 . Lower chiastolite
schist horizon, UZM, Wadi Kid, looking E. Reproduced from Shallaly 2007. f Asymmetric pinch
and swell structure in a mafic exotic block in mylonitized metapelites. UZM, Wadi Kid, looking SE

chaotic fabric that is derived from the partial cataclasis of the UZM and MMV, and
is itself overthrust by the HMV (Fig. 5b). The whole pile is intruded by different
pulses of deformed syn-tectonic to undeformed post-tectonic intrusive rocks and
dykes (Fig. 4).
272 N. A. Shallaly and A. S. A. A. Abu Sharib

2.2.1 Lithological and Sedimentological Characteristics

This well-bedded UZM metasedimentary unit is highly foliated and suffers severe
mylonitization and shearing along the major thrusts and along minor internal thrusts.
However, it is affected by polyphase deformation, overturning, and thrusting, yet
the original stratigraphy of this unit is rarely preserved in two locations (Shal-
laly 2007). Stratigraphically, the metasedimentary unit is composed of two fining-
upward sequences that comprise: basal metaconglomerates, pebbly schists, meta-
greywackes, quartzites, schists of pelitic and calc–pelitic origin. The metaconglom-
erates are polymictic that exhibit lateral facies change from clast support to matrix
support and lack internal stratification which attest to debris flow sedimentation. The
metagreywackes are preserved as discrete beds or as huge elongated blocks within
the metapelites and wrapped around by their foliations. The quartzites are foliated
and still preserve primary structures such as bedding/lamination and cross-bedding
(Fig. 5c). The metacalcpelites are thinly laminated garnet hornblende schists that
carry lenses of tremolite/wollastonite diopside marbles. The metapelites constitute
the principal lithology in the pile. They are rhythmically layered (Fig. 5d), often
laminated, well-foliated mica schists that bear chiastolite, garnet, biotite, staurolite,
and cordierite porphyroblasts, the variable proportions of which distinguish these
rocks into five varieties. The metapelites include two conspicuous marker horizons
of porphyroblastic chiastolite schists. This unit encloses variably sized foliated to
massive mafic magmatic blocks and lenses (Fig. 5e). These exotic blocks have no
equivalent outcrops in the study area.

2.2.2 Tectono-Metamorphic Evolution

This belt suffers a complex array of tectono-metamorphic history; in addition to


brittle deformation, three ductile deformational episodes (D1 –D3 ) accompanied by
three metamorphic events (M1 –M3 ) are recorded (Shallaly 2007; Abdel Wahed and
Shallaly 2007).
The D1 /M1 event is recorded in UZM and MMV with the development of S1
composite bedding foliation defined by biotite–muscovite–graphite and biotite–acti-
nolite, respectively. This was accompanied by intrafolial, rootless, and isoclinal F1
minor folds (Fig. 6a) that might affect the crystals themselves (Fig. 6b). L1 mineral
lineation is defined by chiastolite and hornblende. On progression, pophyroblastesis
of garnet, chiastolite, staurolite, and cordierite in the UZM and hornblende in the
MMV were crystallized at different stages of D1 witnessing the development of
amphibolite facies metamorphism, at the end of this phase (Hafez et al. 2006).
The D2 /M2 is recorded in UZM, MMV, and TM and produced mylonitization
and intensive shearing along two major thrust faults between these units and several
minor internal thrusts within each unit. This phase produced four major coaxial SE-
plunging overturned folds (Fig. 4) that are coaxial with tight overturned F2 minor
folds (Fig. 6c). Axial planar S2 foliation that is mylonitic near major thrusts intersects
the S1 , is accompanied by shear band cleavage (Fig. 6d, e), and is traced mainly by
Low Baric Metamorphic Belts in the Northern Tip … 273

Fig. 6 a Tight, rootless F1 fold closures traced by chiastolite-rich layers. Chiastolite schist, UZM,
Wadi Kid, looking SE. b Tightly folded chiastolite porphyroclasts associated with numerous much
finer-grained garnet crystals in chiastolite schist. Wadi Kid; looking NW. c Gently SE-plunging tight
overprinted F2 fold. UZM, Wadi El Cord, looking SE. Reproduced from Shallaly 2007. Photograph
(c) and line drawing e show a well-developed S-C fabric in mylonitic chiastolite schist. The C planes
are almost horizontal, and the S planes are oriented top left to bottom right with a dextral sense of
shear. Lower chiastolite schist horizon, UZM, Wadi Kid, looking E. f, g Photograph (f) and sketch
(g) of asymmetric mushroom interference pattern formed due to superposition of F1 and F2 folds.
L2 staurolite (shimmerzied) mineral lineation lies within the S3 plane parallel to the axis of F3 fold.
UZM, Wadi Kid, looking W
274 N. A. Shallaly and A. S. A. A. Abu Sharib

biotite. Interference pattern of folding resulting from the superposition of the second
and first phases of deformation is occasionally seen in the UZM (Fig. 6f, g). The
D2 event appears to have been most prolonged and effective, and it was associated
with initially increasing temperature reaching amphibolite facies conditions, with
the crystallization of a second generation of chiastolite, staurolite in addition to silli-
manite in metapelites, and hornblende in the metavolcanic. L2 is expressed by mineral
and stretching lineations; the former is defined by chiastolite and staurolite in UZM
(Fig. 6f) and hornblende MMV. The latter is exemplified by deformed pebbles in
metaconglomerates, pebbly schists (Fig. 7a), and meta-agglomerates, as well as boud-
inaged quartz vines, exotic blocks, and chiastolite porphyroclasts (Figs. 5f, and 6d).
Typical Abukuma and Buchan LP/HT belts are composed of cordierite and andalusite
(Spear 1993), so the copresence of staurolite as well suggests that these high-alumina
metapelites form a low-pressure intermediate-type metamorphism (Miyashiro 1961).
D3 /M3 event affected all units but is best recorded in HMV, which overthrust
the UZM, MMV, and TM along NNE–SSW-trending major thrust fault. A well-
developed mylonitic zone is recorded at the base of HMV unit, specifically meta-
agglomerate and meta-volcanic conglomerate (Fig. 7b). Overprinting pattern of F3
phase on F2 phase is well documented in MMV and UZM (Fig. 7c, d). F3 minor folds
are overturned that plunge to NE–NNE and followed by kink and open minor folds
(Fig. 7e); they are easily detected in the banded tuffaceous rocks of the HMV due to
different colours. Kink and open folds are either related to a late stage of D3 or may
represent a separate fourth deformational event (Abdel Wahed and Shallaly 2007).
This phase is marked by the development of SSE-dipping S3 mylonitic foliation
that is defined by chlorite and biotite, and L3 lineation is represented by kink axes
(Fig. 7f). Other mylonitic fabrics are best recorded in HMV such as oblique foliation,
boudinage structure, and mantled porphyroclasts (Fig. 7b). The D3 is accompanied
by the greenschist facies metamorphism that has reached the biotite isograd (Hafez
et al. 2006).
In addition to minor and mesoscopic normal and strike-slip faults affecting the
area, master faults that control the configuration of the rock distribution in the mapped
area are mainly strike-slip faults that are associated with a dip-slip component. Sinis-
tral NNE–SSW-trending faults conjugate with dextral WNW–ENE-trending faults
(Fig. 4). They were most probably formed during the same deformation event, related
to the Gulf of Aqaba–Dead Sea tectonics. The possibility that such faults represent
old Precambrian trends rejuvenated at later times cannot be excluded (Abdel Wahed
and Shallaly 2007).
Low Baric Metamorphic Belts in the Northern Tip … 275

Fig. 7 a NW-plunging clast lineation in highly deformed, polymictic, metaconglomerate. Some


deformed clasts (centre of photograph) show a well-developed asymmetric strain shadow. A later
boudinaged quartz vein (horizontal) traverses the rock. UZM, Wadi El-Cord, looking N. b A myloni-
tized metavolcanic conglomerate shows a sigmoidal-shaped granitic porphyroclast with asymmetric
tails wrapped by strong mylonitic foliation S3 , HMV, W. El Cord, looking S. c, d A photograph
(c) and sketch (d) showing F3 /F2 mushroom interference pattern, MMV, W. Madsus, looking NE.
e Overturned F3 fold with NNE-plunging axis, banded metatuffs, HMV near the thrust plane, Wadi
Madsus, looking SW. g Gently SW-plunging F3 open folds in banded metatuffs with alternating
pinkish and greyish bands. HMV, Wadi Madsus, looking SW
276 N. A. Shallaly and A. S. A. A. Abu Sharib

2.3 Samra–Hatmiya Belt

2.3.1 Lithological and Sedimentological Characteristics

The Samra–Hatmiya belt is an allochthonous low baric NE–SW-trending metamor-


phic belt that is structurally sandwiched between two mega-tectonic units, and consti-
tutes the extreme south-eastern part of the Tarr metamorphic complex. It is bounded
by two major NE–SW-striking and NW-dipping thrusts separating it from the struc-
turally higher volcanic-dominated Heib Formation and the structurally lower Khashm
El-Fakh metagreywacke (Fig. 8), and itself dissected by minor thrusts sub-parallel
to and synthetic with the major thrusts forming a characteristic pattern of imbricate
thrusts. The volcano-sedimentary succession is intruded by granite and structurally
controlled diorite and albitite bodies, and dissected by mafic sills and by older NW–
SE-trending mafic and younger NE–SW-trending acidic to intermediate dykes. The
relative timing of the dykes is proved by the fact that the former trend is recorded
only in the volcano-sedimentary succession whereas the latter trend cuts through the
intrusive granite.
Lithologically, the belt is differentiated into a lower pelitic-dominated and an
upper meta-rudite-dominated horizons. The lower horizon is composed mainly of
pelitic and subordinate psammitic schists intercalated with pebbly schist and thin
medium-bedded metaconglomerate a few tens centimetre thick. The pelitic schists are
porphyroblastic containing andalusite porphyroblasts up to 2 cm long (Fig. 9a) and
consist mainly of andalusite and biotite schists, in a decreasing order of abundance,
whereas the psammitic schist is composed mainly of schistose quartzite. The upper
horizon is made up dominantly of very thick, ~350 m total outcrop width, meta-rudite

Fig. 8 A simplified geologic map of Samra–Hatmiya area shows the distribution of the different
rocks as well as the major tectonic contacts
Low Baric Metamorphic Belts in the Northern Tip … 277

Fig. 9 a Textural and compositional lamination in andalusite schist shown by alternation of


andalusite-free and andalusite-rich laminae with the latter showing great variation in grain size.
North of Wadi Samra, looking NW. b Thinly bedded structure in the andalusite schist. The southern
side of Wadi Samra, looking SW. c Tabular cross-bedding in biotite schist. The northern side of
Wadi Ghorabi–El Hatmiya, looking NE. d Tangential cross-bedding in biotite schist. The northern
side of Wadi Ghorabi–El Hatmiya, looking NW. e Panoramic view shows thick primary bedding
in the pelitic-dominated lower horizon, featuring differential weathering. Northern side of Wadi
Ghorabi–El Hatmiya, looking NW. f Matrix-supported lateral extension of the thick metaconglom-
erate bed dissected by quartz veins and stained with iron oxides along fractures. The northern
side of Wadi Ghorabi–El Hatmiya, looking NE. g Thick, poorly sorted, clast-supported, polymictic
metaconglomerate bed. The southern side of Wadi Samra, looking due north
278 N. A. Shallaly and A. S. A. A. Abu Sharib

topped by pelitic, calc–pelitic, and pebbly schists. The metapelites are composed
of porphyroblastic andalusite schist and biotite schist. The calc–pelitic and pebbly
schists are composed of hornblende–biotite schist and thick pebbly biotite schist,
respectively.
From the sedimentological point of view, the two sedimentary horizons repre-
sent upward coarsening followed by upward finning depositional cycles. The schists
preserve well-developed primary sedimentary structures, S0, represented by textural
and/or compositional lamination and bedding (Fig. 9a and b), graded and rarely
cross-bedding of the tabular and tangential types (Fig. 9c and d). Thick mafic sills
are intruded along the bedding planes of thickly bedded andalusite schist (Fig. 9e).
A conspicuous laminated structure is preserved in the andalusite schist and shown
by the alternation of andalusite-rich and andalusite-poor, and biotite-rich laminae
containing coarse- and fine-grained andalusite porphyroblasts, respectively (Fig. 9a).
On the contrary, the meta-rudites are paraconglomerates that are internally chaotic,
and lack stratification and other sedimentary structures. They are polymictic and
very poorly sorted and contain rounded to sub-rounded clasts. Rudites of the lower
horizon are matrix-supported containing pebble- to cobble-sized clasts (Fig. 9f),
whereas the thick conglomerate bed of the upper horizon change along strike from
clast-supported chocked with very coarse boulders up to 40 cm in diameter (Fig. 9g),
to matrix-supported containing much smaller clast sizes. The prolific crystallization
of andalusite porphyroblasts in the pelitic-dominated beds of the upper and lower
horizons reflects, in general, highly aluminous muddy protoliths. In the metased-
imentary succession, a general abundance of thinly laminated schists is recorded.
The metaconglomerates, on the other hand, suggest “rapid deposition on relatively
steep slopes as lags” (see Abu Sharib 2003), in a proximal environment tentatively
a channel or canyon into the depositional basin. They were most probably deposited
by a process of subaqueous mass gravity transport, namely sediment gravity flow,
and in view of their above-mentioned characteristics, specifically by debris flow.
Conglomerate with boulder-sized clasts was deposited as a proximal facies at the
bottom of the feeder channel and the foot of the slope, whereas conglomerate with
smaller clasts would represent a distal facies of a debris flow. Additionally, deposi-
tion of pebbly mudstones originated as dispersed clasts in a muddy matrix has also
been ascribed to debris flows. The feature of upward coarsening followed by upward
fining depositional cycles has been attributed to renewal of relief by faulting of the
basin margin, providing coarser detritus to the downthrown side, followed by a return
to deposition of finer sediments as the uplifted source is worn down (see Abu Sharib
2003).
Rocks exhibiting contrasting laminated beds interbedded with beds that are
completely devoid of internal stratification have the sedimentological characteristic
features of debris flows. Turbidity flow, producing laminated deposits, is likely to
occur together with debris flow during a single event of mass gravity transport. In
conclusion, chaotic deposits which were emplaced by debris flows, and which contain
exotic clasts older than the enclosing sedimentary sequence, are called olistostromes.
Their presence indicates tectonically active phases of ancient basins, and they may be
Low Baric Metamorphic Belts in the Northern Tip … 279

produced by either tectonic shearing during subduction in trenches or gravity gliding


of nappes (Abu Sharib 2003 with references).

2.3.2 Deformation Events and Metamorphism

The low baric metasedimentary succession of the Samra–Hatmiya belt has been
multiply deformed through three phases of ductile (D1 –D3 ) followed by a phase
of brittle deformation. During the D1 and D2 , the rocks have been metamorphosed
up to the biotite zone of the greenschist facies conditions of the low-pressure type
as indicated by the growth of the characteristic low-pressure metamorphic phases,
cordierite and andalusite. The absence of garnet in all the beds is noted and attributed
to the grade of metamorphism, in conjunction with the low-pressure conditions,
which would favour the growth of the less dense cordierite (Miyashiro 1973) rather
than the “denser equivalent garnet” in rocks of appropriately low Fe2+ /Mg ratio.
Accordingly, the tectono-metamorphic history of the belt has been interpreted into
two tectono-metamorphic events (D1 /M1 and D2 /M2 ), together with a third weak
phase of deformation D3 , all of which are overprinted by a later phase of brittle
shearing.

Ductile Deformation

D1 /M1
The earliest tectonic event resulted from a period of NE–SW shortening and is
restricted to the metasedimentary succession. It produced a group of mesoscopic
SW-plunging overturned and nearly recumbent F1 folds (Fig. 10a and b) (Abu Sharib
2003; Abdel Wahed et al. 2006) and associated bedding parallel axial plane cleavage
(Fig. 9a, b) (S1//S0). Fold vergence indicates top-to-SE tectonic transport (Fig. 10a).
To the south of the Samra–Hatmiya belt, spectacular overturned cross-bedding has
been detected in Khashm El-Fakh metagreywacke, the extreme south-eastern expo-
sure of the Tarr complex (Shimron 1984a, b), and in a reconnaissance study done by
Abu Sharib (2003). Low baric greenschist facies metamorphic conditions prevailed
and favoured the growth of first-generation andalusite and cordierite porphyroblasts,
And1 and Crd1, respectively. Growth of the porphyroblasts is interpreted to have
taken place at the onset of D1 event (early syn-tectonic). However, this phase of
syn-tectonic growth is well preserved only at the microscopic scale as evidenced by
the slightly stretched and preferentially oriented andalusite and cordierite porphy-
roblasts (for a detailed microfabric study, the reader refers to Abu Sharib (2003)).
SW-plunging axes of F1 folds form the main linear fabric element, L1 , accompanied
this tectonic event.
However, foliated extraformational clasts (metasedimentary) in the metaconglom-
erates imply an earlier (pre-D1 ) cycle of sedimentation and lithification followed by
deformation (Abu Sharib 2003).
280 N. A. Shallaly and A. S. A. A. Abu Sharib

Fig. 10 a Schistose quartzite exhibits SE-verging, tight overturned folding in the north of Wadi
Samra, looking due south. b F1 overturned minor fold in the andalusite schist. The fold axis plunges
S25°W at 33°. The southern side of W. Samra, looking SW. c Well-developed streaky mylonitic
fabric recorded as a narrowband (50 cm thick) within a shear zone cutting through the pelitic-
dominated horizon. The southern side of Wadi Samra, looking NE. d Well-developed mylonitic
fabric recorded within a schistose quartzite bed. The northern side of Wadi Samra, looking due
east. e Mylonitized schistose quartzite with closely spaced streaky foliation. The entrance of Wadi
Samra on its northern side, looking NE

D2 /M2
This tectonic event resulted from a prolonged period of NW–SE-directed shortening
and produced the penetrative meso- and macroscopic NE-trending parallel to sub-
parallel structures (folds and thrusts and associated fabrics) that dominates the entire
structural grain of the Samra–Hatmiya area. The implication of the very similar
attitude of the fold axial planes and that of the thrusts is that the folds have been
Low Baric Metamorphic Belts in the Northern Tip … 281

formed either synchronous with or later during the thrusting. However, based on the
recorded styles of F2 folds, since they are not overturned (see below), the probability
that these folds formed at the later stages of thrusting is more plausible.
Accordingly, at the early stage, this tectonic event produced a group of NE–SW-
striking and NW-dipping imbricate thrust system (Fig. 8). It resulted in thrusting of
the Heib volcanics on top of the low baric metasedimentary succession, which in
turn thrusts over the underlying Khashm El-Fakh metagreywacke. Well-developed
thrust-related mylonitic fabric forms close to and along the thrust planes with inten-
sity decreases away from the thrust zones. In the metasediments, the earlier S0 //S1 is
overprinted by the mylonitic fabric, Sm , fabric, which despite the strong later shearing
(see below) is still locally preserved in a few narrow zones (Fig. 10c, d and e). Close
to the thrusts, the conglomerate and pebbly schist become mylonitized with intensely
deformed pebbles forming a remarkable NW-plunging clast lineation (Fig. 11 a and
b). In the metavolcanics, SM is the oldest recognized fabric that formed parallel the
primary layering and lamination, S0M, (Fig. 11c) of the rocks. It is well developed
in the tuffaceous varieties whereas much less pronounced in the flows. Myloni-
tized tuffs are highly sheared (Fig. 11d and e). At later stages, the defamation event
produced the pervasive, moderately tight to open, asymmetrical NE-plunging F2 folds
(Fig. 12a) and the associated penetrative axial plane cleavage, S2 , that overprints the
S0 //S1 , Sm , and S0m fabrics. The attitude of F2 axial planes is similar to that of the
thrusts. F2 has been recorded at all scales in the metamorphosed volcano-sedimentary
succession in the Samra–Hatmiya area. The low-pressure greenschist facies condi-
tion has been, still, underway, during which second-generation andalusite And2 and
cordierite Crd2 porphyroblasts grew syn-tectonically. NE-plunging mineral and clast
lineations are the prominent linear fabric, L2 . The former is represented by andalusite
mineral lineation (Fig. 12b), whereas the latter by the axes of deformed pebbles in
the metaconglomerate (Fig. 12c).

D3 /M3
The third phase of deformation was mild and resulted from NE–SW shortening.
It produced NW–SE-trending open symmetrical to asymmetrical NW- and SE-
plunging folds, F3 (Fig. 12d and e), associated with very weak, locally developed
axial planar cleavage, S3 . Linear fabric, L3 , is represented by locally developed NW-
plunging clast lineation in the deformed metaconglomerate (Fig. 13a). The local
development of this event, the widespread development of F3 kink folds, and the
absence of low baric metamorphic phases point to the much shallower conditions,
and hence much lower T, at which this tectonic event formed. Superposition of L3
kink axes on L1 mineral lineation (Fig. 13g), and of F3 on F1 and F2 folds (Fig. 13c,
d and e), is very locally preserved. For detailed structural data including steroplots
of the penetrative S0 //S1 foliation, and of the different lintations, L1 , L2 , and L3
throughout the different structural domains within the area, the reader refers to Abu
Sharib (2003) and Abdel Wahed et al. (2006).
282 N. A. Shallaly and A. S. A. A. Abu Sharib

Fig. 11 a Mylonitized thin matrix-supported metaconglomeratic bed with highly strained clasts
forming stretched clast lineation trending N34°W and plunging at 18°. Entrance of Wadi Samra,
looking NE. b Mylonitized pebbly schist with well-developed foliation and intense flattening and
stretching of the clasts forming a conspicuous clast lineation plunging NE at moderate angle.
Wadi Samra, looking SE. c Well-foliated and bedded metavolcanics, steeply dipping towards the
south-east. Notice the ridge forming coherent flows interbedded with the less competent volcani-
clastics. Closure of Wadi Ghorabi–El Hatmiya, looking NE. d Sheared tuffaceous schists show
well-developed vertical foliation. The southern side of Wadi Ghorabi–El Hatmiya, looking SW.
e Mesoscopically crenulated and fractured tuffaceous schist. The crenulation axes plunge at 25°
towards N38° E and west of the closure of Wadi Ghorabi–El Hatmiya, looking NE
Low Baric Metamorphic Belts in the Northern Tip … 283

Fig. 12 a F2 open fold in the andalusite schist. The fold axis plunges towards N30°E at 15°.
The northern side of W. Ghorabi–El Hatmiya, looking NE. b L2 andalusite mineral lineation in
andalusite schist. The andalusite crystals plunge towards the NE at 18°. The northern side of W.
Samra, looking NW. c Oriented deformed clasts in metaconglomerate. L2 clast lineation plunges
at 21° to the NE. The entrance of W. Samra on its northern side, in the vicinity of a fault, looking
SE. d F3 open fold in biotite schist. The fold axis plunges by 17° to the NW. The entrance of Wadi
Samra at its southern side, looking NW. e F3 open fold in the andalusite schist. The fold axis plunges
towards S60°E at 35°. The southern side of W. Samra, looking SE

Brittle Deformation

The Samra–Hatmiya belt is intensively affected by a brittle shearing event that is


manifested by sub-parallel NE- to ENE-oriented major shear zones. The shear zones
are characterized by the arrangement of a remarkable quartz stockwork, ~100–150 m
284 N. A. Shallaly and A. S. A. A. Abu Sharib

Fig. 13 a Well-foliated thin metaconglomeratic bed exhibits stretched clast lineation, L3 . The clast
lineation plunges NW at moderate angle. The northern side of Wadi Samra, looking NE. b F3
kinking superimposed on L1 mineral lineation which is well developed on the S1 kinked planes.
The kink axes plunge at 33° towards S65°E. The northern side of W. Ghorabi–El Hatmiya, looking
NW. c Superposition of steeply NW-plunging F3 folds (red circle denoted e) on gently NE-plunging
F2 folds (red circle denoted d) in the metavolcanics. The southern side of W. Samra, looking NE.
d, e Schematic diagrams of the folds shown in (c), the red circles labelled d and e show the shape
and orientation of F2 folds prior to (d) and after (e) the superposition
Low Baric Metamorphic Belts in the Northern Tip … 285

outcrop width (Fig. 14a and b), and of albitite bodies and red gossans along linear
outcrops (Fig. 14 c and d). The narrow, not more than 60 cm wide, mylonite zones
that can be barely recognized within the sheared rocks along some of the thrusts
(Fig. 10c, d and e) point to the fact that shearing is a later event overprinted on the old
thrusts (Fig. 14e). The latter conclusion is further supported by the fact that the shear-
related quartz stockwork cuts through the small structurally controlled diorite bodies
intruding the metasedimentary succession (Fig. 14f). The albitite and diorite bodies
are very close-associated, spatial related to the thrusts, and are fractured and affected
by brittle shearing. However, the absence of ductile deformation in these bodies would
imply a post-thrusting and pre-shearing intrusion. However, intrusion during the early
stages of the brittle shearing cannot be excluded. Some quartz veins within the shear
zone are displaced by compressional faults (Fig. 14g), which implies the complexity
and reactivation of the shear zone. Shear zone-related copper mineralization in the
form of secondary malachite and azurite associated with secondary iron has been
recorded (Fig. 14h and i).

2.4 Geological Setting of Central Eastern Desert Belts

The Central Eastern Desert comprises two assemblages that are separated by high-
stain mylonitized zones and/or intrusive contact: an infracrustal and a supracrustal
assemblages. The Cryogenian and Tonian supracrustal assemblage is ensimatic and
consists of ophiolitic rocks, arc volcanics and immature volcaniclastic, banded
iron formation, and diamictite that are metamorphosed under greenschist facies
conditions. This assemblage is characterized by brittle deformation and intruded
by Cryogenic to Ediacaran granitoids. The infracrustal assemblage is affected by
ductile deformation and metamorphism under upper amphibolite facies conditions
is composed of granitic gneisses and amphibolites (Fritz et al. 2013; Stern 2018,
Hamimi et al. 2019).

2.4.1 W. Um Had Belt

Lithological Characteristics

In Um Had area, located almost half away of the famous Qift–Qusir road, ~60 km
west of Qusir city, Central Eastern Desert (Fig. 3), a NW–SE-trending low baric
metamorphic belt is exposed. The belt represents a thrust sheet that is separated
from the overlying Dokhan volcanics and ophiolitic amphibolite, and the underlying
Hammamat molasse sediments by two sub-parallel SW-trending thrusts (Fig. 15).
Close to the thrusts, the rock units are highly sheared (Fig. 16a, b and c). All the rock
units are intruded by the post-orogenic Um Had granite (Fig. 16d), a nearly circular
~10 km in diameter pluton that yielded a crystallization age of ca. 596 Ma (Andersen
et al. 2009).
286 N. A. Shallaly and A. S. A. A. Abu Sharib

Fig. 14 a A general view of the NE-striking brittle shear zone with a spectacular quartz stockwork
cuts through the lower pelitic-dominated horizon. The northern side of W. Ghorabi–El Hatmiya,
looking NNE. b Close-up view of intense quartz stockwork developed near a thrust fault, in the
lower pelitic-dominated horizon. The southern side of Wadi Samra, looking SW. c Panoramic
view of the lower pelitic-dominated, traversed by an almost linear, NE-trending shear zone (ShZ)
occupied by whitish albitite bodies (Ab). South of the entrance of Wadi Samra, looking NW.
d ENE-trending shear zone (ShZ) with well-developed albitites (Ab) cuts through the lower pelitic-
dominated horizon. The southern side of W. Ghorabi–El Hatmiya, looking SW. e NW-dipping shear
zone separates the lower pelitic-dominated horizon from the lower metagreywacke. The southern
side of W. Samra, looking south. f A dioritic mass (Di) dissected by the shear-related quartz veins
and intrudes the andalusite schist (And) of the lower pelitic-dominated horizon. The Heib volcanic
(Vol) appears in the background. The southern side of Wadi Ghorabi–El Hatmiya, looking SW.
g Later minor compressional faults affect the quartz veins of the stockwork cutting through the
metaconglomerate. The southern side of W. Samra, looking due W. h Abandoned shaft in sheared
metagreywacke used for exploiting malachite. i Close-up view of secondary iron associated with
malachite. South of Wadi Ghorabi–El Hatmiya
Low Baric Metamorphic Belts in the Northern Tip … 287

Fig. 15 A general geologic map shows the extension and major tectonic contact of the low baric
metapelite-dominated belt in Um Had area
288 N. A. Shallaly and A. S. A. A. Abu Sharib

Fig. 16 a Highly sheared foliated amphibolites, Am, near the contact with the metapelites. End of
W. Um Had, looking NW. b Close-up view of the foliated amphibolites. The topographically high
rock in the background is the north-western end of Atalla felsite, Fs. W. Um Had, looking NW.
c Highly sheared acidic metavolcanics near the contact with metapelites. End of W. Um Had, looking
NNW. d Um Had granite, Gr, intrudes the pelitic-dominated rocks. W. Um Had, looking NE. e Plan
view of porphyroblastic andalusite and cordierite schist. W. Um Had. f Andalusite schist exhibits
a well-developed banded structure where porphyroblast-rich alternates with porphyroblast-poor
bands. W. Um Had, looking due north
Low Baric Metamorphic Belts in the Northern Tip … 289

Lithologically, the belt comprises a succession consisting of dominantly porphy-


roblastic pelitic schists, subordinate metagreywacke, and a very few thin calc–silicate
beds. Porphyroblasts, some of which exceed 2 cm in length, of cordierite, staurolite,
andalusite, biotite, and garnet characterize the metapelites (Fig. 15e). Bedding is
the most pronounced primary sedimentary structure, S0 . It is represented by alter-
nation of metapelitic and metapsammitic layers, and of porphyroblastic-rich and
porphyroblastic-poor bands in the more pelitic varieties (Fig. 16f).
Based mainly on the geochemical discrimination diagrams, the protolith of the
pelitic metasedimentary rocks was forearc mainly pelitic and subordinate silicic sedi-
ments deposited in sedimentary basins developed within active continental island arc
tectonic setting (Abu Sharib et al. 2019). Minimum depositional age of the sediments
is constrained, based on geochronological data of deformation and metamorphism,
see below, to pre-620 Ma, whereas the maximum depositional age is suggested to be
Late Cryogenian comparable to metapelites from Sinai (Feiran–Solaf metamorphic
complex and Um Zariq Formation) deposited in a similar tectonic setting recorded
(Abu El-Enen and Whithouse 2013; Moghzi et al. 2012; El-Bialy 2013).

Metamorphism and Deformation History of Um Had Belt

The metapelite-dominated succession is multiply deformed and metamorphosed


through a succession of three phases of regional deformation, D1 /M1 , D2 /M2 , and
D3 overprinted by a third phase of thermal metamorphism, M3 . Low-pressure condi-
tions prevailed during the three metamorphic phases with temperature fluctuated from
medium during M1 to low during M2 and back to high during M3 . For a detailed
microfabric study including the interplay between deformation, metamorphism, and
porphyroblasts, the reader refers to Abu Sharib et al. (2019).

D1 /M1
The first deformation event took place at medium grade amphibolite facies meta-
morphic condition and has been dated at 625 ± 5 Ma (Abu Sharib et al. 2019).
Metamorphic phases that grew during this event include first-generation staurolite,
cordierite, biotite, and garnet, all of which are well preserved mainly at the micro-
scopic scale as inclusions enclosed within andalusite and cordierite porphyroblasts of
the later events. Planar fabric elements include bedding parallel foliation, i.e. S1 //S0
that is very locally preserved in less strained outcrops (Fig. 17a). This tectonic event
overlaps well with the 630 Ma Um Ba’anib gneissose granite that occupies the core
of the nearby Meatiq gneissic dome (Abu Sharib et al. 2019).

D2 /M2
This is the most penetrative event in the metapelites, in particular, and Um Had
area, in general. Metamorphic condition dropped to greenschist facies during this
290 N. A. Shallaly and A. S. A. A. Abu Sharib

Fig. 17 a Andalusite schist exhibits thinly bedded primary sedimentary structure shown by alter-
nation of psammitic and pelitic thin beds, and shows a well-developed bedding parallel foliation,
S1 //S0 . W. Um Had, looking NW. b The penetrative foliation, S2 , in metapelites intersects the
composite S1 fabric at a small angle. W. Um Had, looking SE. c L2 moderately NE-plunging
andalusite mineral lineation in andalusite schist. W, Um Had, looking NE. d Gently SE-plunging
andalusite mineral lineation, L3 . W. Um Had, looking NE. e SE-plunging F3 fold/crenulation in
andalusite schist. W. Um Had, looking SE. f Moderately SE-plunging crenulation lineation, L3 , in
metapelite. W. Um Had, looking NW. Gently plunging kinks and open fold in Dokhan volcanic
(g) and Hammamat sediments (h), respectively
Low Baric Metamorphic Belts in the Northern Tip … 291

tectono-metamorphic event. It is represented by growth of syn-tectonic second-


generation cordierite, andalusite, garnet, and biotite porphyroblasts. Deformation
related to D2 includes the penetrative foliation, S2MS , that dominates the entire struc-
tural grain of Um Had area, and cross-cuts the S0 //S1 composite foliation at a moderate
angle (Fig. 17b), and NE-plunging cordierite and andalusite mineral lineations, L2
(Fig. 17c). At the microscopic scale, S2 is a crenulation foliation crenulating the
composite S1 fabric (Abu Sharib et al. 2019).

D3
The third phase of deformation accompanied an episode of NE–SW shortening and
is considered the most prominent phase to affect Um Had affecting all the exposed
rock units. It produced the macro- and mesoscopic NW–SE-trending and gently
SE-plunging upright folds, F2 (Fig. 17d), and resulted in folding and crenulation of
the penetrative S2 fabric about NW–SE-trending axes. Linear fabric elements, L3 ,
include SE-plunging cordierite and andalusite mineral lineations (Fig. 17e), axes of
mesoscopic folds (Fig. 17d), and crenulation lineation (Fig. 17f). D3 is well developed
in the structurally upper and lower units. It resulted in gently SE-plunging mesoscopic
folds in the Dokhan volcanics and Hammamat sediments (Fig. 17g and h) with a
very local axial planar cleavage in the latter and in the pronounced clast lineation
in the Hammamat conglomerate (Abu Sharib et al. 2019, Fig. 2d, e and f). For
steroplots of the penetrative foliation and the different linear fabric elements within
the metapelites, the reader refers to Abu Sharib et al. (2019).
D1 /M1 and D2 /M2 are interpreted to be pre-collision events (i.e. collision of East
and West Gondwana) linked to a system of Late Ediacaran subduction that shared a
common shortening direction. D1 /M1 is related to an early <630 Ma subduction, the
magmatic activity of which produced Um Ba’anib granite ~630 Ma. D2 /M2 is related
to a later, shortly before 610 Ma, subduction system and resulted in gneissification
of Um Ba’anib granite. D2 is correlated with the strong NW–SE bulk horizontal
shortening event well documented in the Central Eastern Desert and responsible for
the NW-/NNW-directed thrusts in Um Had and the nearby areas (Fowler and Osman
2001; Andersen et al. 2009; Abu Sharib et al. 2019). The footwall block of the
thrust occupies the amphibolite facies Meatiq gneissic dome, whereas the hanging
wall occupies a greenschist facies tectonic nappe comprising Dokhan volcanics,
Hammamat sediments, and ophiolitic mélange (Fowler and Osman 2001; Andersen
et al. 2009; Abu Sharib et al. 2019, and the references therein). Locally developed
mesoscopic thrust-related folds account for the few NE-plunging cordierite mineral
lineation. D3 has been interpreted to be a collisional-related event that is correlated
with the dome-forming event in the nearby Meatiq gneiss.

M3
The last metamorphic phase to affect Um Had area is a thermal metamorphism, M3 ,
that has been dated at ~599 Ma and well developed in the contact aureole of the
post-orogenic Um Had granite (Neumayer et al. 1998; Fowler and Osman 2001;
292 N. A. Shallaly and A. S. A. A. Abu Sharib

Fig. 18 A geologic map of the W. El Miyah belt, El-Bakriya area, Central Eastern Desert, Egypt
(after Shallaly 2019)

Abu Sharib et al. 2019). It is of the medium to high temperature type attaining the
sillimanite grade of the hornblende hornfels facies. Porphyroblasts related to this
metamorphic event include cordierite, andalusite, biotite, and garnet of the third
generation together with a second-generation staurolite. It is noteworthy that this is
the only metamorphic event to affect the Hammamat sediments (see Abu Sharib et al.
2019).

2.4.2 W. El Miyah Belt

Lithological and Sedimentological Characteristics

This belt is located in the El-Bakriya area, Central Eastern Desert (Fig. 18). It consti-
tutes one of the ophiolitic mélange–island arc belts that are widespread in the northern
ANS (e.g. Hafez and El-Amin 1983; Mohammed 1999; Zoheir and Lehmann 2011;
Shallaly 2019). This area comprises a fold–thrust belt that is composed of thrust
stacks of arc-related metamorphosed volcano-sedimentary sequence with an ophi-
olitic mélange (Fig. 18) that is flooded with numerous magmatic rocks (Fig. 19):
syn-orogenic gneissose diorite and tonalite that are concordant to the tectonic fabric
of the country rocks (El-Amin 1975; Hafez and El-Amin 1983; Shallaly 2019). Late-
to post-orogenic granitoids (i.e. alkali granite) intrude the metamorphosed volcano-
sedimentary rocks and the ophiolitic mélange as well as the syn-orogenic pulses, with
Low Baric Metamorphic Belts in the Northern Tip … 293

Fig. 19 A 5/7-4/5-3/1 enhanced ratio ETM+ landsat image showing the distribution of the granitoids
in the Wadi El Miyah area (created by Assoc. Prof. M. Abdel Wahed, Cairo University). WMB:
W. El Miyah belt BOB: Barramiya ophiolitic belt SOG: Syn-orogenic granite POG: Post-orogenic
granite NSS: Nubia sandstone

all intrusive features: sharp contact, apophysis, roof pendant, and xenoliths from the
older rocks (Mohammed 1999). The belt is covered to the west by the Cretaceous
Nubia sandstones. The ophiolitic mélange constitutes the northern extension of the
Barramiya belt (Fig. 19). This belt is a classic example of regional contact meta-
morphism (Shallaly 2019), where the emplacement of large amounts of magmatic
bodies during accretion produces a very high thermal gradient responsible for the
development of LP/HT metamorphism (Spear 1993).

Metamorphosed Volcano-Sedimentary Sequence (MVS)

This sequence, which is juxtaposed along a low-angle thrust fault with an ophiolitic
mélange (Fig. 20a), consists of a clastic metasedimentary sequence that is interca-
lated with metavolcanics. The metasedimentary rocks are composed of metapelites,
294 N. A. Shallaly and A. S. A. A. Abu Sharib
Low Baric Metamorphic Belts in the Northern Tip … 295

Fig. 20 a A WNW–ESE-trending low-angle thrust fault that juxtaposes the ophiolitic melange
(OM) against the metamorphosed volcano-sedimentary (MVS) unit. b An internal minor thrust fault
between the metagreywackes (GW) and metapelites (MP) that in turn juxtaposed against ophiolitic
melange along a parallel major low-angle thrust fault looking NE. c A highly sheared yellowish
foliated chiastolite schist with a prominent boudinaged chiastolite porphyroclasts. d Chiastolite
forming two mineral lineations: a ESE-plunging L1 intersected by a younger WNW-plunging L2 .
Metapelites, looking NE (reproduced from Shallaly 2019). e Diamond-shaped staurolite porphyrob-
lasts wrapped around by S2 foliation. Metapelites, looking NW. f A rounded serpentinized ultramafic
block embedded in a highly sheared talc serpentine schist, ophiolitic melange. g Differently sized
elongated blocks of metavolcanic blocks in ophiolitic melange

quartzites, metagreywackes, metaconglomerates, and rare calc–silicate rocks. The


metaconglomerates are at the base of the succession. It is polymictic, poorly sorted,
of angular to subangular clasts of para-gneisses and schists, metavolcanic and
granitic origin. They are embedded in a schistose micaceous matrix (Hafez and El-
Amin 1983). The metagreywackes are interbedded with and/or overthrusted by the
metapelites along minor thrusts that are parallel to the main melange–sediment thrust
(Fig. 20b). The clasts are mainly of volcanic origin and are wrapped around by the
schistose micaceous matrix. The quartzite is rarely seen in the south-west of the area
as thinly bedded yellowish intruded by granitic batholith that may carry them as roof
pendant. The calc–silicate rocks occur in the form of thin greenish to greyish white
bands. The metapelites form the most conspicuous porphyroblastic-rich foliated
schists that on shearing turn into yellowish papery foliation (Fig. 20c). These porphy-
roblasts are mainly chiastolite, staurolite, garnet, and less commonly cordierite;
from these, only the first two minerals define conspicuous mineral lineation that
are recorded in these schists (Fig. 20d, e; Shallaly 2019). The metavolcanics are
in the form of schistose intermediate to acidic lava flows and tuffaceous equiva-
lents that are intercalated with the metasediments. To the north of the study area,
these arc-related metavolcanics are exposed in Wadi Abu-Muawwad as huge belts
of acidic–intermediate and less commonly basic porphyritic rocks and pyroclastics
(Mohammed 1999).

Ophiolitic mélange (OM)


This unit is composed of variably sized blocks of native and exotic nature that are
embedded in a highly sheared matrix. The blocks include proper ophiolitic rocks such
as serpentinite, metagabbros, and metabasalts, and arc metavolcanics and quartzites
and quartzofeldspathic (Fig. 20f, g). Huge fragments of serpentinite and metagabbro
from mountainous masses in the central part of the area are in thrust contact with
the mélange matrix (Fig. 18). Generally, the matrix of the mélange is brown foliated
metapelites, except at the base of the huge serpentinite masses where the matrix is
highly sheared greenish white serpentineous talc schist.
296 N. A. Shallaly and A. S. A. A. Abu Sharib
Low Baric Metamorphic Belts in the Northern Tip … 297

Fig. 21 a A Google Earth image illustrating one of the major ESE-plunging F1 isoclinal folds. MVS.
b An ESE-plunging isoclinal F1 minor fold in metagreywacke, along Wadi El Miyah, looking NW
(reproduced from Shallaly 2019). c A Google Earth image showing the asymmetric WNW-plunging
major F2 fold, ophiolitic mélange. d A WNW-plunging asymmetric F2 fold in metasediments,
with amphibolitic enclave, looking NE (Reproduced from Shallaly 2019). e A NE-dipping rock
fragments defining stretching lineation that are parallel to and embedded in a mylonitic foliation
Sm. metagreywacke. f A NW-plunging kink fold belonging to D3 in quartzite, MVS, looking W
(Reproduced from Shallaly 2019). g A minor NW-plunging open F3 fold in talc schist, ophiolitic
mélange, looking WNW. h A quartz mantled porphyroclast, with stair-stepping indicating a sinistral
sense of shear, metasediments, looking W (Reproduced from Shallaly 2019)

2.4.3 Tectono-Metamorphic History of W. El Miyah Belt

The tectono-metamorphic history of this area could be summarized as follows: the


three deformational episodes (D1 -D3 ) were accompanied by two prograde metamor-
phic events (M1 -M2 ). For microfabrics and monazite dating, see Shallaly (2019).
The MVS is affected by D1 to D3 , whereas the OM is affected by D2 and D3 only.

D1 /M1

During D1 , N–S shortening produced major and coaxial minor ESE-plunging over-
turned F1 folds (Fig. 21a and b). This is accompanied by moderately to steeply
dipping S1 composite bedding foliation//S0 , which is expressed in the lamination
of the metapelites. S1 is defined in the metapelites by muscovite + biotite and in
amphibolite by chlorite +biotite. Moreover, SE-plunging mineral lineation L1 is
defined mainly by chiastolite and staurolite in the metapelites (Fig. 20d). At an age
of 689–92 Ma, M1 reached lower amphibolite facies conditions (Shallaly 2019) that
witnessed the crystallization of garnet, staurolite, and chiastolite in the metapelites
and hornblende in the associated intermediate metavolcanics. This phase probably is
correlated with the accretion of island arc stage (740–660 Ma), produced by NNW
thrusts and associated folds (Abd El-Wahed 2008; Fritz et al. 1996, Loizenbauer
et al. 2001; Hamimi et al. 2019; Shalaby et al. 2005). In this belt, only folding was
recorded, without thrusting.

D2 /M2

D2 is accompanied by a major asymmetric WNW-plunging F2 fold recognized in the


ophiolitic mélange rocks (Fig. 21c and d) contemporaneous with the development
of N- to NE-dipping thrust stacks. Extensive mylonitization is recorded along these
thrust faults, where NE-dipping Sm = S2 (Fig. 20e), and partially crenulated foliation
is axial planar to F2 minor folds. Associated L2 lineation is defined by staurolite and
chiastolite (Fig. 20e), crenulation lineation, and stretched clasts and boudins. D2 was
298 N. A. Shallaly and A. S. A. A. Abu Sharib

accompanied by the prograde M2 metamorphic event at an age of 620–581 Ma (Shal-


laly 2019). Depending on the depth at which metamorphism took place, two kinds
of facies conditions were recorded. Along the major D2 -related thrust, juxtaposition
of greenschist facies belt (i.e. OM at higher level) and amphibolite facies belt (i.e.
MVS at deeper level) took place. The ophiolitic mélange reached the biotite isograd
with S2 defined by biotite in the metapelitic-rich part and by antigorite + talc in
the serpentineous facies. The metavolcano-sedimentary pile reached the sillimanite
isograd with an assemblage of garnet + staurolite + chiastolite + sillimanite. Local
thermal effect of the granites on the studied rocks is exemplified by the crystalliza-
tion of cordierite + chiastolite in the metapelites and hornblende within ophiolitic
mélange. This phase is related to the stage of oblique convergence of the arc and
back-arc assemblage onto the Saharan metacraton that are documented in partic-
ular in the Central Eastern Desert (e.g. Fritz et al. 2002; Abd El-Wahed et al. 2016;
Hamimi et al. 2019).

D3

D3 is expressed only by the development of macro-minor kinks and open F3 (Fig. 21a,
f, g), and NW axial planar weak S3 foliation. L3 lineation is expressed only by kink
axes and porphyroclasts. Additionally, the belt was affected by a NE–SW sinistral
strike-slip fault that developed rotated mantled porphyroclasts (Fig. 21h). D3 was
accompanied mainly by retrograde metamorphism that is best recorded in the MVS
in the form of chlorite fish with a sinistral sense of shear (see Shallaly 2019).This
deformational phase is related to the Najd Fault System affecting the northern ANS
and resulted in deposition of the Hammamat sediments and exhumation of gneissic
domes (e.g. Stern 1985; Johnson et al. 2011; Fritz et al. 1996).

Acknowledgements As part of this work is based on their Ph.D. and M.Sc. theses, the authors
are indebted to Prof. A. Hafez, Assoc. Prof. M. Abdel Wahed, Cairo University, and Prof. Dr.
M. G. Shaheen, Beni-Suef University, for their critical reviewing, and comprehensive and fruitful
discussions during supervision on both theses on Wadi Kid and Wadi Samra areas. The first author
is grateful for Assoc. Prof. M. Abdel Wahed for assessing her in Landsat image processing for the El
Miyah area. The tectono-metamorphic evolution of Um Had area was a part of a paper A. Abu Sharib
published with Prof. A. Maurice, Helwan University, and Assoc. Prof. Y. Abdel-Rahman, Cairo
University, to whom he is greatly indebted for valuable discussion, critical reviewing, and help during
the field trip. Soumyajit Mukherjee reviewed this article in two rounds and handled as an editor.
Springer proofreading team is thanked. Thanks to Marion Schneider, Annett Buettener, Boopalan
Renu, Alexis Vizcaino, Doerthe Mennecke-Buehler, and the proofreading team (Springer). Dutta
and Mukherjee (2021) summarize this article.
Low Baric Metamorphic Belts in the Northern Tip … 299

References

Abdallah, S. E. (1989). Geology, petrology and geochemistry of Wadi Samra area, southeastern
Sinai, Egypt (p. 200). M.Sc. thesis. Egypt: Zagazig University.
Abd El-Wahed, M. A. (2008). Thrusting and transpressional shearing in the Pan-African nappe
southwest El-Sibai core complex, Central Eastern Desert Egypt. The Journal of African Earth
Sciences, 50, 16–36.
Abdeen, M. M., Greiling, R. O., & Dardir, A. A., (1996). Two phases of Pan-African thrusting in
the Proterozoic basement, North Eastern Desert of Egypt. In Abstract of the Geological Survey
Egypt Center Conference (pp. 5–6). Cairo, Egypt.
Abdel Wahed, M., & Shallaly, N. A. (2007). Arc-continent collision and subduction reversal in
the Central Wadi Kid Precambrian rock complex (pp. 37–50). Egypt: Southeast Sinai, Egyptian
Annual Geology Survery.
Abdel Wahed, M., Hafez, A. M. A., Shahin, M. G., & Abu, Sharib A. (2006). Structural framework
of Wadi Samra area, SE Sinai, Egypt. Abstract, 8th International Conference on Geology of the
Arab World (pp. 94–95). Cairo: Egypt.
Abu El-Enen, M. M., & Whitehouse, M. J. (2013). The Feiran-Solaf metamorphic complex, Sinai,
Egypt: Geochronological and geochemical constraints on its evolution. Precambrian Research,
239, 106–125.
Abu Sharib, A. S. A. A. (2003). Structural and petrographical studies on the volcanic sedimentary
sequence in and around Wadi Samra, in Wadi Kid area. SE Sinai, Egypt. Unpublished MSc
Thesis, Cairo University.
Abu Sharib, A., Maurice, A., Abd El-Rahman, Y., Sanislav, I., Schulz, B., & Bakhit, B. (2019).
Neoproterozoic arc sedimentation, metamorphism and collision: Evidence from the northern tip
of the Arabian-Nubian Shield and implication for the terminal collision between East and West
Gondwana. Gondwana Research, 66, 13–42.
Andresen, A., El-Rus, M. A. A., Myhre, P. I., Boghdady, G. Y., & Corfu, F. (2009). U-Pb TIMS age
constraints on the evolution of the Neoproterozoic Meatiq Gneiss dome, Eastern Desert Egypt.
International Journal of Earth Sciences, 98(3), 481–497.
Bell, T. H., & Sanislav, I. V. (2011). A deformation partitioning approach to resolving the sequence
of fold events and the orientations in which they formed across multiply deformed large-scale
regions. Journal of Structural Geology, 33, 1206–1217.
Blasband, B., White, S., Brooijmans, P., De Broorder, H., & Visser, W. (2000). Late Proterozoic
extensional collapse in the Arabian-Nubian Shield. Journal of the Geological Society, London,
157, 615–628.
Bogoch, R. (1990). Mylonites, meta-hornblendite and meta-dolostone fragments in brittle shear-
zone, southeastern Sinai. Egyptian Israel Journal of Earth Sciences, 39, 17–24.
Bregar, M. (1996). The core complex of the Gabal Sibai crystalline dome in the Eastern Desert of
Egypt. In Abstract of the Geological Survey Egypt Center Conference (pp. 25–27). Cairo, Egypt,
pp. 25–27.
Brown, M., & Solar, G. S. (1998a). Shear zone systems and melts: feedback relations and self
organization in orogenic belts. Journal of Structural Geology, 20, 211e227.
Brown, M., & Solar, G. S. (1998b). Granite ascent and emplacement during contractional
deformation in convergent orogens. Journal of Structural Geology, 20, 1365e1393.
Brown, M., & Solar, G. S. (1999). The mechanism of ascent and emplacement of granite magma
during transpression: A syntectonic granite paradigm. Tectonophysics, 312, 1–33.
Bucher, K., & Frey, M. (1994). Petrogenesis of metamorphic rocks (6th ed., p. 320). Berlin: Springer.
De Yoreo, J. J., Lux, D. R., Guidotti, C. V., Decker, E. R., & Osberg, P. H. (1989). The Acadian
thermal history of western Maine. Journal of Metamorphic Geology, 7, 169–190.
Dutta, D., & Mukherjee, S. (2021). Introduction to Structural Geology and Tectonics Field Guide-
book—Volume 1. In S. Mukherjee (Ed.), Structural Geology and Tectonics Field Guidebook—
Volume 1. Springer Nature Switzerland AG. Cham. pp. xi-xvi. ISBN: 978-3-030-60142-3.
300 N. A. Shallaly and A. S. A. A. Abu Sharib

El-Amin H. (1975). Radiometric and geologic investigations of El-Bakriya area (244 p). Eastern
Desert, Egypt. Ph.D. Thesis, Cairo, Egypt: Cairo University.
El-Bayoumi, R., Abdel-Wahed, M., Tobschall, H. J., & Shallaly, N. A. (2006). The counterclockwise
pressure–temperature–time path of the metavolcanics and associated rocks from central Wadi Kid
area, southeast Sinai, Egypt. In Abstract 8th International Conference of the Geology of the Arab
World (pp. 12–13). Cairo, Egypt.
El-Bialy, M. Z. (2013). Geochemistry of the Neoproterozoic metasediments of Malhaq and Um Zariq
formations, Kid metamorphic complex, Sinai, Egypt: Implications for source-area weathering,
provenance, recycling, and depositional tectonic setting. Lithos, 175–176, 68–85.
El Gaby, S., List, F. K., & Tehrani, R. (1988). Geology, evolution and Metallogenesis of the Pan-
African Belt in Egypt. In S. El Gaby, & R. O. Greiling (Eds.), The Pan-African belts of northeast
Africa and Adjacent areas (pp. 17–68). Friedr Vieweg and Sohn.
El Gaby, S., El-Nady, O., & Khudeir, A. (1984). Tectonic evolution of the basement complex in the
Central Eastern Desert of Egypt. Geologische Rundschau, 73, 1019–1036.
El Ramly, M. F., Greiling, R. O., Kroener, A., & Rashwan, A. A. (1984). On the tectonic evolution
of the Wadi Hafafit area and environs, Eastern Desert of Egypt. Faculty Earth Science Bulletin,
King Abdulaziz University, Jeddah, 6, 113–126.
Fowler, T. J., & Osman, A. F. (2001). Gneiss-cored interference dome associated with two phases of
late Pan-African thrusting in the central Eastern Desert Egypt. Precambrian Research, 108(1–2),
17–43.
Fowler, A., Hassen, I. S., & Osman, A. F. (2010). Neoproterozoic structural evolution of SE Sinai,
Egypt: II. Convergent tectonic history of the continental arc Kid Group. Journal of African Earth
Sciences, 58, 526–546.
Fritz, H., Wallbrecher, E., Khudeir, A. A., Abu El Ela, F. F., & Dallmeyer, D. R. (1996). Formation
of Neoproterozoic metamorphic core complexes during oblique convergence (Eastern Desert,
Egypt). Journal of African Earth Sciences, 23, 311–329.
Fritz, H., Dallmeyer, D. R., Wallbrecher, E., Loizenbauer, J., Hoinkes, G., Neumayr, P., et al. (2002).
Neoproterozoic tectonothermal evolution of the Central Eastern Desert, Egypt: a slow velocity
tectonic process of core complex exhumation. Journal of African Earth Sciences, 34, 137–155.
Fritz, H., Abdelsalam, M., Ali, K. A., Bingen, B., Collins, A. S., Fowler, A. R., et al. (2013). Orogen
styles in the East African Orogen: a review of the Neoproterozoic to Cambrian tectonic evolution.
Journal of African Earth Sciences, 86, 65–106.
Greiling, R. O., Kroener, A., El Ramely, M. F., & Rashwan. A. A. (1988). Structural relationships
between the southern and central parts of the Eastern Desert of Egypt: Details of a fold and
thrust belt. In S. El Gaby, & R. O. Greiling (Eds.), The Pan-African belts of northeast Africa and
Adjacent areas (pp. 121–146). Friedr Vieweg and Sohn.
Greiling, R. O., Kroener, A., & El Ramly, M. F. 1984. Structural interference patterns and their
origin in the Pan-African basement of the south Eastern Desert of Egypt. In A. Kroener,
& R. O. Greiling (Eds.), Precambrian tectonics illustrated (pp 401–412). Schweizerbartsche
Verlagsbuchhandlung.
Habib, M. E., Ahmed, A. A., & El-Nady, O. M. (1985). Two orogenies in the Meatiq area of the
CED. Egyptian Precambrian Research, 30, 83–111.
Hafez, A. M. A., & El-Amin, H. (1983). Structure and metamorphism of a Precambrian sequence,
Wadi El Miyah, Barramiya area, Eastern Desert Egypt. International Basement Tectonics
Association, 5, 105–114.
Hafez, A. H., Abdel-Wahed, M., & Shallaly, N. A. (2006). Microfabric, geochemistry and clock-
wise P-T path of the Precambrian Metasediments in the Central Wadi Kid area, Southeastern
Sinai. In. Abstract 8th International Conference of the Geology of the Arab World (pp. 88–89).
Cairo, Egypt.
Hamimi, Z., El-Amawy, M. A., & Wetait, M. (1994). Geology and structural evolution of El-Shalul
dome and environs, Central Eastern Desert Egypt. Egyptian Journal of Geology, 38, 575–595.
Hamimi, Z., Abd El-Wahed, M., Gahlan, H. A., & Kamh, S. Z. (2019). Tectonics of the Eastern
Desert of Egypt: key to understanding the Neoproterozoic evolution of the Arabian-Nubian Shield
Low Baric Metamorphic Belts in the Northern Tip … 301

(East African Orogen). In A. Bendaoud, Z. Hamimi, M. Hamoudi, S. Djemai, & B. Zoheir (Eds.),
Geology of the Arab world—an overview, Springer Geology (pp. 1–81). Springer Geology. https://
doi.org/10.1007/978-3-319-96794-3_1.
Jamtveit, B., Bucher-Nurminen, K., & Stijfhoorn, D. E. (1992a). Contact metamorphism of layered
shale-carbonate sequences in the Oslo rift: I. Buffering, infiltration and the mechanisms of mass
transport. Journal of Petrology, 33, 377–422.
Jamtveit, B., Grorud, H. F., & Bucher-Nurminen, K. (1992b). Contact metamorphism of layered
shale-carbonate sequences in the Oslo rift: II. Migration of isotopic and reaction fronts around
cooling plutons. Earth and Planetary Science Letters, 114, 131–148.
Jamtveit, B., & Andersen, T. (1993). Contact metamorphism of layered shale-carbonate sequences
in the Oslo rift: III. The nature of skarnforming fluids. Economic Geology, 88, 830–849.
Jamtveit, B., Dahlgren, S., & Austrheim, H. (1997). High-grade contact metamorphism of calcareous
rocks from the Oslo Rift, Southern Norway. American Mineralogist, 82, 1241–1254.
Johnson, P. R., Abdelsalam, M. G., & Stern, R. J. (2003). The Bi’r Umq-Nakasib suture zone in
the Arabian-Nubian Shield: A key to understanding crustal growth in the East African Orogen.
Gondwana Research, 6, 523–530.
Johnson, P. R., Andersen, A., Collins, A. S., Fowler, A. R., Fritz, H., Ghebreab, W., et al. (2011). Late
Cryogenian-Ediacaran history of the Arabian-Nubian Shield: A review of depositional, plutonic,
structural, and tectonic events in the closing stages of the northern East African Orogen. Journal
of African Earth Sciences, 61, 167–232.
Kamal El-Din, G. M., Khudeir, A. A., & Greiling, R. O. (1992). Tectonic evolution of a Pan-African
gneiss culmination, Gabal El-Sibai area. Central Eastern Desert, Egypt. Zbl. Geol. Pala¨ont. Teil
1 H 11, 2637–2640.
Kamal El-Din, G. M., Rashwan, A. A., & Greiling, R. O. (1996). Structure and magmatism in a
molasse-type basin Wadi Hammamat-Wadi Atalla areas. In R. O. Greiling, G. M. Naim, & A.
A. Hussein (Eds.), Excursion Across the Pan-African, Neoproterozoic Basement Qena-Quseir
Guide Book (pp. 11–18). Cairo, Egypt: Egyptian Geological Survey & Min Authority.
Kearey, P., Klepeis, K. A., & Vine, F. J. (2009). Global tectonics (p. 482p). UK: Willy & Blackwell.
Kornprobst, J. (2003). Metamorphic rocks and their geodynamic significance (208 p). Kluwer
Academic Publishers.
Kröner, A., Krueger, J., & Rashwan, A. A. (1994). Age and tectonic setting of granitoid gneisses
in the Eastern Desert of Egypt and southwest Sinai. Geologische Rundschau, 83, 502–513.
Loizenbauer, J., Wallbrecher, E., Fritz, H., Neumayr, P., Khudeir, A. A., & Kloetzli, U. (2001).
Structural geology, single zircon ages and fluid inclusion studies of the Meatiq metamorphic core
complex: implications for Neoproterozoic tectonics in the Eastern Desert of Egypt. Precambrian
Research, 110, 357–383.
Meert, J. G. (2001). A synopsis of events related to the assembly of eastern Gondwana.
Tectonophysics, 363, 1–40.
Messner, M. (1996). Basin analysis of Hammamat type molasse basins, Eastern Desert, Egypt. In
Abstract of the Geological Survey Egypt Center Conference (pp. 125–126). Cairo, Egypt.
Miyashiro, A. (1961). Evolution of metamorphic belts. Journal of Petrology, 2(3), 277–311.
Miyashiro, A. (1973). Metamorphism and metamorphic belts (p. 492p). London: George Allen and
Unwin.
Moghazi, A. M., Ali, K. A., Wilde, S. A., Zhou, Q., Andersen, T., Andresen, A., et al. (2012).
Geochemistry, geochronology, and Sr–Nd isotopes of the Late Neoproterozoic Wadi Kid volcano-
sedimentary rocks, Southern Sinai, Egypt: implications for tectonic setting and crustal evolution.
Lithos, 154, 147–165.
Mohammed, N. A. A. (1999) Petrologic, mineralogic and geochemical studies of the basement
rocks and associated mineralization of El Bakriya area (341 p). Central Eastern Desert, Egypt.
Ph.D. Thesis, Cairo University, Cairo, Egypt.
Naim, G. M., Hussein, A. A., Dardir, A. A., Abdeen, M. M., Greiling, R. O., & Warr, L. N. (1996).
The Wadi Queih Area. Lithology and Structure in a Molasse-type Basin. In R. O. Greiling,
G. M. Naim, A. A. Hussein, A.A. (Eds.), Excursion Across the Pan-African, Neoproterozoic
302 N. A. Shallaly and A. S. A. A. Abu Sharib

Basement Qena-Quseir Guide Book (pp. 65–85). Cairo, Egypt: Egyptian Geological Survey &
Min. Authority.
Navon, O., & Reymer, A. P. S. (1984). Stratigraphy, structure and metamorphism of Pan-African
age in Central Wadi Kid Southeastern Sinai. Israel Journal of Earth Sciences, 33, 135–149.
Neumayr, P., Hoinkes, G., Puhl, J., & Mogessie, A. (1996a). Polymetamorphism in the Meatiq Base-
ment Complex (Eastern Desert, Egypt): P-T variations and implications for tectonic evolution.
In Abstract of the Geology Survey Egypt Center Conference (pp. 139–141). Cairo, Egypt.
Neumayr, P., Hoinkes, G., & Mogessie, A. (1996b). Tectonic implications of different metamorphic
evolution histories in the basement and the cover series in the Meatiq Complex, Central Eastern
Desert, Egypt. In Abstract of the Geology Survey Egypt Center Conference (pp. 156–157). Cairo,
Egypt.
Neumayr, P., Hoinkes, G., Puhl, J., Mogessie, A., & Khudeir, A. A. (1998). The Meatiq dome
(Eastern Desert, Egypt) a Precambrian metamorphic core complex petrological and geological
evidence. Journal of Metamorphic Geology, 16, 259–279.
Rice, A. H. N., Osman, A. F., Abdeen, M. M., Sadek, M. F., & Ragab, A. I. (1993). Preliminary
comparison of six late- to post-Pan-African molasse basins, E. Desert, Egypt. In U. Thorweihe &
H. Schandelmeier (Eds.), Geoscientific Research in Northeast Africa (pp. 41–45). Netherlands:
Balkema, Rotterdam.
Ries, A. C., Shackleton, R. M., Graham, R. H., & Fitches, W. R. (1983). Pan-African structures,
ophiolites and mélanges in the east Desert of Egypt: a traverse at 26° north. Journal Geological
Society London, 140, 75–95.
Reymer, A. P. S., & Yogev, A. (1983). Stratigraphy and tectonic history of the Southern Wadi Kid
metamorphic complex Southeastern Sinai. Israel Journal of Earth Sciences, 32, 105–116.
Sanislav, I. V. (2010). A long lived metamorphic history in the contact aureole of the Mooselookme-
guntic pluton revealed by in-situ dating of monazite grains preserved as inclusions in staurolite
porphyroblasts. Journal of Metamorphic Geology, 29, 251e273.
Sanislav, I. V., & Bell, T. H. (2011). The inter-relationships between long-lived metamorphism,
pluton emplacement and changes in the direction of bulk shortening during orogenesis. Journal
of Metamorphic Geology, 29, 513–536.
Shalaby, A., Stüwe, K., Makroum, F., Fritz, H., Kebede, T., & Klotzli, U. (2005). The Wadi Mubarak
belt, Eastern Desert of Egypt: a Neoproterozoic conjugate shear system in the Arabian-Nubian
Shield. Precambrian Research, 136, 27–50.
Sandiford, M., & Powell, R. (1986). Deep crustal metamorphism during continental extension:
Modern and ancient examples. Earth and Planetary Science Letters, 79(1–2), 151–158.
Shallaly, N. A. (2007). Geology of central Wadi Kid area, southern Sinai, Egypt (298 p). Ph D.
thesis (unpubl.), Cairo University, Giza, Egypt.
Shallaly, N. A. (2019). Metamorphic evolution of Pan-African Wadi El Miyah Metasediments,
Central Eastern Desert, Egypt: A distinctive LP/HT metapelitic sequence from the northern
Arabian-Nubian Shield. Arabian Journal of Geosciences, 12, 54.
Shimron, A. E. (1980). Late phase deformation and mylonite belts in Sinai: Pan African thrust
tectonics. Geological Survey of Israel, Current Research, 75–80.
Shimron, A. E. (1983). The Tarr Complex—revisited: Folding, thrusts and mélanges in the southern
Wadi Kid region, Sinai Peninsula: Israel Journal of Earth Sciences. Israel Journal of Earth
Sciences, 32, 123–148.
Shimron, A. E. (1984a). Evolution of the Kid Group, Southeastern Sinai Peninsula: thrusts, mélange,
and implications for accretionary tectonics during the Late Proterozoic of the Arabian-Nubian
Shield. Geology, 12, 242–247.
Shimron, A. E. (1984b). Metamorphism and tectonics of a Pan-African terrain in southeastern
Sinai-a discussion. Precambrian Research, 24, 189.
Shimron, A. E. (1987). Pan-African metamorphism and deformation in the Wadi Kid region, SE
Sinai Peninsula: evidence from porphyroblasts in the Um Zariq Formation. Isr. J. Earth Sci., 36,
173–193.
Low Baric Metamorphic Belts in the Northern Tip … 303

Solar, G.S., & Brown, M. (1999). The classic high-T e low-P metamorphism of west central Maine,
USA: Is it post-tectonic or syn-tectonic? Evidence from porphyroblast-matrix relations. Canadian
Mineralogist, 37, 311e333.
Solar, G. S., & Brown, M. (2001). Deformation partitioning during transpression in response to
Early Devonian oblique convergence, northern Appalachian orogen, USA. Journal of Structural
Geology, 22, 1043–1065.
Spear, F. S. (1993). Metamorphic phase equilibra and pressure–temperature–time paths. Mineralog-
ical Society of America Monograph Series, 1, 1–799.
Stern, R. J., & Hedge, C. E. (1985). Geochronologic and isotopic constraints on Late Precambrian
crustal evolution in the Eastern Desert of Egypt. American Journal of Science, 285, 97–127.
Stern, R. J. (1994). Arc assembly and continental collision in the Neoproterozoic east African
Orogen: Implication for the consolidation of Gondwanaland. Annual Review Earth Planetary
Science, 2(2), 319–351.
Stern, R. J. (2018). Neoproterozoic formation and evolution of Eastern Desert continental crust—
The importance of the infrastructure-superstructure transition. The Journal of African Earth
Sciences.
Sturchio, N. C., Sultan, M., Sylvester, P., Batiza, R., Hedge, C., El Shazly, E. M., et al. (1983a).
Geology, age, and origin of the Meatiq Dome – implications for the Precambrian stratigraphy and
tectonic evolution of the Eastern Desert of Egypt. King Abdulaziz University, Faculty of Earth
Sciences Bulletin, 6, 127–143.
Sturchio, N. C., Sultan, M., & Batiza, R. (1983b). Geology and origin of Meatiq Dome, Egypt: A
Precambrian metamorphic core complex? Geology, 11, 72–76.
Wickham, S., & Oxburgh, E. (1985). Continental rifts as a setting for regional metamorphism.
Nature, 318, 330–333.
Wickham, S. M., & Oxburgh, E. R. (1987). Low-pressure regional metamorphism in the Pyre-
nees and its implications for the thermal evolution of rifted continental crust. Philosophical
Transactions of the Royal Society of London, 321, 219–242.
Zoheir, B., & Lehmann, B. (2011). Listvenite–lode association at the Barramiya gold mine, Eastern
Desert Egypt. Ore Geology Reviews, 39, 101–115.
Review of the Geometric Model
Parameters of the Main Himalayan
Thrust

Kutubuddin Ansari

Abstract This article reviews geometric model parameters (length, width, depth,
dip angle, and strike) of the Main Himalayan Thrust (MHT) in Eastern, Western,
Central, and other portions of the Himalaya. It is expected that when a fieldworker
works on the Main Frontal Thrust (MFT), these compiled data will be useful in
generating models involving subsurface and surface data sets.

1 Introduction

The Indian plate is the part of southern Asia borders with the Eurasian plate on its
eastern and northern boundaries. The continent–continent collision and the conver-
gence between the Eurasian plate and the Indian plates resulted in the Himalayan
mountain belt with significant topographic variations such as crustal shortening
and large-scale folding and thrusting (Mukherjee et al. 2012, 2015, 2017). The
Himalayan convergent boundary is recognized as wedge-shaped in cross section
and having a continuous arc between 77°E and 89°E longitudes (Bilham et al. 2001;
Mukul et al. 2018). The base of this wedge is defined by a north-dipping decolle-
ment known as the Main Himalayan Thrust (MHT) that separates the down going
Indian plate from the overriding Himalayan wedge. The arc of the Himalaya ranges
>2500 km containing several longitudinal thrust faults/zones into the basal detach-
ment of the MHT (Fig. 1). The Main Frontal Thrust (MFT) marks the beginning of
the Himalayan foreland toward north, lies in the southern boundary of the Himalayan
thrust belt, and coincides with the surface trace of the MHT at the foothills (Jade
et al. 2017). The Main Boundary Thrust (MBT) is the second major fault situated
at the north of the MFT and southern edge of Lesser Himalaya domain. The third
major fault known as Main Central Thrust (MCT) marks at the southern boundary
of the Higher Himalayan domain. The South Tibetan Detachment (STD) outlines
the Tethyan Himalaya domain in the northwest Himalaya. See Bose et al. (2019a,
b) for updated review on timing of activation of these major faults in the Himalaya

K. Ansari (B)
Integrated Geoinformation (IntGeo) Solution Private Limited, New Delhi, 110025, India
e-mail: kdansarix@gmail.com

© Springer Nature Switzerland AG 2021 305


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_10
306 K. Ansari

Fig. 1 Sectional view of the Himalayan wedge (Mukul 2010). The symbol stands like this: SKT
for South Kalijhora Thrust, MCT for Main Central Thrust, RT for Ramgarh Thrust, MBT for
Main Boundary Thrust, and MFT for Main Frontal Thrust. The legends symbols are like this: TSS
for Tethyan Sedimentary Sequence, LHS for Lower Himalayan Sequence, and GHS for Greater
Himalayan Sequence

and Martin (2017) reviewed article for MCT. The northern limit of the Himalaya is
represented by the Indus Suture Zone (ISZ). The ISZ separates Indian and Tibetan
continental blocks which is considered as remnant of the Tethys Ocean. The Trans-
Himalayan domain lies in the north of the ISZ. The litho-tectonic units toward north
bound by these faults are called Sub-Himalaya, Lesser Himalaya, Higher Himalaya,
Tethyan Himalaya, and Trans-Himalaya (Jade et al. 2017). Despite several decades
of research on the Himalayan geology (Mukherjee et al. 2013, 2015), many tectonic
issues remain yet to be solved.
The Global Positioning System (GPS) sites established in the Himalaya contribute
for geoscientific studies in Himalaya. These include tectonic movement, earth-
quakes cycle, and currently kinematic boundary conditions. GPS velocities based on
measurements approved the convergence of the Himalayan Arc up to ~50 mm yr−1 in
NNE direction at the eastern boundary near to the Himalayan Syntaxis. This conver-
gence decreased up to ~35 mm yr−1 at the northern limit of the Indian plate (Demets
et al. 2010). There were several studies conducted by using GPS India-fixed veloci-
ties with respect to pole of rotation confirmed the movement of Indian plate toward
south (e.g., Mukul et al. 2014; Ansari 2018). Mukul et al. (2009) used ITRF05 veloc-
ities of GPS measurements and revealed that movement of Indian plate drops from
~59.7 mm yr−1 near to the Maldives (southern end) to ~32.0 mm yr−1 near to the
Indus–Tsangpo suture. This kind of reduction in Indian plate velocities from south to
north is the cause of compression and accumulation of strain in the Himalayan arc,
occasionally released as the Himalayan earthquakes. Jade et al. (2017) used long-term
GPS measurements and estimated rotation pole of Indian plateau (51.698 ± 0.271°N;
11.853 ± 1.790°E) with an angular velocity of 0.553276 ± 0.005520 Myr−1 . These
modeled angular velocities predicted convergence rates along the Himalayan Frontal
Thrust showed ~14% slower compared to the predicted rate by the NUVEL-1A.
Overall seismicity patterns in the Himalayan arc has been well established from
decades (Ni and Barazangi 1984) where the bulk of the Himalaya has been dominated
by thrust earthquakes (Parameswaran and Rajendran 2017). The entire Himalaya
continuously shows seismic activity due the active thrusting and out-of-sequence
Review of the Geometric Model Parameters of the Main Himalayan … 307

deformation (Mukherjee 2015) of the Indian plate. Cluster analysis of the earth-
quake points out that most of the epicenters occurred between the surface trace of
the MBT and the MCT related to the MHT (Monsalve et al. 2006; Mahesh et al.
2013). The strain energy that has been generated due the collision of the Indian and
the Eurasian plates accumulates in the MHT, cause of several great and major earth-
quakes such as Kangra earthquake in 1905, Bihar–Nepal earthquakes in 1934, and
most recently the Gorkha (Nepal) earthquake with Mw 7.8 in 2015 (Duputel et al.
2016). There were several studies concluded that the presence of ramp structures in
the MHT is responsible for generating earthquakes in the Himalaya (Caldwellet al.
2013; Hazarika et al. 2017). However, no earthquake (Mw > 8.0) has been reported in
the Himalayan region in the recent past (https://www.emsc-csem.org). We presume
that the Gorkha (Nepal) earthquake (Mw 7.8) was the largest earthquake. This recom-
mends that deformation of Himalayan region does not act as an elastic body, but it
behaves like a rigid body (Ansari and Park 2019; Ansari and Bae 2020). There-
fore, it is a very crucial subject to recognize the geometry of the MHT at depth and
its connection with seismicity. The summary of dislocation models reviewed in the
current chapter is depicted in Fig. 2.

Fig. 2 Sketch geological map of the Himalayan orogen (modified after Zhang et al. 2015, Fig. 1).
In this work, the location of surface trace of the modeled MHT is marked with red circles and
red texts including the references of previous publications. The symbol stands like this: WH for
Western Himalaya, CH for Central Himalaya, EH for Eastern Himalaya, WCH for Western and
Central Himalaya, CEH for Central and Eastern Himalaya, WB for Western Bhutan, CB for Central
Bhutan, EB for Eastern Bhutan, WAs for Western Assam, EAs for Eastern Assam, Ar for Arunachal,
WAr for Western Arunachal, MCT for Main Central Thrust, MBT for Main Boundary Thrust, MFT
for Main Frontal Thrust, STD for South Tibet Detachment, and ITS for Indus-Tsangpo Suture Zone
308 K. Ansari

2 Solution of the Fault Geometry

Faults are the fracture zone between the rocks which slides in different direction
(Mukherjee 2015) due to any natural phenomenon such as earthquake, landslide, or
tectonic plate motions. This type of rock movements creates friction force between the
rocks and generates the strain energy. Building of strain could take times of decades
or centuries until their internal rigidity is exceeded. However, at a particular moment,
the strain becomes maximum, the rock suddenly separates with a rupture along the
fault line. At the moment, earthquake occurs and releases accumulated energy. Later,
the rocks start to recover their original shape concentrating a motion along the fault.
This whole phenomenon is called the slow elastic deformation (Fig. 3).
Let us suppose that the GPS sites were located on the surface of rocks, and the
suddenly broken or movement of rock affects the dislocation of sites. By using the
dislocation, we can model several kinds of parameters about the fracture plane such
as length, width, depth of hypocenter, dip angle, slip, and moment magnitude of
earthquakes (Fig. 4). This type of modeling is called dislocation modeling. There
were various theoretical and numerical methods that have been investigated for the

Fig. 3 Slow elastic deformation and rebound. P1 and P2 stands for two surface fault planes on the
Earth where GNSS sites were established
Review of the Geometric Model Parameters of the Main Himalayan … 309

Fig. 4 Finite rectangular fault signal at GPS site located on the surface of the crust

deformation study of an isotropic homogeneous medium and simplification of source


geometry. The derivation of the surface displacements due to point source of inclined
horizontal or vertical fault in a solid was arranged with arbitrary elastic constants.
Analytical solution for surface deformation due to strike, dip (normal and reverse)
faults in a half-space given by Okada (1985) is important in dislocation modeling to
work out the causative plane during earthquake occurrences. Here, we are going to
introduce the summary of this formulation for the beginners, so that they can under-
stand it very well. To revise fault kinematics, refer to few recent papers: Mukherjee
and Tayade (2019) and Mukherjee and Chakraborty (2020).
Consider that the displacement field is ui (x 1 , x 2 , x 3 ) due to dislocation of uj (ξ1 ,
j
ξ2 , ξ3 ) across the surface in an isotropic medium. Let ui is the ith displacement
component at (x 1 , x 2 , x 3 ) due to the point force in jth direction with magnitude F at
 
(ξ 1 , ξ2 , ξ 3 ). If (ξ , η ) is an arbitrary coordinate on a fault surface, then for a finite
rectangular fault with dip angle δ, depth d, length L, and width W, the deformation
  
field can be derived by taking x = x–ξ , y = y–η cos (δ) and d = d–η sin (δ).
By applying the integration equation, the Green function (G) can be given in the
following form (Okada 1985):

L W

G= dξ dη (1)
0 0


where λ and μ are Lame’s constants and δ is dip angle. Let x–ξ / = ξ, p–η = η, then
above equation becomes:
310 K. Ansari

x−L 
p−W

G= dξ dη (2)
x p

The analytical solution for G can be understood from Okada (1985)


Suppose G(m) is a function of finite rectangular fault parameters, the observed
displacement is D, and s is the slip on the fault. In case of thrust fault, the slip will
be only reverse slip, and in case of strike-slip faults, the slip will be only strike slip.
If the fault is oblique, then slip is given by (Ansari et al. 2014):

S = s · cos α + s · sin α (3)

where α is known as the rake of fault and S is total slip. The relationship between
the fault geometry and dislocation field can be given by:

D = sG(m) + ε

ε = D − sG(m) (4)

The Green function G depends on the fault parameters (length, width, depth, dip
angle, and location of GPS site). The surface of rupture fault cannot be rectangular
in natural way, hence the ε represents an error.
In case of the forward modeling (Cervelli et al. 2001), we tried to minimize the
error or ε → 0, then:

D̂ = ŝG(m̂) (5)

where d̂ represents modeled displacement, ŝ is modeled slip, and m̂ is modeled


parameters.
The forward modeling used for the model is based on Coulomb 3.3 (Toda et al.
2010) using the Boundary Element Method (BEM). The inputs given to Coulomb
are estimates of length, width, dip angle, strike-slip, and dip-slip of the modeled fault
plane as well as the coordinates of the trace of the fault plane. An example of the
Coulomb file with inputs and parameters is given in Fig. 5. During the modeling of
the finite rectangular fault, an initial estimate of the fault parameters (length, width,
depth, dip angle, strike, and dip slip of rupture plane) is made, and the model is run.
The results are subsequently compared with the observed displacements of GPS sites.
By varying the selected modeled fault parameters, the difference between observed
and modeled displacement is made as close to zero as possible. The best possible
solution of finite rectangular fault corresponds to the fault parameters that simulate
the measured displacements best. Figure 6 shows an example involving three different
faults/dislocations and the modeled parameter of the finite rectangular fault number.
Review of the Geometric Model Parameters of the Main Himalayan … 311

Fig. 5 Coulomb input file (Toda et al. 2010)

3 Geometry of MHT

3.1 Background

Some geophysical studies by using active as well as passive source seismic experi-
ments were conducted to investigate the geometry of the MHT in different parts of
the Himalaya. A recent study suggested that the significant lateral variation in the
geometry of the MHT in the west and east of the Bhutan Himalaya (Singer et al.
2017) emphasizing the view that subsurface structure in the Himalaya laterally varies
all along the 2500 km long Himalayan arc (Kayal 2001).

3.2 Nepal Himalaya (Single Dislocation)

Larson et al. (1999) solved dislocation model for the best-fitting rectangular plane
to signify the seismic fault slip at the north of the tip line of the locked system. The
length and width of the fault plane were taken to be 1500 and 1000 km, respectively.
They inverted the fault parameters by using GPS and leveling data and computed the
dislocation with 106.4° strike and 4.4° dip toward north. The computed dip-slip and
strike-slip rate component were ~19 mm yr−1 and ~4 mm yr−1 , respectively, at the
southern edge of the Himalaya at 18.3 km depth. In second attempt, Larson et al.
(1999) used only GPS data and found change parameters of finite rectangular fault.
312 K. Ansari

Fig. 6 Coulomb 3.3 working templates. The measured GPS velocity field is shown on the figure
to the right. The green lines are the surface traces of the faults, and red lines are the projection of
dipping faults on horizontal surface. The initial fault parameters chosen for the model is shown on
the box to the left

These newly computed dip-slip and strike-slip rate components were ~20 mm yr−1
and ~4 mm yr−1 , respectively, at 18.7 km depth. In next attempt, Larson et al. (1999)
used only the leveling data and computed the fault parameter by fixing the strike of
106.1° with a dip-slip rate ~15 mm yr−1 at ~14.0 km depth. These results indicate
that leveling data estimate a slower and shallower source of deformation. These kind
of difference in the results may be possible due the systematic errors in the leveling
data. Table 1 presents the results of Larson et al. (1999).
Boucher et al. (2004) reported that the slip rate of MHT ~19 mm/yr and a shallow
dip 5° from Nepal. According to their model, the MHT is locked at ~17 km, which was
almost consistent with later studies (Table 2). Ader et al. (2012) used 24 continuous
GPS site network of Nepal and inverted the convergence slip rate of 18–20 mm yr−1 .
The modeled fault was locked from surface to a depth of 15–20 km over a width of
100 km with a dip of 10° (Table 2). Ansari (2018) used vertical displacements of
GPS sites located in Nepal and modeled them by Auto-regressive Moving Average
(ARMA) method (Fig. 7). They used ARMA modeled vertical displacements and
Review of the Geometric Model Parameters of the Main Himalayan … 313

Table 1 These results are presented by Larson et al. (1999) for the best fitting of single-dislocation
model of MHT
Data Depth Dip (°) Strike (°) Dip slip (mm yr−1 ) Strike slip Model misfit
(mm yr−1 )
G+L 18.3 4.4 106.4 19.1 ± 1.0 3.6 ± 1.1 1.07
G 18.7 7.7 106.1 20.1 ± 1.1 3.5 ± 1.2 1.23
L 14.0 14.8 106.1 15.4 ± 1.6 – 0.93
They used GPS (G) and leveling (L) data during the modeling. The symbol from the table stands
like this: G for GPS, L = Leveling and G + L for both GPS and leveling data

Table 2 Numerical dislocation model of MHT presented by Ansari (2018) by using vertical
velocities field
Region Length (km) Width (km) Depth (km) Dip (°) Dip slip Model
(mm yr−1 ) reference
Nepal 700 180 20.0 9.5 19 Ansari (2018)
Himalaya
Nepal – 100 15–20 10 18–20 Ader et al.
Himalaya (2012)
Nepal – – 17 5 19 Boucher et al.
Himalaya (2004)
List of two other studies which were carried out in Nepal Himalaya

Fig. 7 The map-view geometry of the single-dislocation model used to simulate measured vertical
velocities in Nepal (Ansari 2018)
314 K. Ansari

computed the fault parameter of single dislocation of MHT. The estimated single-
dislocation model reverse slip rate was 19 m yr−1 , located at a depth of 20 km with dip
of 9.5°. The calculated length of finite rectangular was 700 km with 277° strike. These
modeled parameters were consistent with the earlier published modeled parameter
estimated by Jouanne et al. (2004) from the Central Himalaya.

3.3 Central Nepal Himalaya

The existence of MHT segmentation by using dislocation model was tested by


Jouanne et al. (2004) in Central Nepal Himalaya (Table 3). They estimated that the
dislocations model in Central Nepal is extending down-dip from locked zone bottom
to a lower edge. The dislocation strike was presumed to parallel the normal structural
direction of the other faults like the MBT and the MFT and as well as to the direction
of the micro seismicity clusters (between 82.5°E and 86.5°E) is orientated along
108°NE. The best dislocation solutions were found with dip angle of 9–10°, a thrust
component of 19–20 mm yr−1 , a dextral strike-slip component of 0–2 mm yr−1 , and
a locking depth ~17–21 km (Table 3). In the same year 2004, the Himalayan Arc in
Central Himalaya was modeled by Chen et al. (2004). They determined the geometry
of the central fault and found that the Indian tectonic plate is locked ~80 km far from
Kathmandu in north with ~102° strike. The locked zone width was ~100 km, similar
to the width estimated by Banerjee and Burgmann (2002) from the NW Himalaya.
Chen et al. (2004) suggested a shallower dip ~3° and a slip rate of 12.2 ± 0.4 mm yr−1
in Central Himalaya between the longitudes of 83°E to 88°E. Bettinelli et al. (2006)
used GPS and DORIS measurements in Central Nepal Himalaya and reported MHT
modeled mean rate of convergence around 19 mm yr−1 . They tried to model the
rate of convergence by taking geodetic data including its horizontal and vertical
measurements separately. The combined geodetic data yield a slip rate of 16.2 ±
0.8 mm yr−1 while separate horizontal and vertical geodetic measurements yield a
slip rate of 16.3 ± 0.7 mm yr−1 and 13.4 ± 5 mm yr−1 , respectively. Continuing the
same discussion, Bettinelli et al. (2006) assessed to fit to the horizontal data and the
levelling data separately. The analysis shows that the best-fitting convergence rate
lies between 16 and 19 mm yr−1 depending on the data set which is considered. The
best-fitting rate of shortening is modeled to be 19 ± 2.5 mm yr−1 when all data was
combined. In the other study which was done by Ponraj et al. (2011) by using three
year of campaign, measurements across beneath the Central Himalaya estimated a
slip of 10 mm yr−1 . They suggested that the MHT to be locked with the depth of
20 km from the surface in this area. In a separate study, Cattin and Avouac (2000)
suggested a dip-slip rate of MHT of ~20 mm yr−1 with strike 108° (depth is not given
model) in the Central Himalaya.
Review of the Geometric Model Parameters of the Main Himalayan … 315

Table 3 Results from this table are mainly presented by Bettinelli et al. (2006) with other studies
Data Region Depth Dip (°) Strike Dip slip Model
(mm yr−1 ) reference
GPS Np 20 ± 4 4±4 N131 20.5 ± 2 Bilham et al.
(1997)
GPS + L NpW 25 4.5 N120 21.3 ± 1.6 Larson et al.
NpE 14.9 3.4 N101 19.6 ± 1.1 (1999)
GPS + L NpC – – N108 20 Cattin and
Avouac.
(2000)
GPS InW 15 6 N133 14 ± 1 Banerjee and
Burgmann
(2002)
GPS NpC 17–21 9–10 N108 19–20 Jouanne et al.
NpW 20–21 9–10 N117 19 (2004)
GPS NpC 14.3 3° N102 12.4 ± 0.4 Chen et al.
NpW 18.3 9° N115 17 ± 1 (2004)
NpE 20.3 3° N84 19 ± 1
GPS in 2D NpW 12 4.5 N120 13.4 ± 5 Bettinelli et al.
GPS + NpE + NpC 20 10 N113 16.3 ± 0.7 (2006)
L+cGPS
GPS in 3D NpW 12 4.5 Variable 13.4
GPS + NpE + NpC 20 10 Variable 19 ± 2.5
L+cGPS
GPS NpC 20 – – 10 Ponraj et al.
(2011)
GPS NpE 19 ± 2 7 ± 2° – 18 ± 1 Vernant et al.
(2014)
Bettinelli et al. (2006) used campaign mode and continuous GPS observation data in his study.
The symbols from the table stand like this: GPS for campaign mode GPS observation; cGPS
for continuous GPS observation; L for vertical velocities determined from repeated leveling
measurements; Np for whole Nepal; NpW for Western Nepal; NpC for Central Nepal; NpE for
Eastern Nepal; InW for Western India, 2D for two-dimensional data, and 3D for three-dimensional
data

3.4 Western Nepal Himalaya

The segment of the Western Nepal Himalaya is modeled differently from the Eastern
Himalaya by several authors, which is situated in the hanging wall of the 1934 Bihar–
Nepal earthquake rupture. They assumed that the slip rate in the Western and the
Eastern Himalaya could differ due to the different stages of the regional earthquake
cycle or may be arise from the structural segmentation of the fault system mentioned
in (Table 3). Larson et al. (1999) modeled the western fault segment parameters and
suggested ≥25 km depth with dip-slip of 21.3 ± 1.6 mm yr−1 . The model fault plane
316 K. Ansari

rotates into approximate parallelism with the Himalayan arc striking 120° and the
western faults dip of 4.5° to the north. Bilham et al. (1997) advised a deeper locking
depth in Western Nepal compared to the Eastern Nepal with the locking depth of 20
± 4 km along the Himalayan Arc. The dislocation dipping was varying from around
4° in the seismogenic portion to around 9° northward. Chen et al. (2004) determined
the geometry of the western fault along the strike of 115° and found the slip rate 17
± 1 mm yr−1 . Their models locking depth was around 18–20 km and steeper dip
angle of around 9°, consistent with previously estimated observations (e.g., Bilham
et al. 1997). The existence of MHT segmentation by using dislocation model is tested
by Jouanne et al. (2004) in the Western Nepal Himalaya. The dislocation strike was
expected by them to be parallel of the normal structural direction of the microseis-
micity clusters along a 117°NE. The best dislocation solution was found with dip
angle of 9–10°, thrust component of 19 mm yr−1 , dextral strike-slip component of
0–1 mm yr−1 , and a locking depth of ~20–21 km. Jouanne et al. (2004) obtained the
existence of 2–3 mm yr−1 misfit for the Siwaliks points and advised an aseismic creep
localized in space. This hypothesis would suggest that aseismic creep evidenced from
comparison of spirit levelling in the external part of Himalayas (Jackson and Bilham
1994; Gahalaut and Chander 1997) could be a very local phenomenon and does not
give an accurate estimate of an aseismic component on the scale of the Himalayan
belt (Jouanne et al. 2004). Bettinelli et al. (2006) projected their model by using
GPS and DORIS measurements, normal to the mean azimuth of the microseismic
midcrustal cluster in Western Himalaya along 120°N striking section. They reported
MHT modeled slip rate of 13.4 ± 5 mm yr−1 with dip angle of 4.5° and depth of
12 km.

3.5 Eastern Nepal Himalaya

Larson (1999) modeled the segment of Eastern Nepal Himalaya and reported the slip
rate of 19.6 ± 1.1 mm yr−1 with a 3.4° dip to the north (Table 3). The estimating
result showed that the fault was located at 14.9 km depth. Later on, in 2004, the
dislocation model of Western Himalaya was presented by Chen et al. (2004). They
accepted that their results are poorly determined due to lack of data. They suggested
that the modeled slip rate of 19 ± 1 mm yr−1 for India beneath Tibet in the Eastern
Himalaya with 84° strike. Although their modeled dip angle was shallower, the slip
rate was lower and locked zone was slightly wider compared to the model presented
by Vergne et al. (2001), but their results were consistent with earlier estimations
from the GPS measurements based on more limited data (Bilham et al. 1997; Larson
et al. 1999). Bettinelli et al. (2006) did dislocation modeling by taking GPS and
DORIS measurements. Their modeled mean rate of convergence was common with
Central Himalaya, which we have already discussed in the section of Central Nepal
Himalaya. Vernant et al. (2014) estimated convergence rate of Eastern Himalaya
including the other part of Himalaya. They noticed that the Eastern Himalaya slip
rate was 18 ± 1 mm yr−1 with dip of 7 ± 2°. They suggested that in Eastern Nepal,
Review of the Geometric Model Parameters of the Main Himalayan … 317

the MHT was locked over a width of 100 ± 10 km from the surface to a depth of 19
± 2 km.

3.6 Other Portions of the Himalaya

There were several authors who collected GPS data from other parts of the Himalaya
and estimated the dislocation modeled parameters of the MHT. They compared their
result to the previous published model parameters and gave important suggestions.
Mukul et al. (2018) used long-term GPS measurement (1997–2012) from Darjeeling–
Sikkim (India) and reported the model parameters of the MHT. They also estimated
the missing parameters of pervious published result and presented it in very simple
form (Table 4, Fig. 8). They chose the strike of MHT parallel to Dharan salient
(~160 km) between the longitudes 87.3°E and 89°E. The selected width of the dislo-
cation was infinite with unconstrained bottom depth because it was most realistic
(e.g., Feldl and Bilham 2006). Mukul et al. (2018) modeled the dislocation locking
depth of 16 ± 0.1 km with a reverse slip of 18 ± 0.1 mm yr−1 . Their model reported

Table 4 These numerical of studies of MHT has been investigated by Mukul et al. (2018) in Sikkim
Himalaya
Region Code Length Width Depth Dip Reverse Strike slip Model
(km) (km) (km) angle (°) slip (mm yr−1 ) reference
(mm yr−1 )
WCH A 160 ∞ 20.0 10.0 20 1 Jouanne
et al.
(2004)
WH B 400 150.0 12.0 4.5 18 0 Bettinelli
CEH C 500 115.2 20.0 10.0 20 0 et al.
(2006)
WH D 800 ∞ 18.3 9.5 17 2 Chen
CH E 485 ∞ 14.3 2.5 13 2 et al.
(2004)
EH F 660 ∞ 20.3 3.1 20 4
Sikkim G 200 105.0 20 9 18 ± 1.5 0 Jade et al.
(2014)
Sikkim H 160 ∞ 23 ± 9 5±2 17 ± 1 3±1 Vernant
et al.
(2014)
Sikkim I 160 100 16 ± 0.1 6 ± 0.2 18 ± 0.1 0 Mukul
et al.
(2018)
They remodeled the dislocation models given by previous studies and presented the missing
parameters. The symbol stands like this: WH for Western Himalaya; CH for Central Himalaya;
EH for Eastern Himalaya; WCH for Western and Central Himalaya; ECH for Eastern and Central
Himalaya, and Sikkim for Sikkim Himalaya or Northeast Himalaya
318 K. Ansari

Fig. 8 The map-view geometry of the single-dislocation models (Table 4) used to model measured
convergence velocities in the Himalaya. The modeled references related of A-I has been given in
Table 4. The models have a locking depth of ~12–20 km below the surface trace of the Main Central
thrust (Mukul et al. 2018)

that the frontal around100 km width of the MHT was locked over the 160 km length
of the Dharan salient between 27.76°N, 87.36°E and 27.61°N, 88.99°E. Jade el
(2014) and Vernet et al. (2014) also took GPS measurements in Sikkim Himalaya
and reported some varying parameters. According to Jade et al. (2014), the MHT
was locked at a depth of 20 km, and width of 105 km with dip of 9°. They reported
the reverse slip rate 18 ± 1.5 mm yr−1 (Table 4). Vernant et al. (2014) estimated
(27.46°N; 87.2–88.8°E) that the MHT was locked at 23 ± 9 km depth, with a dip of
5 ± 2, º and had an associated reverse slip rate 17 ± 1 mm yr−1 (Table 4).
Mukul et al. (2014) reported that the slip rate of 15–20 mm yr−1 was accom-
modating in the Northeast Himalayan wedge based on their campaigns mode GPS
measurement from 1997 to 2006. Jade et al. (2014) used GPS derived measure-
ments from 1997 to 2008 and presented best fit dislocation model in Kumaon and
Garhwal Himalaya and modeled the slip rate (Table 5). They reported in their Kumaon
Himalaya dislocation model that the MHT which coincides with the MFT locked
from the surface to a depth of 20 km over a width of 110 km with allied dip-slip rate
of 18 ± 1.5 mm yr−1 and strike-slip of 1 ± 1.5 mm yr−1 , while in Garhwal Himalaya
the MHT was found to be locked to a depth of 16 km over a width of 109 km with
allied dip-slip rate of 16 ± 1 mm yr−1 and strike-slip of 1 ± 1 mm yr−1 . In the same
discussion, they estimated the dislocation model in western Arunachal and reported
that MHT was locked from the surface to be a depth of 17 km over a width of 130 km
with allied dip-slip rate of 16 ± 2 mm yr−1 and strike-slip of zero mm yr−1 . Jade
et al. (2014) concluded that the modeled velocities are either low or high compared
to the observed velocities of GPS sites that means additional complex models are
needed to model the slip rates in these regions.
Vernet et al. (2014) collected GPS data from Bhutan and estimated slip rate in
East, West, and Central Bhutan with the average slip rate of 14–17 mm yr−1 . They
Review of the Geometric Model Parameters of the Main Himalayan … 319

Table 5 Dislocation modeling of MHT has been presented in several parts of Himalaya by Jade
et al. (2014) and Vernant et al. (2014)
Region of Length (km) Width Depth Dip (°) Dip slip Strike slip Model
Himalaya (km) (km) (mm yr−1 ) (mm yr−1 ) reference
Kumaon 200 110 20 8 18 ± 1.5 1 ± 1.5 Jade et al.
Garhwal 160 109 16 5 16 ± 1 1±1 (2014)
Western 250 130 17 10 16 ± 2 0
Arunachal
Western 89.0–90.0 °E 98 ± 23 ± 5 7±2 16.5 ± 1.5 2±2 Vernant
Bhutan 10 et al.
Central 90.4–91.4°E 80 ± 23 ± 4 7±2 15 ± 1.5 4±2 (2014)
Bhutan 10
Eastern 91.5–93.0°E 60 ± 14 ± 5 7±5 17.0 ± 1 –
Bhutan 10
Western 92.0–94.0°E >120 ± 24 7±3 14 ± 1 –
Assam
Arunachal 94.0–95.0°E – – 11.5 ± 1 4±1
Eastern >95°E ± 10 – 14 ± 2 –
Assam
They studied mostly campaign mode GPS network and compared their outcomes with other existing
studies

reported varying locking depth and dip angle in East, West, and Central Bhutan,
which can be seen clearly from Table 5. On the basis of these results, they concluded
that Bhutan is not immune from great earthquakes. They had few GPS points in the
Central Arunachal Pradesh segment of the Himalaya, and none of them was able
to suitably determine the locking depth or its accurate position. By taking the two
locations from Tibet to the northwest of the locking line, they concluded that a slip
rate of 11.5 ± 1 mm yr−1 prevails with a sinistral shear velocity of 4 ± 1 mm yr−1 .
They noted that the great width of the locked decollement in Arunachal Pradesh is
attended by an associated shortening in the average accretionary slope from 2.2°
close to the Bhutan and 1.5° farther to the east. The estimated rate of slip of 14 ±
2 mm yr−1 (27.39°N; 91.5–93°E) with locking depth of ± 10 km in Eastern Assam
was little bit higher than Arunachal slip rate of 11.5 ± 1 mm yr−1 (Latitude was
missing; 94–95°E).

4 Summary

The dislocation modeling of the Main Himalayan Thrust (MHT) has been reviewed
in Nepal Himalaya including Central, Western, Eastern, and other portion of the
Himalaya.
320 K. Ansari

• The dislocation parameter in Nepal Himalaya concluded that the MHT was locked
from the surface to 15–20 km depth over a ~100 km width with allied dip-slip
rate of 18–20 mm yr−1 and no strike-slip component.
• In Central Himalaya, the MHT was estimated to be locked at a depth of 14–20 km
containing width of about 115 km with related to dip-slip rate of about 20 mm yr−1
and strike-slip of 1–2 mm yr−1 .
• The estimated parameters of Western Himalaya are reported a depth of about
12–18 km, associated width of 150 km together a dip-slip rate of 17–18 mm yr−1
and strike-slip of 1–2 mm yr−1 .
• The Eastern Himalaya the MHT was found to be locked a surface depth of 20 km,
width of about 115 km, and the dip-slip and strike-slip rate of 20 mm yr−1 and
2 mm yr−1 , respectively.
• The Sikkim Himalaya showed a locking surface depth MHT of 16–23 km
and ~105 km width. The slip rate of Sikkim Himalaya was reported to be
17–19 mm yr−1 as dip-slip and ~3 mm yr−1 as strike-slip.

Acknowledgements Soumyajit Mukherjee (IIT Bombay) invited, handled, and reviewed this
article. Thanks to Marion Schneider, Annett Buettener, Boopalan Renu, Alexis Vizcaino, Doerthe
Mennecke-Buehler, and the proofreading team (Springer). Dutta and Mukherjee (2021) had
summarized this work.

References

Ader, T., Avouac, J. P., Liu-Zeng, J., Lyon-Caen, H., Bollinger, L., Galetzka, J., et al. (2012).
Convergence rate across the Nepal Himalaya and interseismic coupling on the Main Himalayan
Thrust: Implications for seismic hazard. Journal of Geophysical Research: Solid Earth, 117(B4),
https://doi.org/10.1029/2011JB009071.
Ansari, K., Mukul, M., & Jade, S. (2014). Dislocation Modelling of the 1997–2009 High-Precision
Global Positioning System Displacements in Darjiling-Sikkim Himalaya, India. In International
Conference on Earth Science and Climate Change (ICESCC 2014) (Vol. 8, No. 10, pp. 609-612).
Dubai, United Arab Emirates.
Ansari, K. (2018). Crustal deformation and strain analysis in Nepal from GPS time-series measure-
ment and modeling by ARMA method. International Journal of Earth Sciences, 107(8),
2895–2905. https://doi.org/10.1007/s00531-018-1633-7.
Ansari, K., & Park, K. D. (2019). Contemporary deformation and seismicity analysis in Southwest
Japan during 2010–2018 based on GNSS measurements. International Journal of Earth Sciences,
108(7), 2373–2390. https://doi.org/10.1007/s00531-019-01768-w.
Ansari, K., & Bae, T. S. (2020). Contemporary Deformation and Strain analysis in South Korea
based on long-term (2000–2018) GNSS measurements. International Journal of Earth Sciences,
109(1), 391–405. https://doi.org/10.1007/s00531-019-01809-4.
Banerjee, P., & Bürgmann, R. (2002). Convergence across the northwest Himalaya from GPS
measurements. Geophysical Research Letters, 29(13), 30–1. https://doi.org/10.1029/2002GL
015184.
Bettinelli, P., Avouac, J. P., Flouzat, M., Jouanne, F., Bollinger, L., Willis, P., & Chitrakar, G.
R. (2006). Plate motion of India and interseismic strain in the Nepal Himalaya from GPS and
Review of the Geometric Model Parameters of the Main Himalayan … 321

DORIS measurements. Journal of Geodesy, 80(8–11), 567–589. https://doi.org/10.1007/s00190-


006-0030-3.
Bose, N., & Mukherjee, S. (2019a). Field documentation and genesis of the back-structures from the
Garhwal Lesser Himalaya, Uttarakhand, India. Geological Society, London, Special Publications,
481(1), 111–125. https://doi.org/10.1144/SP481-2018-81.
Bose, N., & Mukherjee, S. (2019b). Field documentation and genesis of back-structures in ductile
and brittle regimes from the foreland part of a collisional orogen: examples from the Darjeeling–
Sikkim Lesser Himalaya, India. International Journal of Earth Sciences, 108(4), 1333–1350.
https://doi.org/10.1007/s00531-019-01709-7.
Boucher, C., Altamimi, Z., Sillard, P., & Feissel-Vernier, M. (2004) The international terrestrial
reference frame (ITRF2000). IERS Tech Note 31 Verlag des Bundesamts fur Kartographie und
Geodasie, Frankfurtam Main
Bilham, R., Gaur, V. K., & Molnar, P. (2001). Himalayan seismic hazard. Science, 293(5534),
1442–1444. https://doi.org/10.1126/science.1062584.
Bilham, R., Larson, K., & Freymueller, J. (1997). GPS measurements of present-day convergence
across the Nepal Himalaya. Nature, 386(6620), 61. https://doi.org/10.1038/386061a0.
Caldwell, W. B., Klemperer, S. L., Lawrence, J. F., & Rai, S. S. (2013). Characterizing the Main
Himalayan Thrust in the Garhwal Himalaya, India with receiver function CCP stacking. Earth
and Planetary Science Letters, 367, 15–27. https://doi.org/10.1016/j.epsl.2013.02.009.
Cattin, R., & Avouac, J. P. (2000). Modeling mountain building and the seismic cycle in the Himalaya
of Nepal. Journal of Geophysical Research: Solid Earth, 105(B6), 13389–13407. https://doi.org/
10.1029/2000JB900032.
Cervelli, P., Murray, M. H., Segall, P., Aoki, Y., & Kato, T. (2001). Estimating source parameters
from deformation data, with an application to the March 1997 earthquake swarm off the Izu
Peninsula, Japan. Journal of Geophysical Research: Solid Earth, 106(B6), 11217–1123.; https://
doi.org/10.1029/2000JB900399.
Chen, Q., Freymueller, J. T., Wang, Q., Yang, Z., Xu, C., & Liu, J. (2004). A deforming block model
for the present-day tectonics of Tibet. Journal of Geophysical Research: Solid Earth, 109(B1).
https://doi.org/10.1029/2002JB002151.
DeMets, C., Gordon, R.G., & Argus, D.F. (2010). Geologically current plate motions. Geophysical
Journal International, 181(1), 1–80; https://doi.org/10.1111/j.1365-246X.2009.04491.x.
Duputel, Z., Vergne, J., Rivera, L., Wittlinger, G., Farra, V., & Hetényi, G. (2016). The 2015 Gorkha
earthquake: a large event illuminating the Main Himalayan Thrust fault. Geophysical Research
Letters, 43(6), 2517–2525. https://doi.org/10.1002/2016GL068083.
Dutta, D., & Mukherjee, S. (2021). Introduction to Structural Geology and Tectonics Field Guide-
book—Volume 1. In S. Mukherjee S. (Ed.), Structural Geology and Tectonics Field Guidebook—
Volume 1. Springer Nature Switzerland AG. Cham. pp. xi-xvi. ISBN: 978-3-030-60142-3.
Feldl, N., & Bilham, R. (2006). Great Himalayan earthquakes and the Tibetan plateau. Nature,
444(7116), 165. https://doi.org/10.1038/nature05199.
Gahalaut, V. K., & Chander, R. (1997). Evidence for an earthquake cycle in NW Outer Himalaya
near 78 E longitude from precision levelling observations. Geophysical Research Letters, 24(3),
225–228. https://doi.org/10.1029/97GL00032.
Hazarika, D., Wadhawan, M., Paul, A., Kumar, N., & Borah, K. (2017). Geometry of the Main
Himalayan Thrust and Moho beneath Satluj valley, northwest Himalaya: Constraints from receiver
function analysis. Journal of Geophysical Research: Solid Earth, 122(4), 2929–2945. https://doi.
org/10.1002/2016JB013783.
Jade, S., Mukul, M., Gaur, V. K., Kumar, K., Shrungeshwar, T. S., Satyal, G. S., et al. (2014).
Contemporary deformation in the Kashmir–Himachal, Garhwal and Kumaon Himalaya: signifi-
cant insights from 1995–2008 GPS time series. Journal of Geodesy, 88(6), 539–557. https://doi.
org/10.1007/s00190-014-0702-3.
Jade, S., Shrungeshwara, T. S., Kumar, K., Choudhury, P., Dumka, R. K., & Bhu, H. (2017). India
plate angular velocity and contemporary deformation rates from continuous GPS measurements
from 1996 to 2015. Scientific Reports, 7(1), 11439. https://doi.org/10.1038/s41598-017-11697-w.
322 K. Ansari

Jackson, M., & Bilham, R. (1994). Constraints on Himalayan deformation inferred from vertical
velocity fields in Nepal and Tibet. Journal of Geophysical Research: Solid Earth, 99(B7), 13897–
13912. https://doi.org/10.1029/94JB00714.
Jouanne, F., Mugnier, J. L., Gamond, J. F., Le Fort, P., Pandey, M. R., Bollinger, L., et al. (2004).
Current shortening across the Himalayas of Nepal. Geophysical Journal International, 157(1),
1–14. https://doi.org/10.1111/j.1365-246X.2004.02180.x.
Kayal, J. R. (2001). Microearthquake activity in some parts of the Himalaya and the tectonic model.
Tectonophysics, 339(3–4), 331–351. https://doi.org/10.1016/S0040-1951(01)00129-9.
Larson, K., Burgmann, R., Bilham, R., & Freymuller, J. (1999). Kinematics of the India-Eurasia
collision zone from GPS measurements. Journal Geophysical Research, 104, 1077–1093. https://
doi.org/10.1029/1998JB900043.
Mahesh, P., Rai, S. S., Sivaram, K., Paul, A., Gupta, S., Sarma, R., & Gaur, V. K. (2013). One–
dimensional reference velocity model and precise locations of earthquake hypocenters in the
Kumaon–Garhwal Himalaya. Bulletin of the Seismological Society of America, 103(1), 328–339.
https://doi.org/10.1785/0120110328.
Martin, A. J. (2017). A review of definitions of the Himalayan Main Central Thrust. International
Journal of Earth Sciences, 106(6), 2131–2145. https://doi.org/10.1007/s00531-016-1419-8.
Monsalve, G., Sheehan, A., Schulte-Pelkum, V., Rajaure, S., Pandey, M. R., & Wu, F. (2006).
Seismicity and one-dimensional velocity structure of the Himalayan collision zone: Earthquakes
in the crust and upper mantle. Journal of Geophysical Research: Solid Earth, 111(B10). https://
doi.org/10.1029/2005JB004062.
Mukherjee, S., Koyi, H. A., & Talbot, C. J. (2012). Implications of channel flow analogue models
for extrusion of the Higher Himalayan Shear Zone with special reference to the out-of-sequence
thrusting. International Journal of Earth Sciences, 101(1), 253–272. https://doi.org/10.1007/s00
531-011-0650-6.
Mukherjee, S., Mukherjee, B. K., & Thiede, R. C. (2013). Geosciences of the Himalaya–
Karakoram–Tibet orogen. https://doi.org/10.1007/s00531-013-0934-0.
Mukherjee, S. (2015). A review on out-of-sequence deformation in the Himalaya. Geological
Society, London, Special Publications, 412(1), 67–109. https://doi.org/10.1144/SP412.13.
Mukherjee, S., Carosi, R., van der Beek, P., Mukherjee, B. K., & Robinson, D. M. (2015). Tectonics
of the Himalaya: an introduction. Geological Society, London, Special Publications, 412(1), 1–3.
https://doi.org/10.1144/SP412.14.
Mukherjee, S., & Tayade, L. (2019). Kinematic analyses of brittle roto-translational planar and
listric faults based on various rotational to translational velocities of the faulted blocks. Marine
and Petroleum Geology, 107, 326–333. https://doi.org/10.1016/j.marpetgeo.2019.04.024.
Mukherjee, S., & Chakraborty, M. (2020). 3-D slip analyses of listric faults with ideal geometries.
Marine and Petroleum Geology, 112, 104092. https://doi.org/10.1016/j.marpetgeo.2019.104092.
Mukul, M., Jade, S., & Matin, A. (2009). Active deformation in the Darjiling-Sikkim Himalaya
based on 2000–2004 Geodetic Global PositioningSystem Measurements. In P. Ghosh & S.
Gangopadhyay (Eds.), IndianStatistical Institute Platinum Jubilee volumes: numerical meth-
odsand models in earth science (pp. 1–28). New Delhi: New India Publishing Agency.
Mukul, M. (2010). First-order kinematics of wedge-scale active Himalayan deformation: Insights-
from Darjiling–Sikkim–Tibet (DaSiT) wedge. Journal of Asian Earth Sciences, 39, 645–657.
https://doi.org/10.1016/j.jseaes.2010.04.029.
Mukul, M., Jade, S., Ansari, K., & Matin, A. (2014). Seismotectonic implications of strike–slip
earthquakes in the Darjiling–Sikkim Himalaya. Current Science, 106(2), 198–210.
Mukul, M., Jade, S., Ansari, K., Matin, A., & Joshi, V. (2018). Structural insights from geodetic
Global Positioning System measurements in the Darjiling-Sikkim Himalaya. Journal of Structural
Geology, 114, 346–356. https://doi.org/10.1016/j.jsg.2018.03.007.
Ni, J., & Barazangi, M. (1984). Seismotectonics of the Himalayan collision zone: Geometry of
the underthrusting Indian plate beneath the Himalaya. Journal of Geophysical Research: Solid
Earth, 89(B2), 1147–1163. https://doi.org/10.1029/JB089iB02p01147.
Review of the Geometric Model Parameters of the Main Himalayan … 323

Okada, Y. (1985). Surface deformation due to shear and tensile faults in a half-space. Bulletin of
the Seismological Society of America, 75(4), 1135–1154.
Parameswaran, R. M., & Rajendran, K. (2017). Structural context of the 2015 pair of Nepal earth-
quakes (Mw 7.8 and Mw 7.3): an analysis based on slip distribution, aftershock growth, and static
stress changes. International Journal of Earth Sciences, 106(3), 1133–1146. https://doi.org/10.
1007/s00531-016-1358-4.
Ponraj, M., Miura, S., Reddy, C. D., Amirtharaj, S., & Mahajan, S. H. (2011). Slip distribution
beneath the Central and Western Himalaya inferred from GPS observations. Geophysical Journal
International, 185(2), 724–736. https://doi.org/10.1111/j.1365-246X.2011.04958.x.
Singer, J., Kissling, E., Diehl, T., & Hetényi, G. (2017). The underthrusting Indian crust and its
role in collision dynamics of the Eastern Himalaya in Bhutan: Insights from receiver function
imaging. Journal of Geophysical Research: Solid Earth, 122(2), 1152–1178. https://doi.org/10.
1002/2016JB013337.
Toda, S., Stein, R., Lin, J., & Sevilgen, V. (2010). Coulomb 3.2, Graphic-rich deformation &
stress-change software. User Guide.
Vergne, J., Cattin, R., & Avouac, J. P. (2001). On the use of dislocations to model interseismic strain
and stress build-up at intracontinental thrust faults. Geophysical Journal International, 147(1),
155–162. https://doi.org/10.1046/j.1365-246X.2001.00524.x.
Vernant, P., Bilham, R., Szeliga, W., Drupka, D., Kalita, S., Bhattacharyya, A. K., et al. (2014).
Clockwise rotation of the Brahmaputra Valley relative to India: Tectonic convergence in the
eastern Himalaya, Naga Hills, and Shillong Plateau. Journal of Geophysical Research: Solid
Earth, 119(8), 6558–6571; https://doi.org/10.1002/2014JB011196.
Zhang, Z., Xiang, H., Dong, X., Ding, H., & He, Z. (2015). Long-lived high-temperature granulite-
facies metamorphism in the Eastern Himalayan orogen, south Tibet. Lithos, 212, 1–15. https://
doi.org/10.1016/j.lithos.2014.10.009.
Traverses Through the Bagalkot Group
from North Karnataka State, India:
Deformation in the Mesoproterozoic
Supracrustal Kaladgi Basin

Shilpa Patil Pillai and Vivek S. Kale

Abstract The Proterozoic Kaladgi Basin records the interplay between eustatic
sea-level changes and tectonics on the northern fringes of the Dharwar craton of
India. It has suffered strong supracrustal deformation that displays several unique
features including synsedimentary deformation, superimposed tectonic deformation
associated with reactivation of basement structures and the variable response of
competent and incompetent strata to external stresses. Field traverses through the
Mesoproterozoic Bagalkot Group from this basin are described that demonstrate
these features with the intention of enabling students recognise the complexities that
exist in the study of structural geology.

1 Introduction

The Kaladgi basin is located on the northern fringes of the Dharwar Craton of
south India (Fig. 1). It is comparable to other Proterozoic ‘Purana Basins’ of Penin-
sular India (Kale and Phansalkar 1991; Chakraborty et al. 2012; Kale 2016) in its
shallow marine, pericratonic sedimentary sequences. It is the only Purana Basin that
displays stronger deformation in its central parts than along the fringes (Nair and
Raju 1987; Kale 1991). This deformation is restricted to the sedimentary succes-
sion of the Mesoproterozoic Bagalkot Group and is not observed in the younger
Badami Group. Besides the axial zone post-depositional deformation and associated
lower greenschist facies / anchizone metamorphism, the sediments of the Bagalkot
Group also contain a significant record of synsedimentary deformation (Patil Pillai
and Kale 2011). These sediments display lateral and vertical sedimentological facies
changes, well-exposed lithological contacts, as well as structural geological features

S. Patil Pillai (B)


Department of Environmental Science, Savitribai Phule Pune University, Pune 411007, India
e-mail: shilpatil24@gmail.com
V. S. Kale
Advanced Center for Water Resources Development and Management (ACWADAM), Pune
411052, India
e-mail: dr.vivekale@gmail.com

© Springer Nature Switzerland AG 2021 325


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_11
326 S. Patil Pillai and V. S. Kale

Fig. 1 a Google Earth image of the region of the main Kaladgi Basin exposures overlain with the
probable extents of the Bagalkot and Badami subbasins. Inset gives the location of Kaladgi Basin
with the other Proterozoic (Purana) Basins in Peninsular India. b Regional geological map of the
Kaladgi Basin and adjoining region (based on GSI 1981; Patil Pillai and Kale 2019). The inset shows
the probable extent of the Bagalkot and Badami subbasins within this polyhistory Proterozoic Basin

(on macro-, meso- and micro-scales) making it an excellent field laboratory to study
the interplay between sea-level changes and tectonics during the evolution of a
supracrustal sedimentary basin.
We focus on the structural aspects of the Kaladgi Basin in this compilation and
propose traverses that will enable the reader to study these characters in the field.
While doing so, we recommend that the readers familiarise themselves with not
only standard procedures of field data collection and analysis in structural geology
(e.g. Allison 2004; Whitmeyer et al. 2019; Mukherjee 2019), but also recognise that
field observations cannot be random, but need to be systematic if robust kinematic
analysis is to be undertaken (e.g. Dίaz-Axpiors et al. 2019). The compilation deals
Traverses Through the Bagalkot Group from North Karnataka State … 327

with stratigraphy, sedimentology and age of the basin cursorily, with the limited
purpose of providing the background information of this basin.

1.1 Geographic Setting And Access

The Kaladgi basin is exposed between the longitudes 73° E and 76° E and the latitudes
of 15°30 N to 17° N. The contiguous exposures of these sediments, occurring in
parts of the Belgaum, Bijapur and Bagalkot districts of Karnataka, from Almatti in
the east to Ajra in the west, across a distance of about 100 km in the E–W direction
are commonly termed as the ‘main basin’ (Fig. 1a). Outliers of these sediments
capping the Archean crystalline basement are scattered south of the main basin around
Saundatti, Gajendragad and Nargund. Inliers of these sediments within the Deccan
Traps are present near Jamkhandi in Bagalkot district of Karnataka, Kallamavdi
in Kolhapur district and small isolated patchy exposures at Phonda, Malvan and
Kankavali of Sindhudurg district of Maharashtra (Fig. 1b).
Belgaum, Bijapur, Hubli and Dharwar are major urban centres in this region with
significant industrial and economic activities. Bagalkot is the largest town within
this basin, followed in size by Saundatti, Gokak and Badami. Jamkhandi, Lokapur,
Mudhol, Nargund, Ramdurg and Yargatti are smaller towns. Hotels and eateries
are available in all these locations catering to visitors and tourists. Although many
government guest/rest houses are present at many places, they can only accommodate
small groups of 6–8 persons at a time. However, lodging and boarding facilities
in budget hotels are available in Bagalkot, Badami and Gokak towns, which can
accommodate groups of up to 80 at a time. Badami and Bagalkot also have some
luxury hotels and also have facilities for hiring vehicles for conducting field traverses
in this basin.
The Solapur–Gadag railway link connects Bijapur–Bagalkot–Badami, while
Belgaum is on the Kolhapur–Miraj–Hubli–Goa rail line. Belgaum has an airport,
which connects with Mumbai and Bangalore. A network of paved roads, which
interconnect between national highways NH4, NH218 and NH13, connects to towns
and larger villages in this region.

1.2 Geomorphology and Drainage

The eastern parts of this region display three geomorphic segments, namely:
• Dissected plateau (representing the southern edge of the Deccan Plateau) in the
north, which gradually is replaced by a hilly terrain with moderate to steep-sided
hills;
• Rugged terrain with low hill ranges interspersed with flat cultivated plains; and
• Gently undulating, low-lying plains with few isolated plateaus or tors in the south.
328 S. Patil Pillai and V. S. Kale

The western parts have the dissected (Deccan) plateau that is followed westward by
the spectacular Western Ghats Escarpment and then the low-lying coastal tract of the
Konkan Coastal Belt.
Eastward draining perennial rivers and their tributaries drain this region. River
Krishna flows along the northern boundary of this basin. Ghatprabha and Malprabha
are its right bank tributaries, which drain the Kaladgi Basin (Fig. 1b). In addition, the
upper reaches of the left bank tributaries of the river Tungabhadra have their sources
in the southwestern parts of this region.

1.3 Land Use and Economics

Most of this terrain is under irrigated cultivation, with local forests and grasslands.
Irrigation from the dams in the region support cultivation of staple food crops, sugar-
cane, groundnuts, tobacco and pulses. Several parts on rocky scrublands/wastelands
are under horticulture, using drip irrigation.
The extensive limestone from this basin are quarried to support numerous cement
factories (in Bagalkot, Bammanbudhni, Lokapur, Mudhol, etc.), and as paving stones
(Jayaprakash 2007). The sandstones are being quarried in the western part of the
basin (near Ramdurg) as silica sands for glass industry. Unconformity-based uranium
mineralisation has been reported from various locations in this basin (Sridhar et al.
2014). Anomalous concentrations of light gaseous hydrocarbons in the surface soils
of the Kaladgi–Lokapur and Mudhol areas of this basin have been reported by
Kalpana et al. (2010). The Hiremagi–Ramthal iron ore mine, spread over an area
of ~100 acres east of this basin, excavates Algoma-type Banded Iron Formations
from the Hundgund–Kushtagi Schist belt. All these quarries/mines can be easily
visited by acquiring prior permission through proper channels.
World heritage sites of Badami, Aihole and Pattadakal in this region host signif-
icant structures and relicts of the early history of the Chalukya Period (around 400
AD). These and other archaeological sites stand testimony to the rich cultural heritage
of this region (e.g. Korisettar and Petraglia 1993; Korisettar 1995; Pappu 2001;
Yogesh et al. 2001), providing an additional attraction to this picturesque countryside
during the field studies.

1.4 Climate

This region experiences a tropical monsoon climate. Summers are extremely hot and
dry, peak months being April and May, with maximum temperatures of 40 °C and
above. The hot summer months are good for studying the exposures, unhindered
by vegetal cover, but require appropriate precautions for the heat. The pre-monsoon
showers occur around the end of May and early June, which bring relief from the hot
Traverses Through the Bagalkot Group from North Karnataka State … 329

climate. The following rainy season till October has a hot and humid climate; and
field studies should be avoided during this period.
The post-monsoon months of October and mid-November experience increasing
heat and humidity, with occasional thunderstorms. November marks the beginning
of cool and pleasant winter season, which lasts till February. December usually is
the coldest month, with temperatures dropping to 13–15 °C. November to February
is a period of a comfortable climate, but shrubs and grass may cover large segments
of the exposures in this period.

2 Regional Geology

This E–W elongate basin (Fig. 1b) lies on the northern fringes of the Western Dharwar
Craton, with an exposed area of around 8000 km2 . This polyhistory basin consists of a
Mesoproterozoic Bagalkot subbasin and Neoproterozoic Badami subbasin (see Inset:
Fig. 1b) and is estimated to have occupied no less than 20,000 km2 before its conceal-
ment under younger cover and erosional removal (Kale 1991). These Proterozoic
sediments rest on the Archean–Paleoproterozoic sequences of the Dharwar craton,
with a profound angular and erosional unconformity. The Deccan Trap basaltic lava
flows of Late Cretaceous–Early Paleocene age cover them across this region. Table 1
gives the outline of the regional stratigraphy of this terrain.

2.1 Basement Complex

The basement for the Kaladgi sediments consists of crystalline and schistose rocks of
Archean to Early Paleoproterozoic age (Table 1). They are a typical cratonic assem-
blage comprising of granites and associated intrusives and the younger basic dykes
intruding into multiple phases of greenstone belts and granitic gneisses (Ramakr-
ishnan and Vaidyanadhan 2010; Valdiya 2016). The basement–cover relationship is
discussed below (Sect. 2.2).
The Peninsular Gneisses (undifferentiated in Fig. 1b) are composed of tonalitic–
trodndhjemitic–granodioritic(TTG) gneisses; and are known to have evolved in two
separate phases (Jayananda et al. 2015), as listed in Table 1. The NNW-SSE trending
Shimoga Schist belt is present to southwest of the basin (south of Belgaum), and
its equivalents are recognised from the coastal belt around Malwan. This belt is
composed of gneisses, conglomerates, phyllites and banded iron formations (ferrug-
inous chert and quartzite). The NW–SE trending Hundgund–Kushtagi Schist Belt
(see: Fig. 1b) is a ~100 km long and 20 km wide belt that forms the basement in the
eastern parts of the Kaladgi Basin, around Bagalkot. This tightly folded sequence of
banded iron-quartzites, pelitic metasediments, interbedded basalts and minor ultra-
mafics display multiple generations of folding (Matin 2006).The folding and meta-
morphism of the HKSB is of Neoarchean age since it preceded the Neoarchean
330 S. Patil Pillai and V. S. Kale

Table 1 Regional stratigraphy of the Dharwar craton (after Ramakrishnan and Vaidyanadhan 2010;
Patil Pillai et al. 2018; Kale et al. 2019; Krapez et al. 2020)
Recent alluvium Holocene
Laterites (with local Bauxite) Neogene—Quaternary
~~~~~~~~~~~~~~~~~~~~~Nonconformity~~~~~~~~~~~~~~~~~~~~~
Deccan Traps Sahyadri Group ~69–61 Ma
~~~~~~~~~~~~~~~~~~~~~Nonconformity~~~~~~~~~~~~~~~~~~~~~
Gulbarga Intertrappean Late Cretaceous (Maastrichtian)
Formation
~~~~~~~~~~~~~~~~~~~Angular and erosional Unconformity~~~~~~~~~~~~~~~~~~~
Kaladgi Supergroup Badami Group Neoproterozoic (<0.9 Ga)
~~~~~~~~~Angular and erosional unconformity~~~~~~~~~
Bagalkot Group Mesoproterozoic (~1.2–1.4 Ga)
~~~~~~~~~~~~~~~~~~Angular and erosional Unconformity~~~~~~~~~~~~~~~~~~~~~~
Basement Complex Newer Dolerites Paleoproterozoic (1.8 ± 0.2 Ga)
-----------------------Intrusive contact----------------
Closepet Granite Late Neoarchean (2.5 ± 0.1 Ga)
-----------------------Intrusive contact-------------------
Peninsular Gneiss-II Neoarchean (~2.5–2.7 Ga)
- - - - - - - - - - - - - -Metamorphic contact- - - - -- - -- - - – - -
Dharwar Supergroup Neoarchean (2.6–2.8 Ga)
- - - - Diffuse metamorphic contact/unconformity at places- - - -
Peninsular Gneiss-I Mesoarchean (~2.9 Ga)
- - - - - - - - - - - - -Metamorphic/structural contact- - - - - - - - -
Sargur Group Paleoarchean (~3.4 Ga)

granitic intrusions. The NW–SE trend of foliation of the Dharwarian greenstone


belts and Peninsular Gneisses swirls towards WNW–ESE orientation in the vicinity
of the Kaladgi Basin (Jayaprakash et al. 1987; Patil Pillai and Kale 2019).
The Closepet Granites and equivalent rocks represent the K-granite activity asso-
ciated with the coalescence of the eastern and western blocks of the Dharwar Craton
around 2.3–2.5 Ga (Ramakrishnan and Vaidyanadhan 2010; Krapez et al. 2020).
Proterozoic mafic dykes that intrude various parts of the Dharwar Craton were earlier
clubbed as ‘newer dolerites’, suggesting that they post-date the cratonization of the
Dharwar block. Recent studies on their petrogenesis and age (e.g. Kumar et al. 2015;
Samal et al. 2019) have demonstrated them to range in age from late Neoarchean
(~2.7 Ga) to the Mesoproterozoic (~1.4 Ga), with a large population clustered around
1.8 Ga.
These crystalline gneissic, schistose and granitic rocks have not only served as
the basement for the Kaladgi sediments, but also their provenance (Jayaprakash
2007; Dey 2015; Patil Pillai and Kale 2019). They are collectively referred to as the
‘Basement Complex’ or simply ‘basement’ in the ensuing description.
Traverses Through the Bagalkot Group from North Karnataka State … 331

2.2 Kaladgi–Basement Unconformity

The contact between the basement and the capping Kaladgi sediments is an angular
and erosional nonconformity. This contact can be studied all along southern, eastern
and northern margins of the Kaladgi Basin, where the rugged ridges of quartzites
and sandstones of the Kaladgi Supergroup stand out in contrast to the peneplained
topography of their basement(Fig. 2a). This contact is buried under the younger cover
of Deccan traps along the western margin of the main basin. It is significant to note
that the basal arenaceous beds of the Bagalkot as well as the Badami Groups rest
directly upon the weathered and peneplained Basement Complex.
The rocks of the Basement Complex display an intensely weathered profile below
the unconformity, and often give the impression of being a paleosol horizon. This
can best be studied north of the road between Mullur (15°52 38 N:75°18 20 E)
and Jalikatti (15°52 18 N: 75°22 13 E). The capping sediments in these loca-
tions display subhorizontal disposition, indicating their undeformed nature. Often,
pebble beds are interlayered within the sandstones/quartzites at such locations,
with clasts derived from the underlying basement (Fig. 2b). However, at other
places like the road section near Vatnal (15°48 46 N: 75°06 45 E), hill section
west of Ugargol (15°44 53 N:75°10 17 E) and the hill near the town of Gokak
(16°09 58 N:74°48 49 E), excellent exposures of the unweathered Basement
Complex underlying the unconformity can be studied.
The contact between the Basement Complex and the sediments is also a struc-
tural one at some places. In such locations, the latter are exposed displaying local
steepening of dips (Fig. 3a), shearing and brecciation of the sediments as well as the
basement rocks and parasitic folds (Fig. 3b), which suggest the proximity of a faulted
contact between the basement and the cover sediments. Such structural contacts
generally display an E-W trend of the causative fault/shear zone. The deformation
in the basement can be traced for some distances away from the limit of the sedi-
ments along the strike of such shear planes. Often, the actual trace of the fault/shear
plane is concealed below a thick soil cover and slope-base gravelly deposits. Exam-
ples of such contacts can be examined near Almatti (16°21 04 N:75°52 50 E),
Amingarh (16°03 43 N:75°56 37 E), Idgal (15°56 09” N:75°21 08 E) and Saun-
datti (15°45 55 N:75°08 33 E). Based on such evidences, several faults along the
margins of the Kaladgi Basin are interpreted (Patil Pillai and Kale 2019) to represent
growth faults that controlled the subsidence in the basin.

2.3 Kaladgi Supergroup

Foote (1876) gave the earliest comprehensive compilation of the Kaladgi basin. The
work of Viswanathiah (1979) and his colleagues brought about several fundamental
changes in the knowledge of these sediments. Jayprakash et al. (1987), Kale and
Phansalkar (1991) reviewed the contemporary state of the art and compiled elabo-
rate citations of the work done till the close of the last century. Jayaprakash (2007)
332 S. Patil Pillai and V. S. Kale

Fig. 2 a Thin (~10 m) cap of Cave Temple Arenites resting unconformably upon the Basement
Complex. Photograph is taken looking (eastward) downstream of the Gokak Falls (16°11 30 N:
74°46 42 E) on Ghatprabha River. The elevation of the crest of the escarpment is 663 m amsl
while that of the riverbed is ~550 m amsl. b Pebble bed interlayered within the basal sandstone
horizons of Saundatti Quartzite exposed near Sogal (15°51 37 N: 74°58 29 E) with hammer-head
for scale. Note the clasts of amphibolite (dark), schists and gneissic rocks akin to the rocks from
the Basement Complex
Traverses Through the Bagalkot Group from North Karnataka State … 333

Fig. 3 a Steeply dipping (60° towards North) Saundatti Quartzite exposed in the ridge south of
Khanapur (15°53 52 N: 75°29 43 E) photographed looking westwards. The truck and electric
pole provide the scale. The basement is exposed on the south (on the left side of the photo). b A
fold within Saundatti Quartzite exposed in the ridge south of Khanapur (15°53 52 N: 75°29 43
E). Photograph is looking eastwards into a quarry. The quarry face is ~50 m across
334 S. Patil Pillai and V. S. Kale

compiled the studies in this basin, focusing mostly on the elaborate mapping under-
taken by the Geological Survey of India. Dey (2015) and Patil Pillai and Kale (2019)
have given the most recent compilations on the knowledge of this basin.
The sedimentary sequence in the Kaladgi Supergroup is divided into
the older Bagalkot and younger Badami Groups based on an angular and
erosional unconformity between them that was first described from Torgal Tanda
(15°55 39 N:75°16 08 E; Fig. 4) and east of Kendur (15°58 12 N:75°45 11 E) by
Viswanathiah (1968). The Bagalkot Group is subdivided into the Simikeri and
Lokapur Subgroups separated by a disconformity. Based on exhaustive mapping of
the basin, Jayaprakash (2007) subdivided the Lokapur Subgroup into six formations
comprising thirteen members with an aggregate stratigraphic thickness of 3094 m,
and the Simikeri Subgroup into three formations with seven members having a
total stratigraphic thickness of 1140 m. The Badami Group was divided into two
formations and six component members with an aggregate stratigraphic thickness of
286 m.
Kale et al. (1996; 1999) used satellite imageries and aerial photographs and exten-
sive field studies to remap the basin based on the modern concepts of sedimentary
facies. They argued that the erstwhile lithostratigraphy needs revision, since the
‘members’ of Jayaprakash (2007) display lateral variation of facies across the basin

Fig. 4 Subhorizontal Torgal Conglomerate (Badami Group) resting with an erosional and angular
unconformity (marked by white wavy line) over the Manoli Ferruginous Shale (Bagalkot Group) at
Torgal Tanda. S0 (light yellow dashed lines) represent the bedding planes. The ferruginous shales
dip by 20°/50°
Traverses Through the Bagalkot Group from North Karnataka State … 335

and have gradational contacts. Patil Pillai and Kale (2019) compiled the revised
lithostratigraphy and depositional environments of the Kaladgi basin (Fig. 5).
Stromatolites (e.g. Viswanathiah and Chandrasekhara Gowda 1970; Viswanthiah
and Sreedhara Murthy 1979; Raha and Sastry 1982; Sharma et al. 1998) and Ichno-
fossils (Kulkarni and Borkar 1997) from these sediments were interpreted to indicate
a Late Paleoproterozoic to Mesoproterozoic age. Patil Pillai et al. (2018) gave a robust
40
Ar/39 Ar whole rock age of 1154 ± 4 Ma for a mafic dyke (belonging to Mallapur
Intrusives) intruding these sediments along the axial plane of folding. Joy et al. (2018)
computed a crystallization age of 1860 ± 4 Ma for this intrusion, which has been
questioned on multiple grounds by Patil Pillai et al. (2019). It may be concluded
that the Bagalkot Group was deposited presumably between 1450 and 1200 Ma
and was subjected to deformation and magmatic intrusions around 1200–1100 Ma.
The Badami Group is of Neoproterozoic age and may have been deposited around
900–800 Ma.

Fig. 5 Lithostratigraphy and depositional environments of the Kaladgi Supergroup (after Patil
Pillai and Kale 2019)
336 S. Patil Pillai and V. S. Kale

3 Structural Characters

The sediments from the Purana Basins of Peninsular India display an undeformed
and unmetamorphosed character, with deformation being restricted to margins where
they have structural contacts with adjoining terrains (Kale 1991). In this context,
the Kaladgi Basin is unique because of intense deformation in the central parts
of the basin. The younger Badami Group is relatively undeformed and occurs with
subhorizontal disposition, while the Bagalkot Group displays significant deformation
(Awati and Kalaswad 1978; Nair and Raju 1987; Gokhale and Pujar 1989; Mukherjee
et al. 2016; Jadhav and Kshirsagar 2017) as depicted in Fig. 6.
The complexity of the folding in this sequence of alternating competent and
incompetent strata makes them an excellent field laboratory for students of structural
geology. The following descriptions provide a glimpse of locations where one may
examine the structural characters of the Bagalkot Group. While doing so, we do not
focus much on the sedimentological characters of the strata in question, although
they also are worth studying.
Three broad categories of the structures in the Bagalkot Group can be identi-
fied, in context of its deformational history. They are: (a) older shear zones from
the Basement Complex; (b) synsedimentary deformation features, and (c) the super-
posed folding and faulting in the Bagalkot Group. The first named are certainly of
Neoarchean–Paleoproterozoic age while the remaining two can be interpreted to
be of Mesoproterozoic age. Excluding the development of the former, at least four
separate events of structural deformation (designated as D0 , D1 , D2 and D3 ) were

Fig. 6 N–S geological cross sections across the Kaladgi Basin along the longitudes of 75°20 E and
75°35 E (from Patil Pillai and Kale 2019) depicting the deformation of the Bagalkot Group
Traverses Through the Bagalkot Group from North Karnataka State … 337

identified by Patil Pillai and Kale (2019). The events D0 and D3 represent extensional
tectonics, while D2 has been interpreted to represent compressional deformation.

3.1 Inherited Structures

The oldest are the shear zones and faults in the Basement Complex that appear to have
originated during its Late Archean–Paleoproterozoic development (see: Table 1).
These WNW–ESE to E–W trending features have subparallel orientations with the
foliation trends of the sequences that they are present in. Shearing and dislocation
of the foliation/metamorphic bands in the Basement Complex and presence of psue-
dotachylitic bands, quartz veins and slicken-sided surfaces (with linear streaks) are
commonly seen in the exposures along them, the Almatti, Anagavadi, Sirur (earlier
described by Jayaprakash (2007) as the Sirur–Anwal fault), Kolchi and Saundatti
faults/shears fall in this category. All of them display evidence of reactivation in the
form of deformation of the Bagalkot Group sediments along them as well. Some
of these regional structures of the Basement Complex have been shown to continue
below the Kaladgi sediments by geophysical studies (Mallik et al. 2012; Sridhar
et al. 2017; Rajaram et al. 2017). The WNW–ESE trending shears are interpreted to
have controlled the geometry of the Kaladgi Basin, besides being active during the
sedimentation in it (Jayaprakash 2007; Patil Pillai and Kale 2019). They represent
an inherited structural ‘grain’ in the Kaladgi basin. Figure 7 is a simplified map of
the Bagalkot Group depicting the large-scale folds, faults and shear zones associated
with it.

3.2 Synsedimentary Structures

The synsedimentary deformation structures (attributed as D0 ) in the Bagalkot Group


(Kale et al. 1998; Kale and Patil Pillai 2011; Patil Pillai and Kale 2011; Mukhopad-
hyay et al. 2019) can be studied on outcrop scale and do not have easily recognisable
imprint on a regional (macro) scale. However, the distribution of the various sedi-
mentological facies in their vicinity differs from that in sectors where they are absent.
This shows that the depositional system during the Mesoproterozoic growth of the
Bagalkot subbasin was significantly controlled by tectonic subsidence rather than
thermal sagging.
D1 was triggered by reactivation of some of the inherited basement shears and
led to the reorganising of the basin geometry particularly in the northern part of
the Bagalkot subbasin. Its imprints are not very clear along the southern fringe of
the Bagalkot Group exposures (Patil Pillai and Kale 2019). A low magnitude defor-
mation of the deposited sediments led to their emergence and a temporary cessa-
tion of sedimentation, manifested as the Lokapur–Simikeri unconformity. This is
followed by an accelerated subsidence along the Anagavadi, Gaddankeri and Sirur
338 S. Patil Pillai and V. S. Kale

Fig. 7 Major structural elements of the Bagalkot Group (after Patil Pillai and Kale 2019). Note the
tightly folded pattern in the central parts of the basin, with only gentle monoclinal dips along the
northern and southern fringes of the basin. The rectangles A, B, C and D outline the locations of
the Gaddankeri, Anagavadi Anticline, Lokapur Syncline and Yadwad traverses, respectively

shear zones enabling sedimentation of the Simikeri subgroup. Manifestations of this


are recorded in the form of seismites and synsedimentary deformation in the Chik-
shellikeri Limestone (Kale et al. 1998; Patil Pillai and Kale 2011; Mukhopadhyay
et al. 2019).

3.3 Superimposed Deformation Structures

The superimposed deformation on the Bagalkot Group, well documented in litera-


ture of this basin, can be assigned to a compressive tectonic regime (D2 ). Stronger
deformation along the central axis of the basin and significantly lesser intensity along
the fringes is akin to the classical ‘geosynclinal’ cross sections (but lacking in higher
metamorphic grades) as seen in Fig. 6. The tight isoclinal folding of the Simikeri
Subgroup yielded doubly plunging synclines near Lokapur, Muchkundi and Mudhol–
Yadwad that stand out in satellite imageries (see Fig. 1a) and aerial photographs very
sharply (Awati and Kalaswad 1978; Nair and Raju 1987).
Traverses Through the Bagalkot Group from North Karnataka State … 339

The folding in the Bagalkot Group was interpreted as being a result of two co-axial
compressive events by Jadhav (1987) based on meso- and micro-scale deformation
of quartizites and associated rocks from the basin. This deformation was attributed
to gravity-induced gliding along deep crustal detachment planes by Mukherjee et al.
(2016). In case of southward gliding, the fold-planes should display a progressive
increasing tilt southwards, which is absent in the folds from the Bagalkot Group.
It is therefore likely that basement shears were reactivated during the compressive
deformation of these sediments as surmised by Patil Pillai and Kale (2019). Jadhav
and Kshirsagar (2017) recorded the presence of a superimposed weak cross-folding,
which is borne out on a regional scale by the outlines of the folds (see: Fig. 9 of
Patil Pillai and Kale 2019). Mukherjee et al. (2019) have noted the diachronous
nature of cleavage development in the folds from the Bagalkot Group. The cross-
folding may be attributed to a strike-slip component of space accommodation during
this compressive deformation D2 . The Lokapur syncline described below (Sect. 4.3)
provides a good example of this.
Mukherjee et al. (2019) interpreted that the thermal maturity of the Bagalkot Group
(based on illite crystallinity index) may be attributed to the anchizone metamorphic
grade reaching up to 300°C at subsidence depths of 5–8 km. Most alterations and chlo-
rite development along slaty cleavage planes is found near the fold hinges and in the
proximity of shear/fault planes, rather than across the entire strata. Frictional heating
(e.g. Mukherjee 2017) during the deformational episode is more likely to yield this
maturity rather than the subsidence. Frictional brittle deformation in competent strata
and elasto-ductile deformation in incompetent strata further support the possibility of
deformation-related alterations rather than subsidence related maturation. The trans-
verse faults that cut across the various regional folds also indicate that a compressive
deformation with a minor shear component was the primary driver of the D2 event.
The contrast in the response of the competent and incompetent strata to the defor-
mation is a significant aspect worth studying in the Bagalkot Group. While the sili-
ciclastic sediments and silica-bearing strata (e.g. cherty limestones) have suffered
brittle deformation (Fig. 8a, b) resulting in brecciation; the argillaceous and carbonate
sediments have suffered ductile warping and buckling (Fig. 8b, c) within the same
domain. This is evident in most exposures of the Bagalkot Group. The structural char-
acters enumerated in the ensuing traverse descriptions are easily accessible examples
of exposures that display various characters of the deformation of the Bagalkot Group.

4 Field Traverses

Before starting the field studies based on the following descriptions, readers are
advised to familiarize themselves with the fundamental studies on folding of rocks by
Ramsay (1967), Huddleston (1973), Twiss (1988) as have been compiled in the some
of the textbooks of modern practices in structural geology (Fleuty 1964; Ramsay
1967; Ramsay and Huber 1987; Fossen 2011; Davis et al. 2012). The terminology
340 S. Patil Pillai and V. S. Kale
Traverses Through the Bagalkot Group from North Karnataka State … 341

Fig. 8 a Brittle deformation manifested by slickensides and brecciation of the Saundatti Quar-
tizite along the NW–SE trending Kolchi fault/shear zone exposed east of Ramdurg at 15°56 36
N:75°20 52 E. Note the blocks of breccia attached to this exposure and the development of streaky
lineations on the fault plane (L 1 : marked by orange dashed lines) indicating the sense of movement
along the quartzite that dips by 58°/105°. The bedding planes of the quartzite are marked by yellow
dotted lines (S0 ). b Ductile deformation leading to tight mesoscopic folding of the bedding planes
(S0 ) and development of axial plane cleavages (Fr1 ) in the Petlur Carbonates exposed near Hebbal
at16°11 29 N:75°21 51 E. Note that the interlayered cherty limestones have suffered brittle brec-
ciation (Br) in the same exposure. The mesoscopic fold axes in this exposure have a plunge 12°/95°.
Hammer used for scale is 30 cm long. c Quasi-ductile deformation and boudinage development (B)
of the cherty bed within the tightly folded Petlur Carbonates exposed at 16°11 29 N:75°21 51 E.
Note geometry of the folding of the bedding planes (S0 ) and axial plane cleavages (Fr1 ) cutting
across the exposure

of geometric description and classification of the folds used in the following text is
based on these publications, and we do not cite them at each point.
The traverses (see: blocks outlined in Fig. 7) described below present a wide
spectrum of primary sedimentary structures, bedding plane characters and the internal
stratification of the sediments that are like any text-book examples. They should not be
disregarded during the field studies, although we have focussed on structural geology.
The stratigraphic nomenclature used in these descriptions is as enumerated in Fig. 5.
The field traverses are intended to provide a glimpse of the diverse structural features
from the Bagalkot Group. Each of the suggested stops along the traverses describes
a typical structural feature. These structures should be understood and appreciated
as local manifestation of regional structural framework in the particular exposure
and not on a standalone basis. Although the locations in each of these traverses are
specified, it is recommended that the exposures should be studied laterally along
the strike. The spellings of location names are used as given in Survey of India
Toposheets, and they may slightly differ in Google Earth images.
The Gaddankeri Traverse (A in Fig. 7) is intended to cover exposures of the phase
D1 , which includes synsedimentary deformation, results of shallowing of the depo-
sitional interface under tectonic influence and the deformation along an intrabasinal
shear zone. The other three traverses focus on the deformation D2 . The Anagavadi
Anticline (B in Fig. 7) is a regional anticline that exposes the Basement Complex
at its core and the cover sediments of the Lokapur Subgroup on its limbs; besides
local parasitic folds, cross-faults and mesoscopic deformational features that are
associated with large amplitude folds.
The next traverse (C in Fig. 7) covers a doubly plunging syncline rimmed by
quartzites ridges exposed near Lokapur. The geometry of the macro-scale folding
including the swirl in the main fold axis induced by cross-folding, associated shear
plane, cross-faults and the geometry of joint planes associated with such folds can be
observed here. The Yadwad Traverse (D in Fig. 7) addresses the tightly folded argilla-
ceous–carbonate sequences of the Petlur Formation that are outlined by the ridges
of Muchkundi Quartzite. The interrelationship between mega-scale and macro-scale
structures, geometric classification of megascopic folds, and evolution of secondary
342 S. Patil Pillai and V. S. Kale

slip-planes during folding can be studied here, besides the response of competent
versus incompetent strata to compressive tectonic forces.

4.1 Gaddankeri Traverse

This traverse is a N–S section exposed (Fig. 9) on either side of the Belgaum–
Bagalkot State Highway KH SH20, ~2 km east of the Gaddankeri Cross (=junction
with National Highway NH218). It is covered in Survey of India (SoI) Toposheet
No. 47P/12. These rubbly exposures are covered by shrubs in the south of the road,
while in the north is open grassland. Barring the ridge of the Muchkundi Quartzite,
the rest of the area has a fairly thick soil cover and is cultivated.
The traverse is located on the northern limb of the Muchkundi synclinorium that is
flanked in the north by a WNW–ESE trending regional (Gaddankeri) shear (Fig. 10).
It covers the unconformable contact between the Lokapur and Simikeri Subgroups,
which is marked by the thin, impersistent bed of Bevinmatti Conglomerate Member
of the Muchkundi Quartzite. The Muchkundi synclinorium has a WNW–ESE striking
doubly plunging, gently southward inclined axial plane; which displays a gentle swirl.
In terms of the geometric classification, each of the terminal closures of this regional
synclinorium may be attributed to class C4/D4 of Huddleston’s classification.

Fig. 9 Locations of interest along the Gaddankeri Traverse overlain on the Google Earth image.
The exposure of Muchkundi Quartzite is highlighted as it marks the base of the Simikeri Subgroup
(exposed in the south), Lokapur Subgroup (exposed in the north) and the major structural elements
are marked for easy reference. Inset depicts the regional structural setting of being on a limb of the
Muchkundi syncline adjoining the Gaddankeri Shear
Traverses Through the Bagalkot Group from North Karnataka State … 343

Fig. 10 Simplified cross section along X–Y shown in Fig. 9 depicting the sequence of lithologies
encountered along the Gaddankeri traverse. The spot locations described below are marked for easy
reference. CL = Chikshellikeri Limestone; JP = Jalikatti Phyllites; PFA = Petlur Formation—
argillites; MQ = Muchkundi Quartzite; AA = Arlikatti Argillite; NC = Niralkeri Chertbreccia.
The Gaddankeri Shear is shown by red dash line

4.1.1 Location GA

This barren exposure of beds dipping by 65°–85° towards south (due 190°) provides
a perfect location to study the Chikshellikeri Limestone across a thickness of >200 m.
It is characterised by alternating rhythms of dirty brown mudstone and bluish grey
sparry limestone. Climbing ripple laminations, load and flames, pseudo-nodules are
some of the commonly observed sedimentary structures in this exposure.
Horizons of Intraformational Limestone Breccias (IFLB) interrupt these alterna-
tions in an apparently systematic manner. The IFLB represent discrete beds (Fig. 11a)
of the limestone that suffered brittle deformation and brecciation before being
covered by the subsequent bed of limestone. Kale et al. (1998) demonstrated that
the frequency and thickness of the IFLB horizons progressively increases towards
the stratigraphic top of this horizon. Patil Pillai and Kale (2011) interpreted them
to represent seismites, while Mukhopadhyay et al. (2019) suggested that they were
products of aseismic tectonism related to basin floor instability. In any case, they can
be ascribed to the synsedimentary event of deformation D0 .
The presence of a series of intraformational faults can be recognised in these
exposures, by the fault planes and resulting dislocation being restricted to a discrete
set of beds and not transmitting into the overlying beds. These faults are synsed-
imentary in character and hence are not related to any major deformation (D1 or
D2 ); therefore, oritentational connotations such as transverse/longitudinal cannot be
used for them. Similar intraformational faults (associated with synsedimentary basin
floor instability and tectonics D0 ) have been recorded from other horizons of the
Bagalkot Group (Patil Pillai and Kale 2011 and 2019). It is significant that in one or
two patches from this exposure, such synsedimentary faults display listric geometry
(Fig. 11b). Such fault arrays (designated F 0 in recognition of their association with
the synsedimentary deformation event D0 ) are restricted to a limited package of beds
and do not extend in the underlying or capping beds. This is a local synsedimentary
manifestation of the regional geometry and framework of the basinal growth faults
of the Kaladgi Basin.
Several N–S trending small-scale faults (F 1 ) cut across this limestone exposure,
displaying millimetre scale displacement along the rhythmic beds (S 0 ) and filled
344 S. Patil Pillai and V. S. Kale

Fig. 11 a Alternating beds of dirty brown mudstones and bluish grey sparry limestone with
interbedded horizons of IFLB. The bedding planes (S0 ) dip by 65°/190° (towards the top of photo).
Note the up-dip increase in thickness of the IFLB horizons. The scale used is 5 cm long. b Listric
fault array (F0 = red dash line) exposed on the strike (100°–280°) face of Chikshellikeri Limestone
in the GA exposure. Note how this fault plane (trending NE–SW) of the main fault swirls towards
the bedding plane strike and flattens out towards the base and merges with the bedding plane (S0 ).
Pen for scale is 14.5 cm in length. c Small-scale faults (F1 : red dashed lines trending N–S) with
millimetre scale dislocations of the subvertical bedding planes (S0 marked by yellow dashed lines)
dipping by 76°/190° (towards top of photograph) from the GA exposure. Note the interbedded IFLB
horizon

by calcite veins, are visible in this exposure (Fig. 11c). Consistent sense of clearly
decipherable drag is not present across the faults. Therefore, the fault and the bedding
planes can together can be called an n-type flanking structure (Mukherjee 2014). They
represent local fracturing resulted from D1 deformation event of folding and faulting.

4.1.2 Location GB

This location is ~100 m south of GA , across the road, SH20. Open shrubs with
weathered rock exposures occur along this stretch. A thin (<2 m) band of greyish
black phyllitic shales (Fig. 12a) crop out with southerly dipping bedding planes (S 0 )
having strike along 100°–280°. These rocks display a close-spaced slaty cleavage
(S 1 ) dipping by 70°–75° due south, that is slightly oblique to the bedding plane.
Traverses Through the Bagalkot Group from North Karnataka State … 345

Fig. 12 a Exposure of the Jalikatti Phyllite displaying a strong slaty cleavage (S1 ) which is slightly
oblique to the bedding planes (S0 ) resulting in the chipping of the rock into small flakes. The hammer
for scale is 35 cm in length. b The Lokapur–Simikeri unconformable contact between the shales
from the Petlur Formation (in the foreground) and the Bevinmatti Conglomerate Member. The
rubble covered disconformable contact is shown by white wavy line. Geologist Shrikant Chitnis
is 174 cm tall. Photograph is taken looking at the strike face (100°–280°) of the quartzite beds.
c Brecciation along the Gaddankeri Shear recorded in a quarry face at location GC . The affected
Muchkundi Quartzite dips by 75°/200° (towards the left of the photograph). Hammer for scale is
33 cm long. d Slickensides developed along a quartz-vein filling a local fracture-plane) in the zone
of deformation along the Gaddankeri Shear. Linear striations are exposed on a fracture surface (Fr1 )
dipping by 80°/205°. Hammer for scale is 33 cm long

This horizon, recognised as the Jalikatti Phyllite (see Fig. 5), occurs interbedded
within the limestones and dolomites of the Petlur Formation. Although here it is
followed upwards by dolomitic horizons, in other exposures of the Lokapur Subgroup
(Jayaprakash 2007), it is interbedded within the (Chikshellikeri Limestone) entirely.
The Chitrabhanukot Dolomite occurs as small disconnected patches with an
elephant skin weathering. Fresh cuts of the rock display it to be a bluish grey algal
laminated dolomite. However, due to the quartzitic rubble that has rolled downhill,
good continuous exposures of this horizon (mapped based on lateral continuity in
the cross section: Fig. 10) are not available along this traverse. Further in the up-dip
direction, the dolomites are followed by the ferruginous shales of the Petlur Forma-
tion (see Fig. 5). They occur as disconnected patches with a weak slaty cleavage
having a lesser intensity than that of the Jalikatti Phyllite. The argillaceous beds
346 S. Patil Pillai and V. S. Kale

(marked as PFA in Fig. 10) and the dolomites display a strike trending 100°–280° to
110°–290°.

4.1.3 Location GC

At this location, one may examine the nature of the Lokapur–Simikeri Unconformity.
This is exposed about 10 m below the crest of the hill (Fig. 12b). Thin beds of
Bevinmatti Conglomerate rest unconformably on the ferruginous shales of the Petlur
Formation. The abrupt change from a passive mudflat depositional environment of
the argillaceous sediments to the high-energy washed sediments of the Muchkundi
Quartzite (with the conglomerates interbedded in its basal beds) connotes a local
regression. Whether this regression resulted from a eustatic sea-level variation or
was driven by tectonic reorganisation of the basin floor is debated. However, the
intraformational deformation (including IFLB and other evidences discussed by Patil
Pillai and Kale 2019) suggests the greater possibility of the latter.
The occurrence of intense brecciation (Fig. 12c), close-spaced fracturing (in both
the shales and quartzites) and slicken-sided surfaces (Fig. 12d) testify to the pres-
ence an intense zone of deformation close to this unconformity. Lateral tracing
of this zone of deformation establishes this to be associated with the WNW–ESE
trending Gaddankeri shear zone (see Fig. 9) that extends at least 50 km adjoining the
Muchkundi Synclinorium along its northern limb.

4.2 Anagavadi Anticline

The locations in the traverse across Anagavadi Anticline fall partly in SoI toposheet
Nos. 47P/11 and 47P/12. The Anagavadi Anticline (Fig. 13) north of Bagalkot town
is one of the best sections where one may clearly observe the angular and erosional
unconformity between the Kaladgi sediments and their basement. This plunging
anticline (or rather anticlinorium, since it has macro-scale parasitic folds along its
limbs: Inset: Fig. 13) has an 110°–290° striking axial plane that is subvertical in the
east and swirls to a nearly E–W direction westwards. It is a slightly asymmetric,
gently inclined plunging (by ~10°/270°) class of fold (Fleuty 1964) and D4 class as
per Huddlestone (1973) classification (Fig. 14).
This plunging anticline is very closely associated with the NW-SE trending
Anagavadi Shear (see Fig. 7) that cuts obliquely along its southern parts. The
Anagavadi Shear displays quasi-ductile deformation in the Basement Complex
further east of Bagalkot and can be traced across nearly 100 km within the basin
as collinear, disconnected segments. Going by the dislocations of beds, it displays a
dominantly left-lateral sense of shear, besides significant vertical dislocations.
This area (continuing up to Almatti further northeast) is recognised as the type
exposures of the Salgundi Conglomerate and Almatti Quartzite Members of the
Saundatti Formation. The interbedded siltstone layers become thicker and eventually
Traverses Through the Bagalkot Group from North Karnataka State … 347

Fig. 13 Google Earth image of the Anagavadi Anticline, north of Bagalkot. The white dotted line
marks the unconformity between the Kaladgi Supergroup and its basement. Several parts of this
image appear distorted due to the back-waters of the Almatti Reservoir that have flooded large parts
of the area and led to the relocation of the old Bagalkot town to the new location further south
called Navanagar. Inset depicts the outline of the exposures of the Saundatti Quartzite and the key
structural elements of the plunging anticline

Fig. 14 Generalised NW–SE cross section across the asymmetric Anagavadi anticline. Some of
the spot locations described below are marked for easy reference in context of the geometry of this
regional anticline. HKSB = Hundgund–Kushtagi Schist Belt (of the Basement Complex), SQ =
Saundatti Quartzite, MC = Mahakut Chertbreccia

merge (see Fig. 5) into beds of the Manoli Ferruginous Shale Member of Yadhalli
Formation upwards in the sequence. In this area, the Mahakut Chertbreccia occurs
between the Saundatti and Yadhalli Formations.
The Anagavadi anticline is a product of the D2 event of regional deformation
suffered by the Bagalkot Group. Even on a regional scale, the response of competent
and incompetent strata during deformation can be visualised. The quartzitic sand-
stones of the Saundatti Quartzite display a dominantly brittle deformation, dragging
along the transverse faults and localised, minor parasitic folding. The interbedded
siltstones display development of close-spaced jointing, weak slaty cleavage and
quasi-ductile deformation. The quasi-ductile deformation is more evident in the
348 S. Patil Pillai and V. S. Kale

surrounding argillaceous sediments (of Yadhalli Argillites) flanking the ridges of


the quartzites and chertbreccias. The result is that the quartzitic sandstones display
thickening at the fold closures and thinning (stretching) of the limbs even in the
regional scale map (see inset Fig. 13).

4.2.1 Location AA

This location provides a view of the metasedimentary sequence of the Hund-


gund–Kushtagi Schist belt (HKSB), which is considered the local equivalent of
the Neoarchean Dharwar Group. This is the local representative of the Basement
Complex of the Kaladgi Supergroup, exposed in the core of the Anagavadi anti-
cline. Banded haematite jasper and banded haematite chert comprise the Banded
Iron Formation (BIF) bands in this sequence that also contains quartz–mica and
quartz–chlorite bearing phyllites. The BIF horizons display tight isoclinal folding
with hinges striking 110°–290° (Fig. 15a, b).
Weakly foliated green-grey phyllites are exposed in this terrain, but fresh expo-
sures are difficult to find, due to their deeply weathered nature. The foliation of these
rocks display subvertical dips striking 100°–280°. The trend of regional folding and
deformation of the Neoarchean HKSB and that of the D2 folding of the overlying
Bagalkot Group during Mesoproterozoic are almost parallel to each other. It is diffi-
cult to explain how this parallelism was achieved, unless one assumes the reactivation
of the structures from the Archean basement during the later event. This could be a
good example of thin-skinned tectonics controlled by the geometry and orientation
of the inherited basement structures (see Jammes and Huismans 2012; ErdÓs et al.
2014; Misra and Mukherjee 2015).

Fig. 15 a BIF from the HKSB displaying isoclinals folds with hinges that have a dip of about
10°/290°. Hammer for scale is 60 cm long. b Close-up view of the tight isoclinal folding in the BIF
exposed at location AA . Note that the bands have been co-axially refolded and dislocated into their
present status
Traverses Through the Bagalkot Group from North Karnataka State … 349

4.2.2 Location AB

This is located ~100 m west of the preceding exposures and requires one to go uphill
towards the crest of the ridge. The angular and erosional unconformity between the
basement of HKSB (striking 100°–280°) and the thickly bedded Salgundi Conglom-
erate (dipping 35°/290°) is encountered at this location. Pebbles sized, oblate to
disc-shaped clasts of phyllites, jasper, vein quartz, BIF are embedded in dark grey,
poorly sorted clastic matrix in these conglomerate beds (Fig. 16a).
Further up in the sedimentary cover, the conglomerates interbedded within coarse-
grained quartzites, become progressively thinner and are eventually replaced by the
well stratified quartzitic sandstones of the Almatti Quartzite Member (Fig. 16b).

Fig. 16 a Polymictic thickly bedded Salgundi Conglomerate exposure showing clasts of BHJ
(banded haematite jasper), BHQ (banded haematite quartzite), Ch (chert) J (jasper), Ph (phyllite),
Q (quartz) and Qtzt (quartzite) embedded in a poorly sorted detrital matrix. Pen is 14 cm long.
b Quartzitic sandstones of Almatti Quartzite displaying thick beds with internal cross-stratification
and occasional gravelly and conglomerate interbeds. These beds dip by 30°/285° near the crest of
the ridge before location AC . The scale used is 5 cm long. c Steepening of the quartzites (striking
50°–230°) in the vicinity of a transverse fault near the Mallikarjun temple at location AC . Note the
quartz veins that strike parallel to the bedding planes in the foreground. Geology student Bhagyaraj
Kumari is 180 cm tall. d En-echelon sinusoidal gash veins filled by quartz displaying a right-lateral
sense of shear along the transverse fault at location AC
350 S. Patil Pillai and V. S. Kale

4.2.3 Location AC

A transverse fault striking 130°–310° cuts across the swirling limb of the Anagavadi
anticline at this location. The fracturing of the bedrock has supported a natural spring
on which the Mallikarjun Temple has been constructed here. The quartzitic beds in
this segment of limb have an attitude of ~40°/320°. This fault with a dominantly
left-lateral sense of shear has resulted in the dragging and steepening of the beds
to an orientation of 70°/340° in its flanks (Fig. 16c). Quartz veins filling up subver-
tical joints trending subparallel to the strike of bedding and en-echelon gash veins
(Fig. 16d) are profusely encountered transecting the quartzite beds along the fault.

4.2.4 Locations AD to AF

The purpose of proposing this location AD (on the fold axis) and the next two locations
AE and AF is to enable students understand how:

(a) The strike and dip of bedding changes from the hinge to the limbs of a fold.
In this case, the dips vary from 40°/320° (at the previous location AC ), through
~8°/360° (at AD ); 35°/250° (at AE ) and 52°/200° (at AF );
(b) Use these dips to generate the geometry of the fold and compute the orientation
of its axis and axial plane;
(c) Appreciate the variances in orientation of sympathetic fractures/joints that are
seen in the competent (quartzite) strata along the fold trace due to differences
in the local orientation of stress fields; and
(d) Provide a traverse along which it is possible to generate data of dip isogons so
as to enable the recognition of the Anagavadi anticline as Class1 C fold in terms
of Ramsay’s classification scheme (Twiss and Moores 1992).

While taking this traverse, sufficient caution should be exercised since the entire
length of ~4 km is along crest of a very rough hilly terrain and the rubbly exposures of
the quartzites can be sometimes treacherously unstable to stand on. Having said that,
this traverse offers an additional benefit of observing the variations in the internal
stratifications (both in terms of geometry and orientations) and a large variety of
primary sedimentary structures including ripple marks, desiccation cracks, etc.

4.2.5 Location AG

This is located ~500 m NW of Kadampur village on the flanks of the ridge and is
intended to study a small (~1.5 km long and 0.5 km wide) doubly plunging syncline
that is easily recognised on the satellite imageries. A careful observation of the dips
of the quartzites shows that they gradually reduce when coming south from the
preceding location AF . Although the strike of the bedding remains essentially NW–
SE in this sector, the dips change from 52° to 30° southwards (towards 200°) at AF ,
Traverses Through the Bagalkot Group from North Karnataka State … 351

and then gentler (~10°) further 100 m to the south. The dip of quartzite beds then
changes to 15°/25° within about 50 m and eventually becomes 28°/40° (at AG ). This
parasitic doubly terminating syncline in the quartzites is flanked further south by the
exposures of the Mahakut Chertbreccia.
The presence of quartz veins, local close-spaced shearing of the quartzites and
small patchy brecciation can also be observed in the quartzitic beds close to the axis
of the syncline. These are manifestations of the Anagavadi Shear, which cuts across
these exposures (as depicted in Fig. 13).

4.3 Lokapur Syncline

The Lokapur Syncline is a doubly plunging synclinorium (hosting several smaller


parasitic folds), which has an axis plunging towards east (in the west) and towards
west (in the east) as depicted in Fig. 17. It is arguably one of the easiest of the
several doubly plunging synclinoria present in the Bagalkot subbasin to study, being
unhindered by younger cover. All these synclinoria (see Fig. 7) host the Simikeri
Subgroup and are surrounded by the unconformably underlying Lokapur Subgroup.
The WNW–ESE trending 18.5 km long axis of this asymmetric syncline curves
towards an E–W trend in its central part before reverting to the original WNW–ESE
trend towards its other closure (see Fig. 17-Inset). This has been interpreted (Jadhav
1987) to be a manifestation of N–S oriented cross-folding. This is a tight fold with
average (based on >50 measurements) dips of the northern and southern limbs being
65°/195° and 58°/10°, respectively. Both the eastern and western closures of this fold
may be classified as 4D-4E transition as per Huddleston’s classification. Based on

Fig. 17 Google Earth image of the Lokapur Syncline. The white dotted line marks the unconformity
between the Lokapur and Simikeri Subgroups. Inset (adapted from Patil Pillai et al. 2018) depicts the
distribution of the formations of the Simikeri Subgroup within the syncline, while the surrounding
Lokapur Subgroup is undifferentiated. Locations of interest are marked as LA –LD
352 S. Patil Pillai and V. S. Kale

the dip isogon method, these hinges belong to the 1C Class with weakly convergent
to subparallel isogons in Ramsay’s scheme.
The thickening of the quartzite beds is evident at all the closures of the main
syncline as well as the parasitic folds. Using orthogonal thickness, Jadhav (1987)
had demonstrated that this synclinorium displays a flattening index of 70% and
belongs to Class 1C of Ramsay (1967). The orientation of the joint planes within the
Muchkundi Quartizites in the Lokapur syncline is a subject of independent study. The
faults with a dominantly vertical displacement and minor horizontal displacement
that cut across this syncline with a subvertical orientation display strike of:

(a) 30°–210° (e.g. fault running east of LC and cutting the southern limb at
16°08 59 N: 75°17 58 E),
(b) 50°–230° (e.g. exposed near a small reservoir at 16°08 33 N: 75°21 43 E),
(c) 140°–320° (e.g. exposed at 16°08 59 N: 75°21 43 E).

Such faults display brittle deformation (brecciation and fracturing) in the quartzites.
Several joint planes are lined by thin veins of silica cutting across them. The quasi-
ductile shear zones are mostly aligned parallel to the fold axis (WNW–ESE to E–
W) and appear to display a dominantly sinistral sense of movement. They display
close-spaced jointing in the quartzites, with conjugate sets of shear fractures whose
bisecting plane is aligned parallel to the shear plane. Based on the available data, it
can be shown that a homogeneous, NNE–SSW directed compressive strain to have
resulted in the Lokapur syncline. The parasitic folds and transverse faults suggest
a sinistral sense of deformation that is likely to be related to the NW–SE trending
regional shear zone/fault occurring north of this syncline.
Its geomorphic expression as a narrow elongated depressed ‘basin’ (Fig. 18)
fringed by low rugged ridges of the Muchkundi Quartzite can be visualised if one
stands on crest of the hill which is cut by the Belgaum–Bagalkot highway (KH SH20)
near Mallapur at location LA . In many ways therefore, this synclinorium offers and
excellent opportunity for students of structural geology to examine various features
on a macroscopic scale that can contribute to the kinematic analysis of ‘similar’
folding where competent (quartzite and chertbreccia) beds are interlayered between
incompetent (argillites and carbonates) strata. It is therefore recommended that (if
possible), a traverse be taken along entire quartzitic ridge surrounding the Lokapur
Syncline (as was done in case of the Anagavadi anticline), rather than studying
specific spots. The argillaceous and carbonate sediments are entirely covered with
a thick soil cover, and hence, it is difficult to encounter their fresh exposures for
studying structural characters. Rare exposures of these structurally incompetent strata
(in well-cuttings, road cuts and canal cuttings) provide clues to their quasi-ductile
deformation into mesoscopic folds with slaty cleavage development. Therefore, the
traverse LA to LB is essentially for the purpose of providing a glimpse of the rela-
tions between various lithostratigraphic units exposed in this synclinorium, while LC
describes an exposure of the mafic intrusive (Mallapur Intrusives) that cut into the
sediments. All these locations fall in SoI Toposheet No. 47P/8.
Traverses Through the Bagalkot Group from North Karnataka State … 353

Fig. 18 Perspective view of the Lokapur syncline (generated using the 3D views of Google Earth)
looking towards due east from the western closure of the fold. The trace of the axial plane of the
syncline and dips of quartzites recorded at different locations are marked for easy reference

4.3.1 Locations LA to LB

This traverse follows the Belgaum–Bagalkot state highway where it cuts across
the southern limb of the Lokapur syncline and continues north-eastwards towards
Lokapur town. The disconformable contact between the Chitrabhanukot Dolomite
Member (Lokapur Subgroup) and the basal beds of Tulasigeri Quarztite Member
(Simikeri Subgroup) is exposed on the southern side of the low hill range (Fig. 19).
Both these units dip 55°/10° in this exposure.

Fig. 19 Generalised SW–NE cross section across the asymmetric Lokapur Syncline along the
Belgaum-Bagalkot highway between Mallapur (on the left) and Lokapur (on the right). Some of
the spot locations are marked for easy reference in context of the geometry of this regional fold.
The Lokapur Subgroup is represented by the Petlur Carbonates comprising CD = Chitrabhanukot
Dolomite Member, CL = Chikshellikeri Limestone Member. The Simikeri Subgroup includes MQ
= Muchkundi Quartzite Formation, NC = Niralkeri Chertbreccia, AA = Arlikatti Argillite, LD =
Lakshanhatti Dolomite. The subvertical mafic dyke of Mallapur Intrusives (MI) occurs close to the
axial zone of the syncline
354 S. Patil Pillai and V. S. Kale

The former display presence of stromatolitic columns and algal laminated nature
on the banks of a canal that runs parallel to the quartzite ridge, south of it. The
Tulasigeri Quartzites have interbedded conglomerate lenses at its base. It is essen-
tially thickly parallel bedded sandstone (Fig. 20a) that has undergone extensive
diagenetic recrystallisation leading to its glassy appearance.
About 100 m north of the exposure of the disconformity, the rubbly and weath-
ered exposures of Niralkeri Chertbreccia are seen with fragmented clasts of quartzites
and dolomites embedded in cherty matrix. This rock displays very poor stratifica-
tion and has been mistaken for a fault breccia by few earlier workers. However,
its interbedded nature, conformable to the beds of the Simikeri Subgroup has been
established (Jayaprakash et al. 1987; Jayaprakash 2007) by subsequent studies. The
Niralkeri Chertbreccia occurs in these exposures between the Muchkundi Quartzite
and Arlikatti Argillite, but in other parts of the basin, it may be encountered at the

Fig. 20 a Thick parallel–planar bedded Tulasigeri Quartzite exposed in the ridge surrounding the
Lokapur syncline. The beds display dips of ~50° in most of its exposures in the syncline, except
along the fold closures. b Arlikatti Argillite exposed in a canal cutting displaying bedding planes
(S0 : yellow dashed lines) superimposed with slightly oblique slaty cleavages (S1 : white dashed
lines). Two sets of fractures (Fr1 & Fr2 ) are superimposed on the fissile shales. The photograph is of
an N-S oriented face of the canal cutting, with north on the right. c Algal laminated Lakshanhatti
Dolomite displaying parasitic mesoscale, asymmetric folding of the bedding planes (S0 : yellow
dashed lines). Note how the fold axes (F1 ) display a diversity in orientations indicating that the rock
has suffered a ductile deformation. The pen for scale is ~12 cm long
Traverses Through the Bagalkot Group from North Karnataka State … 355

top of the latter, due to its impersistent, interfingering nature of contact with the
enveloping lithologies (Kale and Patil Pillai 2011).
About 50 m across the strike, the chertbreccia beds are followed in the down-
dip direction by ferruginous siltstones and mudstones that display a pervasive slaty
cleavage (Fig. 20b). The bedding planes (S 0 ) and the slaty cleavage (S 1 ) have domi-
nantly E–W strike directions and steep dips of >50°. Along the limbs of this syncline,
they are almost parallel in orientation to each other, but as one approaches the
closures/hinge zones of the (main/parasitic) folds, the orientation of S 1 remains
essentially E–W, but the bedding planes swirl in their strike directions.
The Lakshanhatti Dolomite that follows next is an algal laminated dolomite
(Fig. 20c) with parallel laminated character. At places, they display chertification
of some of their layers, while stromatolitic columns have been known from them for
a long time (Viswanathiah and Chandrasekhara Gowda 1970). Interbedded detrital
siltstones and mudstones are encountered towards the base as well as the top of this
formation. These dolomitic rocks can be examined on the sides of the highway (LB )
and occur along the axis of the Lokapur syncline.

4.3.2 Locations LC

The canal cut west of the village of Dadanhatti exposes weakly foliated, chlorite–
quartz (±muscovite) bearing strata of the Arlikatti Argillite. They have a subvertical
disposition along the canal and display tight isoclinal folding with axial plane striking
in the 95°–275° direction. It is significant that although these exposures are close to
the axial trace of the Lokapur syncline, they appear to display a subvertical dispo-
sition. This is a good example of transposition of bedding and foliation of fissile
fine-grained rocks during intense deformation.
About 50 m north of the bridge on this canal is the exposure of the meta-basic dyke
(that has a whole rock 40 Ar/39 Ar age of 1154 ± 4 Ma) of the Mallapur Intrusives,
trending almost E–W, displaying a sharp contact with the foliated shales (Fig. 21).
The emplacement of this mafic dyke along the axial plane of the Lokapur syncline
has been interpreted to represent a post-deformational event, followed by a second
phase of deformation and anchizone metamorphism of the Bagalkot Group.

4.3.3 Location LD

This location is suggested to demonstrate how the competent (quartizitic) rocks


suffer intense brittle deformation at the closure of a plunging syncline tight isoclinal
fold. It may be approached by the village road that connects Dadanhatti–Hoskoti
to Yadwad. As visualised in Fig. 18, the western closure of the Lokapur anticline
is marked by ridges of Muchkundi Quartzite standing on either side of the narrow
valley where the Arlikatti Argillite are exposed. The southern limb dips by 75°/200°
to 81°/210°; while the northern limb has slightly less steeper dips (65°/190°) but
in the same direction. The exposures of quartzites along the fold axis have been
356 S. Patil Pillai and V. S. Kale

Fig. 21 Contact between the Arlikatti Argillite and the dyke from the Mallapur Intrusives. Note the
spheroidal weathering and large blocks of the mafic dyke material in contrast to the flaky exposures
of the shales with weak foliation. Chlorite development has been observed in thin sections of these
shales. Photograph from Patil Pillai et al. (2018)

intensely brecciated and hence are easily eroded away. Anthropogenic activity has
further destroyed the continuity between the northern and southern limbs of this tight
isoclinal syncline at this location.
The axial plane of the fold at this location has an orientation of 72°/200°, and
the axis dips by 10°/85°. These values of the fold axis and axial plane have been
deciphered by plotting the dips measured in the quartzites along the ridge line in
the vicinity of this fold closure on a stereonet. This exercise of how to interpret the
fold orientations using field measurements and plotting them on the stereonet can be
conducted at this location.

4.4 Yadwad Folds

This sector SW of Mudhol in the Bagalkot subbasin is one where very tightly folded
sequence of argillites and carbonates of Lokapur Subgroup is exposed (Fig. 22).
These exposures are largely discontinuous in nature, due to the cover of the Deccan
Trap basalts (see: Fig. 1) and the extensive capping of the black-cotton soil that makes
this a very fertile terrain. The extensive agricultural activity has obliterated the conti-
nuity in the exposures significantly. However, low, narrow ridgelets of the resistant
Muchkundi Quartizite crop out as curvilinear bands within this terrain (Fig. 23) and
help understand the nature of the folding and deformation that can be studied and
validated by field studies. These exposures fall in SoI Toposheet Nos. 47P/3 and
47P/4.
Traverses Through the Bagalkot Group from North Karnataka State … 357

Fig. 22 Enlarged geological cross section (from Fig. 6) between Dadanhatti (from the Lokapur
Syncline) and Mudhol depicting the regional setting of the Yadwad folds. The folds described in
details here (#4.3) occur about 5 km west of this section line. Colour index for the lithostratigraphic
units is the same as in Fig. 6

Fig. 23 Bedding plane traces (S0 : white dotted lines) and fold axes (S1 : orange dashed lines)
highlighted by exposures of the Muchkundi Quartzite, in Google Earth image of the area around
Yadwad. Transverse shear fractures (Fr1 ) cutting across the folded sequence are represented by
psuedotachylitic and silica (often quartz) veins and local dragging

4.4.1 Competency Contrast

This sector provides excellent examples of how competent (quartzites) and incompe-
tent (argillites and carbonates) strata have responded differently to the deformation.
Without suggesting specific spot locations in this region, this aspect is described
below and can be studied during the field studies in the terrain around Yadwad.
Even on a mappable scale, the amplitudes and wavelengths of the fold geom-
etry vary significantly. To appreciate this, one may compare the fold patterns that
are mappable in the quartzitic bands (Fig. 23) with those in the Petlur Carbonates
(Fig. 24) exposed south of Yadwad. Folds in the quartzites display (mostly plunging
by 10°–15° towards the east, unless they are doubly plunging) wavelengths of 1–
2 km and amplitudes with similar magnitudes. Their limbs are normally stretched
358 S. Patil Pillai and V. S. Kale

Fig. 24 Bedding plane traces (S0 : white dotted lines) and fold axes (S1 : orange dashed lines)
observed within the Petlur Formation south of Yadwad. Field studies have validated these tight
folds faintly outlined in the Google Earth image

across several kilometres in contrast to their closures with diameters of <2 km in


this area. They display thickening at the closures and stretching along the limbs. The
response of the quartzitic bands to the deformation is no different in this sector than
that elucidated earlier (e.g. Figure 3, 8a) and described in the previous traverses. The
exposures of the quartzites are largely obliterated by cultivation, and hence good,
insitu features are not easily encountered in this sector. On the other hand, folds in the
incompetent strata are smooth, with mappable wavelengths of 400–700 m and rarely
display crestal thickening. They have a more symmetric and harmonic geometry in
comparison (as evident in Fig. 24). Their fold axes, with congruent strikes in the
100°–280° orientations, normally plunge by up to 25° towards east or west.

4.4.2 Mesoscopic Folding

The following are typical examples of the mesoscopic folding encountered in the
argillaceous and carbonate sediments around Yadwad, which can also be seen in other
exposures across the folded Bagalkot Group. A traverse across any such exposure will
enable students to examine similar features that are easy to observe in the Yadwad
area.
Each of the following examples (Figs. 25, 26, 27 and 28) that display diversity in
the geometry of the folding is depicted by a photograph and explanatory title with
it. It is suggested that the orientations and geometry of these features of ductile and
quasi-ductile deformation be studied while taking traverses around Yadwad and a
compilation of the same undertaken for the purpose of kinematic analysis of the
deformation.
Traverses Through the Bagalkot Group from North Karnataka State … 359

5 Concluding Remarks

It will be evident from the traverses described above that the Bagalkot Group presents
a wide spectrum of regional, macroscopic, mesoscopic and megascopic structures
that make it an excellent field laboratory for teaching/learning structural geology. It
is noted that only a sample of these structures is elucidated in the foregoing chapter.
Quality of exposures and the break in their lateral continuity due to soil cover or
anthropogenic activity is normal to expect in any region. The challenge is to record
observations on various scales and compile and collate them into a coherent structural
analysis. The exposures of the Bagalkot Group in the Kaladgi Basin is no different
from any other terrain in that respect, but has the benefit of easy access and facilities
for camping during the field studies. It is hoped that future studies will add to what
is already recorded and enable an exhaustive, comprehensive kinematic analysis of
a Mesoproterozoic sedimentary sequence that has suffered a regional deformation.

Fig. 25 Slab of Chikshellikeri limestone displaying mesoscopic folds (F1 ) with an average axial
plunge towards 15°/280°. The average dips of the bedding planes (S0 ) is 20°/265° at the closure of
a macroscopic fold. Note the convergent arrangement of the axial plane fractures (Fr2 )
360 S. Patil Pillai and V. S. Kale

Fig. 26 Sinusoidal folds developed in strata of common competency interlayered with less compe-
tent layers that show a different class of folding. The class of each of the smaller fold closures
is determined using dip isogons. Such variance is typical of similar folds encountered in the core
zones of orogens (Twiss and Moores 1992)

Fig. 27 Dissimilar folds developed in alternate bands of pure and impure limestones. Note the
competent bands (C) of impure limestones display close-spaced axial plane cleavage development
(Fr1); while the incompetent (I) strata displays pure ductile folding. Simplified sketch given as inset
for reference. The scale is 11 cm long
Traverses Through the Bagalkot Group from North Karnataka State … 361

Fig. 28 Sinusoidal folding (pseudo-ptygmatic) in layers of pure limestones interbedded within


fissile beds of impure limestones. Note how the geometry of folds has changed in the thinly bedded
impure limestones on tip. Disruption of the underlying strata leading to development of boudinages
in the lower strata is also evident. Compare this with the deformation observed in Fig. 8c

Acknowledgements The authors thank Soumyajit Mukherjee for inviting this contribution, and
for coordinating two rounds of review that helped improve the contents significantly. Pradeep
Jadhav is particularly thanked for providing his insights and some of his (doctoral thesis) personal
data and photographs for this compilation. Besides Devdutt Upasani, several of our colleagues
and co-workers and our teachers Prof. V. V. Peshwa and late Prof. A. V. Phadke are thankfully
acknowledged for their support during field studies and discussions. Studies in the Kaladgi Basin
were funded from time to time by agencies like Department of Science and Technology (New Delhi),
Atomic Minerals Directorate (Hyderabad), and Council for Scientific and Industrial Research (New
Delhi) under various projects in the past. Thanks to Marion Schneider, Annett Buettener, Boopalan
Renu, Alexis Vizcaino, Doerthe Mennecke-Buehler and the proofreading team (Springer).

References

Allison, D. T. (2004). Structural geology laboratory manual (3rd ed.). NY: Prentice Hall.
Awati, A. B., & Kalaswad, S. (1978). Stucture of the Kaladgis around Yadwad, Belgaum district
(A study based on Landsat-1). Bulletin Earth Sciences, 6(1), 43–47.
362 S. Patil Pillai and V. S. Kale

Chakraborty, P. P., Sarkar, S., & Patranabis-Deb, S. (2012). Tectonics and sedimentation of
Proterozoic basins of Peninsular India. Proceedings Indian National Science Academy, 78,
393–400.
Davis, G. J., Reynolds, S. J., & Kluth, C. F. (2012). Structural Geology of rocks and regions (3rd
ed., p. 839). Wiley.
Dey, S. (2015). Geological history of the Kaladgi–Badami and Bhima basins, south India: sedimen-
tation in a Proterozoic intracratonic setup. In R. Mazumder, P. G. Eriksson (Eds.), Precambrian
Basins of India: Stratigraphic and Tectonic Context. Geological Society London Memoir, 43,
283–296. https://doi.org/10.1144/m43.19.
Dίaz-Axpiors, M., Fernández, C., & Czeck, D. M. (2019). Are we studying deformed rocks in
the right sections? Best practices in kinematic analysis of 3D deformation zones. Journal of
Structural Geology, 125, 218–225. https://doi.org/10.1016/j.jsg.2018.03.005.
Erdős, Z., Huismans, R. S., van der Beek, P., & Thieulot, C. (2014). Extensional inheritance
and surface processes as controlling factors of mountain belt structure. Journal of Geophysical
Research: Solid Earth, 119, 9042–9061. https://doi.org/10.1002/2014jb011408.
Fleuty. M. J. (1964). The description of folds. Proceedings of Geological Association, 75 (4),
461–492.
Foote, R. B. (1876). Geological features of the south Maharatta country and adjoining districts.
Memoir Geological Survey of India, 12, 70–138.
Fossen, J. (2011). Structural Geology (p. 463). Cambridge University Press.
Geological Survey of India. (1981). Geological and Mineral Map of Karnataka and Goa on 1:0.5
million scale. Calcutta: GSI.
Gokhale, N. W., & Pujar, G. S. (1989). Bedding plane faults in the Kaladgi rocks, Basidoni, Belgaum
district, Karnataka state. Current Science, 58, 1088–1089.
Hudleston, P. J. (1973). Fold morphology and some geometric implications of theories of fold
development. Tectonophysics, 16, 1–46.
Jadhav, P. B. (1987). Microstructures of quartzites of Yadwad—Lokapur—Bagalkot area, Bijapur
district, North Karanataka (p. 204). Unpublished Ph.D. thesis, Pune University.
Jadhav, P. B., & Kshirsagar, L. K. (2017). Analysis of Folding in the Proterozoic Sedimentary
Sequence of Bagalkot Group Exposed around Yadwad, N Karnataka, India: A Study Based on
Remote Sensing Techniques and Field Investigations. Journal Indian Society of Remote Sensing,
45(2), 247–258. https://doi.org/10.1007/s12524-016-0583-4.
Jammes, S., & Huismans, R. S. (2012). Structural styles of mountain building: Controls of litho-
spheric rheologic stratification and extensionalinheritance. Journal of Geophysical Research:
Solid Earth, 117, B10403. https://doi.org/10.1029/2012JB009376.
Jayananda, M., Chardon, D., Peucat, J.-J., Tushipokla, T., & Fanning, C. M. (2015). Paleo- to
Mesoarchean TTG accretion and continental growth in the western Dharwar craton, Southern
India: constraints from SHRIMP U-Pb zircon geochronology, whole-rock geochemistry and
Nd-Sr isotopes. Precambrian Research, 268, 295–322. https://doi.org/10.1016/j.precamres.2015.
07.015.
Jayaprakash, A. V. (2007). Purana Basins of Karnataka. Memoir Geological Survey of India, 129,
136.
Jayaprakash, A. V., Sundaram, V., Hans, S. K., & Mishra, R. N. (1987). Geology of the Kaladgi-
Badami Basin, Karnataka. Memoir Geological Society of India, 6, 201–225.
Joy, S., Patranabis-Deb, S., Saha, D., Jelsma, H., Maas, R., Soderlund, U., et al. (2018). Depo-
sitional history and provenance of cratonic ‘Purana’ basins in southern India: a multipronged
geochronology approach to the Proterozoic Kaladgi and Bhima basins. Geological Journal.
https://doi.org/10.1002/gj.3415.
Kale, V. S. (1991). Constraints on the evolution of the Purana Basins of Peninsular India. Journal
Geological Society of India, 38, 231–252.
Kale, V. S. (2016). Proterozoic basins of Peninsular India: status within the global Proterozoic
systems. Proceedings Indian National Science Academy, 82(3), 461–478. https://doi.org/10.
16943/ptinsa/2016/48461.
Traverses Through the Bagalkot Group from North Karnataka State … 363

Kale, V. S., Dole, G., Shandilya, P., & Pande, K. (2019). Stratigraphy and correlations in Deccan
Voclanic Province, India: Quaovadis? Geological Society America Bulletin, 132, 588–607. https://
doi.org/10.1130/B35018.1.
Kale, V. S., Ghunkikar, V., Paul Thomas, P., & Peshwa, V. V. (1996). Macrofacies architecture of
the first transgressive suite along the southern margin of the Kaladgi Basin. Journal Geological
Society of India, 48, 75–92.
Kale, V. S., Nair, S., & Patil, S. (1998). Testimony of intraformational limestone breccias on Lokapur-
Simikeri disconformity, Kaladgi Basin. Journal Geological Society of India, 51, 43–48.
Kale, V. S., & Patil Pillai, S. (2011). A Reinterpretation of two Chertbreccias from the Proterozoic
Basins of India. Journal Geological Society of India, 78, 429–445. https://doi.org/10.1007/s12
594-011-0112-6.
Kale, V. S., Patil Pillai, S., Nair Pillai, S., Phansalkar, V. G., & Peshwa, V. V. (1999). A process-
responsive litho-stratigraphic classification of the Kaladgi Basin, Karnataka. In Abstract Volume,
Field workshop on “Integrated evaluation of the Kaladgi & Bhima Basins (pp. 4–7). Bangalore:
Geological Society of India.
Kale, V. S., & Phansalkar, V. G. (1991). Purana basins of Peninsular India: A review. Basin Research,
3, 1–36.
Kalpana, M. S., Patil, D. J., Dayal, A. M., & Raju, S. V. (2010). Near Surface Manifestation of
Hydrocarbons in Proterozoic Bhima and Kaladgi Basins: Implications to Hydrocarbon Resource
Potential. Journal Geological Society of India, 76, 548–556. https://doi.org/10.1007/s12594-010-
0115-8.
Korisettar, R. (1995). Review of ‘Man-land relationships during Palaeolithic times in the Kaladgi
basin, Karnataka’, by R.S. Pappu and S.G. Deo. Man and Environment XX, 1, 123–125.
Korisettar, R, & Petraglia, M. D. (1993.) Explorations in the Malaprabha Valley (XVIII, pp. 43–48)
Karnataka: Man and Environment XVIII.
Krapez, B., Srinivasa Sarma, D., Ram Mohan, M., McNaughton, N. J., Rasmussen, B., & Wilde,
S. (2020). Tectonostratigraphy of the Late Archean Dharwar Supergroup, Dharwar craton, India:
defining a tectonic history from spatially linked but temporally disconnected intracontinental and
arc-related basins. Earth Science Reviews, 201, 102966. https://doi.org/10.1016/j.earscirev.2019.
102966.
Kulkarni, K. G., & Borkar, V. D. (1997). Ichnogenus Paleophycus Hall from the Bagalkot Group,
Karnataka state. Journal Geological Society of India, 49, 215–220.
Kumar, A., Parasharumulu, V., & Nagaraju, E. (2015). A 2082 Ma radiating dyke swarm in the
Eastern Dharwar Craton, southern India and its implications to Cuddapah basin formation.
Precambrian Research, 266, 490–505. https://doi.org/10.1007/s12040-018-0940-5.
Mallick, K., Vasanthi, A., Sharma, K. K. (2012). Kaladgi—Badami Basin, Karnataka—Maha-
rashtra, India: In Bouguer Gravity Regional and Residual Separation: Application to Geology
and Environment (pp. 60–67). New Delhi: Capital Publishing Co., Springer.
Matin, A. (2006). Structural anatomy of the Kushtagi schist belt, Dharwar craton, south India—An
example of Archaean transpression. Precambrian Research, 147, 28–40. https://doi.org/10.1016/
j.precamres.2006.01.013.
Misra, A. A., & Mukherjee, S. (2015). Tectonic inheritance in continental rifts and passive margins
(p. 88). Springer. https://doi.org/10.1007/978-3-319-20576-2.
Mukherjee, S. (2014). Review of flanking structures in meso- and micro-scales. Geological
Magazine, 151, 957–974. https://doi.org/10.1017/S0016756813001088.
Mukherjee, S. (2017). Shear heating by translational brittle reverse faulting along a single, sharp
and straight fault plane. Journal of Earth System Science, 126(1), https://doi.org/10.1007/s12040-
016-0788-5.
Mukherjee, S. (Ed) (2019). Teaching methodologies in structural geology and tectonics (p. 251).
Springer. https://doi.org/10.1007/978-981-13-2781-0.
Mukherjee, M. K., Das, S., & Modak, K. (2016). Basement-cover structural relationships in the
Kaladgi Basin, southwestern India: indications towards a Mesoproterozoic gravity gliding of the
364 S. Patil Pillai and V. S. Kale

cover along a detached unconformity. Precambrian Research, 281, 495–520. https://doi.org/10.


1016/Jour.precamres.2016.06.013.
Mukherjee, M. K., Modak, K., & Ghosh, J. (2019). Illite crystallinity index from the Mesoprotero-
zoic sedimentary cover of Kaladgi basin, southwestern India: Implications on crustal depths of
subsidence and deformation. Journal of Earth System Science, 128(101), 15. https://doi.org/10.
1007/s12040-19-1124-7.
Mukhopadhyay, S., Choudhuri, A., Chakraborty, N., & Sarkar, S. (2019). Aseismic tectonism
induced soft-sediment deformation in a tranquil palaeogeography: Chikkshelikere Limestone
Member, Proterozoic Kaladgi basin, southern India. In M. E. A. Mondal (Ed.), Geological evolu-
tion of the Precambrian Indian Shield (pp. 351–371). Springer. https://doi.org/10.1007/978-3-
319-89698-4_15.
Nair, M. M., & Raju, A. V. (1987). A Remote Sensing approach to basinal mapping of the Kaladgi and
Badami sediments of Karnataka state. In B. P. Radhakrishna (Ed.), Purana Basins of Peninsular
India (Middle to Late Proterozoic). Memoir Geological Society of India, 6, 375–381.
Pappu, R. S. (2001). Acheulian culture in peninsular India: an ecological perspective (p. 170). DK
Printworld: New Delhi.
Patil Pillai, S., George, B. G., Ray, J. S., & Kale, V. S. (2019). Comments on paper: Deposi-
tional history and provenance of cratonic “Purana” basins in southern India: a multipronged
geochronology approach to the Proterozoic Kaladgi and Bhima basins. Geological Journal.
https://doi.org/10.1002/gj.3589.
Patil Pillai, S., & Kale, V. S. (2011). Seismites in the Lokapur Subgroup of the Proterozoic Kaladgi
Basin, South India: A testimony to syn-sedimentary tectonism. Sedimentary Geology, 240, 1–13.
https://doi.org/10.1016/j.sedgeo.2011.06.013.
Patil Pillai, S., Kale, V. S. (2019). Interplay between tectonics and eustacy in a Proterozoic epicra-
tonic polyhistory basin, North Dharwar Craton. In S. Mukherjee (Ed.), Tectonics and Structural
Geology: Indian context (pp. 75–114). Springer. https://doi.org/10.1007/978-3-319-99341-6_4.
Patil Pillai, S., Pande, K., & Kale, V. S. (2018). Implications of new 40Ar/39Ar age of Mallapur
intrusives on the chronology and evolution of the Kaladgi Basin, Dharwar Craton India. Journal
of Earth System Science, 127(32), 18. https://doi.org/10.1007/s12040-018-0940-5.
Raha, P. K., & Sastry, M. V. A. (1982). Stromatolites and Precambrian stratigraphy of India.
Precambrian Research, 18, 293–318.
Rajaram, M., Anand, S. P., Erram, V. C., & Shinde, B. N. (2017). Insight into the structure below the
Deccan Trap covered region of Maharashtra, India from geopotential data. In S. Mukherjee, A. A.
Misra, G. Calvès, M. Nemčok (Eds.), Tectonics of the deccan large igneous province: Geological
society (Vol. 445, pp. 219–236). London, Special Publications. https://doi.org/10.1144/SP445.8.
Ramakrishnan, M., & Vaidyanadhan, R. (2010). Geology of India (Vol. 1, p. 556). Bangalore:
Geological Society of India.
Ramsay, J. G. (1967). Folding and fracturing of rocks. New York: McGraw Hill.
Ramsay, J. G., & Huber, M. I. (1987). The techniques of modern Structural Geology, Vol.2: Folds
and fractures. London: Academic Press.
Samal, A. K., Srivastava, R. K., Ernst, R. E., & SÓnderlund, U. (2019). Neoarchean—Mesoprotero-
zoic mafic dyke swarms from the Indian shield mapped using Google Earth images and ArcGIS,
and links with large igneous provinces. In R. K. Srivasata (Ed.), Dyke swarms of the World: A
modern perspective (pp. 355–390). Springer. https://doi.org/10.1007/978-981-13-1666-1_9.
Sharma, M., Nair, S., Patil, S., Shukla, M., & Kale, V. S. (1998). Occurrence of tiny digitate
stromatolite [YelmaDigitata Grey, 1984], Yargatti Formation, Bagalkot Group, Kaladgi Basin,
Karnataka India. Current Science, 75(4), 360–365.
Sridhar, M., Chaturvedi, A. K., & Rai, A. K. (2014). Locating new Uranium occurrences by inte-
grated weighted analysis in Kaladgi basin, Karanataka. Journal Geological Society of India, 84,
509–512. https://doi.org/10.1007/s12594-014-0159-2.
Sridhar, M., Markandeyulu, A., & Chaturvedi, A. K. (2017). Mapping subtrappean sediments and
delineating structure with the aid of heliborne time domain electromagnetics: Case study from
Traverses Through the Bagalkot Group from North Karnataka State … 365

Kaladgi basin, Karanataka. Journal of Applied Geophysics, 136, 9–18. https://doi.org/10.1016/j.


jappgeo.2016.10.024.
Twiss, R. J. (1988). Description and classification of folds in single surfaces. Journal Structural
Geology, 10, 607–623.
Twiss, R. J., & Moores, E. M. (1992). Structural Geology (p. 532). Freeman and Co.
Valdiya, K. S. (2016). The making of India: Geodynamic Evolution (2nd ed., p. 924). Springer
International: Switzerland.
Viswanathiah, M. N. (1968). Badami Series: a new post-Kaladgi formation of Karnatak state.
Bulletin Geological Society of India, 5, 94–97.
Viswanathiah, M. N. (1979). Lithostratigraphy of the Kaladgi and Badami Groups, Karnataka.
Indian Mineralogist, 18, 122–132.
Viswanathiah, M. N., & Chandrasekhara Gowda, M. N. (1970). Algal stromatolites from Kaladgi
(Precambrian) formations near Alagundi, Bijapur district, Mysore state. Journal Geological
Society of India, 11, 378–385.
Viswanathiah, M. N., & Sreedhara Murthy, T. R. (1979). Algal stromatolites from Kaladgi Group
around Bilgi, Bijapur district, Karnataka. Journal Geological Society of India, 20, 1–6.
Whitemeyer, S. J., Pyle, E. J., Pavlis, T. L., Swanger, W., & Roberts, I. (2019). Modern approaches
to field data collection and mapping: digital methods, crownsourcing and the future of statistical
analyses. Journal of Structural Geology, 125, 29–40. https://doi.org/10.1016/j.jsg.2018.06.023.
Yogesh, M., Deo, S. G., & Mathews, S. (2001). The Lithic Assemblage from the Kovalli Site,
Ghataprabha Basin, Karnataka: A Re-Examination. Puratattva, 42, 74–86.
Tectonic Framework of Northern
Pakistan from Himalaya to Karakoram

Asghar Ali, Sajjad Ahmad, Sajjad Ahmad, Mohammad AsifKhan,


Muhammad Irfan Khan, and Gohar Rehman

Abstract Orogenesis at continental margins and within oceanic crusts depends on


the conversion of a subducting plate. The Andean/Cordilleran-type orogenic belts
and magmatism form by subduction of a denser oceanic crust beneath a conti-
nental crust. Ocean–ocean convergence associated plutonic and volcanic magmatism
culminated into an island arc. Complete subduction of an intervening oceanic crust
leads to collision orogenic belts and suture zones. Suture zones contain remnants
of almost completely consumed intervening oceanic crust in the form of ophio-
lites. Tectonic exhumation colligated with orogenic belts juxtaposes different grade
metamorphic, upper mantle and deep crustal rocks following the closure of the
intervening ocean. The 2500 km long E–W trending Himalaya, Kohistan Island
Arc and adjacent Karakoram Orogenic Belt that rampart the northern edges of
Pakistan provide an ideal opportunity to examine Andean/Cordilleran, Island Arc
and continent–continent orogenic belts and diverse geological processes associated
with their formation. In this chapter, we communicate to the readers to canvas the
Sub-Himalaya, Greater Himalaya, remnants of the southern and northern Neo–Tethys
that have been respectively closed along the Main Mantle Thrust/Indus Suture Zone
and Main Karakoram Thrust/Shyok Suture Zone, entire Kohistan Island Arc, Main
Karakoram Thrust, Karakoram microplate, the most dynamic/active plate boundary
processes, regional shear zones, the continent–ocean transition, the influence of
mantle dynamics paradigm and the role of tectonic inheritance.

A. Ali (B) · G. Rehman


Department of Geology, University of Peshawar, Khyber Pakhtunkhwa, Pakistan
e-mail: asghar.ali@uop.edu.pk
S. Ahmad
Associate Professor, Department of Geology, University of Peshawar, Khyber Pakhtunkhwa,
Pakistan
S. Ahmad
Professor, Department of Geology, University of Peshawar, Khyber Pakhtunkhwa, Pakistan
M. AsifKhan
National Centre of Excellence in Geology, University of Peshawar, Khyber Pakhtunkhwa, Pakistan
M. I. Khan
MOL Pakistan Oil and Gas, Islamabad, Pakistan

© Springer Nature Switzerland AG 2021 367


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_12
368 A. Ali et al.

Abbreviations

MMT Main Mantle Thrust


KIA Kohistan Island Arc
UHP Ultra–high pressure
MKT Main Karakorum Thrust
MBT Main Boundary Thrust

1 Introduction

Around 2500 km long E–W trending Himalaya and adjacent Karakoram Orogenic
Belt that rampart the northern edges of Pakistan provide an ideal opportunity to
examine diverse geological processes associated with continent–continent collisional
belts. This study is an attempt to understand tectonic imprints of the Indian and
the Kohistan Island Arc (KIA) on the northernmost passive continental margin of
the Indian Plate and southernmost margin of the KIA as a consequence of final
closure of the Neo–Tethys at ~50 Ma and development of the Main Mantle Thrust
(MMT), which is equivalent of the Indus Suture Zone that extends from Afghanistan
to Myanmar (Fig. 1). These tectonic imprints will be considered in a practicable
and interesting way to learn about crustal thickening related magmatism, tectonos-
tratigraphy, middle/lower crust softening, channel flows, metamorphism, doming,
exhumation of very deep-seated rocks in the regional syntaxial structures, basal
detachment faults (decollements) and gravitational collapse, etc. The field will also
strengthen the technical knowledge of readers of diverse geological processes associ-
ated with the closure of the intervening Neo–Tethys between the KIA and Karakoram
microplate at Main Karakoram Thrust (MKT) in the Late Cretaceous (Fig. 2). The
Indus and the Nanga Parbat (~8125 m) syntaxial structures that actively culminate
along the MMT are the ideal locations to study the lower crustal orthogneisses
and paragneisses basement rocks of the Indian Plate that have been travelled all
the way up in the crust by exhumation processes, ductile to brittle shear zones,
younger leucogranites, regional unconformities and wrapping of the MMT around
these syntaxes.
The readers will explore the links between crustal thickening, rate of erosion,
metamorphism, magmatism, thermal and tectonic processes that control the ongoing
orogenesis of the gigantic Himalayas and Karakoram. This chapter is important for
bringing together new and longtime researchers of the Himalayas and Karakoram
region from Pakistan and elsewhere, facilitating the sharing of ideas as well as helping
to bind together the national and international multidisciplinary researchers actively
working on active tectonics, tectonic related hazards, seismology, earth surface
processes and many more associate with this orogenic belt.
From Islamabad to Khunjerab along the ~775.5 km Karakorum Highway,
the participants will be able to examine the Sub-Himalaya, Greater Himalaya,
Tectonic Framework of Northern Pakistan from Himalaya … 369

Fig. 1 a, b Regional geological map showing the main stratigraphic and structural components of
the Himalayan Orogenic Belt. The map is the outcome of ~150 years of active research on the
Himalayan Range(c) South to the north of the Himalayas has been divided into the Sub-Himalaya,
the Lesser Himalaya and the Greater Himalaya. These three regional tectonic terrains crop out as
imbricated thrust slices, which collectively accommodated ~2000 km shortening since the Indian
Plate collision with the southernmost Cimmerian blocks and Eurasian Plate. The Main Frontal Thrust
carries southward the rocks of the Sub-Himalaya and thrust them over the Ganga Foreland Basin.
The MBT separates the Sub-Himalaya from the Lesser Himalaya. North of the MBT is the Main
Central Thrust, which thrusts the Greater Himalayan rocks over the Lesser Himalaya. The Indus
Suture Zone in the north separates the southernmost Cimmerian blocks from the Greater Himalaya
rocks. The northernmost margin of the Greater Himalaya is overlain by the Tethyan Himalayan
Sequence. The basement rocks of the Indian Plate comprising of orthogneisses, paragneisses and
eclogites expose in the Nanga Parbat and in the Indus Syntaxes. North to south foreland fold-thrust
system propagation is eminent from timing of progressive stacking of imbricate system associated
with the ongoing Himalayan Orogeny (after Yin 2006; Frisch et al. 2011)

remnants of the Southern and Northern Neo–Tethys that have been respectively
closed along the MMT/Indus Suture Zone and MKT, entire KIA, MKT, Karakoram
microplate, the most dynamic/active plate boundary processes, regional shear zones,
the continent–ocean transition, influence of mantle dynamics paradigm and the role
of tectonic inheritance (Figs. 3a, b).
370 A. Ali et al.

Fig. 2 Full globe projection of the Pangaea from the Early Permian to Early Triassic, Panthalassa
Ocean, Paleo to Neo–Tethys Oceans and opening of the Atlantic Ocean. It explains separation
of the Iran, Afghan, Karakoram and Qiangtang Cimmerian blocks in the Early Permian from the
pre-Gondwana fringe and opening of the Neo–Tethys. These Cimmerian blocks accreted to the
southernmost active margin of the Eurasian Plate and closed the Paleo–Tethys in the Late Triassic to
Early Jurassic. Therefore, the sedimentary cover sequence classically records the Paleo–geographic
transit of the Karakoram (after Zanchi and Gaetani 2011)
Tectonic Framework of Northern Pakistan from Himalaya … 371

Fig. 3 a Detailed geological road map from Islamabad to Chilas (~576.5 km). The map showing
major stratigraphy and structural features of the Indian Plate and parts of the KIA (modified after
Khan et al. 2000). b Detailed regional geological road log map from Chilas to Khunjerab Pass
(~199 km). The map shows a major stratigraphy and structural features of the northern part of the
KIA and Karakoram (modified after Khan et al. 2000)
372 A. Ali et al.

Fig. 3 (continued)
Tectonic Framework of Northern Pakistan from Himalaya … 373

2 Regional Geology

2.1 Indian Plate

Triassic to Eocene cover and basement sequence along the northernmost passive
continental margin of the Indian Plate are multiply metamorphosed and deformed
as a result of the Indian Plate collision with the KIA in Pakistan and Gondwana
affinity blocks to the east over the past 70–50 ma (Yin and Harrison 2000). These
rocks are culminated into the Himalayan–Tibetan Orogenic Belt in a continent–conti-
nent collision tectonic set up (Fig. 1). The collision caused large scale S–directed
thrusting of the KIA along the north-dipping MMT, ophiolite emplacement, mid-
crust softening, gravitation collapse, ~470 km shortening and exhumation of deep-
seated high-grade rocks of the Indian Plate following the closure of the Neo–Tethys
(Figs. 4, 5 and 6; Coward and Butler 1985; Treloar et al. 2003, 2007; DiPietro et al.
2008). South of the MMT the Indian Plate divided into internal zone equivalent of
the Greater Himalaya and external zone equivalent of the Lesser Himalaya. Meta-
morphism and deformation within the internal zone, which is located between the
MMT and Panjal/Khairabad Thrust progressively increase from south to north. The
northern part between the Malakand fault and the Nanga Parbat is characterized by
lower to upper amphibolite grade metamorphic rocks. In the eastern region of the
internal zone, ultra-high pressure (UHP) eclogite bearing gneisses are exposed in
Naran (Fig. 7). The protolith of these UHP rocks that metamorphosed at 580–600 °C
and > 23 Kbar at ~47–46 Ma was the basement rocks of the Indian Plate below the
north-dipping MMT (Fig. 7f; Treloar et al. 2003; Yin 2006). Rapid exhumation of the
UHP rocks are accompanied by greenschist facies metamorphism at ~4 Kbar between
46–40 Ma at ~40–30 mm year−1 exhumation rate (Tonarini et al. 1993; Treloar et al.
2003). The internal zone divided into three blocks, the western Malakand block,
which comprises the Chakdara Granitic Gneisses and Mekhband–Pinjkora schistose
rocks are separated by the Panjal–Khairabad block from the Banna block in the
northeast (Fig. 4; Pogue et al. 1999; DiPietro 2008). The Panjal–Khairabad block
consists of the Early Precambrian to Mesozoic rocks and towards the east includes
the Kotah dome, Loe Sar dome and the Indus Syntaxis. The Upper Proterozoic and
Paleozoic rocks are not evenly distributed across the Panjal–Khairabad block as a
consequence of pre–Himalayan normal faulting and associated erosion (Pogue et al.
1999). The Panjal–Khairabad block is bound by the Indus Melange Zone to the west
and east. The melange zone is disrupted in the centre by the Indus Syntaxis (Fig. 4).
The core of the Indus Syntaxis is characterized by exhumation of the Early Protero-
zoic Besham Complex and Kotla Complex. The Besham Complex to the west is
separated by the ~21 km N–S trending and east verging sinistral Puran Thrust Fault
from the Swat multiply deformed and metamorphosed rocks. The sinistral strike-
slip along the Puran Thrust Fault sheared Precambrian to Late–Middle Mesozoic
rocks of the Indian Plate (Fig. 5; Baig et al. 1989). Along the same fault the Swat
multiplies deformed rocks thrust over the Besham nappe comprising of the Besham
Group rocks. This relationship indicates that the Besham nappe pre-dated the Puran
374

Fig. 4 Regional geological map showing the KIA, Greater Himalaya, Lesser Himalaya and Sub-Himalaya. Cross-section E–E through Thakot Shear Zone,
core of the Indus Syntaxis, Loe Sar dome and Indus Melange Zone showing structural and stratigraphic framework of the Greater Himalaya that are immediately
exposed south of the MMT (after DiPietro et al. 2008)
A. Ali et al.
Tectonic Framework of Northern Pakistan from Himalaya …

Fig. 5 Detailed regional maps of the rocks exposed west and east of the Indus Syntaxis. The grade of metamorphism increases east of the Indus Syntaxis. Rapid
375

exhumation of very deep-seated rocks in the Kaghan valley exposes UHP eclogites (after Williams 1989; Hussain et al. 2004)
376 A. Ali et al.

Fig. 6 MMT, which marks the Indian Plate and KIA collision zone wraps around the Indus Syntaxis.
The upper mantle rocks of the Jijal Ultramafic Complex thrust over the north-dipping basement
rocks of the Indian Plate

Fault. The Besham Group, exposed in the core of the Besham antiform, consists of the
Indian crystalline basement in the form of paragneisses, orthogneisses, amphibolites,
and marbles. These basement rocks are intruded by undeformed tourmaline granites
(Williams 1989). The Besham Group is subdivided into the Thakot Formation and the
Pazang Formation by Baig (1990). The Thakot Formation comprises paragneisses,
orthogneisses and graphitic schist. The Pazang Formation, which consists of marbles,
schists and komatiites (mantle-derived magnesium-rich igneous rocks), overlies the
Thakot Formation. The Early Proterozoic Besham Group rocks are overlain by the
Late Proterozoic Karora Group metasediments (Fig. 8).
These groups are locally separated by the Amlo–conglomeratic bed (Baig 1990).
The Karora Group is subdivided into the Karora Formation and Gandaf Formation.
The Karora Formation, which overlies the Gandaf Formation comprises graphitic
schist, marbles and metapsammite (Fig. 5). The Gandaf Formation consists of
graphitic schist, garnet schist, graphitic slates, quartzite, marble and metapsam-
mite. Episodic thrust events bring younger rocks over the region nappes that expose
higher grade metamorphic rocks. These internally imbricated blocks induced inverted
metamorphic zonation in the region (Fig. 9). The basal Amlo–conglomerate, which
contains metamorphosed clasts of the Besham Group indicates that these rocks are
metamorphosed up to upper amphibolite facies metamorphic conditions before the
Tectonic Framework of Northern Pakistan from Himalaya … 377

Fig. 7 a–c Regional geological map and a cross-section showing coesite-bearing UHP eclogites
exposed in the upper Kaghan Valley. The peak metamorphic conditions in the region are constrained
at 700–770 °C and 27–32 Kbar. d Zircon 206 Pb/238 U ages from the quartz-bearing medium domain
and coesite-bearing rim domain indicate that the peak metamorphism occurred at ~100 km depth.
e The Greater Himalayan gneisses travelled this distance within 7–9 Ma since the Indian Plate initial
collision with the KIA at 14–19° subduction angle at 1.1–1.4 cm year−1 rate (after Kaneko et al.
2003). f Rock recorded the eclogite grade metamorphic conditions

main Indian Plate and KIA collision in the Eocene. The Thakot dextral shear zone cut
off the Besham and Karora group metasediments from the Hazara Nappe to the east
(Fig. 10). 40–50 km wide and 75 km long Hazara Nappe comprises the Late Protero-
zoic Tanawal Formation, Cambrian (516 ± 16 Ma based on the whole rock Rb/Sr
isochron ages) porphyritic two–mica Mansehra granite and Cambrian Abbottabad
Group cover sequence (Fig. 5; Williams 1989).
The north-dipping imbricate sequence of the Hazara Nappe is truncated by the
Thakot Shear Zone to the west and Balakot Shear Zone to the east. Subsequent
378 A. Ali et al.

Fig. 8 Photograph showing the contact between the Early Proterozoic Besham Group and Late
Proterozoic Karora Group Rocks exposed in the core of the Indus Syntaxis

Fig. 9 Geological sketch showing subsequent imbricate faults that inverted deformation and
Barrovian metamorphic sequence of the Hazara Block (after Williams 1989)
Tectonic Framework of Northern Pakistan from Himalaya … 379

Fig. 10 Photograph showing Thakot/Battal dextral shear zone. The Thakot dextral shear zone cut
off the Besham and Karora group metasediments from the Hazara Nappe to the east

imbricate fault sequence inverted deformation and Barrovian metamorphic sequence


within the nappe (Fig. 5). The Tanawal Formation consists of metapelites, quartzite,
psammites and rarely marbles. The metapelites are metamorphosed up to sillimanite
grade metamorphic conditions. The lower and the upper contacts of the Tanawal
Formation with the Precambrian Hazara Slate and Abbottabad Formation are uncon-
formable. The Tanawal Formation is intruded by 2000 km2 Late Cambrian Mansehra
porphyritic granite of K-feldspar, plagioclase, biotite, muscovite and tourmaline
composition (Fig. 5). It contains Proterozoic to Precambrian xenoliths of the Indian
basement rocks (Fig. 11). The main batholith is subdivided into gneisses, pegmatites,
albitites, aplites, tourmaline granite and porphyry granite by Shams (1969). The
Mansehra granitic batholith intruded by the basic dykes of the Panjal magmatic cycle
during the Gondwana break–up in the Permian–Triassic (Papritz and Rey 1989).
The external zone, which is situated between the Panjal/Khairabad Thrust and
Main Boundary Thrust (MBT) comprises slightly pre–Himalayan metamorphosed
Precambrian metasediments, unmetamorphosed Mesozoic to Eocene Neo–Tethyan
shelfal sediments and Miocene molasse deposits (Table 1; Calkins et al. 1975). These
rocks along with the MBT thrust over the foreland Miocene and younger sediments
(Fig. 1).
380 A. Ali et al.

Fig. 11 Photograph showing Tanawal Quartzite xenolith inclusion in the ascended Mansehra
granitic magma. The Indian Plate basement rocks xenolith is assimilated to varying extent into
the Mansehra felsic magma

2.2 Kohistan Island Arc

In the Early Cretaceous northward dipping intraoceanic subduction in the equatorial


part of the Neo–Tethys resulted in the formation of the KIA (Fig. 12). The eastern
continuation of the same intraoceanic subduction beneath Lhasa Block that was
already accreted to the Eurasian Plate generated the Gangdese magmatic belt along
the northernmost margin of the Lhasa Block (Fig. 13). Subduction of the Neo–Tethys
north of the KIA under the Karakoram microplate closed the intervening northern
Neo–Tethys and brought forth the accretion of the KIA to the Karakoram microplate
along the MKT (Shyok Suture Zone; Fig. 12). The northward collision of the arc
with the Karakoram around 104–85 Ma preceded collision of the Indian Plate along
the MMT/Indus Suture Zone at the southern margin of the Arc at ~50 Ma (Fig. 13).
The KIA spectacularly exposes a complete intraoceanic juvenile crustal section that
developed at ~110–50 Ma (U–Pb zircon ages; Jagoutz et al. 2012).
The KIA crustal section comprises basal ultramafic rocks that formed at <55 km
depth, gabbroic/basaltic rocks, calc-alkaline batholith (26 km thick), and 4 km thick
interlayered volcanoclastic sediments (Fig. 15; Jagoutz and Schmidt 2012). Compo-
sitionally the arc is subdivided into 17 units. An average of 594 whole-rock analyses
indicates andesitic composition of the KIA with 56.6–59.3 wt.% silica and 0.51–0.55
XMg ratio. This composition is later affected by initial and subsequent collisions
(Jagoutz and Schmidt 2012). The E–W trending basal ultramafic to gabbroic rocks
of the KIA crop out in the Jijal Ultramafic Complex. These rocks in the near vicinity
of the Jijal Ultramafic Complex are exposed in Tora Tiga, Shangla to the west, and
Sapat–Babusar in the east (Fig. 4). The origin of these ultramafic complexes is a
Tectonic Framework of Northern Pakistan from Himalaya … 381

Table 1 Precambrian to Miocene stratigraphy of the Hazara Fold–Thrust Belt. The stratigraphy is
interrupted by several unconformities (from Shah 2009)
382 A. Ali et al.

Fig. 12 Regional geological map of the arcuate-shaped KIA wedged between the Indian Plate
and Karakoram. Northward obduction of the KIA onto the Indian Plate exposes ophiolitic sutures,
upper mantle fractional crystallized cumulates predominantly of olivine and pyroxene composi-
tion (Southern Complex), magmatic arc that developed from the asthenosphere and north-dipping
subduction zone associated melt (Chilas Complex) and Gilgit Complex. The Gilgit Complex
comprises the Kohistan batholith, volcanics and metasediments (after Jagoutz and Schmidt 2012)

raised question, they could be the off–scraped remnant rocks of the oceanic crust of
the upper mantle and lower Tethyan crustal rocks that are exposed as isolated large
blocks along with the MMT tectonic melange (Fig. 14) or these rocks were formed
in the southward forearc part of the KIA in the Cretaceous (Jan and Howie 1981;
Jan and Windley 1990; Miller et al. 1991; Searle et al. 1999). Up–section of the Jijal
Complex comprises serpentine, ultramafic cumulates ranging from dunites to wher-
lites to olivine–clinopyroxenites and (Ol)–websterites and granulitic gabbros consists
of clinopyroxene–bearing hornblendite and garnetite (Jagoutz and Schmidt 2012).
Granulitic gabbro indicates that some part of the Jijal Ultramafic Complex meta-
morphosed up to higher grade granulite facies metamorphic conditions. These rocks
are metamorphosed at 700–1150 °C and 15–19 Kbar (Miller et al. 1991; Ringuette
et al. 1999). These metamorphic conditions are achieved by subduction of the lower
crustal rocks at ~ 45–50 km depth along with the MMT. The minimum age of the
Jijal Ultramafic Complex is ~95 Ma based on 95 Ma Sm–Nd isochrons cooling age
(Anczkiewicz and Vance 2000; Yamamoto and Nakamura 1996). 700 m thick granite
sheets of 97.1 ± 0.2 Ma to 75.7 ± 1.4 Ma (U–Pb TIMS zircon age) separate the Kiru
Sequence from the Kamila Amphibolite (Schaltegger et al. 2002; Yamamoto et al.
2005). The green units in the top two sections (Fig. 15) show top-to-S ductile shear,
which is a fore shear, and a common structure in the Himalayas at different locations
(e.g., Bose and Mukherjee 2019a, 2019b). The Kamila amphibolite consists of amphi-
bolite facies meta–plutonic rocks of gabbro to tonalite composition, meta–volcanics
and meta–sedimentary rocks. Kamila amphibolites are derived from regional meta-
morphism of the Tethyan oceanic crust that originated along a Mid-Oceanic Ridge
Tectonic Framework of Northern Pakistan from Himalaya … 383

Fig. 13 Early Cretaceous paleogeographic map of the E–W trending, north-dipping subduction
zone. a Subduction of the Neo–Tethys beneath the Lhasa Block that was accreted onto the southern
margin of the Eurasian Plate resulted in the formation of the Gangdese magmatic arc. The intrao-
ceanic subduction culminated in Kohistan and Ladakh Island Arc. The same subduction zone to the
west obducted the Oman Ophiolitic Sequence during this time. The intervening northern Neo–Tethys
was subducting beneath the southern margin of the Karakoram (slightly modified after Rolland 2002;
Frisch et al. 2011). b Schematic cross-section showing intraoceanic subduction, culmination of the
island arc, Andean type magmatism along the southern margin of the Karakoram and northernmost
passive continental margin of the Indian Plate (after Burg et al. 2006)
384 A. Ali et al.
Tectonic Framework of Northern Pakistan from Himalaya … 385

Fig. 14 Cartoons showing tectonic components and sequential evolution mechanisms of the KIA.
The intraoceanic fossil KIA episodically built on the Neo–Tethys crust during the Cretaceous.
The southern part of the arc exposes the upper mantle peridotite that formed at ~55 km depth,
intrusive gabbro–norites magmatic arc, volcanics, and sedimentary sequence. a Intraoceanic north-
ward subduction of the Neo–Tethys metamorphosed the early subduction related Jijal Ultramafic
Complex to granulite facies conditions. The andesitic to rhyolitic Chalt–Dras volcanism and Matum
Das tonalite intrusion took place around 105–95 Ma. b The Late Cretaceous collision of the KIA
with the southern margin of the Karakoram. During this time the Chilas Complex preponderantly
of gabbro–norite composition intruded the Kamila Amphibolite and Gilgit metasediments. The
south verging Patan Shear Zone, exhumation of granulite and blueschist occurred during this time.
c The final tectonic stage is augmented by the Indian Plate collision with the KIA, culmination of
the andesitic Utror–Dir volcanics, stage-2 undeformed gabbros–diorites to granodiorites intrusions,
deposition of the Baraul Banda Slate and formation of the MMT (after Searle et al. 1999)

Fig. 15 Regional geological cross-sections showing the thickness and geobarometric conditions
of the subunits in the Jijal Complex, Kamila Amphibolite and Chilas Complex. The Mohorovicic
discontinuity separates the upper mantle cumulates of the Jijal Complex from the lower oceanic
crustal garnet granulite gabbro. Note the geobarometric conditions decreases from south to north.
Along the southern margin, which is defined by the MMT/Indus Suture Zone brings the upper
mantle rocks onto the Precambrian basement rocks of the Indian Plate (after Jagoutz and Schmidt
2012)
386 A. Ali et al.

(Khan et al. 1989). 40 Ar/39 Arc cooling age of 83–80 Ma indicates that metamorphism
of the Kamila amphibolite plutonic, volcanic, and sedimentary units preceded the
Indian Plate collision with the KIA (Treloar et al. 1989).
The Kamila Amphibolite is intruded by the mafic to ultramafic 300 km long and
40 km wide Chilas Complex at ~86 Ma (Zeitler et al. 1981; Khan et al. 1989; Schal-
tegger et al. 2002). It comprises massive gabbro–norite with minor dyke intrusive
bodies including harzburgites, dunites, troctolites, pyroxenites, hornblende gabbro
and anorthosites (Treloar et al. 1989; Searle et al. 1999). The Chilas Complex intruded
the Kamila Amphibolite and Jaglot Group Metasediments contact zone situated
in the eastern part of the KIA (Fig. 12). The eastern part of the Chilas Complex
contains xenoliths of highly deformed Jaglot Group Metasediments (Treloar et al.
1996). The highly deformed Jaglot Group Metasediments in the Chilas Complex indi-
cates that KIA and Karakoram collision predated the Chilas Complex emplacement
(Treloar et al. 1996). The Jaglot Group Metasediments and Chalt Volcanic Group
are intruded by stage-1 tonalite, stage-2 undeformed gabbros, diorites and granodi-
orites and stage-3 undeformed granitic plutons between 110–90 Ma, 85–40 Ma and
30 Ma Rb–Sr whole rock isochrons ages respectively (Petterson and Windley 1985,
1991). The stage-1 plutons emplaced the KIA prior its suturing at ca. 100 Ma with
the Karakoram along the MKT/Shyok Suture (Bignold et al. 2006). The stage-2 and
stage-3 plutons emplacement posted the suturing (Fig. 16). The MKT suture zone is
defined by serpentinites, Chalt Volcanic Group and Yasin Group Metasediments. The
E–W trending Cretaceous basaltic–andesitic to rhyolitic composition Chalt Volcanic
Group is well exposed in the Hunza valley. Submarine explosive volcanic are frozen
into elongated to spherical pillows. Based on the presence of Radiolaria, Orbitolina
and Rudist, in the Chalt Interbedded sedimentary strata, the Volcanic Group can have
an age ~108–97in the Interbedded sedimentary strata (Pudsey et al. 1986). In Chalt
village along the MKT Chalt Volcanic Group overlies the Yasin Group Metased-
iments and is underlain by the Jaglot Group Metasedimentary to the south. The
presence of boninitic volcanic rocks, which usually associated with forearc tectonic
setting indicates that the Chalt Volcanic Group may have generated by south subduc-
tion of the northern Neo–Tethys beneath the northernmost margin of the KIA (Khan
et al. 1997). According to Khan et al. (1997) subduction of the northern Neo–Tethys
beneath the KIA and Karakoram generated collision between these terrains along the
Shyok Suture around Santonian–Campanian (~85–70 Ma; Searle et al. 1999). The
75 Ma undeformed Jutal dykes of basic composition, which intrude Chalt Volcanic
Group indicate that Shyok Suture formed before 75 Ma (Fig. 17). The active exhuma-
tion of Nanga Parbat of the Indian Plate affinity separates the KIA from the Ladakh
Island Arc.
Karakoram.
The Shyok Suture separates the KIA from the Karakoram Cimmerian Block of
the Gondwana affinity (Fig. 18; Gaetani 1997). The southernmost Paleo–Andean
type plate margin of the Karakoram microplate comprises the Karakoram Metamor-
phic Complex. These preOrdovician crystalline basement rocks are metamorphosed
to upper amphibolite facies conditions at 650–700 °C and 8–9 Kbar by concurrent
Tectonic Framework of Northern Pakistan from Himalaya … 387

Fig. 16 Photograph of stage-3 felsic intrusion in the Jaglot Group Metasediments. The Jaglot Group
Metasediments and Chalt Volcanic Group are intruded by stage-1 tonalite, stage-2 undeformed
gabbros, diorites and granodiorites and stage-3 undeformed granitic plutons between 110–90 Ma,
85–40 Ma and 30 Ma Rb–Sr whole rock isochrons ages respectively. The stage-1 plutons emplaced
the KIA prior to its suturing at ca. 100 Ma with the Karakoram along the MKT/Shyok Suture. The
stage-2 and stage-3 plutons emplacements posted the suturing

deformation stresses and P–T controlled crystallization during Neo–Tethys subduc-


tion and docking of the KIA (Gaetani 1997). Successive prograde appearance of chlo-
rite, biotite, garnet, staurolite, kyanite and sillimanite metamorphic index minerals
indicate classic Barrovian metamorphism at the southernmost active margin of the
Karakoram (Fig. 19). South verging imbricates of the southern crystalline margin of
the Karakoram inverted the normal Barrovian metamorphic sequence. In a prograde
sequence the sillimanite grade rocks thrust over the kyanite, staurolite, garnet and
chloritoid grade rocks. North of the Karakoram Metamorphic Complex Early to Mid-
Cretaceous (ca. 105–95 Ma U–Pb, zircon ages), western Cordilleran type magma-
tism developed the main ~ 600 km E–W trending Karakoram Batholith of calc–
alkaline composition north of the Neo–Tethys oceanic crust subduction beneath the
southernmost continental margin of the Karakoram (Fig. 19; Le Fort et al. 1983;
Coward et al. 1986; Debon et al. 1987; Crawford and Searle 1992; Fraser et al. 2001;
Heuberger et al. 2007). The Hunza Plutonic Unit comprises granodiorites, quartz
diorites, tonalite, adamellite and magmatic migmatites (agmatites). The felsic core
of the batholith is flanked by the rocks of mafic composition. The Hunza Plutonic Unit
388 A. Ali et al.

Fig. 17 Photograph showing the 75 Ma undeformed Jutal dykes of basic composition, which
intruded Chalt Volcanic Group indicate that Shyok Suture formed before 75 Ma

is intruded by the Batura Plutonic Unit of diorite, gabbro and granodiorite composi-
tion during the Paleocene–Early Eocene plutonism (63.4 ± 2–42.8 ± 5.6 Ma Rb–Sr
isochron ages; Debon et al. 1987; Debon 1995; Fig. 19). The Batura Plutonic Unit
culminated during the post-collisional mantle fed crustal thickening and lower crust
melting (Debon, 1995). The Hunza and Batura plutonic units that collectively termed
the Karakoram Batholith intruded by the post-collisional leucogranite and manzo-
granite plutons in the Early Miocene (25–21 Ma, U–Pb zircon, monazite dates; Searle
et al. 1989). The ongoing crustal shortening and tectonic stacking of Eurasian Plate,
Hindu Kush, Karakoram, KIA and India produced large scale crustal anatectic melt
within these tectonic terrains in the Miocene. The Miocene crustal melting processes
produced the Baltro granitic batholith southeast of the Hunza Plutonic Unit. The
Baltro granitic batholith is not exposed in the Hunza valley.
Crustal thickening as a consequence of ongoing collision resulted into 9.2 ± 2 Ma
Sumayar leucogranite north of the Shyok Suture Zone (Fraser et al. 2001). North
of the pre and post-collision plutons the Karakoram comprises Ordovician to Late
Cretaceous sedimentary rocks. These rocks exclusively deposited along the north-
ernmost passive continental margin of the Karakoram. Gaetani (1997) established
the following six tectonostratigraphic cycles in the northern Karakoram (Fig. 20).
Tectonic Framework of Northern Pakistan from Himalaya …

Fig. 18 A series of cross-sections from the Late Carboniferous to Early Permian and Middle Eocene showing the episodic separation of the Hindu Kush and
Karakoram Cimmerian Blocks from the northernmost part of the Gondawana and their accretion along the Tirich Mir Boundary Zone. The northward drift of
these Cimmerian blocks formed the Neo–Tethys. The intra Neo–Tethys subduction zone culminated in the KIA. The docking of the KIA to the Karakoram
closed the northern Neo–Tethys and formation of the Main Karakoram Thrust (Shyok Suture Zone). This period is also characterized by crustal thickening and
melting that crystallized into the granite. The final closure of the southern Neo–Tethys upon collision of the Indian Plate with the KIA led to the formation of the
MMT (Indus Suture Zone) and southward propagation of the deformation front that resulted in the MMT, the MBT, and the Main Frontal Thrust (after Zanchi
389

and Gaetani 2011)


390 A. Ali et al.

Fig. 19 Regional geological maps of different tectonic units of the Karakoram. Metamorphic index
mineral assemblages of the southern Karakoram Metamorphic Complex depicting inverted meta-
morphism. South verging multiple imbricates in the southern crystalline margin of the Karakoram
has inverted the normal Barrovian metamorphic sequence. In a prograde sequence the chlorite-grade
rocks respectively overlain by biotite, staurolite, kyanite and sillimanite bearing rocks. The Hunza
Plutonic Unit separates the Southern Karakoram Metamorphic Complex from the Ordovician to
Mid-Cretaceous northern sedimentary rocks (after Searle et al. 1996, 1999)
Tectonic Framework of Northern Pakistan from Himalaya … 391

Fig. 20 Ordovician to Mid-Cretaceous stratigraphic columnar section of the northern Karakoram.


The whole stratigraphic sequence was deposited in the time span of ~400 Ma. These rocks are
mostly deposited under shallow marine conditions (from Gaetani 1997)
392 A. Ali et al.

2.3 Ordovician to Late Devonian Cycle

Rocks of this age termed as the Baroghil and the Karambar units crop out in the
western part of Karakoram in Chitral and Karambar. The Ordovician to Silurian
sedimentary rocks deposited during the southward transgression of the Paleo–Tethys
onto the Karakoram crystalline basement (Le Fort et al. 1994). These consist of
muddy shelf dolostone and quartz aranitelenses (Gaetani 1997). Coral, bryozoans
and stromatoporoids in the Devonian unit indicate warm tropical climate depositional
conditions. In the Middle Devonian the Karakoram started separating from the Gond-
wanaland, which resulted in the formation of the Neo–Tethys south of the Karakoram
Cimmerian Block in the Permian and a passive rifted northern and southern conti-
nental margins until the Karakoram final collision with the Hindu Kush and KIA
respectively. Crustal thinning that eventually rifted continental margin commonly
associated with volcanic, the presence of basaltic volcanic in Devonian rocks indicate
that the Karakoram rifting from the Gondawana took place in the Middle Devonian
to Earliest Carboniferous (Gaetani 1997).

2.4 Late Devonian to Earliest Permian Cycle

Rocks of these ages are only reported from the Chitral region, which is located
west of Hunza. These rocks consist of terrigenous successions with lesser carbonate
intercalations (Desio and Martina 1972; Gaetani 1999).

2.5 Early Permian to Earliest Jurassic Cycle

The Paleo-Tethys Permian to Jurassic sediment accumulation took place on north-


ward drifting Karakoram craton. The Permian succession is characterized by terres-
trial sediments. The provenance history indicates that these sediments are mostly
derived from the northward drifting Karakoram Cimmerian Block (Gaetani 1997).
Temporary marine incursions in the Middle and Lower parts of the Permian
succession are locally reported from the Chapursan Valley. The presence of
brachiopods in the marine intercalations assigns Latest Asselian to Early Sakmarian
age to marine incursions in the Middle and Lower parts of the Permian succession
(Angiolini 1995). In Hunza, the Permian succession comprises the Gircha Formation,
Kilik Formation, Misgar Slates and Pasu Slates.
Tectonic Framework of Northern Pakistan from Himalaya … 393

2.6 Early Jurassic Cycle

North of Sost The Triassic Aghil Formation, which comprises limestone is over-
lain by the Jurassic Yashkuk and Ashtigar Formations. The Ashtigar Formation
comprises dark grey litharenites and has transitional contact with the Aghil Forma-
tion. The Yashkuk Formation consists of sandstone with intercalated shales and silt-
stones (Gaetani et al. 1993). The presence of serpentine and basalts in the Ashtigar
Formation indicates obduction of an oceanic crust along with a newly evolving
orogenic wedge. This orogenic wedge has been interpreted as the initial collision
of the Karakoram with a Cimmerian block in the north (Gaetani et al. 1993; Zanchi
and Gaetani 2011). The Misgar Fault south of Khunjerab thrusts the Permian Misgar
slate over the Jurassic succession.

2.7 Middle Jurassic–Earliest Cretaceous Cycle

The Middle Jurassic succession comprises Reshit, Tekri (exposed in Shaksgam


Valley, China), Marpo Sandstone (exposed in Shaksgam Valley, China) and Bango–
La Formations (exposed in Shaksgam Valley, China; Gaetani et al. 1990; Kazmi
and Jan 1997; Zanchi and Gaetani 2011). The Reshit Formation is well exposed
west of Sost in the Chapursan Valley. It comprises a ~300 m thick sequence of
oncolitic to oolitic packstone, wackestones, dark grey well-bedded cherty mudstone,
and chert nodules (Gaetani et al. 1990; Kazmi and Jan 1997; Zanchi and Gaetani
2011). Gypsum and thin mineable coal layer have been reported from this formation
(Donnelly 2004). Aalenian–Bajocian ages (Middle Jurassic) have been assigned to
the formation on the basis of Lenticulina, Trocholina, Spirillina, Protopeneroplis,
Kallirhynchia, and Saccocoma (cf. Kazmi and Jan 1997).

2.8 Late Cretaceous Cycle

The Tupop Conglomerate of the Late Cretaceous age lies upon the Jurassic to Early
Cretaceous sequence in the Chapursan Valley, upper Hunza. The channelized clasts
supported conglomeratic beds are intercalated with shale and sandstone (Zanchi and
Gaetani 2011). Clasts of conglomerate contains cobble and pebble of Permian to
Jurassic units. The Permian clasts in the Tupop conglomerate indicate older rocks
exposure in thrust sheets and folded strata that formed as a result of Hindu Kush and
Karakoram collision.
394 A. Ali et al.

2.9 Khunjerab, Sost, Warbin and Shujerab Plutons

These plutons comprise granodiorite, adamellite, quartz monzodiorite, granite and


diorite (Kazmi and Jan 1997). They intruded the Ordovician to Jurassic sedimentary
successions of the northern Karakoram (Fig. 19). K–Ar biotite ages and hornblende
ages indicate that these plutons formed around 97–115 Ma (Treloar et al. 1989;
Debon et al. 1996).

Day 1
Islamabad to Besham

Stop 1
Tanaki Boulders
Contact between the Precambrian Hazara Formation and Cambrian Abbottabad
Group describes by the Early Cambrian Tanaki boulders bed (Fig. 21). Angular
debris clasts that are accumulated in a terrestrial debris flow are derived from the
Precambrian Hazara Formation and Tanawal Quartzite. The clasts are highly meta-
morphosed than the matrix, which confirms the Early Paleozoic to Late Precam-
brian metamorphism of the underlying formations. The Cambrian Mansehra Granite
post-dated the Late Precambrian to Early Paleozoic uplift of the Indian shield rocks.

Stop 2
Panjal Thrust Fault
The Panjal Thrust Fault on the western side of the Hazara Kashmir Syntaxis brings
Precambrian rocks over the Cambrian Abbottabad Group (Fig. 22). The Panjal Thrust
merges with the MBT in the Apex region of the Hazara Kashmir Syntaxis, and adjoin
Precambrian rocks with the Murree Formation. The Precambrian Tanawal Formation
is missing east and south of the Panjal Thrust.

Stop 3
Mansehra Granite, Precambrian Xenoliths and Dykes
The Tanawal Formation is intruded by 2000 km2 Late Cambrian Mansehra
porphyritic granite of K–feldspar, plagioclase, biotite, muscovite and tourmaline
composition. Presence of tourmaline indicates that the crystallization of the magma
took place in the upper crust. It also contains Proterozoic to Precambrian xenoliths
of the Indian basement rocks (Fig. 3; 11). The main batholith is subdivided into
gneisses, pegmatites, albitites, aplites, tourmaline granite and porphyry granite. The
Mansehra granitic batholith intruded by the basic dykes of the Panjal magmatic cycle
during the Gondwana break–up in the Permian–Triassic.
Tectonic Framework of Northern Pakistan from Himalaya … 395

Fig. 21 The Early Cambrian Tanaki bounders define contact between the Precambrian Hazara
Formation and Cambrian Abbottabad Group. The angular to sub-rounded clasts are derived from
the underlying Hazara Formation and Tanawal Quartzite

Fig. 22 The Panjal Thrust Fault brings the Precambrian Hazara Formation and Tanawal Formation
over the Cambrian Abbottabad Group
396 A. Ali et al.

Stop 4
Thakot Shear Zone
The Thakot dextral shear zone cut off the Besham Group metasediments from the
Hazara Block to the east. Along the Thakot Shear Zone the Mansehra Granite and
Tanawal Formation have been mechanically pulverized in brittle deformation condi-
tions. Towards the core of the Indus Syntaxis deformation becomes more ductile.
Dykes, veins and pegmatites have been pulverized during this deformation. The
Thakot Shear Zone is nicely exposed in Battal and Ahal (Fig. 10).

Stop 5
Core of the Indus Syntaxis
The core of the Indus Syntaxis is characterized by exhumation of the Early Protero-
zoic Besham Complex and Kotla Complex (Fig. 8). The Besham Complex to the west
is separated by the N–S trending Puran Fault from the Swat multiply deformed and
metamorphosed rocks. The Besham Group is subdivided into the Thakot Formation
and Pazang Formation. The Thakot Formation comprises paragneisses, orthogneisses
and graphitic schist. The Pazang Formation consists of marbles, schists and komati-
ites (mantle-derived magnesium-rich igneous rock). The Early Proterozoic Besham
Group rocks are overlain by the Late Proterozoic Karora Group metasediments.
These groups are separated by the Amlo–conglomeratic bed. The Karora Group is
subdivided into the Karora Formation and Gandaf Formation. The Karora Forma-
tion comprises graphitic schist, marbles and metapsammite. The Gandaf Forma-
tion consists of graphitic schist, garnet schist, graphitic slates, quartzite, marble,
and metapsammite. The basal Amlo–conglomerate, which contains metamorphosed
clasts of the Besham Group indicate that these rocks are metamorphosed up to upper
amphibolite facies metamorphic conditions before the main Indian Plate and KIA
collision in the Eocene. The Thakot dextral shear zone cut off the Besham and Karora
group metasediments from the Hazara Nappe to the east.

Day 2
Besham to Chilas

Stop 6
Indus Syntaxis, Dubair Granodiorite
Besham Group schist, quartzite and ortho–para gneisses. The Precambrian Dubair
granodiorite intruded the Besham Group rocks and retains xenoliths of the country
rocks. The granodiorite comprises amphibole, biotite, plagioclase, feldspar and
quartz. U/Pb zircon ages of the Dubair granodiorite is 1858.8 ± 7.2 Ma.
Tectonic Framework of Northern Pakistan from Himalaya … 397

Stop 7
Main Mantle Thrust, Jijal Complex
In Jijal the MMT and Jijal Ultramafic Complex are nicely exposed (Fig. 6). The
contact zone is characterized by mylonites and asymmetric structures with SSE
tectonic transport. The E–W trending MMT/Indus Suture is wrapped around in the
apex region of the Indus Syntaxis. Note the Indus Suture extends from Afghanistan in
the west to Myanmar in the east. The Jijal Ultramafic Complex is exposed north of the
MMT. The contact zone of the Jijal and Besham Complex are marked by serpentine.
Concentration of serpentine along suture helps in the subduction and development
of ductile shear zones. Serpentine (Mg6 [Si4 O10 (OH)8 ) form by pyroxene and olivine
reaction with water. The E–W trending basal ultramafic to gabbroic rocks of the
KIA crop out in the Jijal ultramafic complex. These rocks in the near vicinity of
the Jijal Ultramafic Complex are exposed in Tora Tiga, Shangla to the west and
Sapat–Babusar in the east.

Stop 8
Garnetite, Granulite and Hornblendite of the Jijal Ultramafic Complex
Up–section of the Jijal Ultramafic Complex comprises granulitic gabbros consisting
of clinopyroxene–bearing hornblendite and garnetite (Jagoutz and Schmidt 2012).
Granulitic gabbro indicates that some part of the Jijal Ultramafic Complex meta-
morphosed up to higher grade granulite facies metamorphic conditions (Fig. 23).
These rocks are metamorphosed at 700–1150 °C and 15–19 kbar by subduction of
the lower crustal rock at ~45–50 km depth along with the MMT. The Mohorovicic
discontinuity separates the upper mantle cumulates of the Jijal Ultramafic Complex
from the lower oceanic crustal garnet granulite gabbro.

Fig. 23 Garnetite in the Jijal Complex indicating granulite facies metamorphic conditions
398 A. Ali et al.

Stop 9
Patan Shear Zone
North of the Patan market place the Jijal Ultramafic Complex is highly ductilely
sheared (Fig. 24). The 3 km wide Patan shear zones anastomose around the less
deformed Sarangar metagabbro. The Sarangar metagabbro intruded the garnet-
bearing gabbro at 98.9 ± 0.4 Ma. Asymmetries along the Patan Shear Zone indicate
SW vergence. Garnet bearing veins are interpreted to have formed from mylonite
melt.

Stop 10
Kiru Amphibolite
The Sheared Sarangar gabbro in the north is bounded by the Kiru Amphibolite.
The Kiru Amphibolite comprises sheared meta–gabbros, tonalites, diorites and
pegmatites.

Stop 11
Granite Sheets
700 m thick granite sheets of 97.1 ± 0.2 Ma to 75.7 ± 1.4 Ma (U–Pb TIMS zircon
age) separate the Kiru Sequence from the Kamila Amphibolite. The granite comprises
garnet, muscovite, tourmaline, feldspar, plagioclase, quartz and apatite.

Fig. 24 Photograph showing the Patan Shear Zone


Tectonic Framework of Northern Pakistan from Himalaya … 399

Stop 12
Kamila Amphibolite
The Kamila Amphibolite consists of amphibolite facies meta–plutonic rocks of
gabbro to tonalite composition, meta–volcanics and metasedimentary rocks. Kamila
amphibolites are derived from regional metamorphism of the Tethyan oceanic crust
that originated along with the Mid-Oceanic Ridge (Khan et al. 1989). 40 Ar/39 Arc
cooling age of 83–80 Ma indicates that metamorphism of the Kamila Amphibolite
plutonic, volcanic and sedimentary units preceded the Indian Plate collision with the
KIA (Treloar et al. 1989).

Stop 13
Chilas Complex
The Kamila Amphibolite is intruded by the mafic to ultramafic 300 km long and 40 km
wide Chilas Complex at ~86 Ma (Fig. 25). It comprises massive gabbro–norite with
minor dykes intrusive bodies including harzburgites, dunites, troctolites, pyroxenites,
hornblende gabbro and anorhtosites. The Chilas Complex contains inclusions of the
country rocks.

Fig. 25 Unsorted and unconsolidated terrigenous Jalipur alluvial sediments. This sequence is well
exposed between Chilas and Bunji
400 A. Ali et al.

Day 3
Chilas to Hunza

Stop 14
Chilas Ultramafics
Ultramafic rocks, pyroxenites alternating with anorthosites and troctolites intruded
the mafic gabbro–norite of the Chilas Complex. These rocks are formed from the
crystallization of the upper mantle melt.

Stop 15
Jalipur Alluvial Sediments with Glacier Till at the Base
Around 20–25 km north of Chilas the Jalipur alluvial sediments with glacier till are
exposed. The basal part of the Jalipur sediment is diamictite. It comprises unsorted
terrigenous sediments (Fig. 25). The basal part of the Jalipur alluvial sediments does
not contain clasts of the Nanga Parbat Gneiss that are located at ~19 km in the east
(Shroder et al. 1989). It indicates that these sediments predate the active and rapid
exhumation of the Nanga Parbat Syntaxis.

Stop 16
Raikot–Sassi Fault, Hot Springs
The Raikot Fault separates the Kohistan Island Arc from the Nanga Parbat of the
Indian Plate affinity. The active Raikot fault formed due to Neogene rapid exhuma-
tion of the Nanga Parbat. Emergence of geothermally heated groundwater along the
Raikot–Sassi fault resulted in sulfur–smelling hot springs south of the Raikot bridge
(Fig. 26).

Stop 17
Jaglot Group Metasediments, Kohistan Batholith
The Jaglot Group Metasediments and Chalt Volcanic Group are intruded by stage-1
tonalite, stage-2 undeformed gabbros, diorites, and granodiorites and stage-3 unde-
formed granitic plutons between 110–90 Ma, 85–40 Ma and 30 Ma Rb–Sr whole
rock isochrons ages respectively (Fig. 27). The stage 1 plutons emplaced the KIA
prior to its suturing at ca. 100 Ma with the Karakoram along the MKT/Shyok Suture.
The stage-2 and stage-3 plutons emplacements posted the suturing.

Stop 18
Jutal Dykes, Chalt Volcanic Group
The E–W trending Cretaceous basaltic–andesitic to rhyolitic composition Chalt
Volcanic Group is well exposed in the Hunza valley (Fig. 27c). Submarine explo-
sive volcanics are frozen into elongated to spherical pillows. The Chalt Volcanic
Group is dated at ~108–97 Ma based on the presence of Radiolaria, Orbitolina
and Rudist in the Interbedded sedimentary strata. In Chalt village along the MKT
Tectonic Framework of Northern Pakistan from Himalaya … 401

Fig. 26 Emergence of geothermally heated groundwater along the Raikot–Sassi fault resulted in
sulfur–smelling hot springs south of the Raikot bridge. This region was exposed on the western side
of the Nanga Parbat

Chalt Volcanic Group overlies the Yasin Group Metasediments and is underlain by
the Jaglot Group Metasedimentary to the south. The presence of boninitic volcanic
rocks, which usually associated with forearc tectonic setting indicates that the Chalt
Volcanic Group may have generated by south subduction of the Neo–Tethys beneath
the northernmost margin of the KIA. The 75 Ma undeformed Jutal dykes of basic
composition, which intrude Chalt Volcanic Group indicate that Shyok Suture formed
before 75 Ma (Fig. 17). The active exhumation of Nanga Parbat of the Indian Plate
affinity separates the KIA from the Ladakh Island Arc (Fig. 28).

Stop 19
Main Karakoram Thrust/Shyok Sure Zone
The Shyok Suture separates KIA from the Karakoram Cimmerian Block of the Gond-
wana affinity (Fig. 29). The southernmost Paleo–Andean type plate margin of the
Karakoram microplate comprises Karakoram Metamorphic Complex. Subduction
of the Neo–Tethys north of the KIA under the Karakoram microplate closed the
intervening northern Neo–Tethys and brought forth the accretion of the KIA to the
Karakoram microplate equivalent to the Lhasa Block in the Late Cretaceous along
with the MKT. The MKT suture zone is defined by serpentinites, Chalt Volcanic
Group and Yasin Group Metasediments.
402 A. Ali et al.

Fig. 27 a Stage-3 intrusions are surrounded by the Jaglot Group Metasediments. b ~245 Km2
Kohistan batholith intrudes the Jaglot Group Metasediments and Chalt Volcanic Group. c Chalt
Volcanic Group is capped by gossan (possibly underline by sulphide mineralized zone)
Tectonic Framework of Northern Pakistan from Himalaya … 403

Fig. 28 The Nanga Parbat protrusion within the KIA induced high deformation along with the
MMT and shaped the Nanga Parbat Syntaxis

Day 4
Hunza to Khunjerab

Stop 20
Karakoram Metamorphic Complex
These pre–Ordovician crystalline basement rocks are metamorphosed to upper
amphibolite metamorphic facies conditions at 650–700 °C and 8–9 Kbar by concur-
rent deformation stresses and P–T controlled crystallization during Neo–Tethys
subduction and docking of the KIA (Gaetani 1997; Fig. 29). Successive prograde
appearance of chlorite, biotite, garnet, staurolite, kyanite and sillimanite metamor-
phic index minerals indicate classic Barrovian metamorphism at the southernmost
active margin of the Karakoram (Fig. 19). South verging imbricates of the southern
crystalline margin of the Karakoram has inverted the normal Barrovian metamorphic
sequence.

Stop 21
Karakoram Batholith
North of the Karakoram Metamorphic Complex Early to Mid-Cretaceous (ca. 105–
95 Ma U–Pb, zircon ages) western Cordilleran type magmatism developed the main
404 A. Ali et al.

Fig. 29 The MKT separates the KIA from the Karakoram microplate. The same is the type locality
of the MKT on the Karakoram Highway

~600 km E–W trending Karakoram Batholith of calc–alkaline composition north


of the Neo–Tethys oceanic crust subduction beneath the southernmost continental
margin of the Karakoram (Fig. 30). The Karakoram Batholith is made up of Darkot
Pass, Ghamu Bar, Kesu, Hunza, Batura and Muztagh plutonic units. The Miocene
crustal melting processes produced Baltro granitic batholith southeast of the Hunza
Plutonic Unit. The Baltro granitic batholith is not exposed in the Hunza valley.

2.9.1 Early Permian to Earliest Jurassic Cycle

The Permian to Jurassic sedimentary accumulation took place on northward drifting


Karakoram craton through the Paleo–Tethys. The Permian succession is character-
ized by terrestrial sediments. The provenance history indicates that these sediments
are mostly derived from the continental block. Temporary marine incursions in the
Middle and Lower parts of the Permian succession are locally reported from the
Chapursan Valley. The presence of brachiopods in the marine intercalations assigns
Latest Asselian to Early Sakmarian age to marine incursions in the Middle and Lower
parts of the Permian succession (Angiolini 1995). In Hunza the Permian succession
comprises the Gircha Formation, Kilik Formation, Misgar Slates and Pasu Slates
(Fig. 31).
Tectonic Framework of Northern Pakistan from Himalaya … 405

Fig. 30 Karakoram Batholith of calc–alkaline composition is intruded by a large number of


pegmatites

Stop 22
Pasu Slate
Carboniferous Pasu Slate consists of splintery dark slates. The type locality of the
formation is the Pasu village located in the upper Hunza. The presence of fusulinids
(parafusulina sp.) describes Middle Permian age to the formation.

Stop 23
Gircha Formation
Lithology of the Permian Gircha Formation consists of thick-bedded sandstone,
interbedded slate and limestone–dolomite intercalations. Arkosic to quartz arenite
contains igneous lithics. Cross–bedded channelized sandstone fluviatile or deltaic
deposition system in a rapidly subsiding rift (Kazmi and Jan 1997).

Stop 24
Gujhal Formation
The Jurassic to Cretaceous Gujhal Formation comprises dolomite, dolomitic lime-
stone and minor slate (Fig. 32).
406 A. Ali et al.

Fig. 31 Photographs
showing the Permian
succession of the Karakoram
microplate
Tectonic Framework of Northern Pakistan from Himalaya … 407

Fig. 32 Photograph showing the Jurassic to Cretaceous Gujhal Formation, Permian Pasu Slates
and the famous Pasu glacier

Stop 25
Kilik Formation
The Permian Kilik Formation comprises slates, dolomite and dolomite limestone.
408 A. Ali et al.

Fig. 33 Photograph illustrating Khunjerab pluton intrusion into the Permian Misgar Slates

Stop 26
Misgar Slates
The 3500 m thick Permian Misgar Slates exclusively comprises dark grey slate with
subordinate phyllites, limestone and quartzite. The formation is intruded by dolerite,
pegmatite, aplite, quartz syenite and gabbro dykes.

Stop 27
Khunjerab Pluton
These plutons comprise granodiorite, adamellite, quartz monzodiorite, granite and
diorite (Fig. 33). They intruded the Ordovician to Jurassic sedimentary successions
of the northern Karakoram. K–Ar biotite ages, and hornblende ages indicate that
these plutons formed around 97–115 Ma.

Acknowledgements We thank Ayyaz Ahmad, Rashid Jamil Chauhan, Rai Hamood Inam and
Muhammad Sharif, members of Pakistan Association of Petroleum Geoscientists for facilitating
and funding this geological field. We thank Ali Rehman for redrawing some maps and Rafique
Ahmad for providing garnetite photograph. The Springer team (Marion Schneider, Annett Buettner,
Boopalan Renu, Alexis Vizcaino and Doerthe Mennecke-Buehler) is thanked for various assistance.
Soumyajit Mukherjee acted as editor and reviewer. Dutta and Mukherjee (2021) encapsulate this
work.
Tectonic Framework of Northern Pakistan from Himalaya … 409

References

Anczkiewicz, R., & Vance, D. (2000). Isotopic constraints on the evolution of metamorphic condi-
tions in the Jijal–Patan complex and the Kamila Belt of the Kohistan arc, Pakistan Himalaya.
In M. A. Khan, P. J. Treloar, M. P. Searle, & M. Q. Jan (Eds.), Tectonics of the Nanga Parbat
Syntaxis and the Western Himalaya (Vol. 170, pp. 321–331). Geological Society of London
Special Publications.
Angiolini, L. (1995). Permian brachiopods from Karakorum (Pakistan), Part I (with appendix).
Rivista Italiana Paleontologia e Stratigrafia, 101, 165–214.
Baig, M. S. (1990). Structure and geochronology of pre–Himalayan and Himalayan orogenic events
in the northwest Himalaya, Pakistan, with special reference to the Besham area (300pp) (Ph.D.
thesis). Oregon State University, Corvallis.
Bignold, S. A., Treloar, P. J., & Petford, N. (2006). Changing sources of magma generation beneath
intraoceanic island arcs: An insight from the juvenile Kohistan island arc, Pakistan Himalaya.
Chemical Geology, 233(1–2), 46–74.
Bose, N., & Mukherjee, S. (2019a). Field documentation and genesis of the back-structures from the
Garhwal Lesser Himalaya, Uttarakhand, India, Tectonic implications. In R. Sharma, I. M. Villa,
S. Kumar (Eds.), Crustal architecture and evolution of the Himalaya-Karakoram-Tibet Orogen
(Vol. 481, pp. 111–125). Geological Society of London Special Publications.
Bose, N., & Mukherjee, S. (2019b). Field documentation and genesis of back-structures in ductile
and brittle regimes from the foreland part of a collisional orogen: Examples from the Darjeeling-
Sikkim Lesser Himalaya, India. International Journal of Earth Sciences, 108, 1333–1350.
Burg, J. P. (2006). Two orogenic systems and a transform–transfer fault in the Himalayas: Evidence
and consequences. Earth Science Frontiers, 13(4), 27–46.
Burg, P. J., Celerier, B., Chaudhry, M. N., Ghazanfar, M., Gnehm, F., & Schnellmann, M. (2005).
Fault analysis and paleostress evolution in large strain regions: Methodological and geological
discussion of the southeastern Himalayan Fold-and-thrust belt in Pakistan. Journal of Asian Earth
Sciences, 24, 445–467.
Burg, P. J., Jagoutz, O., Dawood, H. & Hussain, S. (2006). Precollision tilt of crustal blocks in rifted
island arcs: structural evidence from the Kohistan Arc. Tectonics, 25(5). https://doi.org/10.1029/
2005TC001835.
Calkins, J. A., Offield, T. W., Abdullah, S. K. M., & Tayyab Ali, S. (1975). Geology of the Southern
Himalaya in Hazara, Pakistan, and adjacent areas. U.S. Geological Survey Professional Paper,
716–C, 29.
Coward, M. P., & Butler, R. H. W. (1985). Thrust tectonics and the deep structure of the Pakistan
Himalaya. Geology, 13, 417–420.
Coward, M. P., Windley, B. F., Broughton, R. D., Luff, I. W., Peterson, M. G., Pudsey, C. J., Rex, D.
C. et al. (1986). Collision tectonics in the NW Himalayas. In M. P. Coward & A. C. Ries (Eds.),
Collision tectonics (Vol. 19, p. 219). London: Special Publication Geological Society.
Crawford, M. B., & Searle, M. P. (1992). Field relationships and geochemistry of pre–collisional
(India–Asia) granitoid magmatism in the central Karakoram. Tectonophysics, 206, 171–192.
Debon, F., Afzali, H., Le Fort, P., & Sonet, J. (1987). Major intrusive stages in Afghanistan: typology,
age and geodynamic setting. Geologische Rundschau, 76, 245–264.
Debon, F. (1995). Incipient India-Eurasia collision and plutonism: The Lower Cenozoic Batura
granites (Hunza Karakoram, North Pakistan). Journal of Geological Society of London, 152,
785–795.
Debon, F., Zimmermann, J.-L., & Le Fort, P. (1996). Upper Hunza granites (North Karakoram,
Pakistan): A syn–collision bimodal plutonism of Mid-Cretaceous age. Comptes Rendu De
l’Academiedes Sciences Paris, II, 323(5), 381–388.
Desio, A., & Martina, E. (1972). Geology of the Upper Hunza valley (Karakorum W. Pakistan).
Bollettino della societá Geologica italiana., 91, 283–314.
DiPietro, J. A., Ahmad, I., & Hussain, A. (2008). Cenozoic Kinematic history of the Kohistan Fault
in the Pakistan Himalaya. Geological Society of America Bulletin, 120(11/12), 1428–1440.
410 A. Ali et al.

Donnelly, L. J. (2004). Geological investigations at high altitude, remote coal mine in Northwest
Pakistan and Afghanistan frontier, Karakoram Himalaya. Coal Geology, 60, 117–150.
Dutta, D., & Mukherjee, S. (2021). Introduction to Structural Geology and Tectonics Field
Guidebook—Volume 1. In S. Mukherjee (Ed.), Structural Geology and Tectonics Field Guide-
book—Volume 1. Switzerland: Springer Nature Switzerland AG. Cham. pp. xi-xvi. ISBN:
978-3-030-60142-3.
Fraser, J. E., Searle, M. P., Parrish, R. R., & Noble, S. R. (2001). Chronology of deformation,
metamorphism, and magmatism inthe southern Karakoram Mountains. Geological Society of
America Bulletin, 113, 1443–1455.
Frisch, W., Meschede, M., & Blakey, R. (2011). Plate tectonics (pp. 1–212). Berlin: Springer.
Gaetani, M., Garzanti, E., Jadoul, F., Nicora, A., Pasini, M., Tintori, A., & Kanwar, S. A. K. (1990).
The north Karakorum side of the Central Asia geo–puzzle. Geological Society of America Bulletin,
102, 54–62.
Gaetani, M., Jadoul, F., Erba, E., & Garzanti, E. (1993). Jurassic and Cretaceous orogenic events in
the North Karakorum: Age constraints from sedimentary rocks. In P. Treloar & M. Searle (Eds.),
Himalayan tectonics (Vol. 74, pp. 39–52). Geological Society of London Special Publication.
Gaetani, M. (1997). The Karakorum Block in Central Asia, from Ordovician to Cretaceous.
Sedimentary Geology, 109, 339–359.
Heuberger, S., Schaltegger, U., Burg, J. P., Villa, I. M., Frank, M., Dawood, H., et al. (2007). Age
and isotopicconstraints on magmatism along the Karakoram-Kohistan Suture Zone, NW Pakistan:
Evidence for subduction and continued convergence after India-Asia collision. Swiss Journal of
Geosciences, 100, 85–107.
Hussain, A., DiPietro, J. A., Pogue, K. R., & Ahmad, I. (2004). Geological Map of the 43B Degree
Sheet, Northwest Frontier Province, Pakistan: Geological Survey of Pakistan Degree Sheet Map
Series, map n. 11, scale 1:250,000.
Jagoutz, O., & Schmidt, M. (2012). The formation and bulk composition of modern juvenile
continental crust: The Kohistan arc. Chemical Geology, 298–99, 79–96.
Jan, M. Q., & Howie, R. A. (1981). The mineralogy and geochemistry of the metamorphosed basic
and ultrabasic rocks of the Jijal Complex, Kohistan, NW Pakistan. Journal of Petrology, 22(1),
85–126.
Jan, M. Q., & Windley, B. F. (1990). Chromium spinel–silicate chemistry in ultramafic rocks of the
Jijal Complex Northwest Pakistan. Journal of Petrology, 31, 667–715.
Kaneko, Y., Katayama, I., Yamamoto, H., Misawa, K., Ishikawa, M., Rehman, H. U., et al. (2003).
Timing of Himalayan ultrahigh–pressure metamorphism: Sinking rate and subduction angle of
the Indian continental crust beneath Asia. Journal of Metamorphic Geology, 21, 589–599.
Kazmi, A. H., & Jan, Q. M. (1997). Geology and tectonics of Pakistan. Karachi: Graphic Publishers.
Khan, M. A., Jan, M. Q., Windley, B. F., Tarney, J., & Thirlwall, M. F. (1989). The Chilas mafic–
ultramafic igneous complex; the root of the Kohistan island arc in the Himalaya of Northern
Pakistan. In L. L. Malinconico & J. Lillie (Eds.), Tectonics of the Western Himalayas (Vol. 232,
pp. 75–94). Geological Society of America Special Paper.
Khan, K. S. A., Latif, M., Fayaz, A., Khan, A. N., & Khan Z. S. M. (2000). Geological roadlog along
the Karakoram Highway from Islamabad to Khunjerab Pass. Government of Pakistan Ministry
of Petroleum and Natural Resources, Geological Survey of Pakistan
Khan, M. A., Stern, R. J., Gribble, R. F., & Windley, B. F. (1997). Geochemical and isotopic
constraints on subduction polarity, magma sources and paleo–geography of the Kohistan Arc,
northern Pakistan. Journal of the Geological Society of London, 154(Part 6), 935–946.
Le Fort, P., Michard, A., Sonet, J., & Zimmermann, J. L. (1983). Petrography, geochemistry
and geochronology of some samples from Karakoram axial batholith (northern Pakistan). In
F. A. Shams (Ed.), Granites of Himalayas, Karakoram and Hindu Kush (pp. 377–387). Lahore,
Pakistan: Institute of Geology, Punjab University.
Le Fort, P., Tongiorgi, M., & Gaetani, M. (1994). Discovery of crystalline basement and Early
Ordovician marine transgression in the Karakoram mountain Range. Geology, 28, 941–944.
Tectonic Framework of Northern Pakistan from Himalaya … 411

Miller, D. J., Loucks, R. R., & Ashraf, M. (1991). Platinum–group element mineralization in the
Jijal layered ultramafic–mafic complex, Pakistani Himalayas. Economic Geology (Bulletin of the
Society of Economic Geologists), 86, 1093–1102.
Papritz, K., & Rey, R. (1989). Evidence for the occurrence of Permian Panjal trap basalts in the
Lesser and Higher Himalayas of the western syntaxis area, NE Pakistan. Eclogae Geologicae.
Helvetiae, 82(2), 603–627.
Petterson, M. G., & Windley, B. F. (1985). Rb–Sr dating of the Kohistan arc–batholith in the Trans-
Himalaya of North Pakistan, and tectonic implications. Earth and Planetary Science Letters,
74(1), 45–57.
Petterson, M. G., & Windley, B. F. (1991). Changing source regions of magmas and crustal growth
in the trans–Himalayas: Evidence from the Chalt volcanics and Kohistan Batholith, Kohistan,
Northern Pakistan. Earth and Planetary Science Letters, 102, 326–341.
Pogue, K. R., Hylland, M. D., Yeats, R. S., Khattak, W. U., & Hussain, A. (1999). Stratigraphic
and structural framework of Himalayan foothills, northern Pakistan. In A. Macfarlane & R.B.
Sorkhabi (Eds.), Himalaya and Tibet: mountain roots to mountain tops (Vol. 328, pp. 257–274).
Geological Society of America Special Paper.
Pudsey, C., Schroeder, R., & Skelton, P. W. (1986). A Cretaceous (Aptiard Albian) Age for Island–
Arc Volcanics, Kohistan, N Pakistan. Contributions to Himalayan Geology (Vol. 3, pp. 150–168).
Hindustani Publication Corp.
Ringuette, L., Martignole, J., & Windley, B. F. (1999). Magmatic crystallization, isobaric cooling,
and decompression of the garnet bearing assemblages of the Jijal Sequence (Kohistan Terrane,
western Himalayas). Geology (Boulder), 27, 139–142.
Rolland, Y. (2002). From intraoceanic convergence to post–collisional evolution: Example of the
India-Asia convergence in NW Himalaya, from Cretaceous to present. Journal of the Virtual
Explorer, 8, 185–208.
Schaltegger, U., Zeilinger, G., Frank, M., & Burg, J. P. (2002). Multiple mantle sources during
island arc magmatism; U-Pb and Hf isotopic evidence from the Kohistan arc complex, Pakistan.
Terra Nova, 14(6), 461–468.
Searle, M. P., Rex, A. J., Tirrul, R., Rex, D. C. & Barnicoat A. (1989). Metamorphic, magmatic and
tectonic evolution of the centralKarakoram in the Biafo–Baltoro–Hushe regions of N. Pakistan.
In: L. Malinconico & R. J. Lillie (Eds.), Tectonics of the Western Himalaya (Vol. 232, pp. 47–74).
Geological Society of America, Special Paper.
Searle, M. P., et al. (1996). Geological map of north Pakistan and adjacent areas of northern Ladakh
and western Tibet.Scale 1:650,000. Oxford, England: Oxford University.
Searle, M. P., Khan, M. A., Fraser, J. E., Gough, S. J., & Qasim, J. M. (1999). The tectonic evolution
of the Kohistan-Karakoram collision belt along the Karakoram Highway transect, North Pakistan.
Tectonics, 18, 929–949.
Shah, I. M. S. (2009). Stratigraphy of Pakistan. Government of Pakistan Ministry of Petroleum and
Natural Resources (p. 381).
Shams, F. A. (1969).The Granites and the associated metamorphic .rocks of the Mansehra–Amb
State area, West Pakistan (Unpublished Ph.D. thesis). Punjab University, West Pakistan.
Shroder Jr., J. F., Khan, M. A., Lawrence, R. D., Madin, I. P., & Higgins, S. E. (1989). Quaternary
glacier chronology and neotectonics in the Himalaya of Northern Pakistan. In L. L. Malinconico
& R. J. Lillie (Eds.), Tectonics of the Western Himalayas (Vol. 232, pp. 275–294). Special Paper,
Geological Society of America.
Treloar, P. J., et al. (1989). K-Ar and Ar–Ar geochronology of the Himalayan collision in NW
Pakistan; constraints on the timing of suturing, deformation, metamorphism and uplift. Tectonics,
8(4), 881–909.
Treloar, P. J., Petterson, M. G., Jan, M. Q., & Sullivan, M. A. (1996). A reevaluation of the stratig-
raphy and evolution of the Kohistan arc sequence, Pakistan Himalaya: Implications for magmatic
and tectonic arc building processes. Journal of Geological Society (London), 153, 681–693.
412 A. Ali et al.

Treloar, P. J., O’Brien, P. J., Parrish, R. R., & Khan, M. A. (2003). Exhumation of early Tertiary,
coesite–bearing eclogites from the Pakistan Himalaya. Journal of the Geological Society, 160,
367–376.
Treloar, P. J., Vince, K. J., & Law, R. D. (2007). Two–phase exhumation of ultra high–pressure and
medium–pressure Indian Plate rocks from the Pakistan Himalaya. In: A. C. Reis, R. W. H Butler,
& R. H. Graham (Eds.), Deformation of the continental crust: The legacy of Mike Coward (Vol.
272, pp. 155–185). Geological Society [London] Special Publication.
Tonarini, S., Villa, I., Oberli, F., Meier, M., Spencer, D. A., Pognante, U., & Ramsay, J. G. (1993).
Eocene age of eclogite metamorphism in the Pakistan Himalaya: Implications for India–Eurasia
collision. Terra Nova, 5.
Williams, P. M. (1989). The Geology of the Besham Area, North Pakistan: Deformation and imbri-
cation in the footwall of the main mantle thrust. Geological Bulletin University of Peshawar, 22,
65–82.
Yamamoto, H., & Nakamura, E. (1996). Sm–Nd dating of garnet granulites from the Kohistan
complex, northern Pakistan. Journal of Geological Society (London), 153, 965–969.
Yamamoto, H., Kobayashi, K., Nakamura, E., Kaneko, Y., & Kausar Allah, B. (2005). U-Pb zircon
dating of regional deformation in the lower crust of the Kohistan arc. International Geology
Review, 47, 1035–1047.
Yin, A., & Harrison, T. M. (2000). Geologic evolution of the Himalayan Tibetan Orogen. Annual
Review of Earth and Planetary Sciences, 28, 211–280.
Yin, A. (2006). Cenozoic tectonic evolution of the Himalayan orogen as constrained by along–strike
variation of structural geometry, exhumation history, and foreland sedimentation. Earth-Science
Reviews, 76, 1–131.
Zanchi, A., & Gaetani, M. (2011). The geology of the Karakoram range, Pakistan: The new
1:100,000 geological map of Central-Western Karakoram. Italian Journal of Geosciences, 130,
161–262.
Zeitler, P. K., Tahirkheli, R. A. K., Naesser, C., Johnson, N., & Lyons, J. (1981). Preliminary fission
track ages from the Swat valley, northern Pakistan. Geological Bulletin University of Peshawar,
13, 63–65.
Structures of Lesser/Greater Himalaya
in and Around an Out-of-Sequence
Thrust in the Chaura-Sarahan Area
(Himachal Pradesh, India)

Rajkumar Ghosh and Soumyajit Mukherjee

Abstract Field and thin-section documentation of structures in detail have been


scarce from the out-of-sequence thrust (OOST) areas in the Himalaya. In this field
guide, we present meso- and micro-scale structures in and around the OOST from
the Chaura-Sarahan area (Himachal Pradesh, India).

1 Introduction

Several review papers present Himalayan tectonics as review (Yin 2006; Mukherjee
2013b). The southern portion of the Greater Himalayan Crystalline rocks of Himachal
Pradesh can be encountered from the south of Jhakri up to the north of Wangtu along
the Sutlej river section, Himachal Pradesh, India, along the National Highway 22A
(Fig. 1). The southern boundary of this study is the Jhakri Thrust (JT) where brittle
deformation and brecciated zones exist (Pandey et al. 2004; Miller et al. 2000; Misra
and Gururajan 1994). The JT can be considered as the Main Central Thrust-Lower
(MCTL ). The active JT plays a key role in the neotectonics of this area. Chambers
et al. (2008) recognized the Sarahan Thrust (ST) around the Sarahan village, in
between the Vaikrita Thrust (VT) at north and the JT at south. Singh (1979) reported
a dislocation zone in the Jeori area. These thrusts JT, Jeori dislocation and ST could
be a part of single larger-scale thrust. Singh (1980) described in great detail fold
morphologies from Sarahan and surrounding regions, but the tectonic interpretation
of these folds was not presented.
The Greater Himalayan Crystallines (GHC) lies between the Main Central Thrust
Zone (MCTZ) in the south and the South Tibetan Detachment System (STDS) in
the north. It comprises medium to high grade metamorphic rocks ranging from
kyanite/staurolite/garnet bearing schist to quartz rich mica schist.
The Jhakri Thrust, active since <4.5 Ma, is also locally named as the Jutogh
Thrust (Misra and Gururajan 1994). The Jakhri Thrust (Zone) dips 50° towards

R. Ghosh · S. Mukherjee (B)


Department of Earth Sciences, Indian Institute of Technology Bombay, Powai, Mumbai 400076,
Maharashtra, India
e-mail: soumyajitm@gmail.com; smukherjee@iitb.ac.in

© Springer Nature Switzerland AG 2021 413


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_13
414 R. Ghosh and S. Mukherjee

Fig. 1 Lithological map of Chaura area, Himachal Pradesh (Higher Himalaya). Reproduced from
Jain et al. (2000). MCT: Main Central Thrust; VT: Vaikrita Thrust; KT: Kulu Thrust. S7, etc.: sample
locations

NE and is presumably 15–17 km deep (Pandey and Virdi 2003). The MCT, or the
MCTU (Mukherjee and Koyi 2010), or the ‘True MCT’ based on difference in ENd
signature (Ahmad et al. 2010; Chambers et al. 2009) are the alternate designations
of the Vaikrita Thrust. The Chaura Thrust, an out-of-sequence thrust (OOST; Jain
et al. 2000; Mukherjee 2015a), is located near the ‘Kinnaur Dwar’ in the Kinnaur
district. The Chaura area lies within the MCTZ. Singh and Jain (1993) reported
ductile shear based on N/NE-dipping C-planes and associated S-planes. They also
reported a flattening type of finite stain from their study area. Main foliations near
the Chaura Thrust dip steeper (65–75°, at places even sub-vertical) than those further
south (45–50°).
Below we present structures from the transect as shown in Fig. 1. The transect
will be quite effective in demonstrating student Himalayan structures and explaining
their significance, such as fore-thrusting and fore folding. Mukherjee et al. (2019)
performed analogue models for out-of-sequence deformation (OOST). Using crustal
channel flow mechanism in dynamically scaled models in the lab, they explained the
genesis of the Greater/Lesser Himalayan OOST such as the Chaura Thrust. Recently,
Dutta and Mukherjee (2019) reported opposite ductile shear senses based on thin-
section study of rocks from this transect and explain it in terms of a general shear
mechanism.
Structures of Lesser/Greater Himalaya … 415

2 Structures

Y- and P-planes: These planes are curved. The brittle deformations are very promi-
nent in the Rampur Quartzite/Manikaran Quartzite. Y-planes in Jhakri dip towards
NE (Fig. 2). The Jhakri Thrust is located in between the quartzite and the mica schist.
This is indicated by highly fractured and crushed nature of the quartzites.
Veins: Quartz veins of various regular and irregular shapes are noted. Thickness of
the vein varies along their length although few cases also exist for uniformly thick
veins. Veins can be (i) parallel C-plane, (ii) fold and cut across C-plane, and they
are not bound by a pair of C-planes, (iii) isolated sigmoid veins show ductile top-
to-S/SW shear. Veins can be faulted or warped (Fig. 3). Sometimes across the vein
as cross-cutting elements, the gneissic foliation of the host rock is locally curved.
Thus, a flanking structure (Mukherjee 2014a) is defined (Fig. 4).

S-C fabric and other structures


Ductile sheared rocks are usually mylonitized. Oriented thin sections were prepared
such that those are perpendicular to main foliation and parallel to lineation. Where
lineations are absent sections were prepared perpendicular to main foliation and
parallel to the dip direction of the foliation planes. The C-planes can be wavy (Figs. 5,
6 and 7). The S- and the C-planes make <45° angles and usually range 20–30°.

Fig. 2 Brittle shear planes in Jhakri area (location: S5 in Fig. 1). a Sense of brittle shearing: primary
shear Y-plane dips towards NE, in quartzite
416 R. Ghosh and S. Mukherjee

Fig. 3 Top-to-S/SW brittle faulted quatz veins. a near Ponda, b near Taranda Devi temple, c, d
Nichar village

Curved and elongated mica grain represents prolonged deformation parallel to C-


plane. One needs to be cautious in interpreting structures in micro-scale as numerous
other features not necessarily related to ductile shear are also documented (Fig. 8).
Symmetric structures (Mukherjee 2017) are also documented that do not indicate
any shear sense. Interestingly, close to the Chaura Thrust, shear sense indicators in
thin section are extremely prominent (Fig. 9). Consult Passchier and Trouw (2005),
Mukherjee (2011, 2013b, 2014b, 2015b, 2020, 2021) and Misra and Mukherje
(2018) as easy texts to study shear senses. Refer to Finch et al. (2020) for recent
understandings on secondary ductile shear C/-planes.

Asymmetric and symmetric boudins and pods (Figs. 10, 11, 12, 13, 14, 15, 16
and 17)
Boudins and pinch-and-swell structures indicate layer-parallel local brittle/ductile
extension. These are found in quartzofeldspathic schistose rock with variation in
thickness and shape. Scar folds associated with boudins develop inside the inter-
boudin space. Boudin trend locally parallel the C-plane. Sometimes, fractures are
found restricted within the boudinaged clasts that do not continue within the host
rocks. Asymmetric boudins indicate ductile shear sense: mostly top-to-SW/SSW.
Symmetric clasts of boudins do not show shear sense. Intra-boudins morphological
variations were encountered: (i) internal foliation parallels to boudins boundary, (ii)
internal foliation at high-angle to the boudins boundary and (iii) curved internal
foliations inside the boudinaged clast.
Structures of Lesser/Greater Himalaya … 417

Fig. 4 Quartz veins of diverse shapes defining cross-cutting elements that sometimes locally swerve
the host fabric elements defined by gneissic foliation. Here, c is the only case where a locally folded
hinge region seems to be intruded by a vein. Near Manglad
418 R. Ghosh and S. Mukherjee

Fig. 5 a, b, d S-C fabric: top-to ~ S ductile shear (in images top-to-right). c Crenulation cleavage,
not to be confused with S-C fabric! No shear sense indicated. Location: S10 (see Fig. 1)

Fig. 6 Mica grains in quartzofeldspathic matrix. Out of the four examples, only a defines S-C
fabric with top-to-SW shear. Location: S10 (see Fig. 1)
Structures of Lesser/Greater Himalaya … 419

Fig. 7 Mica grains in quartzofeldspathic matrix. a, b, d gives ductile shear top-to-S/SW sense.
c micas wrap few quatzofeldspathic grains: overall symmetric shape, no shear sense indicated.
Location: S9 (see Fig. 1)

Deformed asymmetric clasts (Figs. 18 and 19)


Sigma-like structures of quartz and feldspar grains are common in field and under
thin sections. These reveal a top-to-S/SW shear. Few clasts are fractured, but still
then in some cases, the sigmoid shape of the clast can be easy deciphered.
420 R. Ghosh and S. Mukherjee

Fig. 8 Mica grains in quartzofeldspathic matrix. None of them are convincing shear sense indi-
cators. These are not S-C fabrics. Here, d shows strong folding restricted inside mica aggregates.
Location: S8 (see Fig. 1)

Fig. 9 Strong top-to-S shear sense indicated by sigmoid moca fish and curved recrystallized matrix
materials. Location: S8, Chaura Thrust area (see Fig. 1)
Structures of Lesser/Greater Himalaya … 421

Fig. 10 a–c Top-to-SW compressional shear revealed by asymmetric quartz veins, d top-to-SW
extensional shear revealed by a clast

Fig. 11 Top-to-S ductile shear revealed by quartz veins of various geometries


422 R. Ghosh and S. Mukherjee

Fig. 12 Quartz pods/boudinaged clasts of various sizes display top-to-S shear sense. b, d intensely
folded mylonitized rock

Fig. 13 a Boudin train, b, c zooed parts of the previous sub-figure, d slipped boudin of irregular
geometry. Black full arrows: scar folds
Structures of Lesser/Greater Himalaya … 423

Fig. 14 Fractures quartz veins show overall a top-to-S ductile shear. South of Wangtu

Fig. 15 Fractured quartz clasts, b, d: convincing top-to-S shear


424 R. Ghosh and S. Mukherjee

Fig. 16 Top-to-S ductile sheared quartz vein. Towards Majgaon

Fig. 17 Quartz veins showing clear cut boudinaging of various geometries. Black full arrows: scar
folds. Near Sarahan village
Structures of Lesser/Greater Himalaya … 425

Fig. 18 Clasts a, c, d and vein b showing shear sense. For b–d: internal foliation (Si ) defines the
shear sense. Near Chaura. In a, we kept two white boxes. Note the photograph and try to put half
arrows there

Fig. 19 Ductile shear sense deduced from clasts. a Looks like a delta structure (also see Mulchrone
and Mukherjee 2019, 2020). b Biotite inclusion inside quartz body gives a shear sense. See
Mukherjee (2014b) for inclusion of minerals giving shear sense
426 R. Ghosh and S. Mukherjee

Acknowledgements RG was funded by the Department of Sciences and Technology (New Delhi).
SM was supported by the flexible CPDA grant of IIT Bombay. Mohamedharoon A. Shaikh (MS
University Baroda) rearranged figures. The Springer team (especially Marion Schneider, Annett
Buettener, Boopalan Renu, Alexis Vizcaino, Doerthe Mennecke-Buehler) is thanked for numerous
assistance and proofreading. Dutta and Mukherjee (2021) encapsulate this work.

References

Ahmad, T., Harrism, N., Bickle, M., Chapman, H., Bunbury, J., & Prince, C. (2010). Isotopic
constraints on the structural relationships between the Lesser Himalayan series and the high
himalayan crystalline series, Garhwal Himalaya. Geological Society of America Bulletin, 112,
467–477.
Chambers, J. A., Caddick, M., Argles, T., Sherlock, Horstwood M., Harris, S. T., & Ahmad, N.
(2009). Empirical constraints on extrusion mechanisms from the upper margin of an exhumed
high-grade orogenic core, Sutlej valley, NW India. Tectonophysics, 477, 77–92.
Chambers, J. A., Argles, T. W., Horstwood, M. S. A., Harris, N. B. W., Parrish, R. R., & Ahmad,
T. (2008). Tectonic implications of Palaeoproterozoic anatexis and Late Miocene metamorphism
in the Lesser Himalayan Sequence, Sutlej Valley, NW India. Journal of the Geological Society,
165, 725–737.
Dutta, D., & Mukherjee, S. (2019). Opposite shear senses: Geneses, global occurrences, numerical
simulations and a case study from the Indian Western Himalaya. Journal of Structural Geology,
126, 357–392.
Dutta, D., Mukherjee, S. (2021). Introduction to Structural Geology and Tectonics Field Guide-
book—Volume 1. In: Mukherjee S. (Ed.), Structural Geology and Tectonics Field Guidebook—
Volume 1. Springer, Nature Switzerland AG. Cham. pp. xi-xvi. ISBN: 978-3-030-60142-3.
Finch, M. A., Bons, P. D., Steinbach, F., Griera, A., Llorens, M.-G., Gomez-Rivas, E., et al. (2020).
The ephemeral development of C shear bands: A numerical modelling approach. Journal of
Structural Geology. https://doi.org/10.1016/j.jsg.2020.10409.
Jain, A. K., Kumar, D., Singh, S., Kumar, A., & Lal, N. (2000). Timing, quantification and tectonic
modelling of pliocene-quaternary movements in the NW Himalaya: Evidences from fission track
dating. Earth Planet Science Letters, 179, 437–451.
Miller, C., Klotzli, U., Frank, W., Thoni, M., & Grasemann, B. (2000). Proterozoic crustal evolution
in the NW Himalaya (INDIA) as recorded by circa 1.8 Ga mafic and 1.84 Ga granitic magmatism.
Precambrian Research, 103, 191–206.
Misra, A.A., Mukherjee, S. (2018). Atlas of structural geological interpretation from seismic images.
Wiley Blackwell. ISBN: 978-1-119-15832-5.
Misra, D. K., & Gururajan, N. S. (1994). The closure of Rampur window in Satluj valley of Himachal
Pradesh: Some new observations. Journal of Himalayan Geology, 5, 127–131.
Mukherjee, S. (2011). Mineral Fish: their morphological classification, usefulness as shear sense
indicators and genesis. International Journal of Earth Sciences, 100, 1303–1314.
Mukherjee, S. (2013b). Deformation microstructures in rocks. Springer Geochemistry/Mineralogy.
Berlin. pp. 1–111. ISBN 78-3-642-25608-0.
Mukherjee, S. (2014a). Review of flanking structures in meso- and micro-scales. Geological
Magazine, 151, 957–974.
Mukherjee, S. (2014b). Mica inclusions inside host mica grains from the Sutlej section of the Higher
Himalayan Crystallines, India—Morphology and constrains in genesis. Acta Geologica Sinica,
88, 1729–1741.
Mukherjee, S. (2015a). A review on out-of-sequence deformation in the Himalaya. In: Mukherjee,
S., Carosi, R., van der Beek, P., Mukherjee, B.K., Robinson, D. (Eds.), Tectonics of the Himalaya
(412, pp. 67–109). Geological Society, London. Special Publications.
Structures of Lesser/Greater Himalaya … 427

Mukherjee, S. (2015b). Atlas of structural geology. Amsterdam: Elsevier. ISBN 978-0-12-420152-1.


Mukherjee, S. (2017). Review on symmetric structures in ductile shear zones. International Journal
of Earth Sciences, 106, 1453–1468.
Mukherjee, S. (2021). Atlas of structural geology (2nd Edn). Elsevier.
Mukherjee, S., Bose, N., Ghosh, R., Dutta, D., Misra, A.A., Kumar, M., Dasgupta, S., Biswas, T.,
Joshi, A., Limaye, M. (2019). Structural geological atlas. Springer. ISBN: 978–981-13-9825-4.
Mukherjee, S., & Koyi, H. A. (2010). Higher Himalayan shear zone, sutlej section-structural geology
and extrusion mechanism by various combinations of simple shear, pure shear & channel flow in
shifting modes. International Journal of Earth Sciences, 99, 1267–1303.
Mulchrone, K., & Mukherjee, S. (2019). Kinematics of ductile shear zones with deformable or
mobile walls. Journal of Earth System Science, 128, 218.
Mulchrone, K. F., & Mukherjee, S. (2020). Numerical modelling and comparison of the temporal
evolution of mantle and tails surrounding rigid elliptical objects in simple shear regime under
stick and slip boundary conditions. Journal of Structural Geology, 132, 103968.
Passchier, C. W., & Trouw, R. A. J. (2005). Microtectonics (2nd ed.). Berlin: Springer.
Pandey, A. K., & Virdi, N. S. (2003). Microstructural and fluid inclusion constraints on the evolution
of Jakhri thrust zone in the Satluj valley of NW Himalaya. Current Science, 84, 1355–1364.
Singh, K. (1979). Deformation history of the rocks around Sarahan Bushahr, Himachal Pradesh.
In P. S. Saklani (Ed.), Structural geology of the Himalaya (pp. 163–182). New Delhi: Today and
Tomorrow’s Printers & Publishers.
Singh, K. (1980). Deformation history of the rocks around Sarahan Bushair, Himachal Pradesh. In P.
S. Saklani (Ed.), Structural geology of the Himalaya (pp. 163–182). Delhi: Today and Tomorrow’s
Printers & Publishers.
Singh, S., & Jain, A. K. (1993). Deformational and strain patterns of the Jutogh nappe along the
Sutlej valley in Jeori-Wangtu region Himachal Pradesh, India. Journal of Himalayan Geology, 4,
41–55.
Yin, A. (2006). Cenozoic tectonic evolution of the Hima-layan orogen as constrained by along-strike
variation of structural geometry, exhumation history, and foreland sedimentation. Earth-Science
Reviews, 76, 1–131.
Structural Geology Along
the Nainital–Pangot Road (Kilbari
Section), Nainital Lesser Himalaya
(Uttarakhand, India): Focus
on Back-Structures

Mohit Kumar Puniya and Soumyajit Mukherjee

Abstract This chapter presents an interesting track in the Lesser Himalaya aimed
for structural geological fieldwork. We report back-structures, mainly back-faults,
from the Nainital–Pangot road. We expect geological trainers will find this track
interesting and will conduct tectonic fieldwork here. The Department of Geology
(Kumaun University) is located close to this study area, and it will be very easy to
bring the B.Sc. and M.Sc. students from this department in a say three day fieldtrip
to demonstrate and explain these structures.

1 Introduction

The Himalayan mountain belt is an example of continent–continent collision, started


50–55 Ma and resulted in 60–75-km-thick lithosphere (Hauck et al. 1998; Nelson
et al. 1996; Yin and Harrison 2000). This orogenic belt is characterized by compres-
sional thrusts that generally dip toward N, NNE, or NNW directions. Crustal short-
ening and tectonic uplift of the Himalaya is mostly controlled by major thrusts such as
the Main Central Thrust (MCT), the Main Boundary Thrust (MBT), the Himalayan
Frontal Thrust (HFT), or the Main Frontal Thrust (MFT) and other associated struc-
tures within the Indian continental crust (Yin 2006; Mukherjee 2015; Goswami et al.
2020).
Recently, back-structures, back-folds and back-faults have been documented from
several parts of the Himalaya (Mukherjee 2013; Bose and Mukherjee 2019a, b;
Mahato et al. 2019). It appears that such structures are integral parts of collisional
orogens.

M. K. Puniya
National Geotechnical Facility, Survey of India, 192/1 Kaulagarh Road, Dehradun, Uttarakhand,
India
S. Mukherjee (B)
Department of Earth Sciences, Indian Institute of Technology Bombay, Powai, Mumbai,
Maharashtra, India
e-mail: smukherjee@iitb.ac.in; soumyajitm@gmail.com

© Springer Nature Switzerland AG 2021 429


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_14
430 M. K. Puniya and S. Mukherjee

2 Geology of the Study Area

Kumaun Lesser Himalaya has been studied by several workers since long (e.g.,
Auden 1934; Heim and Gansser 1939; Fuchs and Sinha 1974; Hukku et al. 1974;
Pal and Merh 1974; Pande 1974; Valdiya 1980; Kumar et al. 2017; Sah et al. 2018).
The study area lies in the Outer Lesser Himalaya Sequence (LHS) sub-vertical road
section along the Nainital–Pangot road for ~18 km was studied (Fig. 1). This section
is also known as the Kilbari section locally. It is the northeastern part of a strip
of en-echelon basins of the Krol belt. In this traverse, one encounters rocks of the
Mussoorie Group (Valdiya 1988). The Mussoorie Group can be divided into, in
an upward direction in the succession, Blaini Formation, Krol Formation, and Tal
Formation in the investigated area. The main rock types in the area are quartzite of
the Blaini Formation, pyrite-bearing slates of the Kailakhan Member (Infra-Krol),
slates of the Manora Member (calcareous slates, greyish to greenish in color), purple
slates of Hanuman Garhi Member (ferruginous slates), and dolomites of the Upper
Krol Formation (Valdiya 1988).
The MBT is characterized by imbricating thrusts and faults (Valdiya 1984). Other
major faults in this area are the Nainital lake fault (Middlemiss 1890), Bhumiadhar

Fig. 1 Geologic map of the study area with back-structures locations. The map is reproduced from
Valdiya (1980). Red boxes: back-structure locations
Structural Geology Along the Nainital–Pangot Road … 431

Thrust, Kosi Transverse Fault and the Manora Fault (Valdiya 1988) (Fig. 1). The
Nainital Lake Fault passes through the Nainital Lake. This fault has dextrally offset
the MBT near Beluwakhan west of Jeolikote (Valdiya 1984) and the Manora Fault
near Alukhet (Fig. 1). Bhumidhar Thrust passes through the study area near Pangot
village and mark the boundary between the Infra-Krol slate at the hanging wall and
the Blaini quartzite at the footwall (Fig. 1). The Bhumiadhar Thrust is intersected by
the Kosi Transverse Fault and displaced it sinistrally (Fig. 1).

3 The Present Study

A fieldwork was carried out along the Nainital–Pangot road ~18 km long in the
Kumaun Outer Lesser Himalaya, Uttarakhand, India. This study focuses on the field
observations of back-structures (Fig. 2). Also, one can document Himalayan fore-
structures in the form of ductile top-to-south/-SSW shear (Figs. 3, 4). Back-structures
are identified from here based on S-C fabrics, brittle minor faults (Fig. 2a, d, h) and
P- and Y-brittle shears. P-planes slightly swerve close to the Y-planes (Fig. 2c).
In the Krol group of rocks, back shears are present in the form of fault planes
(Fig. 2a; Table 1) and P- and Y-planes (Fig. 2b, c, e; Table 1) with top-to-north shear.
The Blaini Formation rock is brittle top-to-NE back sheared. This is deciphered from
faulted veins and P- and Y-planes (Fig. 2h; Table 1). Except back- structures, top-to-S
shear are also documented in the form of P- and Y-planes and even asymmetric folds
(Fig. 4).
This study was confined in the NNE portion of Nainital syncline of Mussoorie
Group of rocks (Figs. 1 and 3). The Nainital syncline was faulted by the Nainital Lake
Fault (NLF), with NNW trend (Valdiya 1988; Kumar et al. 2017; Sah et al. 2018).
The nature and the geometry of the NLF were described by Valdiya (1988); it is a
pivot type fault with the swing tip is in SW of study area (Fig. 1). The locking point of
the NLF is near the outlet of the Nainital Lake (near the bus stand) in SSE direction
of this traverse (Valdiya 1988; Kumar et al. 2017). There is one more transverse
sinistral fault that displaced the Bhumiadhar fault in SE of the investigated area
(Fig. 1). Our study area lies at the hanging wall block of the Bhumiadhar Thrust and
the NLF. Due to these tectonic setting, there can be an extensional regime developed.
Whether the back-structures in this region developed due to the combined effect of
NLF, transverse fault, and Bhumiadhar Thrust needs to be studied.
432 M. K. Puniya and S. Mukherjee

Fig. 2 Field photographs, a parallel normal faults show top-to-NNW slip. Fault plane attitude:
124°/29°-SW. Manora Slate, b S-C structure: top-to-NE shear. S-plane attitude: 010°/41°-NW;
C-plane attitude: 340°/64°-SW. Manora Slate, c P- and Y-planes: Top-to-NE back-shear. P-plane
attitude: 220°/21°-NW. Y-plane attitude: 170°/12°-SW. Slate, d parallel normal faults show top-
to-NNW shear. Fault attitude: 330°/7°-SW. Slate of Krol Formation, e P- and Y-planes show top-
to-NE back-shear. ‘P’-plane orientation 330°/31°-SW and ‘Y’-plane 340°/38°-SW. Slate, f P- and
Y-planes: top-to-NE shear. P-plane attitude: 330°/45°-SW. Y-plane attitude: 340°/38°-SW. In slate,
g P- and Y-planes show top-to-NE slip. P-plane attitude: 360°/31°-S. Y-plane attitude: 010°/30°-
SW. In slate, h faulted veins in quartzite show top-to-NE back-slip. Normal fault plane’s attitude:
285°/47°-NNE
Structural Geology Along the Nainital–Pangot Road … 433

Fig. 3 Other field locations from where shear has been noted. The map is reproduced from Valdiya
(1980). Red boxes: spot locations

Fig. 4 Field photographs of shear senses, a pointed pen tip show NE shearing direction in Krol
slate at location E-1, b brittle shearing show top-to-SW direction in Infra-Krol slate at location E-4,
c asymmetric fold shows top-to-S slip. Location E-5, d P- and Y- brittle top-to-SSW shear. Location
E-6
434 M. K. Puniya and S. Mukherjee

Table 1 Back-shear data table from the study area


Backthrust field data
Location (see Fig. 1) P-plane Y-plane shear sense Location
B-1 010/41°-280° 340/64°-250° Top-to-40° 29°24 13.3 ,
79°26 54.4
B-2 300/20°-210° 235/18°-325° Top-to-N 29°24 19.5 ,
79°26 40.7
B-3 325/39°-235° N-S/28°-W Top-to-050° 29°24 33.5 ,
79°26 39.6
B-5 062/38°-SE 010/22°-100° Top-to-N 29°24 49 ,
79°26 55.6
B-6 N/7°-W 235/30°-145° Top-to-NNE 29°25 49 ,
79°26 46.
B-7 310/34°-220° horizontal Top-to-north 29°25 19.7 ,
79°26 15.4.
290/30°-200° horizontal Top-to-355° 29°25 19.7 ,
79°26 15.4.
B-7 300/21°-210° 350/26°-260° Top-to-N 29°25 19.7 ,
79°26 15.4.
B-8 330/35°-240° 340/20°-250° Top-to-045° 29°25 03.7 ,
79°25 44.2
B-9 N/31°-S 010/30°-280° Top-to-N 29°25 19.6 ,
79°25 49.3
B-10 260/18°-170° 245/21°-335° Top-to-350° 29°25 28 N,
79°25 39.4
B-11 315/26°-225° 350/34°-260° Top-to-N 29°25 42.9 N,
79°25 23.6
B-12 315/72°-225° 220/47°-310° Top-to-30° 29°25 46.2 N,
79°25 23.3

Acknowledgements CPDA grant of IIT Bombay and the Springer proofreading team. Thanks to
Marion Schneider, Annett Buettener, Boopalan Renu, Alexis Vizcaino, Doerthe Mennecke-Buehler,
and the proofreading team (Springer). Dutta and Mukherjee (2021) encapsulate this work.

References

Auden, J. B. (1934). Geology of the Krol Belt. Records of the Geological Survey of India, 67,
357–454.
Bose, N., & Mukherjee, S. (2019a). Field documentation and genesis of the back-structures from
the Garhwal Lesser Himalaya, Uttarakhand, India. In: Sharma, Villa, I.M., Kumar, S. (Eds.),
Crustal architecture and evolution of the Himalaya-Karakoram-Tibet Orogen (pp. 481, 111–
125). Geological Society of London Special Publications. https://doi.org/10.6084/m9.figshare.c.
4339784
Structural Geology Along the Nainital–Pangot Road … 435

Bose, N., & Mukherjee, S. (2019b). Field documentation and genesis of back-structures in ductile
and brittle regimes from the foreland part of a collisional orogen: Examples from the Darjeeling-
Sikkim Lesser Himalaya, India. International Journal of Earth Sciences, 108, 1333–1350. https://
doi.org/10.1007/s00531-019-01709-7
Dutta, D., & Mukherjee, S. (2021). Introduction to Structural Geology and Tectonics Field Guide-
book—Volume 1. In: Mukherjee S (Ed) Structural Geology and Tectonics Field Guidebook—
Volume 1. Springer Nature Switzerland AG. Cham. pp. xi-xvi. ISBN: 978-3-030-60142-3.
Goswami, T. K., Mahanta, B. N., Mukherjee, S., Syngari, B. R., & Sarmah, R. K. (2020). Orogen-
transverse structures in the eastern Himalaya: Dextral Riedel shear along the Main Boundary
Thrust in the Garu-Gensi area (Arunachal Pradesh, India), implication in hydrocarbon geoscience.
Marine and Petroleum Geology, 114, 104242.
Hauck, M. L., Nelson, K. D., Brown, L. D., Zhao, W. J., & Ross, A. R. (1998). Crustal structure
of the Himalayan orogen at similar to 90degrees east longitude from Project INDEPTH deep
reflection profiles. Tectonics, 17, 481–500.
Heim, A., & Gansser, A. (1939). Central Himalaya: Geological observations of the Swiss
Expeditionin 1936. Zurich: GebruderFretz.
Hukku, B. M., Srivastava, A. K., & Jaitle, G. N. (1974). Evolution of lakes around Nainital and the
problem of hillside instability. Himalayan Geology, 4, 516–531.
Kumar, M., Rana, S., Pant, P. D., & Patel, R. C. (2017). Slope stability analysis of Balia Nala
landslide, Kumaun Lesser Himalaya, Nainital, Uttarakhand, India. Journal of Rock Mechanics
and Geotechnical Engineering, 9, 150–158.
Mahato, S., Mukherjee, S., & Bose, N. (2019). Documentation of brittle structures (back shear
and arc-parallel shear) from Sategal and Dhanaulti regions of the Garhwal Lesser Himalaya
(Uttarakhand, India). In S. Mukherjee (Ed.), Tectonics and Structural Geology: Indian Context
(pp. 411–424). Cham: Springer International Publishing.
Middlemiss, C. S. (1890). Geological sketch of Nainital with some remarks on the natural conditions
governing the mountain slopes. Records of the Geological Survey of India, 21, 213–234.
Mukherjee, S. (2013). Higher Himalaya in the Bhagirathi section (NW Himalaya, India): Its struc-
tures, backthrusts and extrusion mechanism by both channel flow and critical taper mechanisms.
International Journal of Earth Sciences, 102, 1851–1870.
Mukherjee, S. (2015). A review on out-of-sequence deformation in the Himalaya. In: Mukherjee S,
Carosi R, van der Beek P, Mukherjee BK, Robinson D (Eds) Tectonics of the Himalaya (pp. 412,
67–109). Geological Society, London. Special Publications.
Nelson, K. D., et al. (1996). Partially molten middle crust beneath southern Tibet: Synthesis of
project INDEPTH results. Science, 274, 1684–1688.
Pal, D., & Merh, S. S. (1974). Stratigraphy and structure of the Nainital area in Kumaun Himalaya.
Himalayan Geology, 4, 547–562.
Pande, I. C. (1974). Tectonic interpretation of the geology of the Nainital area. Himalayan Geology,
4, 532–546.
Sah, N., Kumar, M., Upadhyay, R., & Dutt, S. (2018). Hill slope instability of Nainital City. Kumaun
Lesser Himalaya, Uttarakhand, India. Journal of Rock Mechanics and Geotechnical Engineering,
10, 280–289.
Valdiya, K. S. (1980). Geology of Kumaun Lesser Himalaya (p. 291). Dehradun: Wadia Institute of
Himalayan Geology Publication.
Valdiya, K. S. (1984). Aspects of tectonics: Focus on south-central Asia. New Delhi: TataMcGraw-
Hill Pub. Co., Ltd.
Valdiya, K. S. (1988). Geology and natural environment of Nainital hills, Kumaun Himalaya.
Nainital: Gyanodaya Prakashan.
Yin, A. (2006). Cenozoic tectonic evolution of the Himalayan orogen as constrained by along-strike
variation of structural geometry, exhumation history, and foreland sedimentation. Earth-Science
Reviews, 76, 1–131.
Yin, A., & Harrison, T. M. (2000). Geologic evolution of the Himalayan-Tibetan orogen. Annual
Review of Earth and Planetary Science, 28, 211–280.
Geology, Structural, Metamorphic
and Mineralization Studies Along
the Mandi-Kullu-Manali-Rohtang
Section of Himachal Pradesh, NW-India

Paramjeet Singh, Aliba Ao, S. S. Thakur, Shruti Rana, Rajesh Sharma,


A. K. Singh, and Saurabh Singhal

Abstract In the Himachal region of the NW-Himalaya almost all the lithotectonic
units of the Himalayan fold-thrust belts. This field excursion guide is designed to
showcase the structural geological marvels of the Mandi-Kullu-Manali section along
the Beas River valley from the Main Boundary Thrust (MBT) to the South Tibetan
Detachment System (STDS). This dip-section along the road provides an excel-
lent opportunity to study the structural imprints of MBT, Panjal/Chail Thrust (also
known as Main Central Thrust), Vaikrita Thrust and the position of STDS. In this
field guide chapter, we mainly present the deformation structures metamorphism
and economic mineral occurrences in the study area. This field excursion guide
encounters the following litho units: (i) metasedimentary rocks of Lesser Himalayan
Sequence (LHS), (ii) low- to medium-grade crystalline rocks of Lesser Himalayan
Crystalline thrust sheets including Chail and Jutogh groups and intrusive Paleozoic
granites, (iii) High-grade crystallines and migmatites of the Higher Himalayan Crys-
talline (HHC) including intrusive Paleozoic leucogranites, (iv) Tethyan Himalayan
Sequence (THS). Based on our field, structural and metamorphic observation, we
try to distinguish the Chail formation, Jutogh and Vaikrita Groups, which have been
considered as a single unit ‘Haimanta group’ (Webb et al. 2011 and reference therein).

1 Introduction

The Himalaya is one of the youngest mountain chains on the Earth formed during the
Cenozoic collision between Indian and Eurasian plate around ~55 Ma ago (Yin 2006;
Singh et al. 2020). This fold-thrust belt offers a unique opportunity to understand
the geological processes of the mountain building event including the structural and
metamorphic processes. Even after several decades of research on this orogen, the
geometry, kinematics and evolutionary history is poorly understood (Mukherjee et al.
2013, 2015). The geodynamic evolution of this mountain belt is largely controlled by
major normal/thrust faults such as South Tibetan Detachment System (STDS), Main

P. Singh (B) · A. Ao · S. S. Thakur · S. Rana · R. Sharma · A. K. Singh · S. Singhal


Wadia Institute of Himalayan Geology, #33 GMS Road, Dehradun 248001, India
e-mail: psinghgeol@gmail.com

© Springer Nature Switzerland AG 2021 437


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_15
438 P. Singh et al.

Central Thrust (MCT), Main Boundary Thrust (MBT), Main Frontal Thrust (MFT)
and their associate thrust systems (Mukherjee 2013, 2015; Patel et al. 2011a, b, 2015;
Singh and Patel 2015, 2017; Singh et al. 2012, 2017, 2020; Yin 2006). During the
Cenozoic, remobilization of the northern margin of the Indian plate transformed the
terrain into a vast metamorphic belt like Jutogh/Munsiari Group of Higher Himalayan
Crystalline (HHC) and the Lesser Himalayan Crystalline rocks (LHC) (Bhargava
et al. 1991; Patel et al. 2015; Singh 2014; Singh et al. 2012, 2020; Webb et al. 2011).
These crystalline rocks are thrust southwards on the metasedimentary zone of the
Proterozoic Lesser Himalayan Sequence (LHS) along the MCT (Martin 2017; Singh
et al. 2020). The kinematic evolution of the Himalayan fold-thrust belt is believed to
have undergone three phases of deformation: (1) the continental collision between
Indian-Eurasian plates during the Cenozoic period (ca. 55–50 Ma) that thrust the
THS over the Indian continental lithosphere (DeCelles et al. 2016; Mandal et al.
2019; Robinson and McQuarrie 2012; Webb 2013; Webb et al. 2011 and references
therein); (2) southward thrusting of the crystalline rocks along the MCT during
the Late Oligo–Early Miocene (ca 25 Ma), (Patel et al. 2011a, b, 2015; Singh and
Patel 2017; Singh 2014; Singh et al. 2012, 2020); (3) compression and folding of
transported crystalline rocks and development of Lesser Himalayan Duplex (LHD)
structures in the footwall of MCT during the late Miocene (Patel et al. 2015; Robinson
and Martin 2014; Robinson et al. 2006; Srivastava and Mitra 1994; Singh and Patel
2015, 2017).
In the Himachal region of NW-Himalaya, the geological and tectonic settings are
complex particularly regarding the north-dipping MCT. This sole thrust plays a key
role in the present-day architecture of the Himalayan fold-thrust belt. It is known with
different thrust names based on the regional identification of the Himalayan Meta-
morphic Belt (HMB). Some of the previous studies considered it as Punjal/Chail
Thrust (Thakur 1992), but recent studies (Webb 2013; Webb et al. 2011) present it
as MCT. The HMB zone consists of inverted metamorphic isograds from kyanite-
garnet-grade down to biotite grade. It is one of the largest ductile shear zones, and the
northernmost limit of the HMB is marked as STDS. The STDS separates the HMB
and the Tethys Himalayan Sequence (THS), which is covered by Late Precambrian
to Eocene sediments (Hodges 2000; Jain 2014). The STDS juxtaposes the unmeta-
morphosed rocks of THS and overrides the high-grade metamorphic rocks of the
HHC.
The field excursion documents structures observed, along the Mandi-Larji-Kullu-
Rohtang section of Beas valley in the Himachal Himalaya. Here, in this field extrusion
chapter, we highlight the following key geological characteristics/features in the
section:
(a) Definition and characterization of the MCT vis-à-vis the Panjal/Chail Thrust
(CT/PT), Vaikrita Thrust (VT) and South Tibetan Detachment System (STDS).
(b) Deformation pattern, shear sense analysis and metamorphic processes of the
HHC.
(c) Inverted metamorphism of the Himalaya, evolution of high-grade metamorphic
rocks and the mineralization in the different tectonic regimes.
Geology, Structural, Metamorphic and Mineralization … 439

2 Geology

We summarize the geological background of the major lithotectonic units of the


Himachal Himalaya, it consists of four major tectonostratigraphic domains along the
Beas River valley (Thakur 1992; Yin 2006; Webb et al. 2011). From South to North,
these are: (a) Sub-Himalayan Cenozoic sequence; (b) Lesser Himalayan Sequence
(LHS); (c) Higher Himalayan Crystallines (HHC); (d) the Tethyan Himalayan
Sequence (THS) (see Fig. 1).
(a) Sub-Himalayan Cenozoic sequence: The outermost part of the Himalaya
towards south comprises mainly of Siwalik fluvial molasses sediments of
Neogene (Cenozoic) time. It comprises lower shallow-marine sediments of
Paleocene–Eocene time and sediments from the continental origin of Miocene–
Pliocene time. The stratigraphic sequence that starts from the Eocene marines
were described as Subathu and Singtali Formations consisting of limestones,
shales, minor fine-grain sandstones and quartz arenite, followed by fluvial
deposits of upper Dharamshala Formation (equivalent to Kasauli Forma-
tion) containing sandstones, minor shales and siltstones. Similarly, the lower
Dharamshala Formation (equivalent to Dagshai Formation) consists of sand-
stones, siltstones, shales and the lower/middle/upper Siwalik rocks with sand-
stones, shales and conglomerates. The sub-Himalaya rises abruptly above the
Indo-Gangetic alluvial plains along the MFT between Nalagarh-Swarghat-
Bilaspur-Sundernagar section and overridden by the LHS rocks (Srikantia and
Bhargava 1998) along the MBT.
(b) Lesser Himalayan Sequence (LHS): The Lesser Himalayan rocks are mainly
autochthonous to para-autochthonous Proterozoic un-fossiliferous low-grade
metamorphic rocks, also known as the Lesser Himalayan Metasedimentary
Sequence (LHMS) Zone. They are tectonically overlain by low- to medium-
grade crystalline rocks, also known as the Lesser Himalayan Crystalline (LHC)
Zone (Gansser 1964; Srikantia and Bhargava 1976; 1998; Thakur 1992). The
LHMS is represented by Shali structural belt comprising of the Sundernagar
Group and Shali Group in the area and mainly comprised of the grey slates
and phyllites in the lowermost part, white-to-purple quartzite, grey, olive green
to purple shales, slates and phyllite (Srikantia and Bhargava 1998; Thakur and
Choudhury 1983; Yin 2006). Following geological features are observed along
the transect in the Beas valley section where the rocks of Lesser Himalayan
Sequence are exposed (Fig. 1): (i) the Proterozoic rocks north of the MBT, i.e.
LHMS zone; (ii) isolated Early Cambrian to Early Permian intruded igneous
bodies and outliers by the Proterozoic sequences; (iii) klippen of crystalline
thrust sheets such as the Chail/Jutogh thrust sheet (Bhargava and Srikantia 2014;
Bhargava et al. 1991); and (iv) windows of the Proterozoic/Paleozoic rocks (i.e.
Kullu-Larji-Rampur window) within the crystalline/Proterozoic rocks along the
antiformal valleys (Srikantia and Bhargava 1976, 1998; Thakur 1992).
(c) Higher Himalayan Crystallines (HHC): The core metamorphic part of the
Himalaya is known by many names such as Greater Himalayan Crystalline,
440 P. Singh et al.

Fig. 1 (a) Topography map based on the GTOPO30 digital elevation model, (U.S. Geological
Survey) of the Himalaya, (b) Geology cum Tectonics map of the Himachal Himalaya. (Modified
after Thakur and Rawat 1992; Pandey et al. 2003; Webb et al. 2011; Bhargava and Srikantia 2014)
black-dotted box indicates the study area and black stars shows the locations.MBT: Main Boundary
Thrust; CT: Chail Thrust; JT: Jutogh Thrust; PT/MCT: Panjal Thrust/Main Central Thrust; VT:
Vaikrita Thrust and STDS: South Tibetan Detachment System
Geology, Structural, Metamorphic and Mineralization … 441

Higher Himalayan Crystalline, Higher Himalayan Crystalline Sequence, Crys-


talline Core and Central Himalayan Crystalline sequences, which occupies the
highest elevation of the Himalayan range (Singh 2014). In the present study area,
it is characterized by a 50–65 km thick sequence of Paleoproterozoic to Ordovi-
cian (Parrish and Hodges 1996; DeCelles et al. 2000) crystalline rocks over-
riding the LHS through north-dipping south-vergent MCT/Punjal Thrust (PT)
in the south (Fig. 1). At a few places, it is also known as the Punjal/Chail Thrust
(Thakur 1992). The rocks of Jutogh Group are low- to medium-grade meta-
morphic rocks comprising of the biotite-chlorite schist, mica schist, gneisses
and with intrusive granites of Proterozoic to Paleozoic ages (1.8–0.5 Ga age)
(Singh et al. 2002, 2006, 2017, 2020; Miller et al. 2000) and are underlain
by the low-grade phyllite of the Chail formation (Mukhopadhyay et al. 1997).
The tectonic contact between them marks the Jutogh thrust which has been
considered to be a continuation of the MCT. The previous studies from this
section named it as Chail/Punjal Thrust (Thakur 1992), MCT/PT (Srikantia and
Bhargava 1998) and MCT (Thiede et al. 2009; Deeken et al. 2011; Webb et al.
2011). In this work, it is termed as MCT/PT based on present and previous
metamorphic studies. Further north, the Jutogh Group of rocks is overridden
by the high-grade metamorphic and migmatites rocks of Vaikrita Group. The
Vaikrita Group of rocks are bounded by the Vaikrita Thrust in south and STDS
in the north, which separates it from the THS (Burg et al. 1984; Searle 1986;
Singh 2014; Herren 1987; Pêcher 1991; Patel et al. 2011a, b). The rock types
consist of the remobilized thick basement of ~30 km having greenschist to upper
amphibolite grade metamorphism. This is exposed throughout the Himalaya.
It is recently described to have evolved due to ductile extrusion during early
Miocene since ~23 Ma along a broad northeast dipping intracontinental ductile
shear zone between coeval MCT/PT in the south and STDS in the north (Burg
et al. 1984; Searle 1986; Herren 1987; Pêcher 1991; Burchfiel et al. 1992; Patel
et al. 1993; Coleman 1998; Dezes et al. 1999; Robyr et al. 2006).
(d) Tethyan Himalayan Sequence (THS): In the north of the Higher Himalayan
Crystallines, a sedimentary sequence of the Tethys Himalaya is exposed along
the Chandra river valley. The rocks of THS mainly consist of a thick pile of
sediments ranging in age between Paleoproterozoic and Eocene (1840–40 Ma).
The siliciclastic and carbonate sedimentary rocks interbedded with Palaeo-
zoic and Mesozoic volcanic rocks and fossiliferous rocks exhibits typical plat-
form sediments of marine facies deposited on the northern passive margin of
the Indian continental plate (Gansser 1964; Baud et al. 1984; Garzanti et al.
1996; Gaetani and Garzanti 1991; Brookfield 1993; Garzanti 1993, 1999; Steck
et al. 1993; Critelli and Garzanti 1994; Liu and Einsele 1994). The THS rocks
either attained low-grade Cenozoic metamorphism or remain unmetamorphosed
(Steck et al. 1993; Yin 2006). The metamorphic equivalents are commonly
known as Haimanta group of rocks, whereas unmetamorphosed parts commonly
referred as Tethys Sedimentary Sequences (TSS).
442 P. Singh et al.

3 Field Excursion

Reaching Mandi Town


To access the area studied in this field excursion can be approached by Air/train/road
from Delhi to Chandigarh and then to Mandi town by road via Chandigarh-
Bilaspur-Sundernagar highway about 180 km further northwest. The journey on
the Chandigarh-Bilaspur-Sundernagar highway provides a glimpse of the Cenozoic
sedimentary deposits of the Siwalik Group of rocks. This journey provides a unique
opportunity to experience the splendid scenic beauty of Himalayan foothills, mesmer-
izing landscapes, elegant peaks and many more in-store to explore. The Mandi town is
situated on the bank of the Beas River, and the northeastward journey in the upstream
of the river provides a wide range of picturesque Beas valley. The entire stretch of the
study area (~180 km) can be covered by road starting from Mandi town and ending
at Rohtang La. The time taken depends on the time spent on the outcrops. In the
present field excursion guide, the field locations (spots) are numbered as location 1,
2, 3…, etc. (Fig. 1), starting from Mandi town (Fig. 2).
Pandoh Syncline (Chail formation of HMB)
Location 1: The observations of MCT/Punjal thrust zone along the Beas valley, just
north of Mandi Town, near the Hanuman Temple is recorded on the first day as
Location 1. Although the outcrop was partially covered with vegetation, the exposed
bedding plane/dipping direction of the strata can be seen. This outcrop belongs to the
Chail Formation and the rock type is chlorite schist with 2–3 m wide beds of foliated
quartzite dipping towards northwest (Fig. 3a). The rocks are very fine-grained, and the
mineral association is: quartz, chlorite, muscovite, epidote and oxides. The foliation
in the rocks is visible as defined by the parallel layering in chlorite + muscovite
association. The GPS location is N 31° 42.04 and E 76° 57.61 with the elevation
~770 m and the strike almost SW-NE with dip ~53° NW. Thakur (1992) suggested that
the crystalline thrust sheet exposed between Mandi-Manali sections is correlatable
with the Chail and Jutogh groups of eastern Himachal Pradesh.
Similar to the chlorite schist, the intruded quartz veins are also folded and indi-
cate southward shearing. Two phases of deformation with top to the south sense of
shearing are apparent in Fig. (3b). Top-to-S/SE ductile sheared sigmoid clasts of
quartz occur within chlorite schist.
Location 2: After ~2–3 km towards the north, next location is well-exposed Paleozoic
intrusive granite, which is known as the Mandi granite (Fig. 3d). The contact between
Paleoproterozoic crystalline rocks and the granite body is clearly visible. The granite
intruded presumably in the southern limb of the Pandoh syncline structure and has
a width of ~5–6 km. The granites contain biotite-muscovite grains and coarse grain
phenocrysts. The shape of these phenocrysts is rectangular or elliptical and elongated
that shows shearing along the direction of intrusion. The core part is massive and
leucocratic in nature with the aplitic veins within this granite (Thakur 1992). The
Mandi granite shows well-developed foliation at the contact zone. It also has some
Geology, Structural, Metamorphic and Mineralization … 443

Fig. 2 Geological and Tectonic map of the study area along the Beas valley of Himachal Pradesh,
NW-Himalaya (Modified after Thakur 1992)
444 P. Singh et al.

Fig. 3 (a) Chail formation outcrop and the rock type is chlorite schist with 2-3 m wide beds of
foliated quartzite, (b) top-to-S/SE ductile sheared sigmoid clusts of quartz were observed within
the chlorite schist and indicates two phases of deformation with top-to-south sense of shearing, (c)
a thin-section shows garnet poikiloblasts containing inclusions of quartz and opaque minerals are
wrapped by the dominant foliation of muscovite + biotite, (d) outcrop image showing porphyroclast
in the Paleozoic granite of Mandi Area

intrusive mafic parts within the medium to fine-grained granite (granite is medium to
fine-grained or mafics?) (Miller et al. 2000; Our field observation). The GPS location
is N 31° 39.75 and E 77° 02.47 with the elevation ~1035 m.
Location 3: After crossing the Mandi granite body, the next location is within the
Chail Formation and the southern limb of the Pandoh syncline structure. At this
location, the garnet zone is present and the rock contains: garnet, biotite, muscovite,
chlorite, plagioclase, epidote, quartz, tourmaline and oxides. Garnet poikiloblasts
contain inclusions of quartz, and opaque minerals wrapped by the dominant foliation
of muscovite + biotite (Fig. 3c). Within this dominant foliation, an older foliation
of chlorite + muscovite association is also observed (Fig. 4d). The quartz veins are
intruded discordant to the foliation plane in the first phase of deformation and folded
in the second phase during shearing in the south (inset Fig. 4a, b). Similarly, the
boudin structures of the quartz veins are also visible and indicate the second phase of
the deformation of the Chail group (Fig. 4a). The GPS location is N 31° 42.22 and
Geology, Structural, Metamorphic and Mineralization … 445

Fig. 4 (a, b) Chail Formation outcrop showing the quartz veins intruded across the foliation plane
in the first phase of deformation and folded in the second phase during shearing in the south, (c)
the northern limb of Pandoh syncline (near Hanogi Mata Temple), we observed that the dipping
sense of the bed has changed by dipping towards south, (d) the thin-section shows the two sets of
foliation plans and an older foliation (S1) of chlorite + muscovite associated with (S2), (e) showing
the first exposure of mylonitic gneiss belongs to the Chail formation near the Pandoh village Area,
(f) the thin-section from the northern limb of Pandoh syncline indicates the biotite flakes are much
larger and aligned parallel to the foliation

E 77° 01.90 with the elevation ~870 m. The strike of the foliation is almost NW-SE
and dip 25° due north/NE.
Location 4: Location 4, just north of the Pondoh village, is the first exposure of
mylonitic gneiss that belongs to the Chail Formation (Fig. 4e). The rock is fine-
grained compact mylonitic gneiss and is divided into mylonitic, proto-mylonitic and
ultra-mylonitic types based on their texture (Thakur 1992). The strike of the mylonitic
446 P. Singh et al.

gneiss is almost NW-SE and dip ~39° NE. The GPS location is N 31° 39.84 and E
77° 03.58 , and the elevation is ~863 m. This location is within the southern limb of
the Pandoh syncline. It is important to observe that the strike of the beds is slightly
changed towards the west and the dip amount is also changed to moderate.
Location 5: This location is after crossing the Pandoh hydroelectric project dam site,
situated in the lower part of the northern limb of Pandoh syncline (near the Hanogi
Mata Temple). It is observed that the dipping sense of the bed has changed as the beds
are dipping towards the south (Fig. 4c). The rocks consist of the quartz, muscovite,
chlorite, biotite and opaques. The rock is fine-grained and dominated by muscovite
+ chlorite flakes. In comparison, the biotite flakes are large and aligned parallel to
the foliation (Fig. 4f). A similar cross-over into the biotite zone in the core of Pandoh
syncline was also observed by Leger et al. (2013) and constrained the temperature
condition to about ~480 °C. The GPS location of this spot is N 31° 41.70 ; E 77°
08.54 , and the elevation is ~940 m. The strike of the bed is nearly SW-NE and dip
37° SW. After a distance of ~2 km, a garnet bearing mica schist can be seen again at
a tunnel construction site exposed along the DwadaNala. The GPS location of this
site is N 31° 41.95 and E 77° 09.79 , and the elevation is ~937 m.
Kulu-Larji-Rampur (KLR) Window: The existence of a tectonic ‘window’ within
the metamorphic zone of HMB was reported around the Rampur area by Berthelsen
(1951), who suggested its extension further westward up to Larji-Kullu area. Herein
he described a thick sequence of chlorite, phyllite and quartzite, which is considered
as part of Shali Structural belts of LHS, tectonically overlain by the metamorphics and
having their roots within the Higher Himalayan Crystallines (Srikantia and Bhargava
1998). The KLR window is characterized as the para-autochthonous structure (Dubey
and Bhat 1991) and an antiformal-fold structure with the axial zone of the fold
trending NNW–SSE. The KLR window is approximately 120 km long and is accurate
in its concavity towards southwest with a width of approximately 25 km. It is divided
into two tectonic zones, viz. lower zone of autochthonous nature is Larji structural
zone, known as Larji window zone. Similarly, the upper zone is the next higher
structural level and known as Rampur window which is para-autochthonous in nature
(Srikantia and Bhargava 1998).
Location 6: After a few km northward, we cross a ~1-km-long Larji tunnel is situ-
ated on the National Highway, and the location-6 is within a tectonic window zone
known as Kullu-Larji-Rampur window. The exposure of quartzite belongs to the
lower structural zone of window. Near the Aut village, the exposure of the rocks
are poor, but recent road-widening work in the nearby areas, help us to find out the
good rock exposures. The rocks exposed are off-white colour quartzite, which is
highly compact, massive, fine to course grained, sharp edges and is highly fractured
(Fig. 5a). The GPS location of this site is N 31° 44.62 and E 77° 12.49 , and the
elevation is ~1108 m. The strike of the bed is almost NW-SE, and dip is 29° SW. After
a few km distance towards the northwest direction, in Bhuntar town, quartzitic rocks
are exposed (Fig. 5b). It is also noticed that the fractured quartzites have developed
shearing sense indicator, i.e. top-to-south shearing (Fig. 5b). The rock is off-white
Geology, Structural, Metamorphic and Mineralization … 447

Fig. 5 (a) The rock exposed in the Kullu-Larji-Rampur window are quartzite with very compact,
massive, fine to course grained, sharp edges with highly fractured and off-white colour, (b) a massive
and highly fractured quartzite, have developed shearing sense indicator, i.e. top-to-south shearing

coloured, massive, compact and has high density. The GPS location of this site is
N 31° 49.72 and E 77° 10.36 , and the elevation is ~1112 m. The strike of the bed
slightly rotates towards the east direction and is almost NE-SW, whereas the dip vary
between 25 and 31° due W/SW.
Jutogh Group (Hanging wall of Jutogh Thrust)
Location 7: After moving further north of Kullu town, near the Kullu Bridge, a
huge cliff is seen on the Kullu-Bhukhali road. The exposed rock is a garnet bearing
mica schist with intrusive quartz veins. This rock is composed of garnet, muscovite,
chlorite, quartz, biotite, tourmaline and opaque. The quartz veins parallel to bedding
planes are folded during the shearing and clearly show top-to-south shearing (Fig. 6a).
The GPS location is N 31° 58.28 and E 77° 07.23, and the elevation is ~1292 m. The
strike of the bed slightly rotates towards the east direction and is almost NE-SW with
the dip amount 35° due NW. At some places, the garnet crystals are large in size and
even visible to the naked eye (Fig. 6b, c). The GPS location: N 31° 58.50 ; E 77°
07.33 and elevation is ~1353 metres. In these rocks, an increase in the size of the
garnet poikiloblast is observed, which are wrapped around by mica flakes defining
the foliation (Fig. 6d). The main foliation in the rock is also folded in some domains
(Fig. 6e).
Location 8: At location 8, towards Bhukali village, the interlayered garnet mica
schist and psammitic schist of the Jutogh Group were observed. The intruded quartz
veins are intensively deformed, sheared and folded (Fig. 7a–d). We observed the
sense of asymmetry of sigma (σ) type ‘tails’ of quartz porphyroclastic material in
the gneisses and showing top-to-south shearing sense is seen (Fig. 7a). Sigma (σ)
porphyroclast—tails extend on either side of the grain in the downstream direction
of the relative shear zone (Mukherjee 2011, 2015; Passchier and Trouw 2005). The
GPS co-ordinates of this location are N31° 58.73 , E77° 07.14 . Similarly, Fig. (7b)
presents the delta (δ)-type ductile sense of shearing (top-to-south) seen here. It also
shows the asymmetrical boudins structure along the Jutogh Thrust near to the Kullu
448 P. Singh et al.

Fig. 6 (a, b) The quartz veins parallel to bedding planes were folded during the shearing and clearly
shows top-to-south shearing sense; location Kullu-Bhukali road, (c) garnet bearing mica-schist, the
garnet crystals are large in size and even visible to the naked eye, (d, e) the thin-sections indicate an
increase in the size of the garnet poikiloblast in Kullu area and host rocks which is wrapped around
by mica flakes defining the foliation. Whereas the main foliation of the host rock is also folded in
some domains of Kullu area
Geology, Structural, Metamorphic and Mineralization … 449

Fig. 7 (a) Showing the sense of asymmetry of sigma (σ) type ‘tails’ of quartz porphyroclastic mate-
rial in the gneiss rocks and showing top-to-south shearing sense. Sigma (σ) porphyroclast—tails
extend on either side of the grain in the downstream direction of the relative shear zone (Mukherjee,
2011, 2015; Passchier and Trouw, 2005), (b, c, d) represent the delta (δ) type ductile sense of
shearing (top-to-south), delta-types (δ) develop when the core-object rotates rapidly and little recrys-
tallized material is produced. The tails of the bend resulted due to rotation and compression during
deformation, it changes its orientation and shape

Town. The delta-type porphyroclasts (δ) develop when the core-object rotates rapidly,
and some recrystallized material is produced (for counterview see Mulchrone and
Mukherjee 2020). Such tails of the bend result due to rotation and compression during
deformation, and consequently, a change in its orientation and shape occurs (Fig. 7b,
c). The GPS location is N31° 58.73 , E77° 07.14 and elevation 1558 m.
Location 9: Next location was near the Raison-Patli Khad, initially south of the
Raison village to document the exposure of mylonitic quartzite interbedded with the
mica-schist was observed (Fig. 8a, b). This location lies within the crystalline rocks
of the Jutogh Group. The phyllitic schists consist of the garnet, biotite, muscovite,
quartz, chlorite, tourmaline and the opaques. Here, the second-generation isoclinal
folds plunge either NW or SE at low angles and were associated with the crenulation
folds. The quartz veins and mineral lineation are stretched along a narrow zone
of mica schist (Fig. 8b). The general trend of beds is NE-SW direction, and the
schistosity shows dip varying between 37 and 45° due NW. The GPS location is
N32° 04.50 , E77° 07.17, and elevation is 1631 m.
450 P. Singh et al.

Fig. 8 (a) Near the Raison-Patli khad, exposure of mylonitic quartzite interbedded with the mica
schist of the Jutogh Group, (b) Quartz veins and mineral lineation represent a stretched along narrow
zone of mica schist

Around Manali Area (Footwall of the Vaikrita Thrust):


Location 10: The first exposure of the mylonitic gneiss and Paleozoic intrusive
granite of Jutogh Group were seen near to the Shanag village in the Manali area. The
mylonitized gneiss is course grained interlayered with thin beds of mica schist and
intruded by quartz vein parallel and/across the bedding planes (Fig. 9a). The intruded
quartz veins are folded due to the compression (Fig. 9b) and form the Phy (ϕ)-type
symmetry (Mukherjee 2017) as shown by porphyroclasts of feldspar in the granitic
rock of the Higher Himalayan Crystalline (Fig. 9c). Section parallel to the stretching
lineation and normal to the foliation show no stair-stepping. The GPS location is
N32° 16.22´, E77° 10.54´ and elevation is 2049 m. The Paleozoic intrusive granite
contains the intruded quartz veins, and the presence of tourmaline/mica at the edges.
In addition to this, some of these quartz veins are folded as well and form the boudin
structures (Fig. 9d). The structural data indicates that the trend of beds is NE-SW
direction and dip amount varies between 25 and 30° NW.
Location 11: A sharp contact in the form of Vaikrita Thrust, between Jutogh Group
rocks of Paleoproterozoic ages and Paleozoic–Ordovician granitic gneiss of Vaikrita
Group (Fig. 10d, e) is observed between the Kothi and Marhi village on the Manali-
Rohtang National Highway. The mylonitic gneisses are exposed along the road
wherein the intruded quartz veins are folded and shows shearing sense towards the
south. The granitic gneiss of Vaikrita Group is compact and mylonitized in nature
with some intrusive leuco-granite and Paleozoic granitoids. The intruded quartz veins
are ~2.0–2.5 cm thick and contain mafic intrusion within the quartz veins. The mafic
part is accumulated within the hinge zone as well as in the limbs of folded quartz
veins (Fig. 10a–c). The strike of beds is NE-SW direction and the dip amount varies
between 25 and 30° towards NW. It is observed that the dip amount of bedding plans
changes as compared to the previous locations
Location 12: This location is situated between Rohtang Pass-Sissu. At this location,
the augen gneiss and granitic gneiss are observed. The augens are stretched in north
south direction (Fig. 11a), and some of them indicate the shearing sense top-to-
Geology, Structural, Metamorphic and Mineralization … 451

Fig. 9 (a) The mylonitized gneiss is course grained interlayered with thin beds of mica schist and
intruded by quartz vein parallel and/across the bedding planes, (b) the intruded quartz veins are
folded due to compression, (c) the Phy (ϕ) type symmetry as shown by porphyroclasts (see also
Mukhejee 2014) of feldspar in the granitic rock of the Higher Himalayan Crystalline, (d) Quaterz
veins in the granitic gneiss are folded as well and form the boudin structures, Location near Manali
town

south. Similarly, another augen gneisses also represent the stretching structure and
have alternative shearing sense, i.e. top-to-south as well as top-to-north (Dutta and
Mukherjee 2019) (Fig. 11b). The gneiss of the Vaikrita Group also has intrusive
granite, which are known as Rohtang granite (Fig. 11c, d). After the ~500 m towards
Khokhsar (i.e. down to Rohtang Pass), the folded limestone beds of Tethys Himalaya
is observed near to Tandi bridge on the Leh-Manali national highway (Fig. 11e, f).
In addition to this, a well-exposed contact between Rohtang Granite and Tethys
Himalayan Sequence exists (Fig. 11g).
Mineralization in Larji-Kulu-Rampur Window:
Noteworthy occurrences of polymetallic sulphides and uranium are reported in and
around Kullu mainly in Larji-Rampur window in adjoining Parvati and Garsah valley
in the Lesser Himalaya. The mineralization in both these formations has been inter-
preted as epigenetic and hydrothermal in nature. However, it is unclear whether the
hydrothermal fluid is exsolved from magmatic rocks or reworked/migrated in shears.
In addition to the sulphides, uranium has also been reported earlier. The host rocks
for sulphides are largely the quartz veins and in quartzite, and uranium occurs in
Manikaran quartzites.
452 P. Singh et al.

Fig. 10 (a–c) The granitic gneiss of Vaikrita group are compact and mylonitized in nature with
some intrusive Leuco-granite and Paleozoic granitoids. The intruded quartz veins are approx. 2.0–
2.5 cm thick and contain mafic intrusion within the quartz veins. The mafic part assembled within
the hinge zone as well as limbs of folded quartz veins, (d, e) a sharp contact in the form of Vaikrita
Thrust, between Jutogh Group rocks of Paleoproterozoic ages and Paleozoic-Ordovician granitic
gneiss of Vaikrita Group, location national highway Manali-Leh road
Geology, Structural, Metamorphic and Mineralization … 453

Fig. 11 (a) On the Rohtang Pass-Sissu road section, The augen are stretched in north south direction
and indicating the shearing sense top-to-south, (b) a exposure of augen gneiss from the STDS
footwall, represents the stretching structure, and have alternative shearing sense, i.e. top-to-south
as well as top-to-north shearing sense, (c, d) represent the gneiss and granite are known as Rohtang
granite of the Vaikrita group, (e–g) represents the folded limestine beds of Tethys Himalaya near
to Tandi bridge on the Leh-Manali national highway and a well-exposed contact between Rohtang
Granite and Tethys Himalayan Sequence
454 P. Singh et al.

Location 13: Just before the Larji tunnel at Aut, the road bifurcation on the right side
connect this Highway to the Sainj and Banjar. A 17-km-long road section is from this
tunnel to Sainj, passing along the Sainj River through a bridge due to new road cutting,
the rock exposures are fresh on the right side of the river. Overall, the rocks are steeply
dipping, folded and lithology varies from purple grey limestone, dolomitic limestone,
massive to calcareous quartzites, slates and phyllites. An important spot is observed
near Behali on the right bank of Sainj River (N 31° 74.01 and E 77° 23.46 ), wherein
iron, sulphur leaching and malachite encrustations are seen indicating the presence of
a Fe-Cu mineralized zone (Fig. 12a). The host rocks are white massive to bedded and
jointed Hurla quartzite, showing thickness of the beds from about 10 cm to >60 cm.
These quartzites are folded, and dipping NW with >70°, belong to Hurla quartzite,
a unit of Larji Formation. Further moving to the left side, highly dipping bands
comprised of calcareous quartzite, phyllite and interlayered carbonaceous phyllites
are present here. These are fine-grained, milky white to greenish in colour. This 200 m
thick zone (N 31° 75.62 and E 77° 25.67 ; elevation ~1140 m) is marked with the
carbonaceous slates and phyllites at one side and by sericitic chlorite schist at the other
side. This section shows copper and iron sulphides mineralization, constrained within
the calcareous quartzite in the form of specs, fine veins, lenses and dissemination.
At the outcrop stains of malachite and iron, sulphur leaching is clearly seen. Careful
observations show that pyrite and chalcopyrite are present both as dissemination and
in the discordant quartz veins present in the fractured calcareous quartzite of Naraul
Formation. The strike of the beds is almost NNW-SSE, and dip is 55°–60° SW.
Location 14: On the road section from Bajaura to Mandi bypass, low-grade metamor-
phic rocks of the Chail Formation mainly carbonaceous and ash grey slate-phyllite,
quartzite, augen gneiss, quartz-chlorite-biotite schist, and at a few places garnetif-
erous chlorite schist. The GPS location is N 31° 82.11 and E 77° 11.15 ; elevation
~1611 m lies west and outside of the Larji-Kulu-Rampur tectonic window. The
outcrop is bedded quartzite rocks with significant leaching of iron and sulphur. The
extensive red, black and yellowish greenish encrustations is seen, and fluid activity
is envisaged in the fold axis and limbs. The high specific gravity of the hand speci-
mens from this outcrop indicates the enrichment of metals (Fig. 12b). After few km,
some evidences of sulphide minerals such as pyrite and minor galena and sphalerite
are found in quartz veins garnitifeours mica-schist. The dip of the bed is 62°SE, ~8
inches wide, and epigenetic quartz vein is highly fractured and sheared consisting
sulphide mineralization.
Location 15: About 22 km SE of Kullu towards east, Naraul is known for the old
working/mine of copper. The rocks exposed are phyllites, slates, white coloured
gritty, conglomeratic and massive quartzites of Chail Formation and GPS location
is N 31° 81.66, E 77° 23.64 at an elevation of 2044 m. The conglomeratic quartzite
consists of boulders and pebbles of gritty quartzite cemented in a siliceous matrix. The
regional strike of the formation varies from NNW-SSE to NW-SE with ~50° dip due
NE. The mineralization is confined to the conglomeratic quartzite, massive quartzite
and along the shear planes of quartzite. Significant mineralization occurs along joint
planes and shear zones, which mainly consists of chalcopyrite, covellite, pyrite and
Geology, Structural, Metamorphic and Mineralization … 455

Fig. 12 (a) Iron and sulphur leaching and the malachite encrustations manifesting Fe–Cu mineral-
ization in the host bedded and jointed Hurla quartzites; location right bank of Sainj River between
Behali-Spangini road section, Sainj Valley, (b) Quartzites exposure in the low-grade metamor-
phics of Chail Formation near Rahla on Bajaura-Mandi bypass, with substantial quartz veins and
leaching of iron and sulphur, (c) malachite, covellite and azurite encrustations in conglomeratic and
massive quartzite at the mineralized zone near Naraul, Garsah valley, (d) iron leaching in uraniferous
Manikaran quartzites in Dharmoar, Parvati valley

cuprite present as specs, aggregates of grains, veinlets and stringers. Disseminations


of sulphides are also present in the massive quartzite. The whole outcrop is stained
and encrusted with malachite and azurite along with limonite (Fig. 12c).
Location 16: From Kullu to Manikaran along the Parvati River valley, the massive
and hard Manikaran quartzites are present within the Larji-Rampur tectonic window.
Red coloured, finely-bedded, steeply dipping, highly jointed, fine-grained quartzites
are well-exposed on the roadside upstream of Parvati River. The fracture planes and
bedding planes within quartzite have patches of chlorite-fuchsite. These rocks bear
456 P. Singh et al.

variably oriented quartz veinlets, and very fine lamellae and veinlets of hematite
which impart a reddish colour to these quartzites. In Chinjjra village, Manikaran
quartzite is overlain by the chlorite phyllites, which contain pebbles of Manikaran
quartzite at various places. These rocks are known for the uranium contents, which
is hosted by the quartzites along their unconformable contacts with phyllites or in
the basal conglomerates. These rocks are trending ENE-WSW to ESE-WNW with
dip ~40° towards N. The GPS location of the area is N 31° 98.07 and E 77° 22.07
and elevation of ~1423 m. After few km Dharmoar area, the quartzites exposed are
massive with obscure bedding to highly jointed, sheared and brecciated quartzite.
Greenish chloritic flakes are observed along the joints and sheared planes (Fig. 12d).
The GPS location is N 31° 97.88 and E 77° 23.01 and elevation ~1673 m. These
reddish, greyish and white quartzite consist specks of copper sulphide mineralization
and are known for uraninite. The specks, streaks and stringers of chalcopyrite and
hematite with encrustations of malachite and azurite are seen at the outcrops.

4 Observations and Conclusions

The section exposed between Mandi-Larji-Kullu-Manali-Rohtang sections belongs


to three different lithotectonic units (i.e. Chail, Jutogh and Vaikrita Groups) of
the Himalayan mountain belts. Our structural, mineralization and metamorphic
studies revealed that all three lithotectonic units have undergone two different phases
of metamorphism and deformation. They also consist of economic minerals. We
conclude
• The structural studies from the Chail Formation (i.e. Mandi-Larji section) reveal
two phases of deformation; (i) the first phase was pre-syn orogeny. During this
phase, the quartz veins intruded and folded as open to close folds. (ii) The
second phase indicates the post-orogenic deformation which is clearly seen in the
deformed or folded quartz veins and the formation of recumbent/overturned folds
(Mukherjee et al. 2015). The second phase deformation also indicates shear-sense
indicator (top-to-south shearing) along the Chail/Punjal and Jutogh Thrusts.
• The mineralization studies suggest that ample iron and sulphur leaching together
with malachite encrustations are widely distributed in various localities in the area.
In the Sainj valley, pyrite, chalcopyrite and sphalerite are present. The mineral-
ization near Naraul appears significant and widespread occurrence of malachite,
azurite, covellite and chalcopyrite, pyrite are visible on the outcrops. The sulphide
mineralization is also seen within the vein. In Chinjjra and Dharmoar village, the
quartzites have patches of chlorite- fuchsite, oriented quartz veinlets, and very
fine lamellae and veinlets of hematite which impart a reddish colour. The reddish,
greyish and white quartzite consist of specks, streaks and stringers of chalcopy-
rite and hematite with localise dencrustations of malachite, azurite in uraniferous
quartzite.
Geology, Structural, Metamorphic and Mineralization … 457

• The metamorphic study reflects that the core part of Pandoh syncline reached up
to biotite grade of metamorphism and both limbs indicate slightly higher grade,
i.e. garnet-grade metamorphism. Subsequently, the Kullu area indicates probably
higher grade of metamorphism compared to the lower section (i.e. Mandi-Larji
section) as indicated by larger garnet grain size (up to 1 cm).
• Additionally, the section exposed between the Manali-Rohtang areas attained up
to the sillimanite grade (e.g. Thakur 1992; Webb et al. 2011 and reference therein)
but we have not observed sillimanite in the rocks in the present study indicating
the rare occurrence of this mineral in the rock. However, sillimanite bearing rocks
have been reported by earlier workers which suggests that a critical sampling
strategy and meticulous field work may be needed to be adopted to identify such
minerals in this area.

Acknowledgements The authors are thankful to Director, WIHG for allowing to carry out the field
work under the research project TAT 1.4 (contribution no WIHG/0020) and providing all facilities
and providing all facilities, infrastructure and moral support. P. Singh is thankful to Professors O. N.
Bhargava and R. C. Patel for fruitful discussion and encouragement. The authors are also thankful
to Soumyajit Mukherjee (IIT Bombay) for invitation, editorial handing and reviewing this chapter.
Dutta and Mukherjee (2021) summarizes this paper.

References

Baud, A. Y. M. O. N., Gaetani, M., Garzanti, E., Fois, E., Nicora, W., & Tintori, A. (1984).
Geological observation in southeastern Zanskar and adjacent Lahul area (northern Himalaya).
EclogaeGeologicaeHelvetiae, 77(1), 177–197.
Bhargava, O.N., & Srikantia, S.V. (2014). Geology and age of metamorphism of the Jutogh and
Vaikrita Thrust sheets, Himachal Himalaya. Himalayan Geology, 35(1), 1–15.
Bhargava, O.N., Bassi, U.K., & Sharma, R.K. (1991). The crystalline thrust sheets, age of
metamorphism, evolution and mineralization of the Himachal Himalaya. Indian Minerals, 45,
1–18.
Berthelsen, A. (1951). A geological section through the Himalaya. A Preliminary report, Bulltain
of geological Society of Danmark, 12, 167–234.
Brookfield, M. E. (1993). The Himalayan passive margin from Precambrian to Cretaceous times.
Sedimentary Geology, 84(1–4), 1–35.
Burchfiel, B. C., Zhiliang, C., Hodges, K. V., Yuping, L., Royden, L. H., Changrong, D. (1992).
The South Tibetan detachment system, Himalayan orogen: Extension contemporaneous with and
parallel to shortening in a collisional mountain belt (Vol. 269). Geological Society of America.
Burg, J. P., Brunel, M., Gapais, D., Chen, G. M., & Liu, G. H. (1984). Deformation of leucogranites
of the crystalline Main Central Sheet in southern Tibet (China). Journal of Structural Geology,
6(5), 535–542.
Coleman, M. E. (1998). U-Pb constraints on oligocene-miocene deformation and anatexis within
the central Himalaya, Marsyandi Valley, Nepal. American Journal of Science, 298(7), 553–571.
Critelli, S., & Garzanti, E. (1994). Provenance of the lower Tertiary Murree redbeds (Hazara-
Kashmir Syntaxis, Pakistan) and initial rising of the Himalayas. Sedimentary Geology, 89(3–4),
265–284.
458 P. Singh et al.

DeCelles, P. G., Carrapa, B., Gehrels, G. E., Chakraborty, T., & Ghosh, P. (2016). Along-strike
continuity of structure, stratigraphy, and kinematic history in the Himalayan thrust belt: The view
from Northeastern India. Tectonics, 35(12), 2995–3027.
DeCelles, P. G., Gehrels, G. E., Quade, J., LaReau, B., & Spurlin, M. (2000). Tectonic implications
of U-Pb zircon ages of the Himalayan orogenic belt in Nepal. Science, 288(5465), 497–499.
Deeken, A., Thiede, R. C., Sobel, E. R., Hourigan, J. K., & Strecker, M. R. (2011). Exhumational
variability within the Himalaya of northwest India. Earth and Planetary Science Letters, 305(1–2),
103–114.
Dezes, P. J., Vannay, J. C., Steck, A., Bussy, F., & Cosca, M. (1999). Synorogenic extension: Quan-
titative constraints on the age and displacement of the Zanskar shear zone (northwest Himalaya).
Geological Society of America Bulletin, 111(3), 364–374.
Dubey, A. K., & Bhat, M. I. (1991). Structural evolution of the Simla area, NW Himalayas:
implications for crustal thickening. Journal of Southeast Asian Earth Sciences, 6(1), 41–53.
Dutta, D., & Mukherjee, S. (2019). Opposite shear senses: Geneses, global occurrences, numerical
simulations and a case study from the Indian western Himalaya. Journal of Structural Geology.
Dutta, D., Mukherjee, S. (2021). Introduction to Structural Geology and Tectonics Field Guide-
book—Volume 1. In: Mukherjee, S. (Ed.), Structural Geology and Tectonics Field Guide-
book—Volume 1 (pp. XX). Springer Nature Switzerland AG. Cham. pp. xi-xvi. ISBN:
978-3-030-60142-3.
Gaetani, M., & Garzanti, E. (1991). Multicyclic history of the Northern India continental margin
(Northwestern Himalaya)(1). AAPG Bulletin, 75(9), 1427–1446.
Gansser, A. (1964). Geology of the Himalayas.
Garzanti, E. (1993). Himalayan ironstones,” superplumes,” and the breakup of Gondwana. Geology,
21(2), 105–108.
Garzanti, E. (1999). Stratigraphy and sedimentary history of the Nepal Tethys Himalaya passive
margin. Journal of Asian Earth Sciences, 17(5–6), 805–827.
Garzanti, E., Critelli, S., & Ingersoll, R. V. (1996). Paleogeographic and paleotectonic evolution
of the Himalayan Range as reflected by detrital modes of Tertiary sandstones and modern sands
(Indus transect, India and Pakistan). Geological Society of America Bulletin, 108(6), 631–642.
Herren, E. (1987). Zanskar shear zone: Northeast-southwest extension within the Higher Himalayas
(Ladakh, India). Geology, 15(5), 409–413.
Hodges, K. V. (2000). Tectonics of the Himalaya and southern Tibet from two perspectives.
Geological Society of America Bulletin, 112(3), 324–350.
Jain, A. K. (2014). When did India–Asia collide and make the Himalaya? Current Science, 254–266.
Leger, R. M., Webb, A. A. G., Henry, D. J., Craig, J. A., & Dubey, P. (2013). Metamorphic field
gradients across the Himachal Himalaya, northwest India: Implications for the emplacement of
the Himalayan crystalline core. Tectonics, 32(3), 540–557.
Liu, G., & Einsele, G. (1994). Sedimentary history of the Tethyan basin in the Tibetan Himalayas.
Geologische Rundschau, 83(1), 32–61.
Mandal, S., Robinson, D. M., Kohn, M. J., Khanal, S., & Das, O. (2019). Examining the tectono-
stratigraphic architecture, structural geometry, and kinematic evolution of the Himalayan fold-
thrust belt, Kumaun, northwest India. Lithosphere, 11(4), 414–435.
Martin, A. J. (2017). A review of definitions of the Himalayan Main Central Thrust. International
Journal of Earth Sciences, 106, 2131–2145.
Miller, C., Klotzli, U., Frank, W., Thoni, M., & Grassemann, B. (2000). Proterozoic crustal evolution
in the NW Himalaya (India) as recorded by circa 1.80 Ga mafic and 1.84 Ga granitic magmatism.
Precambrian Research, 103, 191–206.
Mukherjee, S. (2011). Flanking microstructures of the Zaskar shear zone, west Indian Himalaya.
YES Bulletin, 1, 21–29.
Mukherjee, S. (2013). Deformation microstructures in rocks. Springer Science & Business Media.
Mukherjee, S. (2014). Atlas of Shear Zone Structures in Meso-scale (pp. 1–124). Springer Geology.
Cham. ISBN 978-3-319-0088-6.
Geology, Structural, Metamorphic and Mineralization … 459

Mukherjee, S. (2015). Atlas of structural geology. Elsevier. Mukherjeet, S., 2015. A review on out-
of-sequence deformation in the Himalaya. Geological Society, London, Special Publications,
412, 67–109.
Mukherjee, S. (2017). Review on symmetric structures in ductile shear zones. International Journal
of Earth Sciences, 106(5), 1453–1468.
Mukherjee, S., Mukherjee, B. K., & Theide, R. C. (2013). Geosciences of the Himalaya–Karakoram–
Tibet orogen. International Journal of Earth Sciences, 102(7), 1757–1758.
Mukherjee, S., Carosi, R., vanderBeek, P. A., Mukherjee, B. K., Robinson, D. M. (2015). Tectonics
of the Himalaya. Geological Society of London.
Mukhopadhyay, D. K., Bhadra, B. K., Ghosh, T. K., & Srivastava, D. C. (1997). Ductile shearing
and large-scale thrusting in the Main Central Thrust Zone, Chur-Peak Area, Lesser Himachal
Himalaya. Journal of Geological Soceity of India, 50, 5–24.
Mulchrone, K. F., & Mukherjee, S. (2020). Numerical modelling and comparison of the temporal
evolution of mantle and tails surrounding rigid elliptical objects in simple shear regime under
stick and slip boundary conditions. Journal of Structural Geology, 132, 103968.
Pandey, A. K., Virdi, N. S., & Gairola, V. K. (2003). Evolution of structural fabrics and deformation
events in the Kulu-Rampur and Larji Window zones, NW Himalaya, India. Himalayan Geology,
24(1), 1–24.
Parrish, R. R., & Hodges, V. (1996). Isotopic constraints on the age and provenance of the Lesser
and Greater Himalayan sequences, Nepalese Himalaya. Geological Society of America Bulletin,
108(7), 904–911.
Passchier, C. W. & Trouw, R. A. (2005). Microtectonics (2nd Edn). Springer publications New
York.
Patel, R. C., Singh, S., Asokan, A., Manickavasagam, R. M., & Jain, A. K. (1993). Extensional
tectonics in the Himalayan orogen, Zanskar, NW India. Geological Society, London, Special
Publications, 74(1), 445–459.
Patel, R. C., Adlakha, V., Lal, N., Singh, P., & Kumar, Y. (2011a). Spatiotemporal variation in
exhumation of the Crystallines in the NW-Himalaya, India: constraints from fission track dating
analysis. Tectonophysics, 504(1–4), 1–13.
Patel, R. C., Adlakha, V., Singh, P., Kumar, Y., & Lal, N. (2011b). Geology, structural and exhuma-
tion history of the Higher Himalayan Crystallines in Kumaon Himalaya, India. Journal of the
Geological Society of India, 77(1), 47–72.
Patel, R. C., Singh, P., & Lal, N. (2015). Thrusting and back-thrusting as post-emplacement kine-
matics of the Almora klippe: Insights from low-temperature thermochronology. Tectonophysics,
653, 41–51.
Pêcher, A. (1991). The contact between the Higher Himalaya Crystallines and the Tibetan
Sedimentary Series: Miocene large-scale dextral shearing. Tectonics, 10(3), 587–598.
Robinson, D. M., & Martin, A. J. (2014). Reconstructing the Greater Indian margin: A balanced cross
section in central Nepal focusing on the Lesser Himalayan duplex. Tectonics, 33(11), 2143–2168.
Robinson, D. M., & McQuarrie, N. (2012). Pulsed deformation and variable slip rates within the
central Himalayan thrust belt. Lithosphere, 4(5), 449–464.
Robinson, D. M., DeCelles, P. G., & Copeland, P. (2006). Tectonic evolution of the Himalayan
thrust belt in western Nepal: Implications for channel flow models. Geological Society of America
Bulletin, 118(7–8), 865–885.
Robyr, M., Hacker, B. R., & Mattinson, J. M. (2006). Doming in compressional orogenic settings:
New geochronological constraints from the NW Himalaya. Tectonics, 25(2).
Searle, M. P. (1986). Structural evolution and sequence of thrusting in the High Himalaya, Tibetan-
Tethys and Indus suture zone of Zanskar and ladakh, Western Himalaya. Journal of Structural
Geology, 8(8), 923–936.
Singh, P. (2014). Exhumation history of the Higher and Lesser Himalayan crystallines in
the Kumaon-Garhwal region, NW-India: As revealed from Fission Track Thermochronology,
Unpublished Ph.D. thesis, Department of Geophysics, Kurukshetra University Kurukshetra.
460 P. Singh et al.

Singh, P. & Patel, R. C. (2015). Development of MBT and Tectonic dominance exhumation
within Kumaon-Garhwal Himalaya. Extended abstract in 30th Himalayan-Karakoram-Tibetan
Workshop held at Wadia Insitute of Himalayan Geology, Dehradun, India.
Singh, P., & Patel, R. C. (2017). Post-emplacement kinematics and exhumation history of the Almora
klippe of the Kumaun-Garhwal Himalaya, NW India: revealed by fission track thermochronology.
International Journal of Earth Sciences, 106(6), 2189–2202.
Singh, P., Patel, R. C., & Lal, N. (2012). Plio-Pleistocene in-sequence thrust propagation along the
main central thrust zone (Kumaon–Garhwal Himalaya, India): new thermo-chronological data.
Tectonophysics, 574–575, 193–203.
Singh, P., Bhakuni, S. S., & Singhal, S. (2017). Tectonic implications of U-Pb (zircon)
Geochronology of Chor Granitoids of the Lesser Himalaya, Himachal Pradesh, NW Himalaya.
AGU meeting, 11–14 Dec, 2017 held at New Orleans, USA.
Singh, P., Singhal, S., & Das, A. N. (2020). U-Pb (zircon) geochronologic constraint on tectono-
magmatic evolution of Chaurgranitoid complex (CGC) of Himachal Himalaya, NW India: impli-
cations for the Neoproterozoic magmatism related to Grenvillian orogeny and assembly of the
Rodinia supercontinent. International Journal of Earth Sciences, 109, 373–390. https://doi.org/
10.1007/s00531-019-01808-5.
Singh, S., Barley, M. E., Brown, S. J., Jain, A. K., & Manickavasagam, R. M. (2002). SHRIMP
U-Pb in zircon geochronology of the Chorgranitoid: evidence for Neoproterozoic magmatism in
the Lesser Himalayan granite belt of NW India. Precambrian Research, 118(3–4), 285–292.
Singh, S., Claesson, A., Jain, K., Gee, D. G., Andreasson, P. G., & Manickavasagam, R. M. (2006).
2.0 Ga granite of the lower package of the Higher Himalayan Crystallines, MagladKhad, Sutlej
Valley, Himachal Pradesh. Journal of the Geological Society of India 67(3), 295–300.
Srikantia, S. V., & Bhargava, O. N. (1976). Tectonic evolution of the Himachal Himalaya. Misc.
Publ. Geol. Surv. Ind, 34, 217–236.
Srikantia, S. V. & Bhargava, O. N. (1998). Geology of Himachal Pradesh (Vol. 2(1)). GSI
Publications.
Srivastava, P., & Mitra, G. (1994). Thrust geometries and deep structure of the outer and lesser
Himalaya, Kumaon and Garhwal (India): Implications for evolution of the Himalayan fold-and-
thrust belt. Tectonics, 13(1), 89–109.
Steck, A., Spring, L., Vannay, J. C., Masson, H., Bucher, H., Stutz, E., Marchant, R., Tieche, J.
C. (1993). The tectonic evolution of the Northwestern Himalayan in eastern Ladakh and Lahul,
India. In: Treloar, P. J., Searle, M. P. (Eds.), Himalayan tectonics (pp. 74, 265–276). Geological
Society Special Publication,.
Thakur V. C. & Choudhury, B. K. (1983). Deformation, metamorphism and tectonic relations of
Central Crystallines and Main Central Thrust in Eastern Kumaun Himalaya. In Himalayan Sears
(P. S. Saklani Edn, pp. 45–57). Himalayan Books, New Delhi.
Thakur, V. C. (1992). Geology of western Himalaya. Physics and Chemistry of the Earth, 19, 1–355.
Thakur, V. C., & Rawat, B. S. (1992). Geological map of Western Himalaya. Wadia Institute of
Himalayan Geology, Dehradun India, Scale 1: 10,00,000.
Thiede, R. C., Ehlers, T. A., Bookhagen, B. & Strecker, M. R. (2009). Erosional variability along
the northwest Himalaya. Journal of Geophysical Research: Earth Surface, 114(F1).
Webb, A. A. G. (2013). Preliminary balanced palinspastic reconstruction of Cenozoic deformation
across the Himachal Himalaya (northwestern India). Geosphere, 9(3), 572–587.
Webb, A. A. G., Yin, A., Harrison, T. M., Célérier, J., Gehrels, G. E., Manning, C. E., et al. (2011).
Cenozoic tectonic history of the Himachal Himalaya (northwestern India) and its constraints on
the formation mechanism of the Himalayan orogen. Geosphere, 7(4), 1013–1061.
Yin, A. (2006). Cenozoic tectonic evolution of the Himalayan orogen as constrained by along strike
variation of structural geometry, exhumation history, and foreland sedimentation. Earth Science
Review, 76, 1–131.
Tectonics and Channel
Morpho-Hydrology—A Quantitative
Discussion Based on Secondary Data
and Field Investigation

Mery Biswas, Ankita Paul, and Mostafa Jamal

Abstract Rivers and channels are extremely sensitive in changing their grades, base
levels and degree of meandering. We review a number of morpho-tectonic and hydro-
logical indicators along with minor-scale landforms. It explains the tectonic effect
on river channel and the landscape topography related with landform deformation.
Such indicators are closely associated with deflection zone of (backthrust) tilting,
slope and weak zone alteration, and change in hydrological parameters, e.g., velocity,
discharge, stream power and shear stress. The indicators also include the aggradation
as well as erosional landscape in minor-scale. We decipher 30 geomorphic indices
in assessing the impact of tectonics in channel morphology from three study sites:
the North East foreland basin of North Bengal, the Singbhum Shear Zone (SSZ) and
the Janauri–Chandigarh anticline. Analyses of morpho-hydrological parameters are
also applicable through statistical techniques such as Analytical Hierarchy Process
(AHP) and the Technique for Order of Preference is done by Similarity Ideal Solution
(TOPSIS).

1 Introduction

Structural and geomorphological features are the interlinked branches in geosciences


that can be approached by hydro-morphological and tecto-morphological means
(e.g., Guba and Glennie 1998; Claudio 2013). The exposed structure along the river
channel and the orogen are associated with sensitive indicators of river incision,
steeping of river gradient. Inherent geological characteristics and tectonic state of a
river channel before conducting a field investigation can be determined using satellite
images, aerial photographs, topographical map, geological map and seismic data
from USGS (Nakata 1989). For example, the concave or convex shape of the long
profile of rivers deduced from SRTM data decides the equilibrium condition of the
river due to endogenetic and exogenetic forces (Harmar and Clifford 2007; Sonam

M. Biswas (B) · A. Paul · M. Jamal


Department of Geography, Presidency University, 86/1 College Street, Kolkata 700073, West
Bengal, India
e-mail: mery.geog@presiuniv.ac.in

© Springer Nature Switzerland AG 2021 461


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_16
462 M. Biswas et al.

and Jain 2017). The pre-identification of knick points, break of slope and channel
shape from the long profile analysis formed due to tectonic anomalies are important
exercises. In field, channel geometries such as width, depth, width-depth ratio, bed
slope, valley side slope and valley side height and width ratio before and after the
break of slope can connote tectonic anomalies.
Further, field observation in studying channel morpho-hydrology includes selec-
tion of cross sections along the channel. For example, selected cross sections along
fault lines can identify the morphological anomalies of a channel due to tectonics.
Channel planform width and the wetted perimeter with the maximum depth of
channel from each cross sections are vital for computing its hydrological param-
eters, viz. velocity, discharge, stream power, shear stress (Charlton 2008). Topo-
graphic survey can be conducted using total station (e.g., LeciaFlexLine TS06plus,
FOIF total station OTS685-R500, etc.) and Digital Global Positioning System
(DGPS). Recently, laser scanning technology has been applied in terrain anal-
ysis. Elevation data collected using dumpy level and handheld GPS can be suit-
able for analysis of a terrain after error correction (e.g., Garmin Oregon 650, error
limit: 3.65 m). Extraction of 2D images from Landsat/Sentinel and 3D images
from SRTM/ASTER/Bhuvan images using Geographic Information System (GIS)
platform where a number of morphometric techniques (ruggedness index, drainage
pattern, drainage density, bifurcation ratio, relative relief, form ratio, slope, etc.) are
applied to define the tectonic disturbances.
Validation of such analyzed data is important in terms of actual field-based obser-
vations. We aim to portray the various geomorphic indices for assessing river channel
and its response to tectonics. We have selected three study sites using selective
geomorphic indices for computing data both from secondary and primary sources.

2 Study Areas

The study sites are selected from three different regions of India. We study topo-
sheets, SRTM data and conduct field survey. These are (i) the river Lish and Jayanti
from the north eastern Himalayan foothill zone of North Bengal (Fig. 1a), (ii) the
river Kankuram and Kharswati from Singbhum Craton of Jharkhand(Fig. 1b) and
(iii) the Janauri–Chandigarh anticline in the foothills of the Siwalik Himalaya, NW
India (Fig. 1c).
Lish is a tributary of Tista originated from the Lesser Himalaya flowing along
NNE–SSW that crosses the Main Boundary Thrust (MBT) and the Himalayan Frontal
Thrust (HFT). Jayanti is a river originated after the confluence of Katulum and
Sanchipu near the Indo-Bhutan border. It dissects MBT at its upper course. The
Janauri–Chandigarh anticline developed due to seismic cumulative slip during Late
Quaternary-Holocene (Thakur et al. 2009). Kharswati and Kankuram are the tribu-
taries of Subarnarekha joining its left and right banks, respectively. Both the rivers
dissect the Singbhum Shear Zone (SSZ) near its confluence with Subarnarekha River.
Tectonics and Channel Morpho-Hydrology—A Quantitative … 463

Fig. 1 Brief location map a The Lish in the Tista river system and Jayanti in the Torsa river system
in North East Himalaya. b The river Kankuram and Kharswati; tributaries of Subarnarekha at East
Singbhum District, Jharkhand c The Janauri anticline portraying its geological setting (Malik et al.
2007) Source Google earth Pro using Arc GIS Platform
464 M. Biswas et al.

Fig. 1 (continued)

These study sites have been selected on the basis of their tectonic significance in
controlling the channel morphology.

3 Geological Perspectives in Morpho-Tectonic Analysis

The extended foreland basin of Siwalik is composed of Quaternary sediments, carried


by the rivers flowing through the Greater Himalaya and exists as an undulating
sedimentary terrain from the Indus in the west and the Brahmaputra in the east
termed as the Indus Ganga Brahmaputra Plain (IGBP). The intermontane Soan Dun
is confined between the Lower Siwalik hills at its north and the Upper Siwalik in
south. The northern boundary of the Lower Siwalik and the Dun is known as the
Nalagarh Thrust (NaT; Mukherjee 2015) and the Southern limit as the backthrust
limb as the Janauri anticline (Malik et al. 2010). The 100 km SW–NE long Janauri
anticline is bordered by Beas in the west and Sutlej in the east. The elongated flat top
area is dissected by a number of opposite trending flowing stream, being controlled by
active tectonics. These streams are formed parallel to radial and dendritic drainage
pattern. The river channels are structurally controlled as evident in their drainage
pattern, orientation, shifting, sinuosity, cross section, long profile, riverbed slope,
bed terrain morphology, etc. The Siwalik in the eastern Himalaya is well exposed.
The Cenozoic foreland basin is presently tectonically active (Nakata 1989). The
Tectonics and Channel Morpho-Hydrology—A Quantitative … 465

Siwalik successions are buried below the Late Quaternary sediments (Sinha and
Parker 1996, Singh et al. 1999; Tandon et al. 2008).
Paleo-morphologic processes can be deciphered in terms of morpho-tectonic
indices of a channel and field-based hydrological data. Field data from channel
profile (latitude/longitude/maximum depth and altitude at a cross section), riverbed
topographic survey, meandering, valley side slope to height, etc., in different sections
of Kharswati and Kankuram basin portrays the presence of SSZ (Table 1).

4 Methodology

Two major database are used: secondary source of data based on satellite images
and primary-based data collected during field survey. Advanced spaceborne thermal
emission and reflection (ASTER G-DEM) or SRTM provides a near-global high-
resolution digital elevation model (DEM) with great advantages of homogeneous
quality of 30 m resolution. It may be used along with high-resolution Digital globe
GIS-ready Stereo DEMs of 2 m, 4 m or 8 m accuracy and resolution. Such enhanced
data sets are suitable for applying the tectonic indices in the platform like Arc
GIS 10.4.1, TNT mips2014, Quantum GIS, MATLAB, etc. Table 2 demonstrates
important geomorphologic indices.

5 Discussions

The tecto-morphological indices (Table 2) have been assessed from three different
regions.

5.1 Secondary Source of Data Sets

Among the several open source data, 30 m resolution band SRTM data (C30) DEM
and Google Earth Pro have been obtained. The relief ratio of Jayanti River computed
from the elevation data from its source to mouth indicates the break of slopes. This
implies the tectonic sensitivity. It is revaluated by the normalized long profile analysis
with its different representation as linear function(e.g., Fig. 2a), exponential function
(e.g., Fig. 2b), logarithmic function (e.g., Fig. 2c) and power regression model (e.g.,
Fig. 2d). These are the representation of slope-channel interrelation to detect the
lithology and profile anomalies of a stream in terms of grade. The differences in the
R2 values specify different level of tectonic imprint and variability over the rivers
and their basins.
The semi-logarithmic profile/hack profile (Hack 1973) of a river is an important
indicator of the inherent geological characteristics and deformation in the region.
Table 1 Brief description on geological formations and faults determining channel geomorphology at three individual study sites
466

Location Geological formations Significant Structure


1. North Bengal Himalayan Foothill Zone. • River Lish • Presence of MBT across the channel where the contact zone is
Considered Rivers: Darjeeling Gneiss-Daling Series-Damuda Series-Siwalik Tertiary- identified between Siwalik conglomerate and Schist.
• River Lish Garubathan Surface- Rangamati Surface- Alluvium (Ayaz et al. 2018) 26° 55 45 N/88° 37 30 E (Mukherjee 2013a, Mukherjee et al.
• River Kurti 2013, 2015)
• River Chel
• River Kurti and Chel • Three minor fault lines across the channel of Kurti as Matiali fault
• River Jayanti
Samsing Formation- Matiali Formation-Chalsa Formation- (MBT). Chalsa Fault (HFT) (Nakata 1989, Goswami et al. 2012,
Baikunthapur Formation- Shaugaon Formation. (Chakrabarti et al. Chakrabarti et al. 2012, Ayaz et al. 2018), Baradighi fault (Mullick
2012; Biswas 2015; Biswas and Banerjee 2018) and Mukhopadhyay 2011) (local tectonics)
• River Jayanti • Between Pathorjhora and Maynaguri, SE along the abandoned
Daling Series (Archean)—Buxa Series (Precambrian)—Epidorites- channel of Chel lays a NNE–SSW trending faults, along which the
Gondwana group- Siwalik (Mio-Pliocene)—Older Alluvium rivers like Chel, Lethi and Gish flowed once. Chel between
(Pliestocene to Recent). Ghosh (1968) Raina and Saha (1970), Ghosh Oodlabari and Rajadanga (26° 46 20 N, 88° 43 30 E; the NW–SE
and Ghosh (1980), Mitra (1983, 1985) trending channel reach is controlled by fault) (Bisaria, 1980)
• MBT and a major thrust between Siwalik-Gondwana and
Gondwana-Buxa Series, respectively. Major and minor fault affected
the geological setting of the chosen area for study(Ghosh and Ghosh
1980)
(continued)
M. Biswas et al.
Table 1 (continued)
Location Geological formations Significant Structure
Singhbhum Craton Collision zone between Chotanagpur Gneiss complex and Singhbhum Near the confluence of both the rivers, existence of SSZ has changed
Considered Rivers: Craton forming two major shearing: (A) Singhbhum Shear Zone and the bed slope and joined at an angle of near 90°. Along the Singbhum
• River Kankuram (B) Puruliya Shear Zone. North Singhbhum Mobile Belt and South Shear Zone (SSZ), River Kankuram has formed a waterfall which is
• River Kharswati Singhbhum Mobile Belt are divided by Dalma Ophiolite Belt (DOB) quite unusual in normal confluence of a tributary (Biswas and Biswas
(Middle part of Subarnarekha River) Major Formations in considered Basins: 2015)
• Somplipal Volcanics Secondly, they are under the deformation of Galudhi-Tetuldanga Fold
• Felsic Plutons Belt
• Kolhan and Dhanjhori molasse belt
• Older Metamorphic group
• Older Metamorphic Tonalite Gneisses
• Iron ore basins
This area includes unclassified metamorphic rocks in the Gangpur belt
in the west and north of the Dalma belt. The Dhanjori and Dalma
Groups are shown separately. Abbreviations for geologic units:
DVB—Dalma volcanic belt, DH—Dhanjori basin, ON—Ongarbira
volcanics, M—Malangtoli volcanics, SM—Simlipal volcanics;
Singhbhum Granite batholith (SBG) and adjacently located smaller
granite plutons, e.g., BG—Bonai Granite, CKG—Chakradharpur
Granite, KPG—Kaptipada Granite, MBG—Mayurbhanj Granite,
TG—Tamperkola Granite (remnants of older supracrustal Older
Metamorphic Group (OMG) and Older Metamorphic Tonalite Gneiss
Tectonics and Channel Morpho-Hydrology—A Quantitative …

(OMTG) not shown); GP—Gangpur belt, IOG—Iron Ore Group;


K—Kolhan cover sediments, EGGB—Eastern Ghats Granulite Belt,
NSMB—North Singhbhum Mobile Belt, PGN—Pala Lahara Gneiss,
SSZ—Singhbhum Shear Zone. Abbreviated localities: B—Besoi,
Bh—Bhaunra, H—Hata/Haludpukur, J—Jorapokhar, P—Pala Lahara,
S—Saraikela, Th—Thakurmunda. Index for inset map: B—Bastar
craton, BN—Bundelkhand craton, D—Dharwar craton,
S—Singhbhum North
Orissa craton. (b) Geological map of the study area, showing sample
locations for U–Pb zircon dating
(compiled from Sarkar and Saha 1963, Gupta et al. 1980, Gupta and
Basu 1985; Ray 1990; Saha 1994)
467

(continued)
Table 1 (continued)
468

Location Geological formations Significant Structure


3. Janauri–Chandigarh Anticline area This anticline is characterized by two segments of Siwalik: (A) Inner Both the inner Siwalik and outer Siwalik parts of Janauri anticline and
Siwalik Range and (B) Frontal Siwalik Range, Thakur et al. (2009) Chandigarh anticline are displaced by N–S trend fault line and saddles
Major Formations: of topography created by fold crest of Janauri anticline. It also
• Indo-Gangatic Plain represents the different drainage pattern in Chandigarh anticline. River
• Siwalik (Lower, Middle and Upper) Sutlej is flowing through the Siwalik range following the N–S trend
• Dharamsala Formation fault line where western part of the river is down throw side and the
• Sabathu eastern part is an up throw side. The total area is a compilation of five
• Lesser Himalaya Metasediment (Bilaspur LS) major fault lines, transverse to Himalaya forming four segments of
• High Crystalline/Tethys Himalaya displacements. Both the Siwalik ranges are experienced by tectonic
• Undifferentiated Bed rock sensitivity in major to minor spatial entity determining different
drainage pattern
Delcaillau et al. (2006), Mukherjee et al. (2015), Mukherjee (2012)
M. Biswas et al.
Table 2 Geomorphic indices in assessing channel morphology and anomalies due to tectonics
Sl. No. Indices Geometric Calculation References *
1. Normalize long profile The linear function y = ax + b e.g., Lee and Tsai (2009), Kale et al. MORPHO-TECTONIC
The exponential function y = aebx (2014), Paul and Biswas 2019 ANALYSIS
The logarithmic function y = a ln x + b
The power regression model y = axb
Where y is the elevation (H/H 0 ; H =
elevation of each point, H 0 = elevation
of the source), x is the length of the
river (L/L 0 ; L = distance of the point
from the source, L 0 = total length of the
stream), a and b are the coefficients
derived independently from each
profile. The R2 value determines the
best fit. The curve with highest R2 value
is the best fit curve
(Extracted from SRTM 30 m DEM and
formulated using Arc GIS 10.4 and
TNT mips 2014 platform)
1
Tectonics and Channel Morpho-Hydrology—A Quantitative …

2. Concavity index(º) Ceh = S2−S1 E e.g., Wobus et al. (2006), Whipple et al.
Where S 1 is the channel slope prior to (2007)
disturbance, S 2 is the channel slope
after disturbance (e.g., due to a change
in incision rate E) and E is the
difference between the incision rate
before and after disturbance
(continued)
469
Table 2 (continued)
470

Sl. No. Indices Geometric Calculation References *


f
3. Stream gradient index (SL) SL = ln D2− ln D1 Hack (1957)
Where f = fall in elevation (e2 –e1 )
ln = Natural logarithm of the
cumulative distance
* Higher SL value indicates tectonic

control over stream


4. Cross section analysis The data is collected from the field using e.g., Biswas and Biswas (2015)
5. Valley side slope total station/dumpy level, measuring
tape, staff, clinometers and GPS
Ref. Figure 14 for cross section analysis
6. Drainage basin asymmetry (Af ) Af = 100 × AArt Hare and Gardner (1985), Cox (1994),
Where Ar is the area of the basin to the Keller and Pinter (2002)
right of the trunk stream while facing
downstream. At is the total area of the
drainage basin. Af > 50 or < 50 states
unstable ground setting
(Extracted from SRTM 30 m DEM and
formulated using Arc GIS 10.4
platform)
 
7. Hypsometric curve analysis (HC) h Pike and Wilson (1971)
H
HC = a
( ) A
Where (h/H) = relative height versus
(a/A) = relative area
(continued)
M. Biswas et al.
Table 2 (continued)
Sl. No. Indices Geometric Calculation References *
8. Hypsometric integral (HI) ( H mean−H min) Strahler (1952), Schumm (1956),
HI =
( H max−H min) Andreani et al. (2014)
Where H mean = Mean elevation of
the basin, H min = Minimum elevation
of the basin, H max = Maximum
elevation of the basin.
HI value  0.30 states tectonically
stable basin and  0.30 indicated
tectonically unstable basin
2V f w
9. Valley floor width to height ratio V f = [(Eld − E sc )+(Er d −E sc )] Bull and McFadden (1977)
(V f ) Where V fw is the valley width, E ld is
altitude of the left bank, E rd is altitude
of the right bank and E sc altitude of the
channel
Lower the V f value (<1) indicates active
tectonic activity (V-shaped valley), and
higher V f value (>1) indicated lower
tectonic activity (U-shaped valley) of
Tectonics and Channel Morpho-Hydrology—A Quantitative …

the river
10. Steepness index (k s ) ks = AS−θ e.g., Snyder et al. (2000), Kirby and
Where S = Local slope. A = Drainage Whipple (2001), Vanlaningham et al.
basin area, θ = stream concavity (2006)
(Slope of a basin can be measured from
SRTM 30 m resolution using Arc GIS
software)
Increase in the ks represents anomalous
upliftment rate and vice versa
(continued)
471
Table 2 (continued)
472

Sl. No. Indices Geometric Calculation References *


11. Riverbed slope S = ks A−θ e.g., Noss and Larke (2016)
Where ks = Steepness index, A =
Drainage basin area, θ = stream
concavity
(Slope of a basin can be extracted from
SRTM 30 m DEM)
Da
12. Transverse topographic symmetry T = D d
Sajadian et al. (2015), Takieh et al.
factor (T ) Da = distance between the midline of (2015)
the drainage basin and the active
meander belt midline and Dd = distance
between the midline and the basin
divide
If the river flows through the midway of
the basin, the resulting (T ) would be ‘0’
indicates symmetric basin. If the value
is >0, the river basin is asymmetric
13. Asymmetry factor (Af ) Af = AArt × 100 Hare and Gardner (1985), Hamdouni
Ar is the area of the basin to the right of et al. (2008), Mahmood and Gloaguen
the river flowing downstream. At is the (2012)
area of the entire basin
Af smaller or greater than 50 shows
active tectonic, lithological control and
differential erosion
(continued)
M. Biswas et al.
Table 2 (continued)
Sl. No. Indices Geometric Calculation References *
14. Hack profile (H) H = C − K × log (L) Hack (1973), Chen et al. (2006), Magar
H is altitude of the profile resembling and Magar (2016)
the change in slope demonstration
tectonic activity and topography of the
river. C is constant (regression of y). K
is the SL index, and L is the length of
the stream from its source.
15. Mountain front sinuosity index Smf = Lmf
Ls Bull and McFadden (1977), Keller and
Where Lmf = Mountain front Pinter (2002), Silva et al. (2003), Bull
planimetric length (2007)
L s = Length of the mountain front
measured along the straight line

16. Tilt angle  b 2 2 Pinter and Keller (1995)
β = arc cos 2
a sin a + cos a

Where a =Depositional slope, a =


Maximum half length of fan, b =
Maximum half length of fan
Tectonics and Channel Morpho-Hydrology—A Quantitative …

(continued)
473
Table 2 (continued)
474

Sl. No. Indices Geometric Calculation References *


Channel Length
17. Sinuosity index(SI) SI = Valley length Brice (1964), Schumm and Khan
Straight channel values <1.05 (1972), Miall (1977), Biswas and Dhara
Sinuous values between >1.05–<1.50 (2019)
Meandering channel indicates >1.50
• Channel Index (CI)
CI = CL/AL
Where CL = Length of the channel; AL
= Shortest distance between source and
mouth.
• Valley index (VI)
VI = VL/A
Where VL = Length of the valley
between the base of the valley walls;
AL = Shortest distance between source
and mouth.
• Hydraulic sinuosity index (HSI)
HSI = (CI – VI)/(CI-1)
• Topographic sinuosity index
TSI = (VI − 1)/(CI − 1)
• Standard sinuosity index (SSI)
SSI = CL/VL
(continued)
M. Biswas et al.
Table 2 (continued)
Sl. No. Indices Geometric Calculation References *
18. Channel Network Analysis
a. Basin shape index (Bs ) Bs = BBw1 Bull and McFadden (1977),
Where B1 = Length of the basin and Bw Ramírez-Herrera (1998)
= Width of the basin measured at its
widest part
Greater the Bs value, more tectonically
active is the basin representing its
elongated shape. Lower the Bs value,
lesser the basin is tectonically active
representing the circulatory shape
b. Basin perimeter (Method: Computed by Arc GIS
c. Basin area software 10.3.1 using SRTM data 30 m)
d. Basin length
A
g. Form ratio (Rf ) Rf = L2
Brzezinska-Wójcik and Gawrysiak
(2010)
Where A = Area of the basin (km2 ), L2
= Square of the basin length Rf near 0
Tectonics and Channel Morpho-Hydrology—A Quantitative …

indicates tectonically active basin and


near 1 shows tectonically inactive basin
2
b. Lemniscate coefficient (k) k = π4LA Chorley (1971)
Where L = Length of the river basin
A = Area of the river basin
Tectonically inactive basins have lower
k value than active tectonic basins
(continued)
475
Table 2 (continued)
476

Sl. No. Indices Geometric Calculation References *


c. Fractal dimension of drainage FD = lim Log N (S)/Log(1/s) e.g., Gloaguen et al. (2007), Mahmood
pattern (FD) s−0 and Gloaguen (2011)
Where N(S) is the number of boxes and
s is the length of the box size applied.
The slope of the best fit line for the
log-log plot of N(S) and 1/s is equal to
FD
* Lower FD values denote river basin

highly valuable to neotectonic


deformation and vice versa
19. Velocity (m/s) V = dt Charlton (2008) HYDROLOGICAL
Where D = Distance, T = Time CHARACTERISTICS
20. Discharge (m3 /sec) Q =V ×W ×D
Where V = V elocity, W = Width, D =
Depth
21. Cross-sectional area (m2 ) D×W
Where D = Depth, W = Width
22. Wetted perimeter (m) Depth on the right bank + depth on the
left bank + channel width
23. Hydraulic radius (m) Cross-sectional area/wetted perimeter
24. Stream power () Ω = ρg Q S
Where ρ is water density, g is the
acceleration due to gravity, Q is the
discharge, S is the channel slope
(continued)
M. Biswas et al.
Table 2 (continued)
Sl. No. Indices Geometric Calculation References *
25. Shear stress (Du Boys equation) τ0 = ρgh S
Where ρ is water density, g is the
acceleration due to gravity, h is the
depth of the flow, S is the slope
R 0.67 s 0.5
26. Manning equation v= n
v = Velocity, R = Hydraulic radius, s
= Channel slope, n = Manning
roughness coefficient (Usually
determined from manning’s table)
√ V
27. Hydraulic stream flow (Froude’s (0.93×D)
number) Where V = Velocity, 0.93 =
Gravitational constant and D = Depth
V 12 V 22
28. Energy loss ΔE = y1 + 2g − y2 + 2g

q 2
1 1
ΔE = 2g y12
+ y12
− (y2 − y1)
Where V 1/V 2 = Mean velocity, y1 /y2 =
Tectonics and Channel Morpho-Hydrology—A Quantitative …

Conjugated depth, q =
Discharge/Width, g = Acceleration due
to gravity.
29. River terraces Prepared from total station survey from e.g., Biswas and Biswas (2015) DEFORMED
DEM (SRTM 30 m resolution) near LANDFORMS
Chalsa
(continued)
477
Table 2 (continued)
478

Sl. No. Indices Geometric Calculation References *


30. Riverbed topographic analysis Contouring using total station survey
Micro landform identification (Potholes e.g., Ayaz et al. (2018)
near confluence zone, Fan)
Identification of both potholes and
landforms has been done in field using
measuring tape, staff, GPS and total
station
Software used: Arc GIS 10.3.1,
MATLAB 15
* Resultant analysis from combination of the geomorphic indices
M. Biswas et al.
Tectonics and Channel Morpho-Hydrology—A Quantitative … 479

Fig. 2 Long Profile of Jayanti River a linear curve with a best fit curve with R2 value 0.887
b Exponential curve with R2 value 0.977 c logarithmic curve with R2 value 0.804 d normalized
long profile—power and regression curve with R2 value 0.917

Fig. 3 Langbein curve showing the concavity of Jayanti River

Presence of lineament or fault incites adjustments in its profile causing deviation


in the ideal profile of the river. It is a straight line in a semi-logarithmic scale.
Brookfield (1998) calculated the convex hack profiles of major rivers of south and
south-east Asia that are caused by the tectonic processes during the Cenozoic era
due to collision of Indian and the Eurasian plates. Convex-upward long profiles of
the rivers are observed in areas of general uplift as seen in South Carolina Coastal
plain (Maple and Talwani 1993; Rhea 1989). Whereas concave profiles for the river
basin depict quasi-equilibrium profile as an indicator of disturbance in the region
(Kale et al. 2014). Hack profiles are convex in high uplifted areas and almost straight
or slightly concave in low uplifted areas (Merritts and Vincent 1989). In Central
foothill region of Taiwan, hack profile of the rivers is convex indicating its early
stage adjusting to fault movements. While in the south-western foothills, the profiles
480 M. Biswas et al.

are convex–concave indicating later stage of adjustments to fault movements (Chen


et al. 2003a; Chen 2004). Gradient and concavity is largely included as one of the
attributes to discuss the neotectonic activity of any study area. The Langbein curve of
the Jayanti River in Fig. 3 with a gradient of 0.01 is partly in an equilibrium state. At
a stretch of 3 km from its source, the river is highly aggraded exceeding degradation
rate of sediments.
Stream length gradient index (SL index) of Jayanti River reconfirmed the insta-
bility in relationship among stream power, tectonics and rock resistance. The riverbed
slope of the Jayanti River induced by tectonic anomalies or uplift in relation with rock
resistance, i.e., lithological setup enhances the increased value of the stream power.
It detects the knick points along the river due to tectonic uplift resulting in vertical
incision and valley side cutting having recompilation and alteration of hydrological
parameters. Figure 4 shows the SL index of the river Jayanti. Hypsometric curve
(e.g., Fig. 5) plays an active role to measure the area of the respective watershed
with its elevation. The convex and the concave curve determine the activeness of

Fig. 4 Stream length gradient (SL) index of Jayanti River (Based on SRTM data)

Fig. 5 Hypsometric curve portraying different stages of the river Kankuram (Based on SRTM
data) (Biswas and Biswas 2015)
Tectonics and Channel Morpho-Hydrology—A Quantitative … 481

tectonics, whereas hypsometric integral displays a magnitude within 0–1 that signi-
fies the tectonics. The youthful stage is characterized by rugged relief, deep valley
incision with vertical cutting, and the old basin is sub-denuded with wide valley.
Drainage basin asymmetry factor is an important geomorphic indices commonly
used to determine the tectonic tilt of both small and large basin area irrespective to
the regional and local tilt. The valley floor width–height ratio displays the valley
shape. It is a non-dimensional index, and the value near 0 means V-shaped valley
indicating tectonic activeness and near 1 as U-shaped valley with an attainment of
erosion to its base level. Transverse topographic symmetry factor (T) defines the
erosive surface of a drainage basin in respect to channel cohesiveness. The results
of this exercise along the basin focus on alternative disseminate nature of tilting that
decodes the regional instability with respect to tectonics. Mountain front sinuosity
(Smf) reflects the balance between the erosional forces that tend to cut embayment
into a mountain front and tectonic forces that tend to produce straight mountain front
coincident with an active bounding range. A lower value (<1.25) for Smf indicates
the fact that active tectonics act over the erosional processes of the area portraying
tectonically active fronts (Keller 1986). A higher value (>1.25) of Smf indicates more
dominant erosional process (Bull 2007). Similarly, tilt angle demonstrates the tilting
of a landmass controlled by tectonic activities. The higher level of tectonic activity is
responsible for greater values of tilt angle and vice versa. Figure 6a, b are examples
from the Janauri anticline depicting the mountain front sinuosity (Smf) index from
49 mountain segments.
Considering all the aforesaid indices, sinuosity index is one of the most important
parameters to assess tectonic vicinity of an area. Sinuous character of a channel
and its movement is confined by the terrain character of the drainage basin. In hilly
regions, sinuosity value is generally low as rivers lack opportunity for free moving.
But in the lower parts of the basin characterized by plain region, river moves freely
due to little control of terrain character on the flowing river. As a result, the sinuosity
value of the river channel becomes high comparatively in the upper part of the basin.
Figure 7 depicts a conceptual diagram of the Kankuram River with a sinuosity index
of 1.27. Along with this, there are some correlated geometric attributes as channel
index (CI), valley index (VI), hydraulic sinuosity index (HSI), topographic sinuosity
index (TSI) and standard sinuosity index (SSI).
Proper visualization of channel networking system and explication of their spatial
placement, geometry and connectivity make significant tectonic imprint on basin
physiography on a drainage aligned in a same direction over same slope and topog-
raphy. It interprets the equal basin shape, which is prominent in the Janauri–Chandi-
garh anticlines (Delcaillau et al. 2006). This parallel arrangement of basins also
reanalyzes the front of the propagation thrust. Along the variation of strike line
and the tectonic induced drainage network system, a parallel pattern of drainage
has developed over the Upper Siwalik conglomerate, where dense dendritic drainage
network coincides with the steep flank to the Janauri fault-propagation anticline(e.g.,
Fig. 8 a, b). Here, neotectonic activeness is distinctly furnished by drainage anomalies
as well as pattern analysis like linear gullies along steep walls and radial drainage
pattern signifying the recent uplift. However, somewhere dry valleys portray the
482 M. Biswas et al.

Fig. 6 a Mountain front sinuosity, Smf calculation from DEM b assessment of Smf index from 49
mountain front segments. (Mahmood and Gloaguen 2012)
Tectonics and Channel Morpho-Hydrology—A Quantitative … 483

Fig. 7 Conceptual diagram calculating the sinuosity index showing micro-level features like
meanders and terraces (Biswas and Biswas 2015)

paleo-drainage system being designated as antecedent drainage channels. They were


interrupted by tectonic and lateral propagation of movement. This may induce the
stream to abandon its course where past channels may act as a water gap on the
rising anticline ridge. Figure 8 c portrays a 3D elevation profile of Janauri anticline
extracted from a 30 m DEM.
Trend of tributaries joining another river/stream depicts the presence of linea-
ments or faults. The Subarnarekha basin has been joined by number of small tribu-
taries. Among which, Kharswati originated from Belpahari and Aamjharna, whereas
Kankuram originated from Samardehi and Netrabera meets Subarnarekha at its right
and left bank, respectively, orthogonal at its lower course due to the presence of
the SSZ (Fig. 9a). Geologically, Kankuram originated in the Singbhum granite zone
flowing through the Dhanjhuri-Jadugara belt. It creates a plunge pool and a water-
fall of 12 m at its mouth as it dissects SSZ (Fig. 9b, c). SSZ is the major structural
discontinuity between the Chotanagpur plate and the Singhbhum plate. Formation
of a water fall and a plunge pool at its mouth is rare. It signifies rejuvenation of a
river due to SSZ causing tectonic anomalies. Fractal dimension of drainage pattern
(FD) is considered as another parameter to focus on tectonic vicinity where slope
and length of drainage lines are taken to calculate. The best fit log plot deciphers the
tectonic instability.

5.2 Based on Primary Survey and Data Collection Using


Different Instruments

Applying the secondary database on the study sites discussed above, field investiga-
tion was conducted. As the geomorphic and tectonic parameters disclose the tectonic
484 M. Biswas et al.

sensitivity using secondary data, the detail study and data collection for hydrological
parameters of a river, landform deformation, riverbed micro landform analysis and
cross profile analysis were done during the field survey.
Spatial pattern of hydrological parameters of a river displays the key nodal points
that express the anomalies resulting channel width minimization, increase of slope

Fig. 8 Janauri anticline and its drainage patterns. a i, ii, iii, iv and v show different drainage network
pattern b the geological cross sections and longitudinal profiles of two rivers (a and h) showing
channel diversion. (Delcaillau et al. 2006). c A 3D elevation profile across Janauri anticline extracted
from 30 m DEM using Surfer 15 platform
Tectonics and Channel Morpho-Hydrology—A Quantitative … 485

Fig. 8 (continued)

and vertical depth. Such changes along the river channel ensure the spatial vari-
ability of hydrological parameters as stream power (e.g., Fig. 10a), shear stress (e.g.,
Fig. 10b), velocity(e.g., Fig. 10c) and discharge (e.g., Fig. 10d). The calculated data
set is based on field survey of Lish River using some selected primary instruments
such as digital current meter, measuring tape, staff, handheld GPS, clinometer and
distance meter. The considered Lish River shows an interesting change of hydro-
logical parameters (Fig. 10a, b, c and d) as it focuses on channel slope and channel
geometric anomalies, indirectly indicate tectonic control. It may be entitled as fault
zone or presence of contact zone between Siwalik and Lesser Himalaya. Use of
total station survey of a riverbed terrain (e.g., Fig. 11) incorporates all the minor
changes of elevation that express riverbed topography. Digital elevation model of
any region at a small scale can be obtained using total station. It helps in identi-
fying the micro landforms and stream pattern. For example, Fig. 12a portrays the
braided nature of the Jayanti River at a small scale on a DEM. It depicts the exis-
tence of micro landforms like alluvial fan (Kalikhola at its confluence with Jayanti)
(Fig. 12b) and potholes (Fig. 13a) near Tetuldanga (300 m upstream from its conflu-
ence with Subarnarekha). The presence of SSZ rejuvenated the river and interrupted
the riverbed slope increasing the velocity, discharge, shear stress and stream power.
Thus, this results in increasing riverbed load size and corrosion generating huge
486 M. Biswas et al.

Fig. 9 a Conceptual diagram showing trend of Kankuram tributary joining river Subarnarekha at
orthogonal depicts the presence of lineaments or faults b formation of a small-scale water fall (12 m)
and a plunge pool as it dissects the SSZ at its confluence with Subarnarekha at its right bank c the
long profile of the river Kankuram depicting a knick point at its confluence with Subarnarekha,
200 m SW corner from Moubhandar (Based on field survey)

potholes of different shape and size. Figure 13b is a photographic evidence of the
potholes having different shapes collected during field survey. Cross section profile
of a river can be conducted in the field using total station. It provides nearly accurate
elevation data mentioning its latitudinal and longitudinal position (Fig. 14a, b).
However, through this discussion, it has been disclosed that the primary and
secondary data enhancement and representation is a combined morpho-hydro-
tectonic approach to enrich the study of tectonic vicinity of an area. The accumulated
tabulated data can be evaluated. The statistical techniques like analytical hierarchy
process (AHP) in Satty scale may be applied to give weightage of each factor that
indicate the tectonic instability followed by Technique for Order Preference by Simi-
larity to the Ideal Solution (TOPSIS) (Fig. 15). It is suitable to rank the responsible
factors in a systematic and scientific method (Hwang and Yoon 1981; Yoon in 1987;
Hwang et al. 1993).

6 Conclusions

Concerned morpho-tectonic indices like basin asymmetry, long profile analysis and
hydrological parameters have distinctly quantified to ensure the paradox of tectonics
Tectonics and Channel Morpho-Hydrology—A Quantitative … 487

Fig. 10 a Stream power and b shear stress map of River Lish indicating morpho-tectonic alteration
along the channel (Based on field survey data). c variation of flow velocity and d Discharge map
along the channel of River Lish following the terrain slope and alteration of channel geometry
(Based on field survey data)
488 M. Biswas et al.

Fig. 11 Riverbed terrain map of the Jayanti River using total station survey data (collected from
field survey)

Fig. 12 DEM of Jayanti River a showing braiding channel pattern in October 2017 b Steep slope
alluvial fan of its left side tributary, Kali khola (based on total station survey data) prepared in
MATLAB 15 platform

in different spatial scales. Such tectonic activeness has also been evident through the
representation of minor-scale landforms, which are well overlain with the tectonic
setup like presence of SSZ in Singbhum Craton area and alternation of stream power,
shear stress, velocity and discharge near MBT along River Lish is progressively
evident as active tectonics. Based on both primary and secondary data set, statis-
tical analysis and GIS application have been considered to reframe the tectonic
significance on channel morphology and minor-scale landform development. Such
controlling indicators may be re-evaluated through applying AHP and TOPSIS like
statistical techniques to rank them according to their weightage alteration.
Tectonics and Channel Morpho-Hydrology—A Quantitative … 489

Fig. 13 a Terrain of the Kharshoti riverbed portraying potholes, prepared from the field survey
using total station and Surfer 15 platform b field photographs of circular, oval and elongated pothole
shapes (classified during field)

Fig. 14 a Cross profile of Kankuram River at different survey points, Ghatshila (field survey using
total station) b Field photograph of Kankuram riverbed (Biswas and Biswas 2015)
490 M. Biswas et al.

Fig. 15 Brief methodology of TOPSIS in a form of flowchart

Acknowledgements The authors thank the Department of Geography, Presidency University,


Kolkata. We extend our sincere thanks to Dr.Soumyajit Mukherjee (IIT Bombay) for providing
multiple rounds of review. We would also like to render our thanks to Edison David, for helping us
during fieldwork. Thanks to Marion Schneider, Annett Buettener, Boopalan Renu, Alexis Vizcaino,
Doerthe Mennecke-Buehler and the proofreading team (Springer). Dutta and Mukherjee (2021)
encapsulate this work.

References

Andreani, L., Stanek, K., Gloaguen, R., Krentz, O., & Domínguez-González, L. (2014). DEM-based
analysis of interactions between tectonics and landscapes in the ore mountains and eger rift (East
Germany and NW Czech Republic). Remote Sensing, 6, 7971–8001. https://doi.org/10.3390/rs6
097971.
Ayaz, S., Biswas, M., & Dhali, K. (2018). Morphotectonic analysis of alluvial fan dynamics:
Comparative study in spatio-temporal scale of Himalayan foothill, India. Arabian Journal of
Geosciences, 11(41), 1–16. https://doi.org/10.1007/s12517-017-3308-2.
Bisaria, B. K. (1980). Report on geomorphological mapping of a part of the foothills of Darjeeling
Himalayas. West Bengal, Geological survey of India report, 1–16.
Biswas, M., & Dhara, P. (2019). Correction to: Evolutionary characteristics of meander cut-off—
A hydro-morphological study of the Jalangi River, West Bengal, India. Arabian Journal of
Geosciences, 12, 739. https://doi.org/10.1007/s12517-019-4971-2.
Biswas, M., & Banerjee, P., (2018). Bridge construction and river channel morphology—A compre-
hensive study of flow behavior and sediment size alteration of the River Chel, India. Arabian
Journal of Geosciences, 11. https://doi.org/10.1007/s12517-018-3789-7.
Biswas, M. (2015). Impact of neotectonism in the discussion of geomorphological processes as a
feedback system: North Bengal Foothills, West Bengal. GSTF Journal of Geological Sciences,
2. https://doi.org/10.5176/2335-6774_2.1.22.
Tectonics and Channel Morpho-Hydrology—A Quantitative … 491

Biswas, M., & Biswas, A. (2015). GIS based semi-quantitative morphological analysis of Kankuram
Basin, Ghatsila. International Research Journal of Natural and Applied Sciences, 2(3), 79–114.
Brice, J. C. (1964). Channel patterns and terraces of the Loup Rivers in Nebraska. Geological
Survey Professional Paper 422-D, Washington, pp. D2–D41.
Brookfield, M. E. (1998). The evolution of the great river systems of southern Asia during the
Cenozoic India-Asia collision: Rivers draining southwards. Geomorphology, 22(3–4), 285–312.
https://doi.org/10.1016/S0169-555X(97)00082-2.
Brzezińska-Wójcik, T., & Gawrysiak, L. (2010). Neotectonic mobility of the Roztocze region,
Ukrainian part, Central Europe: Insights from morphometric studies. Annales Societatis Geolo-
gorum Poloniae, 80.
Bull, W. B. (2007). Tectonic geomorphology of mountains: A new approach to paleoseismology
(p. 328). Oxford: Wiley.
Bull, W. B., & McFadden, L. D. (1977). Tectonic geomorphology north and south of the Garlock
fault, California. Geomorphology in Arid Regions. In D. O. Doehring (Ed.), Proceedings of the
Eight Annual Geomorphology Symposium (pp. 115–138). Binghamton, NY: State University of
New York at Binghamton.
Chakrabarti, C., Mukhopadhyay, D., & Poddar, B. C. (2012). Geomorphology in relation to
tectonics: A case study from the eastern Himalayan foothills of West Bengal, India. Quaternary
International, 298, 80–92. https://doi.org/10.1016/j.quaint.2012.12.020.
Chen, Y., Sung, Q., Chen, C., & Jean, J. (2006). Variations in tectonic activities of the central and
southwestern foothills, Taiwan, Inferred from River Hack Profiles. TAO: Terrestrial, Atmospheric
and Oceanic Sciences, 17(3), 563–578.
Chen, Y. C. (2004). Morphotectonic features of Taiwan mountain belt based on hypsometric integral,
topographic fractals and SL index. (Ph.D. Thesis) Institute of Earth Sciences, National Cheng
Kung University p. 129.
Chen, Y. C., Sung, Q. C., & Cheng, K. Y. (2003). Along-strike variations of morphotectonic
features in the western Foothills of Taiwan: Tectonic implications based on stream gradient
and hypsometric analysis. Geomorphology, 56, 109–137.
Charlton, Ro. (2008). Fundamentals of Fluvial Geography. Routledge, London and New York:
Taylor and Francis Group.
Chorley, R. (1971). The drainage basin as the fundamental geomorphic unit. In R. Chorley (Ed.),
Introduction to fluvial processes (pp. 30–32). London: Methuen and Co. Ltd.
Claudio, B. (2013). Geology and geomorphology. The soil of Italy. pp. 39–56. https://doi.org/10.
1007/978-94-007-5642-7_3.
Cox, R. T. (1994). Analysis of drainage-basin symmetry as a rapid technique to identify areas of
possible quaternary tilt-block tectonics: an example from the Mississippi Embayment. Geological
Society of America Bulletin, 106, 571–581.
Delcaillau, B., Carozza, J. M., & Laville, E. (2006). Recent fold growth and drainage development:
The Janauri and Chandigarh anticlines in the Siwalik foothills, northwest India. Geomorphology,
76(3–4), 241–256. https://doi.org/10.1016/j.geomorph.2005.11.005.
Dutta, D, & Mukherjee, S. (2021). Introduction to Structural Geology and Tectonics Field Guide-
book—Volume 1. In S. Mukherjee (Ed.), Structural Geology and Tectonics Field Guidebook—
Volume 1. Switzerland AG: Springer. Cham. pp. xi-xvi. ISBN: 978-3-030-60142-3.
Ghosh, P. K., & Ghosh, P. C. (1980). Report on the preliminary investigation of dolomite in Jainti
Area, Jalpaiguri District, West Bengal (Progress report for The Field Secession). Geological
Survey of India.
Ghosh, P. K. (1968). Investigation of dolomite in the Mahakal Hill Area, near Jainti, Jalpaiguri
District, West Bengal (Progress report for The Field Secession). Geological Survey of India.
Gloaguen, R., Marpu, P. R., & Niemeyer, I. (2007). Automatic extraction of faults and fractal
analysis from remote sensing data. Nonlinear Processes in Geophysics, 14, 131–138.
Goswami, C., Mukhopadhyay, D., & Poddar, B. C. (2012). Tectonic control on the drainage system
in a piedmont region in tectonically active eastern Himalayas. Frontiers of Earth Science, 6(1),
29–38. https://doi.org/10.1007/s11707-012-0297-z.
492 M. Biswas et al.

Guba, I., & Glennie, K. (1998). Geology and geomorphology. In S. A. Ghazanfar, M. Fisher (Eds.),
Vegetation of the Arabian Peninsula (Vol. 25, pp. 39–62). Dordrecht. Geobotany: Springer. https://
doi.org/10.1007/978-94-017-3637-4_3.
Gupta, A., & Basu, A. (1985). Structural evolution of Precambrians in parts of North Singhbhum,
Bihar. The Geological Survey of India, 113(3), 13–24.
Gupta, A., Basu, A., & Ghosh, P. K. (1980). The Proterozoic ultramafic and mafic lavas and tuffs
of the Dalma greenstone belt, Singhbum, eastern India. Cnadian Journal of Earth Sciences, 17,
210–231. https://doi.org/10.1139/e80-017.
Hack, J. T. (1973). Stream profile analysis and stream gradient index. Journal of Research of the
United States Geological Survey, 1(4), 421–429.
Hack, J. (1957). Studies of longitudinal stream profiles in Virginia and Maryland. (Geological
Survey Professional Paper) 294–B, 45–95.
Hamdouni, R. E., Irigaray, C., Fernandez, T., Chacón, J., & Keller, E. A. (2008). Assessment of
relative active tectonics, southwest border of Sierra Nevada (Southern Spain). Geomorphology
96, 150–173.
Hare, P. H., & Gardner, T. W. (1985). Geomorphic indicators of vertical neotectonism along
converging plate margins, Nicoya Peninsula, Costa Rica. In M. Morisawa & J. T. Hack (Eds.),
Tectonic geomorphology (pp. 75–104). Boston: Allen and Unwin.
Harmar, O. P., & Clifford, N. J. (2007). Geomorphological explanation of the long profile of the
Lower Mississippi River. Geomorphology, 84, 222–240.
Hwang, C. L., Lai, Y.-J., & Liu, T. Y. (1993). A new approach for multiple objective decision
making. Computers and Operations Research, 20, 889–899. https://doi.org/10.1016/0305-054
8(93)90109-V.
Hwang, C. L., & Yoon, K. (1981). Multiple attribute decision making: Methods and applications.
New York: Springer. http://dx.doi.org/10.1007/978-3-642-48318-9.
Javed, N. Malik, C. Mohanty. (2007). Active tectonic influence on the evolution of drainage and
landscape: Geomorphic signatures from frontal and hinterland areas along the Northwestern
Himalaya, India. Journal of Asian Earth Sciences 29, (5-6):604–618
Kale, V. S., Sengupta, S., Achyuthanc, H., & Jaiswald, K. M. (2014). Tectonic controls upon Kaveri
River drainage, cratonic Peninsular India: Inferences from longitudinal profiles, morphotectonic
indices, hanging valleys and fluvial records. Geomorphology, 227(15), 153–165.
Keller, E. A., & Pinter, N. (2002). Active tectonics: Earthquakes, uplift, and landscape (2nd ed.,
pp. 1–362). New Jersey: Prentice Hall.
Keller, E. A. (1986). Investigation of active tectonics: use of surficial earth processes. In R. E.
Wallace (Ed.), Active tectonics, studies in geophysics (pp. 136–147). Washington, DC: National
Academy Press.
Kirby, E., & Whipple, K. (2001). Quantifying differential rock-uplift rates via stream profile
analysis. Geology, 29(5), 415–418.
Lee, C., & Tsai, L. L. (2009). A quantitative analysis for geomorphic indices of longitudinal river
profile: A case study of the Choushui River, Central Taiwan. Environmental Earth Sciences,
1549–1558.
Magar, P. P., & Magar, N. P. (2016). Application of Hack’s stream gradient index (SL Index) to longi-
tudinal profiles of the rivers flowing across Satpura-Purna Plain, Western Vidarbha, Maharashtra.
Journal of Geomorphology, 4, 65–72.
Mahmood, S. A., & Gloaguen, R. (2012). Appraisal of active tectonics in Hindu Kush: Insights from
DEM derived geomorphic indices and drainage analysis. Geoscience Frontiers, 3(4), 407–428.
Mahmood, S. A., & Gloaguen, R. (2011). Fractal measures of drainage network to investigate
surface deformation from remote sensing data: A paradigm from HinduKush (NE-Afghanistan).
Journal of Mountain Science, 8, 641–654.
Malik, J. N., Shah, A. A., Sahoo, A. K., Puhan, B., Banerjee, C., Shinde, D. P., et al. (2010). Active
fault, fault growth and segment linkage along the Janauri anticline (frontal foreland fold), NW
Himalaya, India. Tectonophysics, 483(3), 327–343. https://doi.org/10.1016/j.tecto.2009.10.028.
Tectonics and Channel Morpho-Hydrology—A Quantitative … 493

Maple, R. T., & Talwani, P. (1993). Evidence of possible tectonic upwarping along the South Carolina
coastal plain from an examination of river morphology and elevation data. Geology, 21(7), 651–
654. https://doi.org/10.1130/0091-7613(1993)021%3C0651:EOPTUA%3E2.3.CO;2.
Merritts, D., & Vincent, K. R. (1989). Geomorphic response of coastal streams to low, intermediate,
and high rates of uplift, Mendocino Triple Junction region, Northern California. Geological
Society of America Bulletin, 110, 1373–1388.
Miall, A. D. (1977). A review of the braided river depositional environment. Earth Science Review,
13, 1–62.
Mitra, S. K. (1985). Investigation of dolomite in the BajeKhola block and other areas around Jainti,
Jalpaiguri District, West Bengal (Progress report for The Field Secession 1967–68). Geological
Survey of India.
Mitra, S. K. (1983). Report on reconnaissance survey of the Jainti Dolomite Belt, Jalpaiguri District,
West Bengal (Progress report for The Field Secession). Geological Survey of India.
Mukherjee, S., Carosi, R., van der Beek, P. A., Mukherjee, B. K., Robinson, D. M. (2015). Tectonics
of the Himalaya: An introduction. In S. Mukherjee, R. Carosi, P. van der Beek, B. K. Mukherjee, D.
Robinson (Eds.), Geological society (Vol. 412, pp. 1–3). London: London, Special Publications.
Mukherjee, S. (2013). Higher Himalaya in the Bhagirathi section (NW Himalaya, India): its struc-
tures, backthrusts and extrusion mechanism by both channel flow and critical taper mechanisms.
International Journal of Earth Sciences, 102, 1851–1870.
Mukherjee, S., Mukherjee, B., & Thiede, R. (2013). Geosciences of the Himalaya-Karakoram-Tibet
Orogen. International Journal of Earth Sciences, 102, 1757–1758.
Mukherjee, S. (2012). Tectonic implications and morphology of trapezoidal mica grains from the
Sutlej section of the Higher Himalayan Shear Zone, Indian Himalaya. The Journal of Geology,
120, 575–590.
Mukherjee, S. (2015). A review on out-of-sequence deformation in the Himalaya. In S. Mukherjee,
R. Carosi, P. van der Beek, B.K. Mukherjee, D. Robinson (Eds.), Tectonics of the Himalaya.
Geological society (Vol. 412, pp. 67–109), London: Special Publications. https://doi.org/10.1144/
sp412.13.
Mullick, M., & Mukhopadhyay, D. (2011). An analysis of GPS-derived velocities in the Bengal
basin and the neighbouring active deformation zones. Current Science, 101, 423–426.
Nakata, T. (1989). Active faults of the Himalaya of India and Nepal. Geological Society of America
Special Paper, 232, 243–264. https://doi.org/10.1130/SPE232-243.
Noss, C., & Larke, A. (2016). Roughness, resistance, and dispersion: Relationships in small streams.
Water Resources Research, AGU, 52(4), 2802–2821.
Paul, A., & Biswas, M. (2019). Changes in river bed terrain and its impact on flood propagation – a
case study of River Jayanti, West Bengal, India. Geomatics, Natural Hazards and Risk, 10(1),
1928–1947. https://doi.org/10.1080/19475705.2019.1650124.
Pike, R. J., & Wilson, S. E. (1971). Elevation-relief ratio, hypsometric integral and geomorphic
area—Altitude analysis. Geological Society of America Bulletin, 82, 1079–1084. http://dx.doi.
org/10.1130/0016-606(1971)82[1079:ERHIAG]2.0.CO;2.
Pinter, N., & Keller, E. A. (1995). Geomorphological analysis of neotectonic deformation, northern
Owens Valley, California. GeolRundsch, 84, 200–212.
Raina, V. K., & Saha, S. S. (1970). Investigation of the dolomite deposits in the hathipotha area of
Jainti, district, Jalpaiguri, West Bengal. Geological Survey of India. UE, 5831, 1–24.
Ramírez-Herrera, M. T. (1998). Geomorphic assessment of active tectonics in the Acambay Graben,
Mexican volcanic belt. Earth Surface Processes and Landforms, 23, 317–332.
Ray, K. K. (1990). The dalmavolcanics—A Precambrian analogue of the Mesozoic-Cenozoic suture.
Group discussion on suture zones, young and old. Geological Survey of India, 17–21.
Rhea, S. (1989). Evidence of uplift near Charleston, South Carolina. Geology, 17(4), 311–315.
https://doi.org/10.1130/0091-7613(1989)017%3C0311:EOUNCS%3E2.3.CO;2.
Saha, A. K. (1994). Crustal evolution of Singhbhum—North Orissa, Eastern India. Memoirs of the
Geological Survey of India, 27, 341.
494 M. Biswas et al.

Sajadian, M., Pourkermani, M., Qorashi, M., & Moghaddas, N. H. (2015). The analysis of transverse
topographic symmetry factor (T Index) in the Chekene-Mazavand, North East Iran. Open Journal
of Geology, 05, 809–820. https://doi.org/10.4236/ojg.2015.511069.
Sarkar, S., & Saha, A. (1963). On the occurrence of two intersecting Pre-Cambrian orogenic belts
in Singhbhum and adjacent areas. Geological Magazine, 100(1), 69–92. https://doi.org/10.1017/
S0016756800055060.
Schumm, S. A., & Khan, H. R. (1972). Experimental study of channel patterns. Geological Society
of America Bulletin, 83(1), 755–770.
Schumm, S. A. (1956). The evolution of drainage systems and slopes in bad lands at Perth, Amboi,
New Jersey. Geological Society of America Bulletin, 67(5), 597–646.
Silva, P. G., Goy, J. L., Zazo, C., & Azcárate, T. (2003). Faulth-generated mountain fronts in
southeast Spain: Geomorphologic assessment of tectonic and seismic activity. Geomorphology,
50, 203–225. https://doi.org/10.1016/S0169-555X(02)00215-5.
Singh, I. B., Srivastava, P., Sharma, S., et al. (1999). Upland interfluve (Doab) deposition: Alternative
model to muddy over bank deposits. Facies, 40, 197. https://doi.org/10.1007/BF02537474.
Sinha, S. K., & Parker, G. (1996). Causes of concavity in longitudinal profiles of rivers. Water
Resources Research, 32, 1417–1428.
Snyder, N., Whipple, K., Tucker, G., & Merritts, D. (2000). Landscape response to tectonic forcing:
DEM analysis of stream profiles in the Mendocino triple junction region, northern California.
Geological Society of America Bulletin, 112(8), 1250–1263.
Sonam, & Jain, V. (2017). Geomorphic effectiveness of a long profile shape and the role of
inherent geological controls in the Himalayan hinterland area of the Ganga River basin, India.
Geomorphology, 1–51. https://doi.org/10.1016/j.geomorph.2017.12.022.
Strahler, A. N. (1952). Hypsometric (area-altitude) analysis of erosional topography. Geological
Society of America Bulletin, 63, 1117–1142.
Takieh, E., Ghorashi, M., & Rezaie, F. (2015). The Transverse Topographic Symmetry Factor of
Darakeh Stream in the North Tehran, Iran. Open Journal of Geology, 5, 770–779. https://doi.org/
10.4236/ojg.2015.511066.
Tandon, S. K., Sinha, R., Gibling, M. R., Dasgupta, A. S., & Ghazanfari, P. (2008). Late Quater-
nary evolution of the Ganga Plains: myths and misconceptions, recent developments and future
directions. Golden Jubilee Memoir of the Geological Society of India, 66, 259–299.
Thakur, V., Jayangondaperumal, R., & Suresh, N. (2009). Late Quaternary-Holocene fold and
landform generated by morphogenic earthquakes in Chandigrh anticlinal ridge in Punjab Sub
Himalayas. Himalayan Geology, 20(2), 103–113.
Vanlaningham, S., Meigs, A., & Goldfinger, C. (2006). The effects of rock uplift and rock resistance
on river morphology in a subduction zone forearc. Earth Surface Processes and Landforms, 31,
1257–1279.
Whipple, K., Wobus, C., Crosby, B., Kirby, E., & Sheenan, D. (2007). New tools for quantitative
geomorphology: Extracting and interpretation of stream profiles from digital topographic data.
Boulder, USA: GSA Annual Meeting.
Wobus, C., Whipple, K. X., Kirby, E., Snyder, N., Johnson, J., Spyropolou, K., … & Sheehan,
D. (2006). (2006). Tectonics from topography: procedures, promise, and pitfalls. In: Tectonics,
climate, and landscape evolution; Special Paper Geological Society of America, 398, 55–74.
Yoon, K. (1987). Reconciliation among discrete compromise situations. Journal of Operational
Research Society, 38, 277–286. https://doi.org/10.1057/jors.1987.44.
Geological Field Guide: Malvan
(Maharashtra, India)

Ashwin Pundalik, Shiba Nikalje, Arnav Samant, and Hrishikesh Samant

Abstract We present a geological field guide for a lesser known location to (Indian)
geologists from Maharashtra (India): Malvan. Geomorphological and structural
geology exercises have been explained, and a few results have been shown from
the 2016 field report of the third author.

1 Introduction

Fieldwork has been conducted by faculty members of the Department of Geology at


St. Xavier’s College (Mumbai) in Malvan for the last 25 years. Yet this exciting field
location is much less known among geologists in India and abroad. One of the reasons
is being that there have been no major research publications on this area despite there
being a very good scope of undertaking structural geology and geomorphological
studies. Various types of geological training can be conducted with students in and
around this area. The place is free from political disturbance and is relatively less
known to tourists. Through this chapter, we publish to the “world” an exciting field
location and encourage researchers and academicians to explore it. One can watch
the YouTube video (https://www.youtube.com/watch?v=Fpd84rADuC8) to have a
more vivid look.
Third Year B.Sc. fieldwork is conducted over a period of 12 days out of which
10 days are allotted for the actual fieldwork and two days for travelling between
Mumbai and Malvan by train. Some M.Sc. part 2 students also go to Malvan if they
choose this area for their dissertation. The fieldwork is mandatory and is spread across
fifth and sixth semester. In fifth semester, students prepare a pre-field report, which
is evaluated out of 100 marks equivalent to 4 credits (spread across all courses). In
the sixth semester, students are evaluated for actual performance during field work
and the field report that they generate later on.
Malvan is situated on the Indian west coast within the Konkan coastal belt and
comes under Sindhudurg district of Maharashtra. The extent of Malvan taluka may be

A. Pundalik · S. Nikalje · A. Samant · H. Samant (B)


Department of Geology, St. Xavier’s College, Mumbai, Maharashtra 400001, India
e-mail: hrishikesh.samant@xaviers.edu

© Springer Nature Switzerland AG 2021 495


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_17
496 A. Pundalik et al.

approximately delineated with Gad River in the north, Kolamb and Tarkarli Beach in
the west, Karli River in the south, and the Konkan railway track in the East (Fig. 1b).
The area has a considerable exposure of continental flood basalts. Vigorous rainfall
has created extensive lateritic (soil) horizons; metamorphic and igneous rocks also
occur locally.
Malvan is connected to adjacent talukas and towns like Kankavli, Ramgad,
Devbag, and Kudal by road. There is no direct rail connectivity for Malvan. By
road, while travelling from Mumbai, take Mumbai–Goa highway (NH-17) and exit
toward Malvan to the right at Kasal. By rail, board the train going toward Goa and get
off at Kudal station, then take a public transport (state transport bus/rickshaw) from
Kudal to Malvan. It takes around 15 hrs to reach Malvan from St. Xavier’s College,
Mumbai. The distance between Kudal and Malvan by road is 30 km. One can also

Fig. 1 a Location of study area. b Local drainage pattern, Malvan and the surrounding areas
Geological Field Guide: Malvan (Maharashtra, India) 497

travel by air, by taking a flight to the newly constructed airport at Chipi, Vengurla.
The distance between Chipi Airport and Malvan by road is 21 km.
St. Xavier’s College students and faculty members stay at Om Shraddha Lodge,
near Chivala beach—this has been a tradition for the past 23 years. Om Shraddha
Lodge is a clean, safe, and reasonably priced accommodation. There are a number
of other hotels and resorts in the area, viz Hotel Chivala Beach and Sea Island Beach
Resort, with dynamic pricing. The locals are cordial and welcoming, and some also
provide home stay facilities for tourists. Malvani cuisine boasts a wide variety of
vegetarian and nonvegetarian foods, especially seafood. Food is readily available in
all local hotels.
Topography of the area under study is relatively rugged, it does house several
detached residual hills with the lowest elevation being MSL and the highest being
≤150–230 m (Gupta and Rai 1987). The hills are outliers that are underlain by basalts.
The first- and second-order river valleys are not very deep, although prominent in
field. Several sand dunes are present along the coast. The area is characterized by
dendritic to sub-dendritic drainage-pattern reflecting thereby impervious nature of
the formations. At places, Karli River shows braided geometry in proximity to the
coast, which may be due to fall in current velocity and increase in sediment load. Just
south of Ramgad, small bar deposits occur on the right bank of Gad River. Drainage
system of major rivers in the district is mostly sub-parallel while their tributaries
have a rectangular drainage pattern (Mishra 2014).
About half the area of Sindhudurg district has a hilly terrain. The district has three
major physiographic divisions: (i) the eastern strip close to Western Ghats highly
dissected with deep valleys; (ii) the middle strip of the district occupied by flat-topped
hills having elevations up to 300 m above the MSL covered by laterite; and (iii) the
coastal plain in the western part with elevation 100–150 m above the MSL (Fig. 2).
The physiographic features have given rise to five characteristic landforms viz (i)

Fig. 2 Local physiography, Malvan and the surrounding areas


498 A. Pundalik et al.

the coast line; (ii) the estuarine alluvial plains; (iii) the lateritic plateau; (iv) highly
eroded remnant hills; and (v) scarp faces of Sahyadri hill ranges (Fig. 2).

2 Geomorphologic Exercises

2.1 Beach Profiling

A beach profile survey is a topographic/bathymetric survey of a beach carried out


perpendicular to the shoreline at various points along the shore. Profiling is repeated
multiple times at the same locations across long time periods. It is usually carried
out during low tide so as to get the longest possible profile.
A beach profile survey serves as a way of cataloging effects on topography of
beaches by long periods of heavy rainfall, floods, strong wave action, construction
of manmade structures, etc. It serves as a means to determine total sand budget of
the beach over a considerable amount of time. Preservation of beaches is important
as it plays a major role in protecting shorelines from strong wave action and thus
preventing the sea from encroaching landward.
Equipments required to carry out beach profiling are (i) measuring tape, (ii)
clinometer/Brunton compass, (iii) pole/staff (taller than the Brunton bearer), and
(iv) marker/bright colored ribbon. The procedure for beach profiling is as follows:
• The pole is held vertical, right against the Brunton bearer’s eye, and a mark is
made with a marker or a ribbon tied on the pole at the same height as the person’s
eyes distinct enough to be spotted from a distance.
• The survey begins with the Brunton bearer who logs readings of gradient also
carrying the loose end of the tape, with the other end held up by the pole bearer
who logs the distance.
• The Brunton bearer initially stands at the low tide line while the pole is held at the
point where gradient changes from that on which the Brunton bearer is standing.
• The Brunton bearer at the low tide line sights the mark on the pole using the
Brunton at eye level and adjusts the clinometers level. The gradient is directly
read in degree and logged. At the same time, the Brunton bearer holds up the
tape’s loose end at his eye level while standing in the same stance as he was
while the pole was marked. The pole bearer notes down the tape reading which
corresponds to the length of patch of coast with a particular gradient (Fig. 3)
• The entire process is repeated after the Brunton bearer stands on the spot that the
pole was placed previously, and the pole is shifted further landward to the very
next point where the slope changes.
• When pole bearer reaches the end of the beach on the landward side and readings
are taken, one set of beach profile is completed. Multiple sets are recorded in a
similar fashion over a decided stretch of the shore line (Table 1).
• A graph of gradient versus distance along profile is plotted (Fig. 4).
Geological Field Guide: Malvan (Maharashtra, India) 499

Fig. 3 Positions of the Brunton bearer and pole bearer during beach profiling exercise

Table 1 Beach profiling readings


SET 1 SET 2 SET 3
Gradient (in Length (m) Gradient (in Length (m) Gradient (in Length (m)
degrees) degrees) degrees)
14 11.10 9 9.64 2 4.55
4 6.50 7 3.83 5 9.30
9 1.35 12 2.40 9 3.32
-4 10.40 −1 9.50 −1 6.27
5 4.45 2 6.30 3 8.2
Sets 1, 2, and 3 consist of readings taken on parallel points on the beach ~50 m apart. A negative
gradient indicates depression which could be caused due to removal of sand on the beach (this is
a probable site for heavy mineral deposition), whereas positive gradient indicates that the land is
elevated

3 Understanding Coastal Geomorphology

Coastal landforms are formed primarily due to wave action and salt
wedging (Davidson-Arnott, 2009). The shore-features observed in Malvan are domi-
nantly created by waves. The following features are observed at the locality in
Malvan called ‘Sarjekot’ (16° 05 12 N 73° 27 48 E), in a 1.5km traverse along
the coastline.
500 A. Pundalik et al.

Fig. 4 Three sets of beach profiles

3.1 Sea Cave

A sea cave is a hollow excavated by waves in a zone of weakness on a cliff face


(Fig. 5). A mature cave is deeper than its entrance width. Sea caves are likely to form
in places of geological weakness such as bed contacts, bedding planes, joints, and
Geological Field Guide: Malvan (Maharashtra, India) 501

Fig. 5 A sea cave and headland

fault planes (Hugget 2007). Once a tiny part of the cliff face is wedged out, the rate
of erosion becomes higher (Fig. 5).

3.2 Headland

A headland is a cliff of resistant rock strata projecting into the sea deeper than the
surrounding coastline, with a tall scarp face and a rocky beach at its base formed
primarily by erosion due to wave action. Such a structure is seen at Sarjekot (Fig. 5).
The cliff is made up of secondary laterite.

3.3 Sea Corridor and Sea Stack

When a sea cave starts forming at two places in close proximity, the caves tend to
meet at a certain distance inside the rock mass. The depth at which these caves meet
is proportional to thickness of rock above the cave. When the caves meet, it forms a
semi-circular hollow called a ‘sea corridor’ (Fig. 6).
The corridor widens temporally leaving a narrow bridge like structure connecting
the main rock mass to the one across the corridor. This narrow connection is called
the ‘sea arch’. A sea arch becomes slimmer at a considerably fast rate and finally
502 A. Pundalik et al.

Fig. 6 Sea corridor and sea


stack at Sarjekot

collapses leaving a rock mass detached from the main cliff called a ‘sea stack’. It
looks like an obelisk with a broad base and narrowed top (Fig. 6). These two structures
discussed above form when height of cliff face is low, roughly, few tens of meters
above the high waterline (Hugget 2007).

3.4 Pothole

Potholes are restricted to fluvial environments. They form due to swirling motion
of stones and pebbles within depressions. The abrasive material is swirled due to
flowing water in the river which enters the depression and induces circular motion
to material within the depression. Over time, the depression deepens and forms a
cylindrical hole that can go several tens of meters deep.
A large number of potholes were observed near Tarkarli in a dried river channel
of the tributary of river Karli (Figs. 7 and 8) near a village named Devli (16° 00 59
N 73° 29 52 E). The tributary follows a NNW–SSE trend along a fault plane.
Geological Field Guide: Malvan (Maharashtra, India) 503

Fig. 7 A pothole north of Tarkarli, length of hammer marker ~25 cm, the pothole has a lenticular
geometry owing to a high aspect ratio

Fig. 8 A giant pothole; students standing in the pothole for scale. Shiba Nikalje at right, waist to
head height ~90 cm

3.5 Beach Barrier—Spit

Spit forms by the combined action of river and sea. When a river meets the sea,
its velocity is countered by the action of waves in the opposite direction from the
504 A. Pundalik et al.

Fig. 9 Satellite imagery of Kolamb Creek’s double spit

sea. This leads to a sudden drop in velocity of water entering the sea and abrupt
deposition of sand, silt, and sometimes clay. In case of the spit at ‘Talashil’ in Malvan,
as deposition progressed, the pile of sediments has been reworked and reshaped by
longshore current (Samant 1989). The final shape of the spit is a linear, elongated
landmass running parallel to coast separated by the river also running parallel to the
coast. This particular type of situation is called a ‘bar-built estuary’.

3.6 Beach Barrier—Paired Spit

A situation similar to the one observed at Talashil (Malvan) is also observed at the
Kolamb Creek (Malvan). Also, the point where the river joins the sea is blocked
partially by two spits projecting from either side of the river bank. Such spits are
known as ‘paired spit’ (Fig. 9). The bay is filled with water throughout the year, even
during low tides.
Paired spits form commonly in bays where frontal waves generate convergent
longshore drifts from the headland on both sides of the bay. Thus, the spit grows
evenly from either side of river bank rather than biased on a single side as in the case
of Talashil. The influence of long-shore currents on Kolamb Creek’s barrier system
is absent as they are obstructed by the headland at Sarjekot (Fig. 9).

3.7 Tombolo

A tombolo connects Sindhudurg Fort to the mainland (Fig. 10). Tombolo forms
due to splitting and deflection of waves around an island of considerable size in
Geological Field Guide: Malvan (Maharashtra, India) 505

Fig. 10 Satellite image of Padmagad Tombolo

close proximity to the mainland. The waves meet again between the island and the
mainland. As waves converge, water velocity suddenly drops leading to deposition
of sand and silt. Continued sedimentation finally forms an elongate stretch of land
connecting the island to mainland.

4 Geological Setting

Sindhudurg District exposes Peninsular Gneissic Complex, Sargur Group,


Bababudan Group, Kaladgi Supergroup, and Deccan Trap Basalts, overlain by
laterite and alluvium (Deshpande and Pitale 2014; Sarkar and Soman 2010;
Duraiswami and Patankar 2011). The basement is defined by feldspathic gneisses
and granites of Peninsular Gneissic Complex along with enclaves of Sargur Group.
Sargur Supracrustals are represented by schistose amphibolites, metaconglom-
erate, quartzite, banded magnetite quartzite), graphite schist, mica-garnet schists,
staurolite-kyanite-biotite schist, and calc-metasediments (Duraiswami and Patankar
2011). These rocks are intruded by chromite-bearing ultrabasic rocks near Vagda
and Janavli (Kamble 1993). The contact between gneisses and Sargur Supergroup
is gradational, indicating that the gneisses are much younger than the Sargur
supracrustals (Koregave 1980). The metaconglomerate-quartzite sequence of the
basal member of Bababudan Group (Walkunje Formation) extends along the coast
of Malvan to Nivti (Sarkar and Soman 2010; Sarkar and Soman 1986) for >20 km.
These rocks have been intruded extensively by basic to ultrabasic intrusives along
with granites and pegmatites (Kamble 1993; Duraiswami and Patankar 2011). The
Peninsular Gneisses and Dhawar Supergroup rocks are overlain by Kaladgi Super-
group with a prominent unconformity. The Kaladgi sediments are represented by
506 A. Pundalik et al.

Fig. 11 Quartz veins in staurolite schist, Ramgad. Scale is 15cm

pebbly conglomerate, sandstones, and shales. Sarkar and Soman (2010) have recog-
nized them as Achra Formation, which is subdivided into three members viz Golvan
Sandstone, Bagayat Shale, and Poyra Sandstone Members in the order if decreasing
antiquity. These rocks are cut by several faults. Faults are seen extensively concen-
trated more toward the coast than in the internal parts. This is also generally true for
Deccan Trap rocks in and around Mumbai (Misra et al. 2014; Misra and Mukherjee
2017; Mukherjee et al. 2017 etc.). The general trend of fault planes is NW. Therefore,
it may be stated that the coastline in Malvan is structure-controlled. The Deccan Trap
Basalts occur in small patches and rest unconformably over the Archean rocks and
Geological Field Guide: Malvan (Maharashtra, India) 507

the Kaladgi sediments. These basalts are sparsely/moderately porphyritic and occa-
sionally contain (sub) rounded volcanic glass blebs. Occasionally, Early Holocene
red and grey clays along with mollusc shells, wood, amber and peat are seen in
dugwell sections along the coast (Kumaran et al. 2004). The area around Malvan is
dominantly covered by lateritic capping, irrespective of the rock types under them.
The alluvium is seen along the banks of rivers, estuaries, and streams. Beach sand
and costal dunes are commonly seen throughout the stretch of coastline. Table 2
presents the stratigraphic succession as given by Gupta and Rai (1987), Samant
(1989) Kumaran et al. (2004), Sen et al. (2012), and Prusty and Sarkar (2014).

5 Field Observations

5.1 Sargur Group

A traverse can be taken along the southerly flowing second-order stream near Ramgad
[16° 14 28 N, 73° 36 33 E. Deformed staurolite-garnet-biotite schist belonging
to Sargur Group is observed all along the stream bed. Staurolite schist shows promi-
nent porphyroblasts of twinned staurolite with 5–12 cm size, with subordinate garnet
porphyroblasts of ~5 mm size (Fig. 14). Schistosity is defined by parallel orienta-
tion of biotite flakes, and staurolite porphyroblasts make an angle to the schistosity.
The foliation dips 30–40° toward ESE. The schist is intruded by two varieties of
amphibolites namely schistose and granulose. Intrusions of quartz and pegmatite
veins (Fig. 11) cut across the foliation of staurolite schist, and all intrusions trend
E–W. Schistose amphibolites almost entirely consist of hornblende, while granulose
amphibolites exhibit mineral composition Hornblende + Plagioclase + Quartz +
Garnet. Garnet porphyroblasts in amphibolites are 5–6 mm in size and seen along
with well-defined acicular hornblende. An outcrop of migmatized amphibolites of
Sargur Group (?) exists near the Hanuman Mandir (temple), Kunkavle [16° 05
24 N, 73° 34 51 E]. The migmatites exhibit superposed folding: Type-2 and Type-
3 of Ramsay (1967), Fig. 12 and parasitic folds (Fig. 13). Inliers of graphite schist,
migmatized amphibolite, metaconglomerate, metapelites and dumortierite quartzite
belonging to Sargur Group (Duraiswami and Patankar 2011) are seen in the areas
around Sangam Maruti temple, Kankavli [16° 15 10 N 73° 39 56 E; ~48Km NE
of Malvan] along Gad River Channel.

6 Peninsular Gneissic Complex

Feldspathic augen gneiss representing Peninsular Gneisses is well exposed at


Ghumda [16° 03 57 N, 73° 30 13 E] near Kumbharmathi, in the stream bed
of Anand Vahal (Fig. 15). The gneisses contain augens of feldspar (e.g., Mukherjee
508 A. Pundalik et al.

Table 2 Regional stratigraphy of the Sindhudurg district, Maharashtra (Gupta and Rai 1987;
Samant 1989; Kamble 1993; Kumaran et al. 2004; and Prusty and Sarkar 2014)
Alluvium (Holocene)
River terraces, beach dunes and sand ridges.
Local clay deposits with peat, plant remains, and mollusk shells.

Laterites (Plio-Pliestocene)
Primary laterite, reworked or secondary laterite.

Deccan Trap (Cretaceous-Paleogene)


Compact massive basalt, amygdaloidal basalt,
sparsely to moderately porphyritic basalt,
development of pillows at few localities.

-----------------------Unconformity-------------------------

Kaladgi Supergroup (Meso-Neoproterozoic)


Quartzite,
Ferruginous and carbonaceous shale
Ferruginous quartzite (at places grading to grit)
Conglomerate

-----------------------Unconformity-------------------------

Late intrusives Acidic rocks


Metabasic/ultrabasic intrusives

Bababudan Group (Archean)

Fuschite quartzite, Metapelites and Metavolcanics


Banded magnetite-quartzite
Polymict conglomerate-Quartzite
-----------------------Unknown contact-------------------------
Gneisses and Granites of Peninsular Gneissic Complex
-----------------------Unknown contact-------------------------
Sargur Group (Archean)
Kyanite-Staurolite-Biotite-Garnet Schist
Graphite schist
Ultrabasic intrusive
Dumortierite Quartzite-Metaconglomerate

-----------------------Basement not seen-------------------------


Geological Field Guide: Malvan (Maharashtra, India) 509

Fig. 12 Type ‘2’ Superposed folding pattern in migmatite at Kunkavle. Length of marker ~13 cm

2017). Gneissosity is defined by parallel preferred orientation of biotite and horn-


blende (Fig. 16). Gneissosity trends N12° and dips steeply toward southeast. These
gneisses exhibit three sets of joints trending N182°, N300°, and NE. Several ~ 15–
75 cm wide meta dolerite dykes intrude the gneisses (Fig. 17). The metadolerite
dykes intruding the gneisses at Kumbharmathi have an attitude of N322°/74, and
show schistosity defined by parallel orientation of hornblende (Fig. 18). Another
outcrop of Peninsular Gneiss is exposed near railway bridge at Kankavli [16° 15
14 N, 73° 43 03 E].
510 A. Pundalik et al.

Fig. 13 Folds in migmatized amphibolites (Sargur Group?) seen at Kunkavle. Scale 15cm.

6.1 Dharwar Supergroup

The conglomerate-quartzite sequence of basal member of Bababduan Group (Sarkar


and Soman 2010) is exposed on the coast near Santacruz Chapel, Rajkot [16° 03
19 N, 73° 27 20 E], and near the Rock Garden, Malvan, and can be traced for
~20 km toward S up to the coast of Nivti. Quartz grains in the quartzite show
a preferred orientation defining locally a crude schistosity. These conglomerate-
quartzite sequences are, folded into open, broad, doubly plunging folds (Fig. 19a)
with axes plunging towards NNW and SSE. The folds near Rajkot are tight, while
the folds near Rock Garden are open. Quartzites exhibit a variety of sedimentary
structures, e.g., large-scale tabular and trough cross bedding (Figs. 19b and 20),
hummocky cross bedding, festoon bedding, wedge cross bedding, sub-horizontal
bedding, parallel bedding, climbing ripple laminations, and parallel laminations indi-
cating their deposition in near-shore transition to inner-shelf environment (Sarkar and
Soman 2010). The paleocurrent direction determined from large-scale tabular cross
bedding is northwesterly. These quartzites also show soft sediment deformation struc-
tures such as convolute bedding (Fig. 21), syn-sedimentary faults, and slump folds.
Along the syn-sedimentary faults, drag folds (different from tectonic drag folds as
in Mukherjee 2014; Mukherjee et al. 2015) exist (Fig. 22).
Geological Field Guide: Malvan (Maharashtra, India) 511

Fig. 14 Staurolite schist outcrop in Ramgad, swiss knife for scale, length ~10 cm

Fig. 15 Peninsular Gneiss outcrop, Ghumda


512 A. Pundalik et al.

Fig. 16 Peninsular gneiss


sample, Ghumda

Fig. 17 Metadolerite dyke


intruding in augen gneiss.
Hammer ~25 cm in length
for scale
Geological Field Guide: Malvan (Maharashtra, India) 513

Fig. 18 Metadolerite dyke


sample, Kumbharmathi.
Coin for scale, diameter
~2 cm

Interbedded between the quartzites are thick layers of polymict pebbly conglom-
erates, which show clasts of quartz, amphibolite, banded magnetite-quartzite, granite,
gneiss, and varieties of quartzite, bounded by fine to medium sandy matrix containing
fuschite mica and ferruginous cement (Figs. 23 and 24). The pebbles range in size
from 5 to 10 cm and are preferentially bedding parallel. Pebbles are commonly elon-
gated along NNW–SSE. The conglomerates are mostly parallel bedded, occasionally
exhibiting wedge cross bedding (Sarkar and Soman 2010). Crude foliation defined
by parallel preferred orientation of fracture cleavage is observed in these conglom-
erates. These sequences are cut by numerous E-W trending, near vertical faults. A
strike slip fault can be observed behind Santacruz Chapel, Rajkot (Fig. 25). Alter-
nating beds of these conglomerates and fuschite quartzite are well exposed within
Sidhudurga Fort (Fig. 26) [16° 02 31 N, 73° 27 36 E].
The Dharwar metasediments represented by pebbly metaconglomerates,
quartzites, and chlorite-sericite-schists are seen to be displaced by a NNE-trending
fault, which is well exposed about 500m west of Devli [16° 01 00 N, 73° 29 50 E].
These metasediments dip steeply toward NNW and show a significant deformation.
Another outcrop of quartzites occurs near Deulwada [16° 03 30 N, 73° 28 53 E]
Malvan, where it is overlain by laterite capping. The quartzites here show a large
number of joints and fractures. Near vertical beds of quartzite strike ~N340°. These
are cut by numerous small scale left-lateral strike slip faults trending ~N352° with
up to 10 cm slip.

6.2 Kaladgi Supergroup

The Meso-Neoproterozoic Kaladgi Supergroup sediments are represented by ferrug-


inous quartzites interbedded with ferruginous and carbonaceous shales. These crop
out near Amdos-Wahalwadi [16° 06 09 N, 73° 33 08 E], Kuperi Ghat, and Shravan.
514 A. Pundalik et al.

Fig. 19 a Open, plunging folds in quartzite near Rock Garden, Malvan. Ashwin Pundalik standing
on the outcrop, height 163cm. b Tabular cross bedding, near Rock Garden, 15 cm scale as marker
Geological Field Guide: Malvan (Maharashtra, India) 515

Fig. 20 Large-scale trough cross bedding near Rock Garden, 25 cm hammer for scale

Fig. 21 Convolute bedding, near Rock Garden


516 A. Pundalik et al.

Fig. 22 Syn-sedimentary
faults in quartzite, Rock
Garden, Malvan. Scale is
15cm

Fig. 23 Sample showing quartzite and conglomerate contact, Rajkot


Geological Field Guide: Malvan (Maharashtra, India) 517

Fig. 24 Field photograph of polymict conglomerate outcrop at Rajkot, 15 cm scale

Fig. 25 Strike slip fault, displaced conglomerate bed marked in photograph and inset shows the
box diagram, Rajkot
518 A. Pundalik et al.

Fig. 26 Bedding in fuschite quartzite of Bababuddan Group, Sindhudurg Fort

Laminated shales (Fig. 27) alternating with ferruginous quartzite in turn capped by
a thick laterite body occur near Vahalwadi. These are exposed in open, broad folds
plunging gently toward north. These can easily be inferred from the topographic
map, where the two cuestas can be seen with dip slopes facing toward east and west.
The Kaladgi sediments form the elevated portion of the Kuperi Ghat section, where

Fig. 27 Thinly laminated shale sample, Vahalwadi


Geological Field Guide: Malvan (Maharashtra, India) 519

Fig. 28 Well-preserved
wave ripples, Shravan.
Hammer 25 cm, length

they seem to be resting unconformably on migmatized amphibolites of Sargur Group


(?) exposed near Kunkavle.
The Kaladgi sediments overly sericite phyllites of Dharwar Supergroup with
a prominent nonconformity near Rathiwde and in Gad River bed near Belna and
Shravan. Here, they are represented by thin layers of ferruginous oligomict pebbly
conglomerate passing upward into ferruginous coarse- to fine-grained sandstones
with occasional buff-colored shale. These sandstones show sedimentary structures
viz parallel bedding, tabular and trough cross bedding, and very well-preserved,
well-defined wave ripples (Fig. 28).

6.3 Deccan Trap Basalts

The Deccan basalts occur as patches in the study area. These are massive and mostly
aphyric (Fig. 29).
Locally, basalts with blebs of volcanic glass (Fig. 30) occur in a laterite quarry
near Devli [16° 00 5 N, 73° 30 0 E] where basalts are seen underlying the laterite.

6.4 Primary and Secondary Laterite

Laterites in Malvan are of two types—primary and secondary/reworked. The primary


laterite has formed from varieties of rocks (Fig. 31), and it constitutes a thick capping
520 A. Pundalik et al.

Fig. 29 Basalt sample from


a quarry, Devli

Fig. 30 Basalt sample with


a glassy bleb in it (pointed
by arrow)

on large areas forming flat-topped hills. Due to this laterite capping, underlying rock
types can only be seen in valleys or stream beds. A complete laterite profile can be
seen in abandoned laterite quarries near Deulwada, Malvan, where laterites overly
Dharwar quartzites. However, due to dense vegetation and risk of rockfall, the outcrop
must be approached with great caution. Excellent quarry sections of primary laterites
can be seen in the areas around Amdos (Fig. 32) and Kumbharmathi. Secondary
laterite having conglomerate-like appearance (Fig. 33), due to re-lithification of the
lateritic and other rock debris, has formed all around the coast of Sarjekot [16° 04
58 N, 73° 27 29 E] and is well exposed in coastal cliffs, exhibiting variety of
coastal geomorphic features discussed earlier (Sect. 3) (Fig. 34).
Geological Field Guide: Malvan (Maharashtra, India) 521

Fig. 31 Laterite sample, Devulvada

Fig. 32 Laterite quarry, Amdos

6.5 Holocene Clays

Early Holocene clays are seen in a pile of dug well debris near Ozar [16° 05 43 N,
73° 28 52 E]. Amber, pyrite, and lignite/peat (Kumaran et al. 2004) are found—all
522 A. Pundalik et al.

Fig. 33 Secondary laterite sample, Sarjekot

contained within clay. Plenty of ferruginized pneumatophores along with other plant
remains are also found (Figs. 35 and 36).

Fig. 34 Sea arch formed in secondary laterite, Sarjekot


Geological Field Guide: Malvan (Maharashtra, India) 523

Fig. 35 Amber collected during 2010 fieldwork

Fig. 36 Amber collected during 2018 fieldwork

6.6 Beach Sand

The Malvan area has extensive stretches of scenic beaches at Tarkarli, Malvan, Chivla,
Talashil, and Deobag. The beaches show well-developed beach profiles as already
described in Sect. 2.1. Various types of primary sedimentary structures such as hori-
zontal parallel bedding, horizontal parallel lamination, ripple lamination, and wave
ripples can be observed and studied in these sands (Figs. 37 and 38). On Tarkarli
beach, a small stream merges with the sea and exhibits variety of current ripple
524 A. Pundalik et al.

Fig. 37 Wave ripples,


Tarkarli Beach, 15 cm scale

Fig. 38 Current ripples,


Tarkarli Beach, 15 cm scale
Geological Field Guide: Malvan (Maharashtra, India) 525

Fig. 39 Syn-sedimentary
deformation structures in a
dug up trench, Tarkarli
Beach, 30 cm scale

geometries such as straight crested, linguoid, cuspate, lunate, etc. (Fig. 37). The phys-
ical processes leading to formation of these structures can be studied and observed
by the students here.
Besides, the beach sands show a variety of syn-sedimentary deformation struc-
tures, e.g., convolute bedding, syn-sedimentary faults, load casts, and ball and pillow
structure (Fig. 39). These structures can be seen in beach dune deposits at Tarkarli
and Talashil, and have been studied occasionally in the trenches dug by students at
Talashil Beach. The beach sands around Malvan host a rich assemblage of heavy
minerals, which can be observed in dark sands, with help of hand lens, or with a
binocular microscope in laboratory. These heavy minerals include zircon, tourma-
line, rutile, magnetite, garnet, ilemnite, augite, occasional kyanite, staurolite, biotite,
hornblende, and chromite (Gujar et al. 2010), indicating a mixed provenance of
Precambrian metamorphics, granites, gneisses, and Deccan Trap Basalts for the beach
sand. Lithic fragments of laterite are also commonly seen, at places imparting red
color to the sand.
In this fieldwork, students also carry out Rf /φ, center-to-center, and Fry method
of strain analyses of deformed clasts of conglomerates of Bababudan Group and
Sargur Group. They also map the area using Plane table, auto level, RTK-GPS, Total
Station. A permanent benchmark has been established by the graduating batch of
2017, near Rock Garden [16° 03 41.55846 N, 73° 27’21.20767 E] at the rocky
beach, near Malvan Court using RTK-GPS on 27th October 2016 (Fig. 42).
526 A. Pundalik et al.

Fig. 40 Formation of current and wave ripples, Tarkarli Beach, 15 cm scale

Fig. 41 Rocks in Malvan and surrounding areas marked on Google Earth image
Geological Field Guide: Malvan (Maharashtra, India) 527

Fig. 42 Permanent benchmark established by students of St. Xavier’s College, near Malvan Court,
Malvan

Acknowledgements Sincere thanks to Mr. Mayekar whose cooperation every year has greatly
facilitated the fieldwork. Special thanks to Smith D’Britto for making the Malvan fieldwork
movie. The authors acknowledge Riya Bidaye and Lynette Dias for graciously sharing field
photographs. Soumyajit Mukherjee (IIT Bombay) made two rounds of review. Thanks to Marion
Schneider, Annett Buettener, Boopalan Renu, Alexis Vizcaino, Doerthe Mennecke-Buehler, and
the proofreading team (Springer). Dutta and Mukherjee (2021) encapsulate this work.

References

Davidson-Arnott, R. (2009). Introduction to coastal processes and geomorphology (1st ed.). New
York: Cambridge University Press.
Deshpande, G. G., & Pitale, U. L. (2014). Geology of Maharashtra (2nd ed.). Bangalore: Geological
Society of India.
Duraiswami, R., & Patankar, U. (2011). Occurrence of fluoride in drinking water sources from Gad
river basin, Maharashtra. Journal Geological Society of India, 77, 167–174.
Dutta, D.,& Mukherjee, S. (2021). Introduction to Structural Geology and Tectonics Field Guide-
book – Volume 1. In S. Mukherjee (Ed.), Structural Geology and Tectonics Field Guidebook -
Volume 1. Switzerland AG: Springer Nature. Cham. pp. xi-xvi. ISBN: 978-3-030-60142-3.
528 A. Pundalik et al.

Gujar, A. R., Ambre, N. V., Iyer, S. D., Misalankar, P. G., & Loveson, V. J. (2010). Placer Chromite
along South Maharashtra, Central west coast of India. Current Science, 99, 492–499.
Gupta, S. K., & Rai, R. P. (1987). Preliminary appraisal for different minerals in part of Sindhudurg
District, Maharashtra. Pune: Geological Survey of India.
Hugget, R. J. (2007). Fundamentals of geomorphology (2nd ed.). New York: Routledge.
Kamble, D. T (1993). A Preliminary mineral appraisal study of the ultrabasic bodies and laterite at
Kankavli and Janavli villages, Sindhudurg District, Maharashtra, Unpublished Geological Survey
of India report, FSP 1990–1991.
Koregave, M. A. (1980). Precambrian geology of the area around Kankavli, Ratnagiri district,
Maharashtra. Unpubl. Ph.D. thesis, University of Pune.
Kumaran, N., Limaye, R. B., & Shindikar, M. (2004). Fossil record of marine manglicolous fungi
fom Malvan (Konkan) west coast of India. Indian Journal of Geo-Marine Sciencs, 33(3), 257–261.
Mishra, S. (2014). Ground Water Information Sindhudurg District Maharashtra. Central Ground
Water Board, 1–24.
Mukherjee, S. (2014). Review of flanking structures in meso- and micro-scales. Geological
Magazine, 151, 957–974.
Mukherjee, S., Punekar, J., Mahadani, T., & Mukherjee, R. (2015). A review on intrafolial folds and
their morphologies from the detachments of the western Indian Higher Himalaya. In S. Mukherjee
& K. F. Mulchrone (Eds.), Ductile shear zones: From micro- to macro-scales (pp. 182–205). New
York: Wiley.
Misra, A. A., Bhattacharya, G., Mukherjee, S., & Bose, N. (2014). Near N-S paleo-extension in the
western Deccan region in India: Does it link strike-slip tectonics with India-Seychelles rifting?
International Journal of Earth Sciences, 103, 1645–1680.
Misra, A. A., Mukherjee, S. (2017) Dyke-brittle shear relationships in the Western Deccan Strike Slip
Zone around Mumbai (Maharashtra, India). In S. Mukherjee, A.A. Misra, G. Calvès, M. Nemčok
(Eds.), Tectonics of the Deccan large igneous province (Vol. 445, pp. 269–295). Geological
Society, London: Special Publications.
Mukherjee, S., Misra, A.A., Calvès, G., & Nemčok, M. (2017). Tectonics of the Deccan large
igneous province: An introduction. In S. Mukherjee, A. A. Misra, G. Calvès, M. Nemčok (Eds.),
Tectonics of the Deccan large igneous province (Vol. 445, pp. 1–9). Geological Society, London:
Special Publications.
Mukherjee, S. (2017). Review on symmetric structures in ductile shear zones. International Journal
of Earth Sciences, 106, 1453–1468.
Prusty, S., & Sarkar, A. (2014) Report on specialized thematic mapping in precambrian rocks in
Sangeli-Konshi-Kalna area, Sindhudurg District, Maharashtra Unpublished Geological Survey
of India Report (Item Code No:/STM/CR/MH/2012/013).
Ramsay, J. (1967). Folding and fracturing of rocks; structural geology of rocks (2nd ed.). New York:
McGraw-Hill Book Company.
Samant, H. P. (1989). Geomorphology of the area around Malvan, Sindhudurg district of
Maharashtra using remote sensing techniques. (Unpublished MSc thesis).
Sarkar, P. K., & Soman, G. R. (1986). Geology of the area around Katta, Sindhudurg district,
Maharashtra based on aerospace data. Journal Indian Society of Remote Sensing, 14 (edition 2).
Sarkar, P. K., & Soman, G. R. (2010). Environment of deposition of the metaconglomerate-
quartzite sequence (equivalent of basal member of Bababuddan Group) along the Malvan Coast,
Sindhudurg district, Maharashtra, India. In Origin and evolution of the deep continental crust
(pp. 32–42). New Delhi: Narosa Publishing House Pvt. Ltd.
A Field Guide to the Champaner Region,
Southern Aravalli Mountain Belt
(SAMB), Gujarat, Western India

Aditya U. Joshi and Manoj A. Limaye

Abstract Geological fieldwork is the most important part of geosciences training


because it acts a backbone to obtain data, over which the entire knowledge of the
Earth’s evolution, existence and origin of life throughout the geological times are
interpreted. Due to the advancement of technology and on account of the sophisti-
cated instruments, the field science has taken a back seat. It remains vital to train
geoscientist in the field, so that they develop eyes for observing geological features,
which will eventually help them frame and solve geological problems. Here, we
present 15 geological stops divided into 4 traverses, from the Champaner region of
Western India, with an aim to cater various geological disciples such as structural
geology, igneous and metamorphic petrology, basics of stratigraphy, etc.

1 Introduction

The Champaner region, located at the eastern extremity of Gujarat, India, offers a
package of rocks from Paleo-Proterozoic basement gneisses to the recent sediments.
The present region mainly covers rocks that belong to the Champaner Group, a
youngest Group in the Aravalli Supergroup. The entire area can be traced using
Survey of India (SOI) topographic sheet numbers 46F/7, 10, 11, 14 and 15. The area
has a good accessibility from Vadodara district, located 72 km NE of Vadodara city.
The area is approachable by both road and railways (Fig. 1). The nearest airport is
The Maharaja Sayajirao Gaekwad (MSG) International Airport, Vadodara. The most
convenient way to visit the area is by roadways. The State Highway (SH) number
87 from Vadodara to Nurpura (~NE–SW trending) reaches the Champaner region.
Further, the SH 87 meets SH 194 and has direct connection to the Halol town. The
Champaner region lies east of Halol town and has a superior accessibility through

A. U. Joshi (B) · M. A. Limaye


Department of Geology, Faculty of Science, The Maharaja Sayajirao University of Baroda,
Vadodara 390002, Gujarat, India
e-mail: adityaujoshi@gmail.com
M. A. Limaye
e-mail: manoj_geol@rediffmail.com

© Springer Nature Switzerland AG 2021 529


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_18
530 A. U. Joshi and M. A. Limaye

Fig. 1 a India inset specifying location of the study area; b. communication map of the Champaner
region (study area)

SH 5. The good thing about the SH 5 is that the exposures of the representative rocks
belonging to all age group are seen on either side of the road. Moreover, the area has
excellent connectivity through internal roads of metal and non-metal nature.
The area has rail connectivity through Vadodara-Godhra-Delhi (broad gauge).
The nearest railway station is the Champaner Road Junction near Khakhariya village.
SH 150 connect the railway station and leads towards the Halol town. SH 5 joins
Halol, which further guides towards the Champaner region. Mobile phone network
is quite poor in this region, although Jio and Vodafone provide decent connectivity.
The nearest emergency medical facility is at the Halol town. Regarding accommo-
dation within the Champaner region, there are Department of Forest, Government
of Gujarat eco-tourisms available at reasonable rates with maximum occupancy of
35–40 people. Besides, there are several good private resorts available.
Under the course heading of geological fieldwork, the present terrain has been
an official training site for the M. S. University Baroda, Vadodara, under-graduate
and post-graduate students since past three decades. The legacy of training students
including young researchers and faculty members have been initiated by the founder
of the Geology Department, MSUB, Prof. S. S. Merh, followed by their students
and later on colleagues, Late. Prof. Mohan. P. Patel and Prof. Subhash. J. Desai
(Retd.). These professors use to conduct field campsites to teach structural geology
in the low-grade metamorphites. On account of their meticulous teaching style, they
developed a thorough interest of the subject within innumerable students. Many of
them today have secured high positions in teaching, research and industrial sectors.
At present, every year more than 130 students from MSUB and 100 from various
other academic and professional institutes visit the Champaner terrain under the
guidance of Prof. Manoj A. Limaye and Dr. Aditya U. Joshi. Giving more emphasis
A Field Guide to the Champaner Region, Southern Aravalli … 531

on teaching to collect field data, the present pair mentors outcrop and regional scales
mapping sessions during the field season between October to February. Till now,
in total two Ph.D. thesis from MSUB, viz., (i) B. B. Jambusaria (1970) (under the
guidance of Prof. S. S. Merh; (ii) Aditya U. Joshi (2019b) (under the guidance of
Prof. M. A. Limaye), plus more than 20 M.Sc. dissertation have been produced from
the Champaner terrain.
Like other Precambrian rocks across India, the present terrain is economically
important for its manganese and uranium mineralization. Erstwhile, due to the resem-
blance in terms of the rock types the area has often been correlated by the Dharwarian
rocks of South India (Fermor 1909) and represents a classical example of structurally
controlled Mn deposit (Beer 1919). The major Mn ores found within the area are
pyrolusite, psilomelane, braunite and winchite (Roy 1958; Joshi and Limaye 2018).
Atomic Mineral Division (AMD), Department of Atomic Energy, explored the area
for the uranium deposit.

2 Topography

The topography of the Champaner region is undulating. The area is characterized


by linear, S-, C- and V-shaped ridges with near-flat tops alternating with narrow
valleys reflecting lithological variations of the constituent rocks. The eastern part
of the Champaner region comprises a flat and rolling topography with occasional
low to medium elevated circular mounds. Water bodies which drain on this side are
Sukhi and Orsang. The Sukhi is a tributary of Orsang and has been obstructed to
form artificial reservoir known as Sukhi Dam, located at the NE direction of the
Champaner terrain. The extreme western and southern part has by and large similar
topography, except at the north-western part where high elevated circular mound
with flat top resembling a plateau-like feature exists at the hill Pavagadh. Major river
which originates is Dhadhar, while its minor tributaries flow on the western direction
are Dev Nadi and Vishwamitri. The core/central portion holds varieties of ridges
shape, starting from west and the elevation gradually increases progressing towards
the east. Initially, the ridges attend broad “S”-shaped pattern and often bound by
linear ridges on either side. Advancing towards east the ridges acquire compressed
C-shaped pattern followed by a V-shape at the middle portion. There also exist few
domal features at the eastern and the north-eastern end of the Champaner region.
The domes are elliptical in shape and the size of its long axis increases towards
the northern part. The area encompasses eroded peneplain surfaces at the south-
eastern part of the Champaner region and can be very well appreciated through
the Shuttle Radar Topography Mission (SRTM)-derived-shaded relief map (Fig. 2).
The alternating valleys between the ridges are mainly occupied by softer rocks like
phyllites and slates, prone to be eroded faster and buried under soil cover, whereas
the quartzites and granites stand as high ridges and mounds, respectively, due to its
surpassing hardness. The highest benchmark situated at the southern part, recorded
is 354 m, which comprises of quartzites.
532 A. U. Joshi and M. A. Limaye

Fig. 2 Shuttle Radar Topography Mission (SRTM)-derived-shaded relief map of the Champaner
region. Source USGS Earth-explorer

Since the geology of the Champaner region is a part of the Champaner Group,
Aravalli Supergroup, a brief discussion on litho-stratigraphy, sedimentation history,
metamorphism and structure is necessary.

3 Litho-Stratigraphy

The rocks of the Champaner region belong to the Champaner Group, previously
referred to as “Champaner Series” (Blanford 1869) of Western India, are believed
to be a part of the upper Aravalli supracrustals and crop-out at the easternmost
fringe of the Indian state Gujarat. The Champaner Group is located geographically
at the southernmost part of the Aravalli Mountain Belt (AMB) and is considered as a
major Proterozoic stratigraphic unit of the Southern Aravalli Mountain Belt (SAMB)
in Gujarat (Fig. 3).
The overall geological setting around the Champaner Group reflects the Creta-
ceous Deccan Trap Formation along with few infra-trappean rocks and recent sedi-
ments at the west. The other three geographic sides occupy Neo-Proterozoic plutonic
intrusives (955 ± 20 Ma. Rb/Sr Method; Gopalan et al. 1979): the “Godhra gran-
ites”. The Champaner Group forms an “inlier” surrounded by rocks of younger age
and forms roughly a horseshoe-shaped outline. This group has been divided into
A Field Guide to the Champaner Region, Southern Aravalli … 533

Fig. 3 Litho-stratigraphic map of the Aravalli Mountain Belt (AMB) (modified after Gupta et al.
1992). I–VII are the major lineaments present within the AMB, viz., (i) I—Great Boundary Fault
or Bundi Lineament; (ii) II—Delwara Lineament; (iii) III—Rakhabdev Lineament; (iv) IV—
Kaliguman Lineament; (v) V—Phulad Lineament; (vi) VI—Pali Lineament; (vii) VII—Banas
Lineament

six formations, viz., the Lambhia, the Khandia, the Narukot, the Jaban, the Shivra-
jpur and the Rajgarh Formation based on lithological homogeneity and occurrence of
intra-formational conglomeratic horizon (Gupta et al. 1997) (Fig. 4). Srikarni and Das
(1996) gave a detail litho-stratigraphic sequence of the Champaner Group (Table 1).
They divided Champaner Group based on the five conspicuous marker horizons
of which two are conglomerate horizons, which separates the Champaner Group
into lower, middle and upper subgroups. The lower Champaner Subgroup represents
Lambia Formation comprise four main rock types from 1a–d. 1a forms the base of
the Champaner Group having quartzite and basal oligomictic conglomeratic horizon.
1b, 1c and 1d are represented by phyllite and spotted slate, dolomitic limestone rich
in stromatolites and phyllites, respectively. The middle Champaner Subgroup clubs
rocks belonging to the Khandia and Narukot Formation and are divided into four sub-
categories (i.e. 2a–d). These rocks mark a distinct gradation in terms of grain size from
oligomictic conglomerate at the base of Khandia Formation, followed by quartzite
534 A. U. Joshi and M. A. Limaye

Fig. 4 Litho-stratigraphic map of the Champaner region (after, Joshi 2019a)

to meta-arkose variety and finally grades into phyllite. The uppermost subgroup of
the Champaner represents the thickest sequence covering the rest of the formations
of the Champaner Group. The Jaban, the Shivrajpur and the Rajgad Formation are
placed within the Upper Champaner Subgroup and are divided into five chief litho
units: 3a–e. Polymict boulder-cobble conglomerate appears at the base of the Upper
Champaner Subgroup deputed at 3a, followed by quartzite and meta-arkose as 3b,
manganiferous phyllite and quartzite as 3c, siliceous dolomitic limestone as 3d and
thick band of phyllite containing minor bands of siltstone/quartzite with occasional
thin limestone bands constitute 3e.

4 Sedimentation History

Sedimentation history of the Champaner Group suggested by Srikarni and Das (1996)
takes care of the overall thickness of the Champaner metasediments >2100 m.
The thick pile of sequence deposited in three major cycles. Out of these, the
lowermost cycle is incomplete and the other two starts with the conglomeratic
horizon having scoured contact with underlying phyllite. These conglomerates show
evidences of submarine fan complex and deposited due to activation of major
A Field Guide to the Champaner Region, Southern Aravalli … 535

Table 1 Stratigraphy of the Aravalli rocks in and around Champaner region (after Srikarni and
Das 1996). Highlighted cells indicate the part of litho units belonging to respective formations have
been described here
Age Group/sub-group Formation Member Lithology
Cretaceous to eocene Deccan traps Basic flows, basic and
alkali dykes
Cretaceous Infratrappean Sandstones and
(Bagh beds) limestones
Neo-proterozoic Godhra granite Post-tectonic and
syntectonic granite and
granodiorite
Meso-proterozoic Champaner Group Rajgarh 3E Slates and slaty
phyllites with thin
calc-silicate and
quartzite bands
3D Dolomitic limestone
Shivrajpur 3C Manganiferous
phyllites and quartzites
3B Quartzite and
meta-arkose
Jaban 3A Polymictic
conglomerate with
greywacke
Narukot 2D Grey phyllite and
carbon phyllite
2C Meta-arkose and
greywacke with
intraformational
conglomerate
Khandia 2B Quartzite and
meta-arkose
2A Boulder/cobble
conglomerate- clasts
mainly of quartzite with
minor magnetite,
quartzite and phyllite
Lambia 1D Phyllite/biotite schist
(in the eastern part)
1C Dolomitic marble and
calc-silicate rock
1B Phyllite/biotite schist
and spotted slate
1A Quartzite and
impersistant
conglomerate
(continued)
536 A. U. Joshi and M. A. Limaye

Table 1 (continued)
Age Group/sub-group Formation Member Lithology
Palaeo-proterozoic Pre-champaner Granite gneiss,
gneissic complex quartzite and pelitic
gneiss

marginal/intrabasinal faults, which promoted coarser clastic supply to the basin. Pale-
ocurrent data of Srikarni and Das (1996) suggests that the provenance was towards
the south with an extension towards north due to several episodes of block faulting.
Basement exposed along the southern part of the Champaner Group display complex
deformation with higher grade of metamorphism than the Champaner metasediments.
Based on the homogeneity in terms of rock types the Champaner Group show simi-
larity with the Debari Group of the Middle Aravalli and Lunavada Group of the
Upper Aravalli Supergroup.

5 Structure and Metamorphism

The Champaner Group deformed thrice, viz., D1 , D2 and D3 , which has developed F 1 ,
F 2 and F 3 folds, respectively (Joshi 2019a). The first two phases of deformation are
coaxial, to develop F 1 ~ ESE-WNW and F 2 ~ E–W trending folds. The third episode
of deformation exhibits fold trends ranging from NNW–SSE to NNE–SSW. Apart
from these trends, there are evidences of out-of-sequence deformation displaying F 1
~ NW–SE to N–S and F 2 ~ NE–SW. Combinations of various folds have generated
interference fold patterns on regional and mesoscopic scale. Superimposition of F 1
and F 2 fold has resulted into Type-III interference fold pattern, where as combination
of F 3 has developed the Dome and Basin geometry (i.e. Type-I). In case of out-of-
sequence deformation, Type-II interference pattern has been generated over Type
0.
Followed by fold events, the rocks of the region have been subjected to brittle
faulting ranging from E-W to N–S direction in the form of radial pattern (Fig. 4).
These faults have developed along the pre-defined axial traces due to the granitic
emplacement in the SE and E extremity of the Champaner region.
As far as metamorphism is concerned, the region has witnessed regional meta-
morphism up to greenschist facies condition by the development of chlorite and
biotite during D1 deformation. The D2 deformation produced garnets within the
meta-pelites. There is no mineral development during D3 deformation. The grade
of regional metamorphism increases towards east (i.e. from Rajghar Formation to
Narukot Formation) by the development of chlorite-biotite-garnet. Due to the effect of
granitic intrusion the rocks of the area has developed, contact metamorphic textures
overprinting the regional metamorphic rocks. The minerals indicative of thermal
metamorphism are andalusite, cordierite, sillimanite (fibrolite), with few metaso-
matic changes to develop minerals such as muscovite and tourmaline. The overall
A Field Guide to the Champaner Region, Southern Aravalli … 537

grade of thermal metamorphism reached up to pyroxene-hornfels facies condition


(Joshi 2019b).

6 Field Traverses-Preamble

The field traverses presented in this chapter covers lithology with special reference
to primary structures, secondary structures, regional and contact metamorphism. In
addition, these traverses aim to elucidate the occurrence of linear and planar fabrics
associated with various rocks exposed within the Champaner terrain. The majority
of the stops in these traverses lies in the SOI topographic sheet number 46F/10
and 11. These stops show many explicit textbook examples, which will be highly
helpful to teach field exposures, specifically for under-graduate and to some extent
for post-graduate students.
In all, 15 stops are covered within 4 traverses by providing exact locations (co-
ordinates) for each stop. Continuous numbers have been given to the stops to avoid
the confusion. Traverse 1 consists of Stops 1 and 2; Traverse 2 consists of Stop from
3 to 7; Traverse 3 covers Stops from 8 to 13 and Traverse 4 has two Stops 14 and 15.
Moreover, each stop has been supplemented by a landmark, which will be helpful to
reach the precise destination.

7 Traverse 1—Objective and Overview

Traverse 1 mainly covers the igneous rocks present within the Champaner region.
This traverse is designed based on accessibility/connectivity of the road available in
the terrain and covers volcanic and plutonic igneous rock. These two stops are at the
road side and contains widen road for all sized vehicles (54 seater bus).

7.1 Stop 1: Base of Mount Pavagadh Volcanic Suite


(Abandoned Quarry Section)—(22° 29 7.48 N;
73° 31 7.82 E)

Mount Pavagadh is an important volcanic suite present within the Champaner region.
It mainly consists of sub-alkaline massive basalts at the base with picrite and rhyolite-
dacite flow as capping (Sheth and Melluso 2008). The maximum altitude is about
810 m above M.S.L and 12 flows have been identified (Hari et al. 1999).
The present stop is located about 450 m prior to the Champaner gate, with a
landmark of Helical Stepped Well (Fig. 5). There lies a pedestrian walk for ~50 m,
538 A. U. Joshi and M. A. Limaye

Fig. 5 Landmark for the Stop 1, Helical Step Well board, 450 m prior to the Champaner gate

which leads towards an abandoned quarry section. The quarry consists about 25–
30 m steep excavated scarp having ~150 m width (Fig. 6). The dominant rock type
exposed is massive melanocratic, fine to medium grained porphyrictic basalt, which
belongs to the Cretaceous Deccan trap. Predominantly, two to three flows can be
indentified at this location, with the evidence of amygdaloidal basalt at the top of the
flow. There also exist isolated outcrops of ignimbrite and volcanic breccia, which
connotes a pyroclastic flow (Figs. 7 and 8). The out skirts of the quarry or along the
main highway a road section comprising of olivine-rich basalts is presumably the
base of the Mt. Pavagadh (Fig. 9).

Fig. 6 Panoramic view of the abandoned quarry section located at base of Mt. Pavagadh
A Field Guide to the Champaner Region, Southern Aravalli … 539

Fig. 7 Field photograph of ignimbrite outcrop indicating pyroclastic flow. The earlier fragments
were seen to be reacted with the newer melt during solidification (exposed pen length = 6 cm)

Fig. 8 Field photograph showing volcanic breccia as isolated outcrops within an abandoned quarry
(pen as a marker having length = 13.5 cm)
540 A. U. Joshi and M. A. Limaye

Fig. 9 Close up view of olivine-rich basalts at the base of the Mt. Pavagadh. Typical euhedral
olivine phenocrysts altered to serpentine (exposed pen length = 4 cm)

7.2 Stop 2: Boulder Granites Near Ranjitnagar


Village—(22° 31 3.61 N; 73° 36 16.87 E)

The granites exposed within the Champaner region belong to the youngest plutonic
intrusive igneous activity pertaining to the Aravalli orogeny, and popularly referred
to as “Godhra Granites”. This granite is the important Precambrian (precisely Meso
to Neo-Proterozoic) stratigraphic unit of Southern Aravalli Mountain Belt (SAMB),
which gave out-of-sequence deformation by modifying pre-existing structural grain
of the Champaner metasediments, suggestive of prolong emplacement record even
after regional deformation (Joshi and Limaye 2018, 2020). Recent work carried out
by Joshi et al. (2018a, b) suggest that these granites have a widespread existence;
responsible for deforming the country rocks (i.e. the Champaner Group) and have
a continuous trail of subsurface occurrences below the Champaner metasediments.
Two pluses have been identified, viz., 1. syn-tectonic granites (of deformed variety)
and 2. post-tectonic pulse (of un-deformed variety with the presence of xenoliths of
earlier granites and metasediments: Joshi 2019b).
The second stop within this traverse is located 1 km East of Ranjitnagar village,
with a landmark of Ranjitnagar Primary School (Fig. 10). A single track road heads
towards Chelavada directs to the granitic outcrops. From a distance, the rock shows
boulder appearance on account of tor weathering (Fig. 11). The rock also comprises
tafoni weathering at mesoscopic scale, with the presence of distinct mural joints
(Fig. 12). The rock is leucocratic, medium to very coarse grained, porphyritic,
A Field Guide to the Champaner Region, Southern Aravalli … 541

Fig. 10 Landmark for the Stop 2, Ranjitnagar Primary School, 500 mts interior from the main road

plagioclase rich and non-foliated. It also consist variable amount of unoriented xeno-
liths (fine grained granites + quartzites + phyllites) indicating post-tectonic pulse
(Fig. 13).

8 Traverse 2—Objective and Overview

This traverse consists of five stops and mainly covers the top four Formations of
the Champaner Group, viz., Rajghar, Shivrajpur, Jaban and Narukot. The traverse
along this route passes through a major axial trace of regional folds. The traverse
is marked opposite to the plunge direction of these regional scale folds, which
helps to encounter consecutive older rocks sequentially along a single array. The
present traverse is useful to record and teach basic linear and planar fabrics in the
low-grade meta-sedimentary and display aesthetically the increment in the grade of
regional metamorphism from younger to older Formations of the Champaner Group,
respectively.
542 A. U. Joshi and M. A. Limaye

Fig. 11 Panoramic view of the granite country, showing prominent boulders (Mr. Sandeep as a
marker having ht. 5.3 ft.)

8.1 Stop 3: Vadatalav Quartzites (Road


Section)—(22° 29 37.17 N; 73° 33 42.14 E)

Quartzites ubiquitous in the Champaner Group occur as hard, massive, compact


at times flaggy and whitish to dark grey and/or black in colour. These are highly
jointed with variable grain size from fine to coarse and display varieties of primary
sedimentary structures at different locations. Quartzites occur as high ridges and
erratically replicate complex deformation pattern in the form of curvilinear to circular
shapes. This stop of the traverse 2 comprises the rocks that belong to the Rajghar
Formation of the Champaner Group. The chief rock types of this formation are
thick phyllites with thinly bedded quartzites accompanied by occasional presence
of meta-greywackes intercalated by dolomitic limestones. This present formation is
located along the western direction of the Champaner region and covers maximum
geographical extent within the Champaner Group. It is characterized by low mounds
and rolling topography. Most of the outcrops are covered by soil or alluvium due to
its susceptibility to physical weathering.
The quartzites exposed at Stop 3 near Vadatalav are located at the road section,
exactly at the junction of Ghoghamba-Halol bypass road (Fig. 14). The rock is fine
to medium grained and display buff-brown colour on weathered surface, while over
A Field Guide to the Champaner Region, Southern Aravalli … 543

Fig. 12 Field photograph showing mural joints within granites (Ms. Parita as a marker having ht.
4.9 ft.)

Fig. 13 Outcrop photograph showing unoriented xenoliths within granites (hammer as a marker
having length = 32 cm)
544 A. U. Joshi and M. A. Limaye

Fig. 14 Landmark for the Stop 3, sign board of Ghoghamba-Halol bypass road

fresh surface it exhibit steel grey to whitish tint. The fold exposed show first-order
open type F 2 folds of amplitude lesser than its wavelength (Mukherjee et al. 2020)
(Fig. 15). These folds show prominent cleavage refraction on account of different
competency within these quartzites. Noticeable occurrences of asymmetric ripples
along limbs portions suggest folding of primary bedding (Fig. 16). The fold trends
ESE–WNW direction with westerly plunging fold axes. Apart from mesoscopic
fold evidences, the fault gauge and slickensides on account of normal and reverse
faulting are seen at the hinges and along the limbs of the mesoscopic fold, respectively
(Figs. 17 and 18). However, these faults are localized and restricted along this specific
section. The quartzites also display significant goethite leaching on the upper surface
(Fig. 19).

8.2 Stop 4: Bhat Mines (Abandoned Mine


Section)—(22° 25 28.36 N; 73° 37 45.83 E)

This stop holds one of the most important outcrops to display structural features and
structurally controlled mineralization. Some of them are explicit textbook examples
in terms of guides to ores and shear sense indicators. Students can measure the fold
attributes by simply sitting over this fascinating outcrops and can develop a lot more
A Field Guide to the Champaner Region, Southern Aravalli … 545

Fig. 15 Outcrop-scale first-order open type F 2 folds. The fold consist of one anticline and syncline
with wavelength greater than amplitude (Dr. A. Joshi as a marker having ht. 5.7 ft.)

understanding about the basics of ductile deformation. The dominant rock types
exposed are intercalated manganese-rich quartzites with thin phyllitic layers that
belong to Shivrajpur Formation of the Champaner Group. The Shivrajpur Formation
always proved to be a highly economic zone, as far as the Precambrian’s of the
eastern Gujarat is concerned. In all, there were three mines initially started in this
area, viz., (i) Shivrajpur including Bhat, (ii) Bamankuva and (iii) Pani. Out of which,
Pani Mines is active and located at the NE part of the Champaner terrain, while rest
two are abandoned.
The rocks exposed at Stop 4 represent an abandoned mine section, popular in
the type locality as “Bhat Mines” located NE of Shivrajpur town or N of Navi-Bhat
village. This area is under jurisdiction of Gujarat Mineral Development Corporation
(GMDC) and can be traced through two entry points. 1. After Bhat Nala bridge
2 km interior towards Juni Bhat village; 2. another entry through vehicle can be
reached through Shivrajpur-Gujarat State Road Transport Company (GSRTC) Bus
stand crossing, a road leading towards Hatni-Mata or Sarasuva. Along this present
road an old gate heading towards Bamankuva mines routes towards the back side of
the Juni Bhat village (Fig. 20). The rocks are fine to medium grained, carbon black
in colour on account of Mn ores and display prominent box work (guides to ore for
Mn mineralization) (Fig. 21). These rocks show F 1 and F 2 plunging folds. F 1 are
second-order tight folds with interlimb angles <25°. These folds are well exposed
546 A. U. Joshi and M. A. Limaye

Fig. 16 Asymmetric ripple marks along the limbs of the fold, suggest the folding of primary
bedding (hammer as a marker having length = 32 cm)

along the valley regions having trend of the plunging fold axis towards N280°/20°
(Fig. 22). There also exists top-to-N280° shears in cross section, at low angle with
the axial planes of the F 1 folds. This shifted the F 1 hinges (Fig. 23). Well-developed
tensional gashes are good brittle shear sense indicator present along the southern
limbs of F 1 folds and the same climbs along the F 1 fold hinge portions (Fig. 24).
The F 2 folds are mainly restricted to the ridges or the highlands and show first-order
open type folds with a plunge of 21° towards N270–275° (Fig. 25).

8.3 Stop 5: Jaban Uranium Bearing


Meta-Conglomerate—(22° 24 13.04 N; 73° 38 45.89 E)

There are two distinct meta-conglomeratic horizons encountered within the Cham-
paner region, viz., 1. at the base of the Champaner Group located at east of the
Lambhia village and 2. middle horizon located near Jaban. The former is an oligomict
variety having impersistent band and found to be associated with quartzites, while
the latter is of polymict variety which has its wide spread occurrence and possesses
~800 m maximum thickness. The meta-greywackes is found to be intercalated
A Field Guide to the Champaner Region, Southern Aravalli … 547

Fig. 17 Fault gauge at the top left corner of the field photograph on account of brittle faulting (Dr.
A. Joshi as a marker having ht. 5.7 ft.)

with the meta-conglomerates occurring as a middle horizon near Jaban and also
as individual thin bands at the base of the Narukot quartzites.
The present stop can be reached from state highway number 05 near Jaban
village and represents a thick polymict meta-conglomerate horizon which exists
from Malabar in the south till Keshavpura in the north, in a folded sequence. The
rock show intercalations with meta-greywackes and is dark grey and massive having
rounded to sub-rounded clasts of quartzite, granite and gneiss (Fig. 26). The rock also
consists angular to sub-angular clasts of phyllite and calc-silicate rock. The overall
size range of the clasts varies between 2 and 40 cm, which adds the prefix to the rock
as bounder-cobble conglomerate. The entire sequence shows reduction in the matrix
content towards south. There are three distinct varieties of matrix, e.g. siliceous,
calcareous and micaceous. In addition, the matrix possesses uranium traces in ppm
on account of the nearest granitic intrusion present within the Champaner region.
In vicinity to the fault zones (east of Jaban), the rock represents stretching of clasts
along with tensional gashes (Fig. 27) with a top-to-N120° brittle slip.
The meta-greywacke crops out as an intercalated sequence with meta-
conglomerate near Jaban and also occurs as individual thin bands at the base of the
Narukot quartzites. The rock is dark grey, massive, matrix-dominated and consists of
angular to sub-rounded clasts of rock fragments of quartzite and phyllites (Figs. 28, 29
and 30). The occurrence of chlorite indicates recrystallization of matrix. At places,
548 A. U. Joshi and M. A. Limaye

Fig. 18 Slickensides on account of brittle faulting. Mineral steps indicate reverse dip slip brittle
fault (pen as a marker having length = 13.5 cm)

clasts are removed to leave behind the pock marks within the rock on account of
weathering.

8.4 Stop 6: Malabar Phyllite—(22° 23 41.56 N;


73° 39 31.94 E)

Phyllites cover major portions of the Champaner region. Based on the geological
map, it is evident that majority of the meta-pelites in the northern portion represent
Rajgarh Formation of the Champaner Group, while those occur within the central
region belong to the Shivrajpur and Narukot Formation. The phyllite exposed at this
stop is the most important to teach and record linear and planar structural fabrics.
The locality can be reached along the same state high way further towards the eastern
direction and just adjacent to the landmark “Parishram” of a hotel at the roadside
(Fig. 31). The rock is metallic grey in colour due to its carbon content, fine to medium
grained, having occasional presence of garnets along with ilmenite ores and clearly
displays compositional banding (Fig. 32). A foliation plane is evident in terms of
preferred orientation of platy minerals and lineations over foliation plane are found on
account of the manifestation of the bedding plane. The variety of lineations preserved
A Field Guide to the Champaner Region, Southern Aravalli … 549

Fig. 19 Quartzites display prominent goethite leaching along the weathered face (hammer as a
marker having length = 32 cm)

here are stripping lineations (a variety of intersection lineations) developed due to


the intersection between the S 0 and the S 1 planes (Fig. 33). The foliation plane dips
30° due south. The lineation pitches 45° towards N300°.

8.5 Stop 7: Narukot, Vertical Dipping Quartzite


and Schist—(22° 23 31.05 N; 73° 42 10.86 E)

Stop 7 is located somewhat interior within the reserved forest premises, but worth
visiting on account of their vertical dipping nature. The stop can be located at north of
the Narukot village at the periphery of the domal-shaped ridge (check dam as the land
mark, Fig. 34). This area represents maximum deformation on a regional scale located
within the Champaner region. Due to its overall lithological and structural setup, Joshi
(2019a) termed this band as “Narukot Hammer-Head Anticline” (NHHA). Dominant
rock type exposed along the periphery is massive quartzite representing Narukot
Formation, having whitish to dull red in colour, coarse grained nature with prominent
granulose structure. The bedding plane in the form of compositional banding is clear
suggesting vertical nature of S 0 within this rock (Fig. 35).
550 A. U. Joshi and M. A. Limaye

Fig. 20 Landmark for the Stop 4, sign board of Juni Bhat village

The core of the NHHA consists of graphite-rich schist and belongs to the older
Formations of the Champaner Group. This rock is exposed 250 m north of the check
dam, along the old working pits prepared by British geologists to excavate graphite
from this area (Fig. 36). Schists predominantly display parallel relationship of S 0 with
S 1 and also preserve secondary schistosity fabric S 2 over S 1 to develop L 2 lineations
(Fig. 37). Extrapolating the nature of fold on the basis of the evidences within the
pit, suggest the fold is of first generation, preserved at the core of NHHA. Based on
the dip of the axial plane, it is steeply inclined and represent F 1 folds developed on
account of D1 deformation of the Champaner Group.

9 Traverse 3—Objective and Overview

Traverse 3 consists of six stops with a prime focus to highlight the thermal effect
of granite over the rocks of the Champaner Group. Most of the outcrops display
the development of contact metamorphic minerals over their regional metamorphic
derivatives. As seen from the geological map of the Champaner Group, the eastern
margin is invaded by the Godhra granites; hence, most of the locations along this
traverse acts as markers to study the rocks present within the contact aureole. Some
of the minerals or even rocks are rare as they require appropriate geological and
A Field Guide to the Champaner Region, Southern Aravalli … 551

Fig. 21 Box work guides to ore for manganese (hammer as a marker having length = 32 cm)

geo-chemical setting to develop in such a short extent as well as beside each other.
The said traverse is very useful to understand the contact metamorphic setting, asso-
ciated evidences in the form of mineral development and structures. Other important
features are the preserved primary sedimentary structures, organogenic bands (stro-
matolitic bands) within the low-grade metamorphites and outcrop-scale superposed
folds.

9.1 Stop 8: Andalusite Hornfels at Wadek—(22° 24 19.44 N;


73° 42 21.02 E)

The present stop can be traced 5 km north of previous location and lies within the
vicinity of Wadek village. The stop is located at the outskirts of Wadek, ~1 km prior of
the formation and is exposed along the valley regions between two quartzitic ridges
on either side. The rock exhibits andalusite mineral of chaistolite variety (Fig. 38).
The typical cruciform and radiating crystals of chaistolite-andalusite have size range
of 5–7 cm in length versus 0.5–2 cm in breadth and show overprinting relationship
with the lineation bearing phyllites of the Wadek region (Joshi et al. 2018a, b). On
account of alkali metasomatism, the mineral shows the development of muscovite rim
around andalusite on a meso-scale. Muscovitization is prominent along the periphery
552 A. U. Joshi and M. A. Limaye

Fig. 22 Tight isoclinal plunging F 1 fold, marker standing in the broken F 1 hinge (Dr. A. Joshi as
a marker having ht. 5.7 ft.)

of the granitic intrusion and diminishes away from the pluton. The same is true for
the reduction in size of the chaistolite-andalusite. Adjacent to the andalusite hornfels,
the low lying phyllites contain spots of garnet porphyroblasts ranging from 0.2 to
0.5 cm in size.

9.2 Stop 9: Jothwad Skarn Rock—(22° 23 40.04 N;


73° 43 33.35 E)

The Stop 9 is located at the Jothwad village, NW of Jambughoda town. The present
spot is popular in the type locality as “Gol Dungri” (Fig. 39). The entire hill is an
exoskarn represented by a calc-silicate rock, on account of adjacent granitic intrusion.
This calc-silicate band is detached from the main Champaner domain and belongs
to the Khandia Formation of the Champaner Group. Due to its isolated nature, the
present area also show evidences of out-of-sequence deformation in the form of
superposed folds, which are unmatched with the deformation trends associated with
the Champaner Group (Joshi and Limaye 2018). The rock is pistachio green in colour,
coarse grained and consists of minerals such as winchite (Mn-rich hornblende),
A Field Guide to the Champaner Region, Southern Aravalli … 553

Fig. 23 Field photograph showing transposed F 1 hinge (observed along the central lower portion
of the photograph). The hinge line is curved

piemontite (Ca+Fe+Mn rich epidote), garnet (of Topazolite-Andradite, Grossularite


variety), epidote and wollastonite (Fig. 40).

9.3 Stop 10: Stratified Stromatolites


at Chalvad—(22° 26 29.25 N; 73° 41 29.85 E)

This is one of the best locations in terms of organogenic bands preserved within the
rocks of the Champaner Group. Sheet like sedimentary rocks that were originally
formed by the growth of cyanobacteria, a single celled photosynthesizing microbe
representing the earliest known form of life on the earth are preserved in calc-silicate
rocks of the region. The present location can be reached along the road that leads to
Bakrol near Chalvad village (Fig. 41). After ~400 m of the sign board of Chalvad, a
narrow single track road routes for small vehicles inside the village and reaches at the
foot hills of the calc-silicate rock. The rock is dark grey in colour and exhibits elephant
skin weathering. The rocks are characterized by interlayer stratified stromatolite
bands of carbon black colour (Fig. 42), which represent remnants of algal growth
during the Precambrian.
554 A. U. Joshi and M. A. Limaye

Fig. 24 Tensional gashes along the southern limb of F 1 (Mr. Ravi as a marker having ht. 5.8 ft.)

9.4 Stop 11: Actinolite and Tremolite Bearing


Calc-Silicate—(22° 27 29.95 N; 73° 41 40.02 E)

Stop 11 is located at Gandhra, after ~800 m long road cut section towards north of
the previous location. The dominant rock is calc-silicate of Khandia Formation and
is characterized by displaced magnesite veins with pearly lustre and low hardness
(Sharma and Golani 2013) (Fig. 43). Due to its vicinity to the granitic intrusion, the
rock shows unoriented actinolite and tremolite needles. The actinolite needles are
light green in colour, while the tremolite exhibits dark grey tone. Both are 2–5 cm long
and occur as stout prism-shaped crystals (Fig. 44). Apart from the evidences of contact
metamorphic structure, the rock also displays beautiful exposures of superimposed
folds: type-I interference pattern above type-III (Fig. 45).

9.5 Stop 12: Ripple Marks at Hatnimata—(22° 28 3.21 N;


73° 43 24.89 E)

The next stop is located at a popular tourist destination: “Hatnimata Waterfall”


(Fig. 46). The stop can be reached near the Sarasvua village. The dominant rocks
A Field Guide to the Champaner Region, Southern Aravalli … 555

Fig. 25 First-order F 2 open fold (Mr. Akash as a marker having ht. 5.9 ft.)

exposed at this location are vertical dipping quartzite of Narukot Formation with
impure dolomitic limestone of Khandia Formation. The quartzite is white in colour,
coarse grained and exhibit prominent asymmetric ripples at the upper surface
(Fig. 47). The other rock in contact with quartzite can be observed here is impure
dolomitic limestone, which is buff-brown in colour and occur intermittently with the
quartzitic bands (Fig. 48).

9.6 Stop 13: Arrow-Headed Folds


at Dharia—(22° 28 36.07 N; 73° 41 35.86 E)

Stop 13 represents magnificent arrow-headed folds preserved within the argillitic


band of the Champaner region (Fig. 49). The present location can be traced near
Dharia village, along a bypass road to Shivrajpur town. The rock is dark grey in
colour, fine grained and also preserves syn-sedimentation deformation signatures.
This rock belongs to the Rajghar Formation of the Champaner Group and is rich in
quartz and biotite. Folds preserved within this rock matches with the F 1 trend of the
Champaner structural setup.
556 A. U. Joshi and M. A. Limaye

Fig. 26 Jaban meta-conglomerate showing rounded to sub-rounded clasts of rock fragments


(hammer as a marker having length = 32 cm)

10 Traverse 4—Objective and Overview

The last traverse tracks down the oldest rock that belongs to the Champaner region. It
consists of two stops with an aim to demonstrate the so-called basement over which
the entire Champaner succession is believed to be resting. Another highlighting
feature of this traverse is an oligomitic meta-conglomerate band, which belong to
the Lambhia Formation of the Champaner Group. This conglomerate is considered
to the basal conglomerate, occurring at the base of the Lambhia Formation.
A Field Guide to the Champaner Region, Southern Aravalli … 557

Fig. 27 Tensional Gashes preserved within the meta-conglomerate. Top-to-N120o slip (scale as a
marker having length = 15 cm)

10.1 Stop 14: Oligomitic Meta-Conglomerate


at Mota-Raska—(22° 21 20.98 N; 73° 41 4.67 E)

The present stop represents basal oligomitic meta-conglomeritic band that belongs
to the Lambhia Formation of the Champaner Group. The present location can be
reached through a road connecting the Jhambughoda-Bodeli state highway, ~5 km
interior to the Mota-Raska village. The landmark for the present stop can be consid-
ered as Mota-Raska Primary School. The outcrops are exposed 200 m north of the
present landmark. The rock type is oligomict meta-conglomerate with dark grey
colour siliceous matrix; the clasts present within them are off-white and composed
of quartzite with 5–8 cm diameter. The rock fragments are rounded to sub-rounded,
558 A. U. Joshi and M. A. Limaye

Fig. 28 Intercalations between meta-conglomerate and meta-greywackes (hammer as a marker


having length = 32 cm)

pebble size and were transported from a longer distance (Fig. 50). In addition, the
conglomerate also preserves N–S trending warps indicating D3 deformation of the
Champaner Group (Joshi 2019a).

10.2 Stop 15: Intermixed Granite-Gneiss


at Jhand—(22° 22 5.87 N; 73° 38 43.65 E)

The last stop is one of the most important outcrops in terms of the basement associated
with the Champaner Group. The location is situated at the “Jhand Hanuman Temple”.
The stop can be reached 4 km northward from the previous location situated at
Mota-Raska (Stop 14) along the same road and accessible for all sized vehicles. The
intermixed granite-gneissic rock is massive, coarse grained, off-white in colour, with
interlayer discontinuous dark minerals (Figs. 51 and 52). The lighter tone minerals
are quartz, plagioclase feldspar, whereas the darker are mainly biotite, which shows
feeble orientation due to its intermixing character with the granite melt. The rock is
devoid of any high-grade mineral and is with the Champaner metasediments (Joshi
and Limaye 2014).
A Field Guide to the Champaner Region, Southern Aravalli … 559

Fig. 29 Fragment of quartzite within meta-greywackes (hammer as a marker having length =


32 cm)

The nearby associated rock is phyllite, which exhibits dark green in colour and
consists of contact metamorphic mineral as cordierite of Indialite variety. The rock
has undergone the process called as pinitization in which the cordierite has altered to
pinite mica along its rims. The rock also consists of minerals such as epidote, quartz,
tourmaline and pyrite (Fig. 53).
560 A. U. Joshi and M. A. Limaye

Fig. 30 Fragment of phyllite within meta-greywackes (pen as a marker having length = 13.5 cm)

11 Concluding Remarks

15 Stops covered under 4 traverses provide a broad idea about the potential of the
Champaner to train geoscientists for fieldwork. The present region did not receive
substantial attention by geologists for ~ five decades. Recent work carried out by
Joshi et al. (2018a, b), Joshi and Limaye (2018), Joshi (2019a, b), Joshi and Limaye
(2020) and Mukherjee et al. (2020) have provided detail insight in terms of structure
and metamorphism. We hope that this chapter will generate an interest among the
students, faculty and researchers across the globe to study in detail the Precambrian
geology of eastern Gujarat. Also, we feel that there is utmost scope for a field enthu-
siast to find more evidences to substantiate the views expressed in the description
for each location or to contradict them!
A Field Guide to the Champaner Region, Southern Aravalli … 561

Fig. 31 Landmark for the Stop 6, photograph of Hotel Parishram located near Jaban village

Fig. 32 Field photograph showing compositional banding preserved within phyllite (10 rupee coin
as a marker having dia = 2.6 cm)
562 A. U. Joshi and M. A. Limaye

Fig. 33 Lineations bearing phyllite located at Stop 6 (camera lens cap as a marker having dia =
6 cm)

Fig. 34 Landmark for the Stop 7, field photograph of check dam at the periphery of the Narukot
Dome
A Field Guide to the Champaner Region, Southern Aravalli … 563

Fig. 35 Quartzite exhibits vertical dipping S 0 as compositional banding (hammer as a marker


having length = 32 cm)
564 A. U. Joshi and M. A. Limaye

Fig. 36 Landmark to locate the graphite-rich schist, old working pits for graphite excavation (Mr.
Sandeep as a marker having ht. 5.3 ft.)

Fig. 37 L 2 lineations within graphite-bearing schist. A variety of intersection lineation developed


on account of S 1 and S 2 (pen as a marker having length = 13.5 cm)
A Field Guide to the Champaner Region, Southern Aravalli … 565

Fig. 38 Chaistolite development over phyllite (exposed pen length = 4 cm)


566 A. U. Joshi and M. A. Limaye

Fig. 39 Panoramic view of Gol Dungri near the Jothwad village

Fig. 40 Skarn rock having coarse grained minerals (hammer as a marker having length = 32 cm)
A Field Guide to the Champaner Region, Southern Aravalli … 567

Fig. 41 Landmark for the Stop 10, photograph of sign board of Chalvad village along the main
Bakrol road
568 A. U. Joshi and M. A. Limaye

Fig. 42 Stratified stromatolite preserved within the calc-silicate rock (exposed pen length = 6 cm)

Fig. 43 Displaced magnesite vein within the calc-silicate rock (exposed pen length = 8 cm)
A Field Guide to the Champaner Region, Southern Aravalli … 569

Fig. 44 Unoriented actinolite and tremolite needles within the calc-silicate due to its vicinity to
the granitic intrusion (pen as a marker having length = 13.5 cm)
570 A. U. Joshi and M. A. Limaye

Fig. 45 Outcrop-scale superimposed folds preserved within calc-silicate rock (Type-I over Type-
III) (hammer as a marker having length = 32 cm)

Fig. 46 Panoramic view of vertical dipping quartzites and impure dolomitic limestones at
Hatnimata, near Sarasvua
A Field Guide to the Champaner Region, Southern Aravalli … 571

Fig. 47 Vertically dipping quartzite showing asymmetric ripple marks on upper surface (Mr.
Sandeep as a marker having ht. 5.3 ft.)

Fig. 48 Contact between quartzite and impure dolomitic limestone at Stop 12 (Mr. Sandeep as a
marker having ht. 5.3 ft.)
572 A. U. Joshi and M. A. Limaye

Fig. 49 Arrow-headed folds within argillite near Dharia village (pen as a marker having length =
13.5 cm)

Fig. 50 Oligomict meta-conglomerate showing N-S warping near the Mota-Raska Primary School
(pen as a marker having length = 13.5 cm)
A Field Guide to the Champaner Region, Southern Aravalli … 573

Fig. 51 Massive intermixed granite-gneisses near Jhand (Mr. Dhrumil as a marker having ht. 5.6
ft.)

Fig. 52 Feeble orientation of biotite grains interlayer with the light coloured minerals in granite-
gneiss (hammer as a marker having length = 32 cm)
574 A. U. Joshi and M. A. Limaye

Fig. 53 Field outcrop of cordierite hornfels adjacent to the granite-gneiss exposures (hammer as a
marker having length = 32 cm)

Acknowledgements The authors are grateful to the In-charge Head, Department of Geology and
Dean, FoS, MSUB, Prof. Haribhai Kataria for providing necessary permissions to carry out geolog-
ical fieldworks. First, author is thankful to the Former Head, Department of Geology, MSUB, Prof.
L. S. Chamyal, for continuous encouragement and support during his Ph.D. work. On account of
which, it has been possible to formulate in the form of chapter. The authors are grateful to Prof.
Subhash J. Desai (Retd.) and Prof. B. S. Deota for providing valuable inputs about the terrain. This
work would have not been possible without all those dissertation students who actively partici-
pated during countless field visits. AUJ and MAL are thankful to the Chief Conservator of Forest,
Government of Gujarat to give necessary permission to work within the limits of the sanctuary area.
Authors are obliged to Soumyajit Mukherjee (IIT Bombay) for reviewing this article. Thanks to
Marion Schneider, Annett Buettener, Boopalan Renu, Alexis Vizcaino, Doerthe Mennecke-Buehler
and the proofreading team (Springer). Dutta and Mukherjee (2021) encapsulate this work.

References

Beer, E. J. (1919). Notes on rocks from Pavagad to Dohad. Min. Geol. Inst. India Trans, 13, 73–127.
Blanford, W. T. (1869). On the geology of the Taptee and lower Nerbudda Valley and some adjoining
districts. Mem Geol. Surv. India, 6(3), 163–207.
Dutta, D., Mukherjee, S. (2021). Introduction to Structural Geology and Tectonics Field Guide-
book—Volume 1. In S. Mukherjee (Ed.), Structural Geology and Tectonics Field Guidebook—
Volume 1. Switzerland AG: Springer. Cham. pp. xi-xvi. ISBN: 978-3-030-60142-3.
A Field Guide to the Champaner Region, Southern Aravalli … 575

Fermor, L. L. (1909). The manganese ore deposits of India. Memoirs of the Geological Survey of
India, 37.
Gopalan, K., Trivedi, J. R., Merh, S. S., Patel, P. P., & Patel, S. G. (1979) Rb-Sr age of Godhra and
related granites, Gujarat (india). Proceedings of the Indian Academy of Sciences-Section A. Part
2, Earth and Planetary Sciences, 88(1), 7–17.
Gupta, S. N., Mathur, R. K., and Arora, Y. K. (1992). Lithostratigraphy of proterozoic rocks of
Rajasthan and Gujarat - areview. Rec. Geol. Surv. India. v., 115, 63–85.
Gupta, S. N., Arora, Y. K., Mathur, R. K., Iqbaluddin, Prasad, B., Sahai, T. N., & Sharma, S. B.
(1997). The Precambrian geology of the Aravalli region, Southern Rajasthan and NE Gujarat.
Memoirs of the Geological Survey of India,123, 1–262.
Hari, K. R., Furuyama, K., & Shabeer, K. P. (1999). The Pavagadh Hill, a unique outcrop of the
Deccan trap, Gujarat, India. Gondwana Research, 2, 676–679.
Jambusaria, B. B. (1970). Geology of the area around Shivrajpur (dist. Panchmahals, Gujarat)
with special reference to the stratigraphy, structure and metamorphism. Unpublished Ph.D, M.S.
University of Baroda. 282 p.
Joshi, A. U. (2019a). Fold interference patterns in Meso-Proterozoic Champaner fold belt (CFB)
Gujarat, western India. Journal of Earth System Science. https://doi.org/10.1007/s12040-019-
1075-z.
Joshi, A. U. (2019b). Structural evolution of Precambrian rocks of Champaner Group, Gujarat,
Western India. Unpublished Ph.D. Thesis, The Maharaja Sayajirao University of Baroda, pp. 1–
190.
Joshi, A. U., & Limaye, M. A. (2014). Evidence of syn-deformational granitoid emplacement within
Champaner Group, Gujarat. Journal of Maharaja Sayajirao University of Baroda, 49, 45–54.
Joshi, A. U., & Limaye, M. A. (2018). Rootless calc-silicate folds in granite: An implication towards
syn to post plutonic emplacement. Journal of Earth System Science, 127(5). ID67.
Joshi, A. U., & Limaye, M. A. (2020). Anisotropy of Magnetic Susceptibility (AMS) studies on
quartzites of Champaner Group, upper Aravallis: An implication to decode regional tectonics of
Southern Aravalli Mountain Belt (SAMB), Gujarat, Western India. In T. K. Biswal, S. K. Ray,
& B. Grasemann (Series Eds.), Society of earth scientist series, under the heading of structural
geometry of mobile belts of the Indian subcontinent (Special Volume, pp. 199–211). ISSN: 2194-
9204. ISBN: 978-3-030-40592-2. https://doi.org/10.1007/978-3-030-40593-9.
Joshi, A. U., Limaye, M. A., & Deota, B. S. (2018a) POTM-cover page photo entitled, Chiastolite-
Andalusite variety encompassing alkali metasomatism within the Champaner Group, Gujarat,
western India. The Geological Society of India, 91(1).
Joshi, A. U., Sant, D. A., Parvez, I. A., Rangarajan, G., Limaye, M. A., & Mukherjee, S., et al.
(2018b). Sub-surface profiling of granite pluton using microtremor method: Southern Aravalli,
Gujarat, India. International Journal of Earth Sciences, 107, 191–201.
Mukherjee, S., Bose, N., Ghosh, R., Dutta, D., Misra, A., & Kumar, M., et al. (2020). Structural
geological Atlas (623p). New York: Springer.
Roy, B. C. (1958). The manganese-ore deposits of Panch Mahals and Baroda districts, Bombay
State. Geological Survey of India Bull. Ser. A, 45(15).
Sharma, B. B., & Golani, P. R. (2013). Magnesite in the Palaeoproterozoic metasedimentary
carbonate sequence of Aravalli Supergroup in Gujarat, western India. Current Science, 104(8),
1013–1015.
Sheth, H. C., & Melluso, L. (2008). The Mount Pavagadh volcanic suite, Deccan Traps: Geochemical
Stratigraphy and magmatic evolution. Jour of Asian Earth Science, 32, 5–21.
Srikarni, C., & Das, S. (1996). Stratigraphy and sedimentation history of Champaner Group, Gujarat.
Journal of the Indian Association of Sedimentologists, 15, 93–108.
Importance of Fracturing in Uranium
Mineralization in Gulcheru Quartzite
Host: A Case from Ambakapalle Area,
Cuddapah Basin, Andhra Pradesh, India

Sukanta Goswami, P. K. Upadhyay, and V. Natarajan

Abstract Fracture and fault zones can tap the uranium-bearing fluid and help to
migrate into the structural trap where the hydrothermal fluids cool and precipitate the
uranium mineral. Such a structural setting gives sufficient time to react the fluid with
receptive country rocks. Fracture zones have played a major role in the mineralizing
event in Ambakapalle area, where uranium anomalies are located along fracture
zone strike for about 1.4 km intermittently. The localization of uranium is influenced
by E-W fault zone which have provided the pathway for uprising hydrothermal
fluids with decreasing heat and pressure. In this context, the structural attributes
are studied by mapping the fault zone and by projecting the subsurface geometry.
Intersections of fractures are observed as better locations of uranium anomalies
and hence given attention for prospecting the uranium mineralization. The block
movement along fault is characterized as diagonal slip normal fault with a net slip
direction parallel to the trace of bedding on the fault plane, which is also called trace
slip fault. Therefore, from plan view, there is no apparent displacement of lithocontact
across the fault. This trace slip fault with respect to the strata is interpreted as the most
important structure from uranium mineralization point of view. The alteration index
values for the host rocks are determined for different types of chemical as well as
hydrothermal alteration, and their bearing with uranium mineralization is interpreted.
High values of chlorite-carbonate-pyrite index [CCPI = (MgO + FeO) × 100/(MgO
+ FeO + Na2 O + K2 O)] indicate intense alteration of Fe- or Mg-rich minerals,
such as chlorite. Low to medium ranges of Chemical Index of Alteration [CIA =
Al2 O3 × 100/(Al2 O3 + CaO + K2 O + Na2 O)] values indicate variable intensity
of chemical weathering due to removal of labile cations (Ca2+ , Na+ , K+ ) relative to
stable residual constituent (Al3+ , Ti4+ ). The correlation among different alteration
indices and uranium is noteworthy. It is interpreted that the initial hydrothermal
alteration along fault zone is masked by later low-temperature alterations. Therefore,
the surficial manifestation of uranium anomalies is disturbed, and intermittent patches
of uraniferous anomalies are seen along the trend of fault.

S. Goswami (B) · P. K. Upadhyay · V. Natarajan


Department of Atomic Energy, Atomic Minerals Directorate for Exploration and Research,
Bangalore 560072, India
e-mail: sukantagoswami.amd@gov.in

© Springer Nature Switzerland AG 2021 577


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_19
578 S. Goswami et al.

1 Introduction

The Proterozoic rifting associated Cuddapah Basin is an important uranium province


(Parihar and Rao 2012; Goswami 2014; Goswami et al. 2017, 2018, 2019, 2020).
Such basins show few characteristic features favourable for uranium mineralization.
Globally, Athabasca Basin, Canada, is one of such well-recognized basins where
extensive uranium mineralization is observed. Fertile uraniferous basement granite
with fractures and dyke swarms can be considered as a favourable provenance set-
up. The deformed and faulted unconformity contact between Archaean basement
granitoids and Proterozoic Gulcheru quartzite is expected to be a possible zone of
interest. Detailed examination of the 100°–280° fault zone, located at around 2 km
WNW of Ambakapalle village (Fig. 1a, b), is the main aim of the present paper.
Significant uranium mineralization is located along this fault zone. Reconnaissance
survey carried out by AMD in SW part of Cuddapah Basin during the year of 2001–
02 has lead to the identification of uranium anomaly with up to 0.13% U3 O8 over
a strike length of ~500 m and a thickness up to 20 m along a scarp fractured face
in Gulcheru quartzite. Subsequently, during 2002–03, detailed geological mapping
(1:2000), shielded probe logging of the scarp face along vertical lines was carried
out around this anomaly. During 2014–15, we found that only the northern wall
of the vertical scarp face of the fault zone exhibits mineralization but no manifes-
tations of uranium in southern wall. General bedding plane strikes N400 W with a
dip of 100 towards NE with steepening near fault plane. The mineralized quartzite
is very hard, massive and compact. A similar geological setting is also found near
Gandi area located at about 50 km SE of Ambakapalle, where the southern face of
E-W fracture exhibits mineralization (Umamaheswar et al. 2001; Goswami et al.
2016). Undoubtedly, the fault zone has played important role for getting this miner-
alization as the intrinsic/general uranium content of Gulcheru quartzite is usually
low. However, enrichment up to 15 times more radioactivity along the fracture trend
is significant. Therefore, the objective of this re-examination was to find out the
role of fracturing/faulting and to understand the subsurface occurrence along with
further tracing of continuity of mineralization and to fix target for fracture-controlled
uranium mineralization.
A new detail map is prepared in this context for the first time, and systematic
structural analysis with hypothetical model for subsurface exploration is presented.

2 A Backdrop of Findings

Natural uranium content in quartzite (0.5 ppm) and greywacke (3 ppm) is very less
(Taylor 1965). However, Gulcheru quartzite in Ambakapalle area shows anomalous
uranium content along the vicinity of fault trend. Initially in 2001–02, ground radio-
metric survey works have recorded significant radioactivity of the order of 5–15
times of general background counts (5 to 15xbgc) associated with grey quartzite
Importance of Fracturing in Uranium Mineralization in Gulcheru Quartzite … 579

Fig. 1 a Geological map of the Cuddapah Basin showing the study area (Modified after Goswami
2014). b Google image showing the study area west of Ambakapalle village (TS no. 57J/3/NW)
580 S. Goswami et al.

Fig. 1 (continued)

of Gulcheru formation. Radioactivity could be traced discontinuously for a strike


length of about 500 m and a thickness of about 20 m along scarp face. Grab samples,
collected from this anomaly, have yielded up to 0.130% U3 O8 (Chemical). Shielded
probe logging was carried out along five lines at an interval of 100 m, over the
anomaly. Low order but correlatable bands of mineralization were recorded in all
the five lines. Based on this preliminary encouraging results, this area was taken up
for detailed radiometric checking and geological mapping subsequently in 2002–
03. Analysis of selected samples recorded 0.01–0.031% U3 O8 and <0.01% ThO2 .
Petrographic observation revealed that the anomalies were due to secondary uranium
minerals occurring as surficial encrustations, fracture filling and lesser irregular
patches. Later during the year 2014–2015, the extension of fault zone towards east
was radiometrically checked in detail along with mapping and sampling. Five (5)
new anomalies were recorded in the eastern extension of the fault zone in footwall
block. The anomalies in the faulted Gulcheru quartzite were ranging from 0.08 to
0.9 mR/h (8 × to 90 × bgc) with dimension range from 5 m × 5 m to 2 m ×
2 m. The bands were continuing as lenses for over 10-80 m with barren zone in
between for 30 and 100 m. Therefore, the anomalous zone is intermittently contin-
uing for about 1.4 km in total (Fig. 2a). The physical assay results of representative
grab samples (n = 28) are shown in Table 1. Apart from that, two private borewells
(T4, T1 in map in Fig. 2) were investigated by gamma ray logging where probably
the effects of fracture-controlled mineralization were manifested in terms of higher
than expected values for the corresponding depth (up to 0.016% eU3 O8 at 94.3 to
94.4 m depth). Although the borehole is located away from the fracture zone but
due to younger subsidiary pathway, which have not shown any surficial manifes-
tation, possibly some portions have been moved and given indication. Therefore,
the surrounding younger faults also must be checked by subsurface exploration for
cross-checking the idea. Hypothetical section A-B (Fig. 2b) indicates the Gulcheru
quartzite can be expected at around 100 m depth from surface. The entire borewell
T4 has been logged for 137 m depth, and chips of Gulcheru quartzite indicate frac-
tured nature of quartzite. Although the T4 borewell is away from main E-W fault
Importance of Fracturing in Uranium Mineralization in Gulcheru Quartzite … 581

zone, the influence of remobilization along sympathetic and subsidiary fracture can
be expected.

3 Geology

The intracratonic rift-related Cuddapah Basin comprises Cuddapah Supergroup


followed by Kurnool Group of sedimentary rocks. The Cuddapah Supergroup is
developed in Papaghni, Nallamalai and Srisailam sub-basins, whereas Kurnool
Group occurs in the Kurnool and Palnad sub-basin. The lowermost arenaceous
Gulcheru Formation of the Papaghni sub-basin is our target host rock in this context.
The stratigraphy, structure, tectonics and basin evolution are extensively described
in Nagaraja Rao et al. (1987) and also in Ramakrishnan and Vaidyanadhan (2008)
and Goswami et al. (2020). The Eastern Ghat orogeny (~1570 Ma) which is related
to the formation of Eastern Ghats Mobile Belt (EGMB) in the east of the Cuddapah
Basin has possibly provided the crescent shape to the basin and that time is regarded
as a metallogenic epoch. Near the western margin of basin tectonic imprints of
the EGMB is manifested as E-W faults. The time stratigraphic relationships with
reference to the geology of this uranium mineralization are matching with lower to
middle Proterozoic sub-unconformity epimetamorphic type (Dahlkamp 1993). The
local geology of the area was studied during detailed geological mapping of the
fracture zone which is located at around 2 km WNW of Ambakapalle village comes
under 1:25,000 scale toposheet no. 57 J/3/NW. Lithologically Gulcheru formation

Fig. 2 a Geological map of the study area (this work). b Sections along X–X  , Y –Y  and A-B
582 S. Goswami et al.

Fig. 2 (continued)

dominated by quartzite with minor shale and lowermost part of Vempalle Dolostone
occur in normal sequence. The lithounits show generalized trend NW–SE with a
dip amount of 10°–15° towards NE with localized variations at places near fault
and intense fractured zones. The map (Fig. 2a) has been made in detail with the
help of tape and compass and GPS readings, which were cross-checked for all the
locations. All the lithocontacts are traced along with fracture trends and plotted in
the map. Gulcheru quartzite followed by shale and then Vempalle dolostone occur
in normal sequence, and at places shale and dolostone are observed as cap above
quartzite at higher elevation hillocks. Several nala has been developed due to intense
fracturing and faulting and along those nala huge amount of broken brecciated pieces
of quartzite boulders are transported and spreaded over shale.
Importance of Fracturing in Uranium Mineralization in Gulcheru Quartzite … 583

Table 1 Physical assay result of the grab samples


Sample No. UTM UTM Rock eU3 O8 U3 O8 Ra eq ThO2
latitude (N) longitude types
(m) (E) (m)
ABP-1 1,595,887 187,310 Gulcheru 0.011% 0.016% 0.011% <5 ppm
ABP-2 1,595,887 187,290 quartzite 0.006% 0.017% 0.006% <5 ppm
ABP-3 1,595,887 187,260 0.037% 0.022% 0.037% <5 ppm
ABP-4 1,595,887 187,240 0.047% 0.033% 0.047% <5 ppm
ABP-5 1,595,887 187,220 0.013% 0.005% 0.013% <5 ppm
ABP-6 1,595,887 187,200 0.009% 0.009% 0.009% <5 ppm
ABP-7 1,595,887 187,010 0.007% 0.009% 0.007% <5 ppm
ABP-8 1,595,887 186,900 0.05% 0.048% 0.05% <5 ppm
A1 1,595,766 187,635 7 ppm <20 ppm 5 ppm <5 ppm
A2 1,595,784 187,620 6 ppm <20 ppm 4 ppm 6 ppm
A3 1,595,776 187,612 7 ppm <20 ppm 6 ppm 5 ppm
A4 1,595,766 187,607 6 ppm <20 ppm 2 ppm 6 ppm
A5 1,595,776 187,595 2 ppm <20 ppm 3 ppm <5 ppm
A6 1,595,789 187,587 2 ppm <20 ppm 2 ppm <5 ppm
A7 1,595,776 187,677 5 ppm <20 ppm 2 ppm 5 ppm
A8 1,595,766 187,667 6 ppm <20 ppm 3 ppm 5 ppm
A9 1,595,762 187,554 3 ppm <20 ppm 2 ppm 4 ppm
A 10 1,595,776 187,527 21 ppm <20 ppm 3 ppm 21 ppm
A 11 1,595,776 187,507 17 ppm <20 ppm 17 ppm <5 ppm
A 12 1,595,786 187,497 6 ppm <20 ppm 6 ppm <5 ppm
A 13 1,595,781 187,490 7 ppm <20 ppm 6 ppm <5 ppm
A 14 1,595,776 187,477 <2 ppm <20 ppm <2 ppm <5 ppm
A 15 1,595,764 187,462 <2 ppm <20 ppm <2 ppm <5 ppm
ABP-9 1,595,776 187,627 0.100% 0.032% 0.110 <5 ppm
ABP-10 1,595,762 187,544 <0.010% – – <5 ppm
ABP-11 1,595,772 187,522 <0.010% – – <5 ppm
ABP-12 1,595,784 187,518 0.032% 0.037% 0.032% <5 ppm
ABP-13 1,595,771 187,505 0.003% 0.003% 0.003% <5 ppm

4 Structural Analysis

The major fault zone trending almost E-W occurs as 100 m deep nala with vertical
scarp faces in the west, and near the eastern extreme the fault plane is more promi-
nently showing ~100° strike with a dip of 70° towards south. The step like slicken-
side lineations (Doblas 1998) are identified on fault plane as evidence of diagonal
sip movement towards east (Fig. 3). Towards east the scarp face gradually becomes
584 S. Goswami et al.

Fig. 3 Step like slickenside lineation shows 10° = > 100° movement. The direction of motion
in the brittle sheared zone is in shear plane/fault plane. This movement must be parallel to fault
striations which have been found on 100°–280° fault plane with 10°–15° pitch due 100°

gentler and showing younger shale in the southern block to be lying topographi-
cally above older quartzite in the northern block (Fig. 4a, b). In the western part of
the fault, the vertical scarp faces are developed due to subsequent sympathetic E-W
vertical joint plane sets which had been removed as layers to form wide zone. The
fault can act as conduit for hydrothermal fluids to circulate and mix among surfi-
cial and juvenile waters. They are hot metal-rich fluids eventually concentrated at
suitable portions of fault plane to become cool with uprising movement. Therefore,
fault plane can give clue towards solution movement where a specific volume of
host rock contains anomalous concentration of element. Presently, faulted Gulcheru
quartzite has provided loci for supersaturation and precipitation of uranium from
the fluid. For a normal fault, the hydrothermal fluids can move over the footwall
block due to gravity, and hence, it can percolate through wherever possible along

Fig. 4 Normal faulting with southern downthrown hang wall block identified from occurrence of
younger Gulcheru shale in the downthrown side and older Gulcheru quartzite in the upthrown side
at same RL. Dip of the fault plane measured in the field as 70° due south
Importance of Fracturing in Uranium Mineralization in Gulcheru Quartzite … 585

Fig. 5 Diagrammatic representation of progressive changes in the fault zone and associated
alteration and uranium mineralization in Ambakapalle fault zone

permeable zones. But in case of hanging wall block solution can just show mineral-
ization suficially, and it cannot pass through rockmasses against gravity. Therefore,
replacement or alteration or any cavity filling mineralization in all probability is
more in footwall block than hanging wall. The normal fault of Ambakapalle area has
been affected by several sympathetic fractures which have removed layers of rock
parallel to fault, and thus the fault plane is eroded and widen to form fault zone.
The schematic representation (Fig. 5) of the entire mechanism is demonstrated with
sketches. It is known that the maximum movement of any fault is taken place in the
middle, and gradually it reduces near tip, so the vertical scarp indicates central part of
the fault. The nala floor is covered mainly by quartzite boulders (99%) and few basic
rock pieces (1%) indicating probable basic dyke along nala in subsurface similar as
Madyalabodu–Gandi–Rachakuntapalle areas where drilled cores had shown pres-
ence of basic dyke without any surficial expressions. However, absence of any in situ
basic rock exposure may be due to their higher rate of erosion and weathering. Struc-
tural analysis has been carried out around the fault zone based on fracture trends and
their interrelationships, stereonet analysis and satellite imagery data. Although the
formations are affected by faulting, the contact between Gulcheru shale (GS) and
Gulcheru quartzite (GQ) is not showing apparent displacement. However, a NW–SE
trending older fracture is displaced against E-W fault (Fig. 6). In fact, the older NW–
SE fracture is displaced with a strike separation AB. Therefore, all the attitudes of
planes, viz. fault (100°, 65° due 190°), older fracture (120°, vertical) and Gulcheru
quartzite/Gulcheru shale contact (135°, 15° due 45°) are plotted in stereonet (Fig. 7).
The pitch/rake of older NW fracture and GS/GQ contact on the E-W fault plane are
found as 10° = >100° and 40° = >100°, respectively. After this, the intersection
points of fracture and lithocontact are plotted on E-W fault trace as A, B and C, where
A and B are fracture intersections in hanging wall and foot wall, respectively, of E-W
fault. Since, GS/GQ does not show any displacement, there is a single intersection
586 S. Goswami et al.

Fig. 6 Satellite image shows the 100°–280° fault displaced older NW–SE fracture zone. However,
the contact between Gulcheru quartzite (GQ) and Gulcheru shale (GS) is not showing apparent
displacement

Fig. 7 Net slip calculation


with the help stereonet to
find rake of trace of bedding
and earlier fracture on fault
plane and intersection of
bedding and fracture on
hanging wall and footwall
block of the fault.
Radioactivity up to 0.9 mR/h
has been recorded in
footwall block (modified
after Goswami et al., 2019)

C. The corresponding pitch values are drawn from A, B and C. Now, the intersection
of the two traces (i.e. GS/GQ and NW fracture) on hanging wall as well as foot
wall of E-W fault is demarcated as C and C , respectively. This C C is the net slip
which is 89 m, and pitch of this slip is measured as 10° = >100° which is same
as the pitch ob GS/GQ on E-W fault. This typical phenomenon is called trace slip
fault, and therefore, no apparent displacement of lithocontact is seen. Henceforth,
Importance of Fracturing in Uranium Mineralization in Gulcheru Quartzite … 587

Fig. 8 Sense of shear has been understood in the field from the feebly developed oblique foliation
and fractures cut across the main fault zone developed due to shear offset like in deck of cards slip
between individual cards. S-C shear band also indicates sinistral sense of shear

the rake of the traces of bedding planes on the fault plane is parallel to the net slip
of fault movement (Fig. 7). Broadly, a sinistral sense is reflected from oblique folia-
tions as additional indicators (Fig. 8). In studying these deformed rocks, an attempt
is made to determine the strain path because stress and deformation are important
with respect to their resource potential. It is a fact that the rupture deformation can
create visible fractures in response to stress, resulting in the loss of cohesion of rock
particles. The shear stress is confined to shallow depths to show dominantly brittle
phenomenon. The diagonal slip normal fault is related to shearing stress evidenced
from some brittle shear fabrics (Fig. 8). Quartz grains are deformed mainly by brittle
processes, thus the temperature must be <300 °C. Fractures are developed along both
previously intact rock and pre-existing weak bedding surfaces. Systematic fracture
sets display orientations suggesting a common origin in response to directed stress.
Systematically collected fracture data shows a dominant E-W trend and different
irregular non-systematic trend (Fig. 9a, b) leads to brecciation. At places conjugate
shear joint sets are also found to develop with E-W and WNW-ESE trends showing
acute bisector direction along 120° as maximum compression (Fig. 9c). Two major
extension joint sets are found: Earlier generated joint trend ranges from N–S to
20°–200° sub-vertical related to N-S compression. The later set has crosscut and the
earlier set along 100°–280° to 130°–310° with vertical plane (Fig. 9d).
588 S. Goswami et al.

Fig. 9 Different types of fractures and associated brecciation

5 Petrography and Geochemistry

Under microscope, Gulcheru quartzite shows fine- to medium-grained arenitic nature


with about 95% quartz and few fine chert clasts also present as minor phase. Sub-
rounded to sub-angular quartz grains show textural and mineralogical maturity with
some effects of diagenesis and later phase of structural deformations. The sutured
grain contacts with an average size from 0.4 to 0.5 mm are notable. Mild strain effects
can be seen in the form of medium undulose extinction. The interstitial spaces of the
quartz clasts are often filled by secondary silica overgrowth. Biotite is also present
in trace amount in the matrix along with fine muscovite flakes. Heavy minerals, viz.
tourmaline and zircon occur as inclusions within the quartz clasts. Iron oxides and
hydroxides in the form of limonite and goethite are also present in the interstitial
spaces as well as disseminated grains in matrix. Anatase is another mineral found
as disseminated grains in interstitial spaces. Cellulose nitrate (CN) film studies are
carried out for alpha radiography on the thin sections. The alpha tracks have indi-
cated adsorbed uranium on iron minerals. The polished sample surface was used
for chromogram test, which showed positive result, and brown-coloured spots were
noted on photo paper (Fig. 10a, b). Such brown colour is indicative of leachable
uranium. Mostly sparse to low density alpha tracks are developed after 4 days of
exposure of CN film on the sample surface. The source of such radioactivity is mainly
Importance of Fracturing in Uranium Mineralization in Gulcheru Quartzite … 589

Fig. 10 a Adsorbed uranium on goethite giving sparse alpha (α) tracks 10×, TL, 1 N. b Positive
chromogram test indicates presence of leachable uranium with characteristic brownish colour

found as adsorbed uranium on anatase and goethite. These minerals are present as
disseminated grains in the matrix part of the rock (Fig. 11a, b).
Total thirteen (13) samples of Gulcheru quartzite from fault zone were studied by
WDXRF analysis (Table 2). The result interprets about a geochemical association of
some trace elements like Sr, Ba and Cu which connote hydrothermal alteration since
these elements are linearly varying with uranium. Some elements also indirectly give
clue towards presence of late-magmatic hydrothermal fluid evidenced by depletion
of compatible elements like Ni, Cr, etc. (Fig. 12). The alteration indices (Table 3)
were calculated to analyse the hydrothermal alteration intensities. The alteration
indices imply a combination of chemical and hydrothermal alteration, and effects of
former are masked by the later. This is due to dominance of chemical weathering and
alteration in surficial environment subsequent to hydrothermal alteration. However,
the bivariant plots of hydrothermal indices against uranium exhibit linear positive
trend unlike the chemical index which is non-systematic in nature (Fig. 13). The

Fig. 11 a Adsorbed uranium on goethite with inset of sparse alpha tracks 20×, TL, 1 N. b Adsorbed
uranium on anatase with inset of sparse alpha tracks 20×, TL, 1 N
Table 2 Major element oxides (%) and selective trace element concentrations (ppm) of Gulcheru quartzite samples with U3 O8 (%) data
590

S.No ABP-1 ABP-2 ABP-3 ABP-4 ABP-5 ABP-6 ABP-7 ABP-8 ABP-9 ABP-10 ABP-11 ABP-12 ABP-13
SiO2 94.56 94.63 94.4 95.11 93.41 93.64 94.41 94.4 89.03 93.66 91.75 94.94 93.36
Al2 O3 1.05 1.38 2.98 1.59 4.02 2.71 2.92 2.57 2.16 1.49 1.62 0.95 1.35
TiO2 0.02 0.04 0.04 0.03 0.05 0.02 0.03 0.04 0.05 0.02 0.03 0.03 0.04
Fe2 O3 0.005 0.005 0.18 0.18 0.57 0.14 0.39 0.005 3.94 1.16 2.63 0.73 0.82
FeO 0.94 0.58 0.58 0.36 0.43 0.72 0.22 0.72 0.22 0.79 0.65 0.72 0.72
MnO 0.01 0.02 0.01 0.02 0.01 0.01 0.01 0.01 0.05 0.04 0.08 0.02 0.05
MgO 0.75 0.18 0.19 0.25 0.23 0.22 0.18 0.2 0.61 0.37 0.18 0.09 0.29
CaO 0.34 0.2 0.09 0.36 0.08 0.1 0.15 0.06 0.3 0.19 0.09 0.11 0.25
Na2 O 0.04 0.08 0.08 0.08 0.11 0.1 0.08 0.03 0.05 0.06 0.04 0.05 0.05
K2 O 0.09 0.22 0.4 0.24 0.5 0.36 0.39 0.27 0.06 0.07 0.1 0.07 0.11
P2 O5 0.14 0.04 0.09 0.15 0.07 0.04 0.1 0.02 0.04 0.06 0.03 0.12 0.06
U3 O8 0.034 0.045 0.08 0.12 0.02 0.029 0.019 0.085 0.105 0.004 0.007 0.11 0.007
Ba 65 75 130 130 130 115 110 100 640 120 200 255 100
Co <10 <10 <10 <10 <10 <10 <10 <10 15 20 <10 <10 20
Cr 60 80 75 60 75 60 65 60 70 110 115 115 130
Cu 100 15 20 10 25 45 15 90 705 45 55 325 35
Mo 55 50 125 145 25 10 20 315 1465 45 115 125 175
Nb <10 <10 <10 <10 <10 <10 <10 <10 <10 <10 <10 <10 <10
Ni 15 20 20 15 45 30 30 20 65 50 55 25 40
Pb 65 90 45 105 75 75 100 160 6350 100 550 2125 305
Rb <10 <10 15 <10 15 <10 <10 <10 <10 <10 <10 <10 <10
(continued)
S. Goswami et al.
Table 2 (continued)
S.No ABP-1 ABP-2 ABP-3 ABP-4 ABP-5 ABP-6 ABP-7 ABP-8 ABP-9 ABP-10 ABP-11 ABP-12 ABP-13
Sr 20 <10 25 30 20 10 10 <10 16 12 11 29 13
Th 10 10 20 25 <10 <10 10 20 10 <10 <10 <10 <10
V 70 95 205 160 220 170 135 210 455 265 175 275 225
Zn 10 10 15 15 10 10 15 10 275 15 15 15 10
Zr 20 40 55 30 40 25 30 35 40 20 25 50 25
Importance of Fracturing in Uranium Mineralization in Gulcheru Quartzite …
591
592 S. Goswami et al.

Fig. 12 Selective trace element plots with respect to uranium. Rb is not showing any correlation
with uranium but strontium shows positive correlation with uranium that implying apart from Rb
some other hydrothermal source might have enriched Sr and hence linearity with U3 O8

Table 3 Alteration indices of Gulcheru quartzite samples


CIA IAI HI SI CCPI
57.327 74.462 41.397 99.350 95.201
65.295 58.334 64.558 99.148 77.655
80.361 75.574 68.658 98.174 73.095
60.316 53.155 53.943 99.025 77.822
82.248 77.479 69.696 97.527 72.670
78.641 73.204 68.325 98.323 76.025
77.938 68.453 64.021 98.210 69.567
85.079 83.435 67.018 98.421 81.789
75.714 71.921 77.594 98.591 97.908
74.128 69.490 73.548 99.071 95.302
82.749 71.065 90.386 98.970 96.456
72.625 51.805 89.564 99.414 93.247
67.298 61.363 73.823 99.155 93.299
Importance of Fracturing in Uranium Mineralization in Gulcheru Quartzite … 593

Fig. 13 Different types of alteration indices and their plot with respect to uranium. CIA—Chemical
Index of Alteration (Nesbitt and Young 1982). IAI—Ishikawa Alteration Index (Ishikawa et al.
1976). CCPI—Chlorite-carbonate-pyrite index (Large et al., 2001). HI—Hashiguchi Index (van
Ruitenbeek, 2007). SI—Silicification Index (Pirajno, 2009)

correlation matrix among the different alteration indices and uranium (Table 4) shows
combined effects. In fact, the presence of basic dyke along the fault zone can enhance
the heat and associated hydrothermal effects. But extreme erosion leads to removal of
basic dyke outcrop from the surface. Thus, the grab samples collected from surface
do not show much intense indication of hydrothermal effects apparently. Therefore,
possibility of subsurface mineralization is recognizable especially below the water
table in the zone of reduction.

Table 4 Correlation matrix of uranium and alteration indices


U3O8 CIA IAI HI SI CCPI
U3O8 1
CIA −0.13559 1
IAI −0.28696 0.596072 1
HI −0.02363 0.584844 −0.1376 1
SI 0.107638 −0.68543 −0.62553 0.012297 1
CCPI −0.03176 −0.23622 −0.12396 0.334364 0.679028 1
594 S. Goswami et al.

6 Discussion

An exhaustive understanding of the fracture network system at different scales


from micro to macrolevel is essential to unknot the mystery of uranium-bearing
fluid plumbing mechanism and uranium trapping. The Palaeoproterozoic Gulcheru
quartzite, lying above uraniferous fertile basement granitoid complex and meta-
morphosed volcano sedimentary greenstone belts, is found to host the uranium
mineralization along a broadly E-W fault intermittently for about 1.4 km strike
length. A detailed understanding of the geology in the field has revealed about
complex multiphase fracture network, where E-W fault is most prominently associ-
ated with uranium mineralization. The fracturing has played a major role in locating
the uranium mineralization. Fracture zones with clearly defined displacements of
blocks are termed as faults, which are characterized as tertiary open system. In
such brittle fault zones, dilational force must have provided space for mobilization,
redistribution and enrichment of uranium from hydrothermal fluids. The mineral-
ization is not having any similarity with either primary process related to juvenile
magma as well as anatectic/palingenetic process-related mechanism or secondary
exogenic processes related to syn- and epigenetic emplacements. Unlike primary or
secondary mechanism, presently defined uranium mineralization points to a move-
ment of hydrothermal fluids along fracture during late orogenic phase. Initially, the
hydrothermal fluids have leached out uranium from fertile granitoids along with
other incompatible elements and Ba, Cu, etc. Afterwards, the element-rich fluid
was transported to favourable structures where it got precipitated selectively. In this
context, a broad similarity with vein generation mechanism can be visualized related
to changes of pressure (P), temperature (T), eh and pH conditions. However, the
present process might have been triggered by two opposite phenomenon, viz. the
intrusion of mafic dike along the fracture pathway and diagenetic changes of over-
lying sediments. The former event may produce an additional heat and reductants
like pyrites, galena and other sulphides which led to increase in uranium holding
capabilities. Further, Fe/Ti oxides and chlorites can also adsorb uranium physically.
On the other hand, diagenetic processes can help in remobilization with addition
or removal of uranium. Thus, the most possible reason for intermittent continuity
of mineralization is diagenetic mobility, removal of mineralized zone by mechan-
ical and chemical weathering and alteration. Hence, field observation and ground
radiometric checking are not giving exactly true picture of the possible subsurface
mineralization because recrystallization of precursor mineral assemblage might have
overprinted surficially. Therefore, the subsurface exploration is required to study the
uranium mineralization along the fault and unconformity intersection. The Ambaka-
palle setting exhibits similarity with sub-unconformity epimetamorphic type uranium
mineralization (after Dahlkamp 1993).
Importance of Fracturing in Uranium Mineralization in Gulcheru Quartzite … 595

7 Proposed Genetic Model

Present study proposes that major faulting and associated fracturing might have
provided the required pathways for hydrothermal fluid circulation. Fractures
enhanced permeability of the host rock and controlled fluid movements and
reworking. The mesoscopic spatial relationship between fault and uranium miner-
alization is much better than microscopic scale. Broadly, open fractured space
filling and sealing mechanism by vein materials are not much common micro-
scopic phenomenon in the area. However, intersections of outcrop scale fractures
are observed as significant locations for uranium concentrations. At places high
angle intersections of sub-vertical fracture sets connote better pathways for fluids.
The fundamental requirements like (1) favourable structural set-up for precipitation
and tapping the mineralization and (2) presence of ore-bearing fluids in these struc-
tures imply structurally controlled mineralization. A block diagram is represented as
model to give a 3D visualization of the entire set-up (Fig. 14a). The probable mech-
anism of reaction between ore-bearing fluids with the faulted wall in host rocks and
further shallow chemical weathering and alteration are also depicted. Shaded part is
indicating the depth up to which surficial processes have masked the possible deeper
features. There should be a specific change in chemical properties of both wall rock
and fluid with continuous upward progress of the later. Therefore, the path of fluid
movement is significant to visualize the gradual changes in fluid character. Apart
from this, factors such as fall in temperature (T) and pressure (P) towards surface
also control chemical reactions. The decreased solubilities with upward movement
might have contributed to the deposition of LILE like uranium. Since the solubilities
of different substances in the fluid are different at different P–T conditions, with
progressive movement of hydrothermal fluids, gradual cooling leads to precipitation
of specific materials from supersaturated fluid. With decreasing P–T, concentration
of dissolved volatiles like H2O and CO2 decreases, which in turn affect pH and
eh condition of fluids. Addition or removal of electrons from host environments to
the fluid is main factor for reduction and precipitation of the dissolved uranium. The
ferrous iron (Fe+2 ) is soluble unlike oxidized ferric state (Fe+3 ), which is mostly insol-
uble. Thus, uranium and iron have opposite geochemical behaviour because uranium
in the reduced (U+4 ) state precipitates preferentially than that in oxidized (U+6 ) state.
Therefore, in this closed hydrothermal system iron played a role of electron donor
and uranium gained that electron. Thus, reduction of uranium and oxidation of iron
eventually lead to co-precipitation along fractured wall. The uranium mineralization
at Ambakapalle suggests the hydrothermal fluid-induced mineralization (Fig. 14b).
Further, remobilization and concentration are probably influenced by thermal effects
and reductants like pyrite, chlorite, etc., which were provided by later emplaced
mafic dyke. Subsequent chemical alteration and complex formation are manifested
by the formation of secondary uranium minerals and shallow surficial occurrences
in the vicinity of fault zone.
596 S. Goswami et al.

Fig. 14 a Block diagram showing the three-dimensional setting of Ambakapalle uranium miner-
alization. b Wall rock alteration and associated uranium mineralization mechanism (modified after
Goswami et al., 2019)

8 Conclusion

Considering the encouraging features like (i) thorium free uranium-rich anoma-
lies, (ii) clearly defined spatial association of mineralization along northern wall
of E-W trending fault, (iii) possible emplacement of basic dyke along the faulted
Importance of Fracturing in Uranium Mineralization in Gulcheru Quartzite … 597

pathway and also (iv) presence of mineralization close to the fractured unconformity
contact between basement granitoid and cover sedimentary rocks, further subsurface
exploration in the area is indispensably requisite.

Acknowledgements The authors are thankful to Dr. Soumyajit Mukherjee for invitation, internal
review and handling this article. We express our sincere gratitude to Shri L.K. Nanda, Ex-Director,
AMD, for encouragement and infrastructure support to publish the work. Special thanks to Mr.
Suman Das, Senior Geologist, Geological Survey of India, CHQ, Kolkata, for constructive sugges-
tion and extended supports in map modification. Thanks to Marion Schneider, Annett Buettener,
Boopalan Renu, Alexis Vizcaino, Doerthe Mennecke-Buehler and the proofreading team (Springer).
Dutta and Mukherjee (2021) summarize this chapter. Mukherjee et al. (2019) provide other recent
geological updates from Cuddapah.

References

Doblas, M. (1998). Slickenside kinematic indicators. Tectonophysics, 295, 187–197.


Dahlkamp, F. J. (1993). Uranium ore deposits. Berlin: Springer. ISBN 3-540-53264-1.
Dutta, D., & Mukherjee, S. (2021). Structural Geology and Tectonics Field Guidebook—Volume 1.
In S. Mukherjee (Ed.), Structural Geology and Tectonics Field Guidebook—Volume 1. Springer
Nature Switzerland AG, Cham, pp. xi-xvi. ISBN: 978-3-030-60142-3.
Goswami, S. (2014). Role of organic matter in Uranium metallogeny in Vempalle Formation
(Unpublished M. Tech. thesis). Homi Bhabha National Institute, India.
Goswami, S, Mukherjee, A, Zakaulla, S, & Rai, A. K. (2016). Microbial mat related features
in Palaeoproterozoic Gulcheru Formation and their role in low grade uranium mineralisation.
International Journal of Protochemical Science & Engineering, 1(4), 83–89.
Goswami, S., Bhagat, S., Zakaulla, S., Kumar, S., & Rai, A. K. (2017). Role of organic matter in
uranium mineralisation in Vempalle Dolostone; Cuddapah Basin, India. Journal of Geological
Society of India, 89(2), 145–154.
Goswami, S., Bhattacharjee, P., Ahmad, J., Mukherjee, A., Bhagat, S., Pandey, S. K., et al. (2018).
Insight into Proterozoic organic activity and uranium mineralisation in Vempalle dolostone,
Cuddapah Basin, Andhra Pradesh, India. Indian Journal of Geosciences, 72(4), 291–312.
Goswami, S., Upadhyay, P. K., Saravanan, B., Natarajan, V., & Verma, M. B. (2019). Two types
of uranium mineralization in Gulcheru quartzite: Fracture-controlled in Ambakapalle area and
litho-controlled in Tummalapalle area, Cuddapah Basin, Andhra Pradesh, India. China Geology,
2, 142–156.
Goswami, S., Dey, S., Zakaulla, S., & Verma, M. B. (2020). Active rifting and bimodal volcanism
in Proterozoic Papaghni sub-basin, Cuddapah basin (Andhra Pradesh), India. Journal of Earth
System Science, 129, 21. https://doi.org/10.1007/s12040-019-1278-3
Ishikawa, Y., Sawaguchi, T., Iwaya, S., & Horiuchi, M. (1976). Delineation of prospecting targets
for Kuroko deposits based on modes of volcanism of underlying dacite and alteration halos.
Mining Geology, 26, 105–117 (in Japanese with English Abstract).
Large, R. R., Gemmell, J. B., & Paulick, H. (2001). The alteration box plot: A simple approach
to understanding the relationship between the alteration mineralogy and lithogeochemistry
associated with volcanic-hosted massive sulfide deposits. Economic Geology, 96, 957–971.
Mukherjee, S., Goswami, S., & Mukherjee, A. (2019). Structures and their tectonic implications
form the southern part of the Cuddapah basin, Andhra Pradesh, India. Iranian Journal of Science
and Technology, Transaction A: Science, 43, 489–505.
598 S. Goswami et al.

Nagaraja Rao, B. K., Rajurkar, S. T., Ramalingaswami, G., & Ravindra Babu, B. (1987). Stratig-
raphy, structure and evolution of Cuddapah basin. In B. P. Radhakrishna (Ed.), Purana Basins of
peninsular India (Vol. 6, pp. 33–86). Memoirs Geological Society of India 6.
Nesbitt, H. W., & Young, G. M. (1982). Early Proterozoic climates and plate motions inferred from
major element chemistry of lutites. Nature, 199, 715–717.
Parihar, P. S., & Rao, J. S. (2012). Cuddapah basin—A uranium province. Exploration and Research
for Atomic Minerals, 22, 1–19.
Pirajno, F. (2009). Hydrothermal processes and mineral systems. Berlin: Springer Science &
Business Media. ISBN 140208613X, 9781402086137.
Ramakrishnan, M., & Vaidyanadhan, R. (2008). Geology of India (Vol. 1). Bangalore: Geological
Society of India.
Taylor, S. R. (1965). The applications of trace clement data to problems in petrology. Physics and
Chemistry of the Earth, 6, 133–213.
Umamaheswar, K., Basu, H., Patnaik, J. K., Ali, M. A., & Banerjee, D. C. (2001). Uranium miner-
alisation in the Mesoproterozoic quartzites of Cuddapah Basin in Gandi area, Cuddapah District,
Andhra Pradesh: A new exploration target for uranium. Journal of the Geological Society of
India, 57(5), 405–409.
van Ruitenbeek, F. J. A. (2007). Hydrothermal processes in the Archaean: New insights from imaging
spectroscopy (148p) (ITC dissertation). ITC, Enschede.
Granitic Rocks Underlying Deccan Trap
Along the Margin of East Dharwar
Craton, Mutnyal (Maharashtra)—Bhaisa
(Telangana), India—General Description
and Deformation
R. D. Kaplay, Md. Babar, Soumyajit Mukherjee, Deepak Wable,
and Kunal Pisal

Abstract The contact between the Eastern Dharwar Craton with the Deccan trap,
in the vertical section, is less studied structurally. A field study at the margin of the
Deccan trap with East Dharwar Craton displays strong deformations in the Dharwar
Craton just underneath the Deccan trap. We report small faults, folds, boudins
and shear zones within veins in the Dharwar Craton. These are the first physical
evidence of faults and other structures in granites below the cover of Deccan trap.
The zone along the contact of Dharwar granite with Deccan trap, stretching NE-
SW for ~60 km, is designated as the ‘East Dharwar Margin Deformation Zone’
(EDMDZ). No deformations are observed in the Deccan trap overlying the Dharwar
Craton. Dominant reverse fault tectonics along the contact of South East Deccan
Volcanic Province with East Dharwar Craton is revealed. NW verging reverse faults
are observed in the underlying basement granites. This suggests that the maximum
horizontal compression direction in the study area trends NW–SE.

1 Introduction

The Eastern Dharwar Craton (EDC) covering parts of the states Andhra Pradesh
and Karnataka (India) mainly consists of reactivated young granitoid and belongs to
the unclassified Peninsular Gneissic Complex (PGC). This perhaps marks the last
phase of the evolution of Dharwar Craton (review in Ramakrishna and Vaidyanadhan
2008).

R. D. Kaplay · D. Wable · K. Pisal


School of Earth Sciences, S.R.T.M. University, Nanded, Maharashtra 431606, India
Md. Babar
Department of Geology, Dnyanopasak College, Parbhani, Maharashtra 431401, India
S. Mukherjee (B)
Department of Earth Sciences, Indian Institute of Technology Bombay, Powai, Mumbai,
Maharashtra 400076, India
e-mail: soumyajitm@gmail.com

© Springer Nature Switzerland AG 2021 599


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_20
600 R. D. Kaplay et al.

The northern extension of PGC is reported in part of the Nanded district, Maha-
rashtra, and is referred to as EDC. These granitoids are also called the ‘Nanded
Granitoids’ (NG) (Banerjee et al. 1993; Wesanekar and Patil 2000; Banerjee 2007).
The study area lies SE of Nanded and is just at the contact between the South East
Deccan Volcanic Province (SEDVP) and the EDC (Fig. 1).
The region is a stretch of ~60 km from L1 (Mutnyal: Nanded district, Maha-
rashtra) to L4 (Bhaisa: Nirmal district, Telangana). This segment is predomi-
nantly represented by thin basalt cover, which is underlain by pink and grey
coarse-grained/porphyritic granite.
Granites at L4 also show patches and enclaves of mafic rocks (Banerjee et al.
2012). In some parts of the study area, Upper Cretaceous ‘infratrappean’ limestones
expose in between granitoid and overlying Deccan trap basalt (Banerjee et al. 2012).
Granitic rocks here belong to Palaeoproterozoic age as the oldest rock unit. These
granitic rocks are intruded by quartzo-feldspathic/quartz veins or basic Palaeopro-
terozoic to Mesoproterozoic intrusions. The overlying Deccan trap belongs to Upper
Cretaceous to Lower Palaeogene in age (Banerjee et al. 2012).
The Nanded city lies ~58 km NW of the study area and documents Deccan trap
tectonics (Kaplay et al 2013). The Kaddam region, ~68 km NE of the study area, has
NW trending tectonic element in the form of Kaddam fault/lineament with strike-slip
tectonics (Sangode et al. 2013).
Regional litho-structural studies carried out by Banerjee et al (2008) show
several N, NNE, NE, NW, and E trending lineaments in the granitoid and basalts.
They interpret these lineaments as faults/fractures/shear zones. Frequently minor
faults/fractures, shear zones, and minor folds close to the contact between the South
East Deccan Volcanic Province (SEDVP) and the EDC exist (Banerjee et al. 2012).
The focus of their study (Banerjee et al 2008, 2012) was geochemistry and petroge-
nesis of granitoid, and structural details were missing. The present study first reports
reverse faults, grabens, normal faults, strike-slip faults, boudins, etc. in the granitic
rocks of the EDC underlying the SEDVP. Detail fieldwork in the areas with intrusions
such as quartzo-feldspathic/pegmatite/quartz/epidote veins and other basic enclaves
are targeted as these features act as distinct markers for deformation. In all, nine sites
from L1-L4 are studied in detail.

2 Field Characters of Deccan Trap-Granitic Contact

The Deccan trap in the entire study area is 1.36 m (at L4) to 7.03 m thick (at L2)
(Table 1).
L1 (Mutnyal): the contact between the Deccan and the Dharwar shows red boles
(tachylitic basalt) in contact with granite. This layer is overlain by green tachylitic
basalt, which is followed by the top layer of compact aa-type basalt flow (Fig. 2). The
compact basalt is weathered spheroidally. This layer of compact basalt is covered
by thin black cotton soil, the product of weathering of basalt. Elsewhere at L1, the
red bole shows characteristic pinching and at other places red and green tachylitic
Granitic Rocks Underlying Deccan Trap Along the Margin of East Dharwar … 601

Fig. 1 a Location map. Study area marked by red rectangle. b Magnified view of the study area.
L1 (Mutnyal), L2 (Biloli), L3 (Dharmabad) and L4 (Bhaisa) are the four sites studied near Nanded.
Gray area: South East Deccan Volcanic Province (SEDVP), unshaded area: East Dharwar Craton
(EDC). Red bordered transparent black ellipse: approximate extent of East Dharwar Margin Defor-
mation Zone (EDMDZ). Red solid line: Kaddam fault/lineament. Black lined white ellipse(s) with
arrows indicate maximum compressive stress direction (σ 1 ) of reverse fault tectonics deduced from
this study. Half white-headed arrows: NE-SW compression resulting from the strike-slip faults.
c Summary of our recent geoscientific work (‘b’, ‘c’, ‘d’, ‘e’ and ‘f’) and other related work (‘a’
and ‘g’) in and around Nanded. ‘a’: NW–SE strike-slip faults near Kaddam (Sangode et al. 2012), ‘b’:
~E trending strike-slip faults in granites near Kinwat (Kaplay et al. 2017b, 2019), ‘c’: NE trending
Strike-slip faults (in granites) NW of Kinwat (RD Kaplay’s work), ‘d’: W verging thrusts in Deccan
trap in intracratonic microseismically active Nanded city (Kaplay et al 2013), ‘e’: steeply dipping
normal faults in Deccan trap in Nanded (Kaplay et al. 2017a, b), ‘f’: offset of dykes in Deccan
trap presumably caused by local stress in Aurangabad city (Babar et al. 2017) and ‘g’: a ‘slow-
deforming non-rifted zone (intracratonic seismicity) (Rajendran et al. 1986). ‘SEDVP’—South
East Deccan Volcanic Province. ‘EDC’—East Dharwar Craton. ‘WBEDCSZ’—West Boundary
East Dharwar Craton Strike-slip Zone’, marked by red ellipse, is reported by Kaplay et al. (2017a,
b). White half arrows: deformation style of the faults. Black half arrow: offset direction of dyke in
Aurangabad. Blue filled red bordered ellipse: present study area (at the contact of SEDVP and EDC)
and is designated as ‘EDMDZ’—East Dharwar Margin Deformation Zone’. ‘h’: NW–SE reverse
faults and ‘i’: mormal faults and ‘j’: NW trending strike-slip faults. Red dash circles indicate
seismicity/microseismicity
Table 1 Field character of contact of Deccan trap-granite
602

S. No. Locations Sites Type of Basalt Thickness (m) Total thickness (m) Structures Contact
1 L1 S1 (Mutnyal) Weathered compact 0.97 05.23 Sharp
basalt
Red tachylitic basalt 1.46 Gradational
Green tachylitic basalt 02.8
S2 (Minki) Weathered compact 01.06 01.64 Sharp
basalt
Weathered green 00.58
amygdaloidal basalt
2 L2 S1 (Pokharni) Weathered compact 06.09 06.69
basalt
Amygdaloidal basalt 00.60
S2 (Biloli) Weathered compact 02.43 03.22 Vertical joints in basalt and Sharp
basalt spheroidal weathering
Weathered greenish 00.79 Gradational
amygdaloidal basalt
(continued)
R. D. Kaplay et al.
Table 1 (continued)
S. No. Locations Sites Type of Basalt Thickness (m) Total thickness (m) Structures Contact
3 L3 S1 (Pimpalgaon) Soil – – – –
S2 (Mudhol) Weathered compact 01.24 1.60 Sharp
basalt
Red bole 00.36
4 L4 S1 (Degam) Soil – – – –
S2 (Bhainsa) Weathered basalt 0.73 01.36
Red bole 0.24
Greenish tachylitic 0.39
S3 (Bhainsa) Weathered compact 03.96 04.26 Weathered columnar joints, Sharp
basalt Quartz vein in basalt and
speroidal weathering
Red bole 00.30 Pinching
Granitic Rocks Underlying Deccan Trap Along the Margin of East Dharwar …
603
604 R. D. Kaplay et al.

Fig. 2 Contact between the Deccan trap and the EDC (vertical section). ‘a’: Compact Basalt, ‘b’:
green tachylitic basalt, ‘c’: red tachylitic basalt and ‘d’: granite. Faulting, at the bottom of the
photograph, is observed in granites (Loc: L1, 18° 40 N, 77° 38 E)

basalt is absent. The granite in these sections is directly overlain by compact basalt
and black cotton soil.
L2 (Biloli district, Nanded): Unlike other sections, the contact between the basalt
and the underlying granite does not show red and green tachylitic basalts. The Deccan
trap here is 3.22–7.3 m thick. Here the compact basalt, as in Muynyal, is weathered
spheroidally and jointed vertically.
L4 (Bhaisa Nirmal District, Telangana): Granite expose at the surface while at
other places of road cutting on Biloli to Mirzapur road, beautiful basalt-granite
single contact expose. Granites here, on Biloli-Mirzapur road section, are overlain
by pinched red tachylitic basalt (Fig. 3), which is overlain by green tachylitic basalt
with top compact basalt layer. The compact basalt is weathered spheroidally, and
the vertical columnar joints in basalt are intruded by thin quartz- and zeolitic veins.
Basalt is 1.36–4.26 m thick. The topmost layer of basalt is the compact type. The
Deccan trap here is 1.60 m thick.
At places flowing basalt followed the topography of granite. At specific spots,
basalt entered the crevices of underlying granite (Fig. 4).
Granitic Rocks Underlying Deccan Trap Along the Margin of East Dharwar … 605

Fig. 3 Basalt-granite contact at L4. ‘a’: basalt, ‘b’: pinching red tachylitic basalt, ‘c’: granite with
profuse intrusions of pegmatites. (Loc: L4, 19° 07 N, 77° 57 E)

Fig. 4 Contact between basalt and granite, exposed in a vertical section, where basalt (a) has
followed the undulating topography of granite (b) at L2 (Pokharni, 18° 45 N, 77° 41 E)
606 R. D. Kaplay et al.

3 Structures

3.1 Intrusives (Dykes/Veins/Sills)

At Mutnyal, 0.2–1 m thick 12 vertical veins are found at one spot. The veins are
vertical to gently/moderately dipping (30° due to W). Some of the veins show
offshoots (Fig. 5a). A few places, the veins dip steeply (Fig. 5b). Pegmatite veins
show several offshoots (Fig. 5c). Further, each branch of offshoot folded intricately.
At L4, the pegmatite vein continues as a thick but a bit irregular sill (Fig. 5d).
Total of 62 intrusions are observed in the study area. Most (88%) of the intrusions
are pegmatites. The rest are veins of quarto-feldspathic minerals, pure quartz, and
epidote. The intrusions constitute a swarm. At L3, a basic intrusion is observed. The
basic intrusion here also shows the intrusion of thin veins of younger granitic rock
with similar features as that of the host rock (Fig. 6).
At L3, we observed a rare case of the intersection of three veins in nearly in the
same region (Fig. 7). It shows a cross-cut relation among the three veins. Vein ‘a’ is
the oldest, which is cut by veins ‘b’, and ‘c’ is the youngest.

Fig. 5 a Vertical pegmatite vein swarm at L1 (18° 40 49 N, 77° 38 17 E). The third pegmatite
vein from left shows offshoot. ‘a’: granite (E Dharwar Craton), ‘b’: Pegmatite veins, ‘c’: Sill,
‘d’: Basalt boulders and black cotton soil (Deccan trap). b Inclined pegmatite vein swarm at L1
(18°40 46 N, 77°38 17 E). ‘a’: granite (E Dharwar Craton), ‘b’: pegmatite veins, ‘c’: basalt boul-
ders and black cotton soil (Deccan trap), c inclined pegmatite vein with intricately folded offshoots
on both sides. Loc: L1 18° 40 34 N, 77° 38 21 E. ‘a’: granite, ‘b’: pegmatite vein, ‘c’: intricately
folded offshoots, ‘d’: black cotton soil. d Granite (a) shows intrusion of dyke (b), which look like
a sill just underneath the basaltic cover (c). All snaps are of vertical exposures
Granitic Rocks Underlying Deccan Trap Along the Margin of East Dharwar … 607

Fig. 6 Basic intrusion show intrusion of thin veins of host granitic rock. ‘a’: granite host rock, ‘b’:
basic intrusion, and ‘c’: intrusion of granite in basic rock (sub-vertical exposure at L2; 18°48 24 N,
77° 46 04 E)

Fig. 7 Intersection of three veins. Vertical exposure at L2: 18° 48 27 N, 77° 46 02 E
608 R. D. Kaplay et al.

3.2 Deformation of EDC Underneath the Deccan Trap

3.2.1 Faults

In Fig. 8a, ‘A’ ~3 cm thick sill at L1 is reverse faulted at two places. It shows minute
drag along the fault. At F1 the vein shows the same sense of drag, while at F2 the
drag is seen only at one side of the cross-cutting element. Therefore as per Fig. 19
of Mukherjee (2014a, b), it is a ‘b.1.1’ kind of flanking structure. At F3, the steeply
dipping vein slips along a gently dipping reverse fault (Fig. 8a).
In Fig. 8b, a sill is both normal- (F4) and reverse faulted (F5). At F4, the sense of
drag across the cross-cutting element is the same (case b.1.2.2 of Fig. 19 in Mukherjee
2014a, b) while at F5, there is no drag (case b.2 of Fig. 19 in Mukherjee 2014a, b).
An inclined (50° due ~ S) pegmatite vein is faulted at two places (Fig. 8c). Both F6
and F7 are normal faults (Fig. 8c). One of the faulted blocks of F6 is reverse dragged
(Mukherjee and Koyi 2009). In Fig. 8d pegmatite vein has cut through a sill (a). It

Fig. 8 a Thin sill (a) in host granite (b) shows step-type of reverse fault at F1 and F2 (Loc: L1;
18° 40 53 N, 77° 38 26 E). F1 slips 2.5 cm with 72° dip; while at F2 the slip is 3 cm with 75°
dipping fault plane. Both these faults dip towards N30° W. At F3 an inclined vein slips to 2 cm, at
very low angle (10°) towards N65° W across the fault (F-F). b A sill (a) in host granite (b) shows
faulting at F4 (normal fault dipping 83° towards N35° E), 18 cm slip. F5 (reverse fault with fault
plane dipping 80° towards N70°W, 2 cm slip. Loc: L1; 18°40 55 N, 77°38 28 E. HW: hanging
wall, FW: footwall. c Pegmatite vein (a) in host granite (b), shows normal faulting at F6 and F7
(Loc: L2; 18° 46 40 N, 77°42 40 E). At F6 the fault plane dips 10° towards N85° E, while at
F7 the fault plane dips 20° towards N85° W. (D) Folded sill (a), in host granitic rock (c), shows
reverse fault at F8. Loc: L1; 18° 40 54 N, 77° 38 29 E. The fault plane dips 25° towards N70° W.
Inclined pegmatite vein (b) displace sill (a) at F9. It is a reverse fault that dips 35° towards NW. All
snaps are of vertical exposures
Granitic Rocks Underlying Deccan Trap Along the Margin of East Dharwar … 609

Fig. 9 a Pegmatite vein (a) in host granitic rock (b) shows reverse fault. The fault plane dips 60°
towards N17°E with 2.5 cm slip. Vertical exposure. b A normal faulted quartzo-feldspathic vein
with 3.5 cm slip. The fault plane dips 30° towards N20°W. Vertical exposure at L3; 18° 58 27 N,
77° 54 45 E

is a reverse fault. The sill also shows a minor drag related to the fault plane F8. In
Fig. 8d, a folded sill (a) is cut by a dipping pegmatite vein (b).
The pegmatite vein, at L2, is faulted in a reverse manner along with ~2 cm thick
fault core (Fig. 9a). In Fig. 9b, a quartzo-feldspathic vein, with antithetic shear within
the vein, is normal faulted at L2.
At L2, four faults are observed: F36-39. Here the central block relatively moved
down. These are conjugate faults forming central grabens (Fig. 10a). Along the fault
plane, an epidote vein intrudes (Fig. 10b, magnified view of F37 in Fig. 10a). The
magnified view of F38 and F39 are Fig. 10c, d.
At Pimpalgaon in Dharmabad taluka (L3), thin pegmatite sill (A), in host granitic
rock (B), is multi-faulted at five places (Fig. 11a). It shows conjugate faults. F16 and
F18 are curviplanar fault planes. The same vein, towards its right side, shows two
more faults with graben (F19 and F20 in Fig. 11b).
A thin inclined pegmatite vein at L2 shows a steep normal fault dipping N50°W
and dip 75° (Fig. 12a). While at L4 of the folded pegmatite veins, one of them is
faulted (Fig. 12b). Its magnified view is Fig. 12c shows two faults: one normal and
another reverse. The epidote vein on horizontal exposure at L3 is a sinistral strike-slip
faulted (Fig. 12d).
A folded sill at L1 is faulted along a wavy horizontal plane (Fig. 13a). In Fig. 13b,
a pegmatite vein is a chevron and gentle folded. Its limb on right side is reverse
faulted (F27). The pegmatite vein, at L1 is locally ductile sheared with the shear
plane along the vein itself (where half arrows are placed: F12 in Fig. 13c), while one
more pegmatite vein at L1 to is ductile sheared (F10 in Fig. 13d). Together with its
sharp dragging pattern, the vein overall resembles a fold.
610 R. D. Kaplay et al.

Fig. 10 A Graben structure exposed on vertical section (Loc: L2; 18° 48 23 N, 77° 45 54 E).
At F36 and F39 the fault plane is dipping towards N75° E at an angle of 80° and 65° respectively,
while at F37 and F38 the fault plane is vertical. The displacement at F38 and F39 is 9 cm, the
displacements at F36 and F37 could not be measured due to steepness of the exposure. b Magnified
view of F37. ‘a’: host granitic rock, ‘b’: pegmatite vein and ‘c’: epidote vein. c Magnified view of
F38. ‘a’: host granitic rock, ‘b’: pegmatite vein and ‘c’: epidote and quartz vein. d Magnified view
of F39. ‘a’: host granitic rock, ‘b’: pegmatite vein and ‘c’: epidote vein

We report total of 39 faults in the study area: 17 normal- and 17 reverse-, two
vertical- and three strike-slip faults (Table 2). Out of the eight reverse faults at L1,
seven dip NW, while only a single SE. The maximum stress axis trends NW. At
L2, a single reverse fault dips NE, and the other two dip SW. The maximum stress
axis trends NE. At L3, two reverse faults dip NW, one towards SE and two NE. The
maximum stress axes trend both NW and NE. At L4, the reverse fault dips towards
SE indicating that the maximum stress direction is NW.
Note the σ1 direction for strike-slip faults is ~N–S/NE–SW, which does not match
with the known faults reported from the coastal region or the western part of Deccan
trap around Mumbai (Misra and Mukherjee 2015, 2017; Misra et al. 2014, 2015;
Mukherjee et al. 2017)
Granitic Rocks Underlying Deccan Trap Along the Margin of East Dharwar … 611

Fig. 11 A Horst and graben structure exposed on vertical section at L3 (18° 55 13 N,
77° 51 46 E). A pegmatite vein (a) is faulted at F16, F17, F18, F19 and F20. The displace-
ment at F16, F17 and F18 is 3.5 cm, 1.5 cm and 10 cm respectively. At F16 and F17 the fault planes
are dipping towards N20° E direction at an angle of 81° and 85° respectively. At F18 the fault plane
dips towards N8° W at 56°. ‘a’: pegmatite vein, ‘b’: horst granite rock and ‘c’: curtain of epidote
& quartz vein. ‘g’: graben and ‘h’: horst. (B) The vein shows ‘graben’ structure. Slip on F19 and
F20 are 3.0 and 2.5 cm, respectively. F19 and F20 dip towards S35° E at 78° and 70°, respectively
612 R. D. Kaplay et al.

Fig. 12 a A thin gently dipping pegmatite vein is normal faulted (vertical section at L2;
18° 40 38 N, 77° 42 44 E). b One of the folded pegmatite veins is reverse faulted (vertical
section at L4; 10° 07 16 N, 77° 57 55 E). c Magnified view of F34 in vertical section. The vein
shows normal- and reverse faulting, the fault planes are dipping towards S30° E at an angle of 48°.
d Direction of strike-slip fault is N45°W with 15 cm slip (horizontal exposure at L3; 19° 01 53 N,
77° 56 07 E)

3.2.2 Folding

Folds are observed in individual pegmatite sill and veins. At Mutnyal (L1), a
pegmatite sill is a cuspate syncline (Fig. 14a). At L2, two pegmatite veins show
chevron folding (Figs. 14b, c). The folds show sharp hinges with near-vertical axial
surfaces. At L1, the westward dipping pegmatite vein becomes horizontal as observed
at three places (Fig. 14d). A similar type of folding of pegmatite vein is observed at
L2. In Figs. 15a–c, the axial plane strikes ~N, hence the compression direction is
~E–W.
The maximum stress axis deduced from reverse faults trends NW and NE. This
suggests that the compression directions during folding and faulting mismatch. The
compression direction switched with time from E–W to NW and NE.

3.2.3 Shear Along Veins

Shear sense within quartz, pegmatite and epidote veins developed at L-2 and L3
(Fig. 15a, b). Brittle P-planes develop exclusively within the veins. This is comparable
to P-planes restricted within dykes are reported by Misra and Mukherjee (2017) from
Granitic Rocks Underlying Deccan Trap Along the Margin of East Dharwar … 613

Fig. 13 a A pegmatite vein (‘b’) displaced along/by a sill (‘a’) by 19.6 cm (Loc: L1; 18° 40 50 N,
77° 38 22 E). The fault plane dips gently (16°) towards S20°W. b Chevron folded pegmatite vein
(Loc: L2; 18° 46 39 N, 77° 42 35 E). One of its limbs slips 20 cm, the fault plane dips 68° towards
S20° W (F27: reverse fault). c A pegmatite vein (a) in granitic rock (b) at L1 shows ductile–brittle
deformation (Loc: L1; 18° 40 12 N, 77° 37 22 E). d Ductile deformed pegmatite vein. F10 dips
towards SE (Loc: L1; 18° 40 49 N, 77° 38 21 E).. All snaps are of vertical exposures

Table 2 Location wise different types of the faults


Locations Total No. of Normal faults Vertical faults Reverse faults Strike-slip faults
faults
L1 15 07 – 08 –
L2 09 04 02 03 –
L3 11 03 – 05 03
L4 04 03 – 01 –
Total 39 17 02 17 03

coastal region of Deccan trap. In all the cases the shear within the vein is left lateral
(Mukherjee 2013, 2014a, b, 2015).

3.2.4 Boudins

Pegmatite veins at places (L1 and L4) are boudinaged. These boudins are ‘drawn
boudins’/‘tapering boudins’ (Goscombe et al. 2004). At L1, the boudins are almost
joined with one another (Fig. 16a); while at L4 (Fig. 16b) rather irregular-shaped
614 R. D. Kaplay et al.

Fig. 14 a Cuspate folded in quartz vein (W-E), axial plane dips 40°due 60° (Loc: L1; 18° 40 55 N,
77° 38 22 E). b Chevron folded pegmatite vein (Loc: L2; 18°46 40 N, 77°42 41 E). c Chevron
fold (Loc: L1; 18° 46 41 N, 77° 42 43 E). d Structural terrace at L1; 18° 40 50 N, 77° 38 19 E.
All snaps are of vertical exposures

boudins exist indicating possibly extensional strain at L4 to be more than that at L1.
The boudinaged asymmetric clast at right displays a top-to-left (down) shear. This
indicates this could be an ‘extensional shear boudin’ (Ghosh 1993). Boudins indicate
NW–SE extension at L1 and L4. All the reported boudins are single-layer boudins,
as the boudins developed in a single vein at those locations.

4 Discussions

How much thick is the Deccan trap along the SEDVP is not well known. However,
Patro and Sarma (2007) reported that the Deccan trap thins towards E. Recent drilling
of deeper bore wells (by common people, for drinking purpose in North Nanded) and
recovery of chips of granites underneath Deccan trap in Nanded region revealed the
thickness of basalt ranges 152–200 m. The thickness of Deccan trap overlying the
granites from Dharwar granite ranges from 1.36–7.03 m in the field. The compact
basaltic rocks exposed overlying the granites consist mostly of undeformed columnar
joints.
The granitoid in the study area are cut profusely by quartzo-feldspathic/pegmatitic
veins, quartz reef and mafic enclaves/intrusion, this signifies reactivation and remo-
bilization of granitoid. This feature is also seen in other parts of the EDC (Perraju
and Natarajan 1977; Sarvothaman and Leelanandam 1987, 1992; Sarvothaman 1993;
Gopal Reddy et al. 1998). The reactivated and remobilized terrain of Eastern Dharwar
Craton (EDC) represents northern extensions of the PGC (GSI 1979, 2001). These
Granitic Rocks Underlying Deccan Trap Along the Margin of East Dharwar … 615

Fig. 15 a Inclined P-planes (N65° W) in pegmatite vein at L2 (sub-horizontal surface) (Loc: L2;
18° 46 46 N, 77° 42 50 E). b Inclined P-planes (N35° E) in epidote vein at L3 (horizontal
surface) (Loc: L3; 18° 55 13 N, 77° 51 40 E)

granitic rocks (2385 ± 40 Ma: Banerjee et al. 1993 and 2074 ± 55 Ma: Wesanekar
and Patil 2000) have persevered within them and show deformations in the form of
small-scale faults, folds, shear within vein and boudins. Such deformations within
the Dharwar Craton indicate their antiquity but there is no evidence of their Tertiary
reactivation as no comparable deformations are observed in the Deccan trap.
The study area is bound towards NE by Nanded region, where ~W verging thrusts
and steeply dipping normal faults are reported from the Deccan trap (Kaplay et al.
2013, 2017a). Reverse- and normal faults are dominant deformations observed in
the East Dharwar Craton underlying the relatively thinner Deccan trap as found in
the study. The zone is highly deformed. We designate this deformation zone as ‘East
Dharwar Margin Deformation Zone’ (EDMDZ).
SE and NW vergence are observed for the reverse faults in the underlying basement
granites suggest NW–SE maximum horizontal compression. This direction mismatch
with the E-W horizontal compression direction of the thrusts reported from Deccan
trap in Nanded region (Kaplay et al. 2013), which is ~58 km NW of the present study
616 R. D. Kaplay et al.

Fig. 16 a Tapering boundins of pegmatite in granitic rocks trends N20° W, at L1 (18° 40 50 N,
77° 38 18 E). b Separated boudins (N40° W) at L4 (19° 07 15 N, 77° 57 56 E)

area. Interestingly, the shear direction within the veins from the study area is NW–
SE. The extension direction shown by the boudins is also the same. Fold observed at
L4 has also experienced horizontal compression in NW–SE direction. This indicates
that the study area has dominantly NW–SE stressed. This is significantly different
from ~NNE trending brittle shear Y-planes as observed by Misra et al. (2014) from
the coastal regions of Deccan trap around Mumbai region.
The deformation styles of the NW–SE strike-slip faults from the study area suggest
that the region at L3 has experienced NE-SW compression regime from the strike-
slip movement. The deformation style from the study area corroborates with that of
the strike-slip faults reported from Kaddam region by Sangode et al. (2013), which
is ~125 NE from the study area. This indicates that the NW–SE strike-slip faults we
observe here could be a manifestation of the far field effect of the NE–SW directed
stress resulting from the strike-slip movement along the Kaddam fault.
Apparent higher frequency of minor faults (with no other details of structural
features provided by previous authors) in EDC (as surface exposure of granites),
Granitic Rocks Underlying Deccan Trap Along the Margin of East Dharwar … 617

Fig. 17 Explanatory block diagram (not to the scale) showing minor faults within the EDC, reverse
faults in granites under the cover of Deccan trap and blind faults (?) below the Deccan trap in Nanded
region. Gray shaded area: SEDVP; white area: EDC. Left lateral sheared vein. The P-plane at L1,
L2 and L4 trends NW; at L2 it is NE

nearer to the Deccan trap-granite contact (‘A’ in Fig. 17), is also reported by Bannerjee
et al. (2012). This study reports faults from granite, which is present below the Deccan
trap (‘B’ in Fig. 17). This study thus proves that the granites under the cover of Deccan
trap hold several crucial structural/tectonic information.
Deformation in the form of reverse faults in Nanded, ~58 km NW of the study
area, is reported from weathered amygdaloidal basalts. The compact basalt layer
from Nanded region- usually the topmost layer-looks undeformed. Note that in the
study area, compact basalt exposed as the topmost layer, which to looks underformed.
But leaving aside the reporting of thrusts from Nanded region, no other deformations
are reported from nearby Deccan trap including the present study area. Interestingly
Kaplay et al. (2017a, b) reported fault-line scarp along the Urvashi Ghat Linea-
ment in the Deccan trap. The epicentre of microseismic events in Nanded region
is reported from 4 km depth along this fault-line scarp, which is also a causative
fault (Srinagesh et al. 2012). Also, note that Madhnure et al. (manuscript communi-
cated to Environmental Monitoring and Assessment) report that the Deccan trap in
Nanded is 150–200 m thick. This suggests that causative fault in Nanded, depth of
which is reported as ~4 km by (Srinagesh et al. 2012), is present in basement granite
underneath Deccan trap, as beyond 200 m depth granites are encountered in Nanded
618 R. D. Kaplay et al.

city. Kaplay et al. (2017a, b) have also reported normal faults from basaltic terrain
in Nanded region.
The present study indicates that the reverse faults (Kaplay et al. 2013) and normal
faults (Kaplay et al. 2017a) in Deccan trap could be the manifestation of hidden faults
in basement granites at ~200 m depth in Nanded region. Likewise, Ernst (2014)
referred surface manifestation of structures in other Large Igneous Provinces.
Interestingly, 23 NE and NW trending, lineaments in the form of straight courses
of streams are reported N of Godavari River in Nanded (Kaplay et al. 2017b). Are
these lineaments including UG Lineament manifestation of some hidden structures
(e.g., Bhave et al. 1989)? Detail geophysical research on possible hidden faults below
the Deccan trap in and around the study area would be much required.

5 Conclusions

The surface exposure of Deccan trap is 0.1–2.55 m thick around the contact between
the South East Deccan Volcanic Province and the Dharwar Craton. The granitic rock
of Dharwar Craton underwent minor faulting and folding. The deformation zone
from L1 to L4 is designated here as the ‘East Dharwar Margin Deformation Zone’
(EDMDZ). Deformations from granites of Dharwar Craton indicate its possible antiq-
uity, however, reactivation of these faults is not observed in overlying the very thin
cover of Deccan trap. Faults, fold, boudins, etc. are reported for the first time from
the basement granite (EDC) underlying Deccan trap. Reverse fault and normal fault
tectonics are dominant in the region and the compression direction switched with
time from E–W to NW and NE.

Acknowledgements RDK thanks the Director, School of Earth Sciences (S.R.T.M. University)
for encouragement. Md. Babar thanks the Principal (Dnyanopasak College) for support. SM was
supported by the CPDA grant of IIT Bombay. The research sabbatical granted to him by IIT
Bombay for the year 2017 has been helpful to finalize this work. Thanks to Marion Schneider, Annett
Buettener, Boopalan Renu, Alexis Vizcaino, Doerthe Mennecke-Buehler and the proofreading team
(Springer). Dutta and Mukherjee (2021) summarize this work.

References

Banerjee, R. (2007). Geological, geochemical and geochronological characterisation of crystallines


for uranium and rare metal and rare earth mineralisation in parts of Nanded district, Maharashtra
(Unpublished Ph.D. thesis). RTM, Nagpur University, Nagpur.
Banerjee, R., Jain, S. K., & Shivkumar, K. (2008). Geochemistry and petrogenesis of uraninite
bearing granitoids and radioactive phosphatic cherty cataclasite of Thadiasoli area, Nanded
district, Maharashtra. Memoirs of Geological Society of India, 73, 55–84.
Banerjee, R., Veena, K., Pandey, B. K., & Parthasarathy, T. N. (1993). Rb–Sr geochronology of the
radioactive granites of Nanded area, Maharashtra, India. Journal of Atomic Mineral Science, 1,
111–117.
Granitic Rocks Underlying Deccan Trap Along the Margin of East Dharwar … 619

Banerjee, R., Shivkumar, K., Maithani, P. B. (2012). Major and trace element geochemistry of
palaeoproterozoic Nanded district, Maharashtra: geodynamic and petrogenetic implications.
Journal of Applied Geochemistry 26–58.
Bhave, K. N., Ganju, J. L., & Ram, J. (1989). Origin, nature and geological significance of
lineaments. In M. N. Qureshy, W. J. Hinze (Eds.), Regional geophysical lineaments (Vol. 12,
pp. 35–42). Bangalore: Geological Society of India.
Dutta, D., & Mukherjee, S. (2021). Introduction to structural geology & tectonics field guidebook.
In S. Mukherjee (Ed.), Structural geology & tectonics field guidebook. Switzerland: Springer
Nature Switzerland AG. Cham, pp. xi-xvi. ISBN: 978-3-030-60142-3.
Ernst, R. E. (2014). Large igneous provinces (pp. 1–653). Cambridge: Cambridge University Press.
Geological Survey of India. (2001). District resource map of Nanded district, Maharashtra on
1:300,000 scale with explanatory brochure.
Ghosh, S. K. (1993). Structural geology: Fundamentals and modern developments. Pergamon Press.
Gopal Reddy, T., Suresh, G., & Rao, N. V. (1998). Classification and characterization of the Penin-
sular Gneissic Complex in the Eastern Block of Dharwar craton. In National Seminar on Concep-
tual Models on the Evolution of Granite-Greenstone Belts, Granulite Terrains and Associated
Mineral Deposits, Abstract Volume, the Indian Mineral (Vol. 32, pp. 9–12).
Goscombe, B. D., Passchier, C. W., & Hand, M. (2004). Boudinage classification: End members
boudin types and modified boudin structures. Journal of Structural Geology, 26, 739–763.
Kaplay, R. D., Vijay Kumar, T., & Sawant, R. (2013). Field evidence for deformation in Deccan
Traps in microseismically active Nanded area, Maharashtra. Current Science, 105, 1051–1052.
Kaplay, R. D., Babar, Md., Mukherjee, S., & Kumar, T. V. (2017a). Morphotectonic expression
of geological structures in eastern part of south east Deccan volcanic province (around Nanded,
Maharashtra, India). In S. Mukherjee, A. A. Misra, G. Calvès, & M. Nemčok (Eds.), Tectonics
of the Deccan Large Igneous Province (Vol. 445, pp. 317–335). London: Geological Society of
London Special Publication.
Kaplay, R. D., Kumar, T. V., Mukherjee, S., Wesanekar, P. R., Babar, M., & Chavan, S. (2017b).
E-W strike slip shearing of Kinwat granitoid at South East Deccan Volcanic Province, Kinwat,
Maharashtra, India. Journal of Earth System Science, 126(5), 71.
Kaplay, R. D., Babar, Md., Mukherjee, S., Mahato, S., & Chavhan, S. (2019). Structural features
of Kinwat Peninsular Gneissic Complex along the western margin of Eastern DharwarCraton,
India. Arabian Journal for Science and Engineering, 44, 6509–6523.
Misra, A. A., Bhattacharya, G., Mukherjee, S., & Bose, N. (2014). Near N-S paleo-extension in the
western Deccan region in India: Does it link strike-slip tectonics with India-Seychelles rifting?
International Journal of Earth Sciences, 103, 1645–1680.
Misra, A. A., & Mukherjee, S. (2015). Tectonic inheritance in continental rifts and passive margins.
Berlin: Springerbriefs in Earth Sciences.
Misra, A. A., & Mukherjee, S. (2017). Dyke-brittle shear relationships in the Western Deccan
Strike-Slip Zone around Mumbai (Maharashtra, India). In S. Mukherjee, A. A. Misra, G. Calvès,
M. Nemčok (Eds.). Tectonics of the Deccan Large Igneous Province (Vol. 445, pp. 269–295).
London: Geological Society of London Special Publication.
Misra, A. A., Sinha, N., & Mukherjee, S. (2015). Repeat ridge jumps and microcontinent separation:
Insights from NE Arabian Sea. Marine and Petroleum Geology, 59, 406–428.
Mukherjee, S. (2013). Deformation microstructures in rocks (pp. 1–111). Berlin: Springer
Geochemistry/Mineralogy.
Mukherjee, S. (2014a). Review of flanking structures in meso- and micro-scales. Geological
Magazine, 151, 957–974.
Mukherjee, S. (2014b). Atlas of shear zone structures in meso-scale (pp. 1–124). Cham: Springer
Geology.
Mukherjee, S. (2015). Atlas of structural geology. Amsterdam: Elsevier.
Mukherjee, S., & Koyi, H. A. (2009). Flanking microstructures. Geological Magazine, 146, 517–
526.
620 R. D. Kaplay et al.

Mukherjee, S., & Koyi, H. A. (2010a). Higher Himalayan Shear Zone, Sutlej Section—Structural
geology & extrusion mechanism by various combinations of simple shear, pure shear & channel
flow in shifting modes. International Journal of Earth Sciences, 99, 1267–1303.
Mukherjee, S., & Koyi, H. A. (2010b). Higher Himalayan Shear Zone, Zanskar Section—
Microstructural studies & extrusion mechanism by a combination of simple shear & channel
flow. International Journal of Earth Sciences, 99, 1083–1110.
Mukherjee, S., Misra, A. A., Calvès, G., & Nemčok, M. (2017). Tectonics of the Deccan Large
Igneous Province: An introduction. In S. Mukherjee, A. A. Misra, G. Calvès, & M. Nemčok
(Eds.), Tectonics of the Deccan Large Igneous Province (Vol. 445, pp. 1–9). London: Geological
Society of London, Special Publication.
Patro, B. P. K., & Sarma, S. V. S. (2007). Trap thickness and the subtrappean structures related to
mode of eruption in the Deccan Plateau of India: Results from magnetotellurics. Earth Planets
Space, 59, 75–81.
Perraju, P., & Natarajan, V. (1977). Peninsular gneiss in the northern parts of Andhra Pradesh.
Journal of Geological Society of India, 18, 224–232.
Ramakrishna, M., & Vaidyanadhan, R. (2008). Geology of India (Vol. 1). Bangalore: Geological
Society of India.
Sangode, S. J., Mesharm, D. C., Kulkarni, Y. R., Gudadhe, S. S., Malpe, D. B., & Herlekar, M.
A. (2013). Neotectonic response of the Godavari and Kaddam Rivers in Andhra Pradesh, India:
Implications to quaternary reactivation of old fracture system. Journal of Geological Society of
India, 81, 459–471.
Sarvothaman, H. (1993). The molar Al2O3/ (CaO + Na2 O + K2 O) ratios. Discriminant contrasting
for oceanic plagio—granites and continental Trondjemites. Journal of Geological Society of India,
42, 513–522.
Sarvothaman, H., & Leelanandam, C. (1987). Petrography and major oxide chemistry of the
Archaean granitic rocks of the Medak area, Andhara Pradesh. Journal of Geological Society
of India, 30, 194–209.
Sarvothaman, H., & Leelanandam, C. (1992). Peraluminous, metaluminous and alkali granites from
parts of A.P. and Karnataka in Dharwar craton: A critical reappraisal of existing data. Journal of
Geological Society of India, 39, 279–292.
Srinagesh, D., Srinivas, T. V. N., Solomon Raju, P., Suresh, G., Murthy, Y. V. V. B. S. N., Saha, S.,
et al. (2012). Causative fault of swarm activity in Nanded City, Maharashtra. Current Science,
103, 366–369.
Wesanekar, P. R., & Patil, R. R. (2000). Rb–Sr dating of pink Granites of Deglur, Nanded district,
Maharashtra, India. In Proceedings of National Seminar on Tectonomagmatism, Geochemistry
and Metamorphism of Precambrian Terrains (pp. 181–187).
Structural Analyses
of the Lunavada–Santrampur Area
(Gujarat, India) Using Remote Sensing
Images

Geetika H. Chauhan, G. S. Rao, and Soumyajit Mukherjee

Abstract The Lunavada Group of rocks occupies parts of Sabarkantha and


Panchmahal districts in Gujarat (India) and have undergone polyphase deforma-
tion in the Kadana Formation at NE Gujarat. This study analyzes folds and linea-
ments to visualize the tectonics using remote sensing images of Santrampur and the
surrounding areas. The DEM data was obtained from ASTER and satellite images
from Sentinel-2. Analytical technique such as Topographic openness is used to visu-
alize the openness and closeness of the topography. Various hill shades were gener-
ated to understand the area from different sun angles, and the best two directions were
merged to interpret the lineaments and were latter plotted on a rose diagram. The
study uses Google Earth for identifying various fold geometries. With this chapter,
we draw attention to structural geologists, the area Santrampur as an excellent place
for fieldwork, training and tectonic research.

1 Introduction

We introduce to the reader a hitherto unknown area to structural geologists-


Santrampur (Gujarat, India). Google Earth image reveals spectacular mega-scale
structures (folds and faults) from field and interpreted for tectonics. We hope this
chapter will give the first impetus to structural geologists to undertake detail works
on this terrain. Santrampur is one such area that the instructor can display on a large
screen to even the beginners on structural geology and encourage to interpret struc-
tures. In this way, the instructor can develop a very interactive session with students.

G. H. Chauhan (B)
Department of Geology, K. J. Somaiya College of Science and Commerce, Vidyanagar,
Vidyavihar, Mumbai, Maharashtra 400077, India
e-mail: geetikamacs@gmail.com
G. S. Rao
Prithvi Geospatial, A-9, Arjun Center, Govandi (E), Mumbai, Maharashtra 400088, India
S. Mukherjee
Department of Earth Sciences, Indian Institute of Technology Bombay, Powai, Mumbai,
Mahatrashtra 400076, India

© Springer Nature Switzerland AG 2021 621


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_21
622 G. H. Chauhan et al.

The structures are so obvious in the image that even the shy students would start
interpreting, right or wrong!
The southernmost part of Aravalli Mountain belt is known as Southern Aravalli
Mountains Belt (SAMB). This area is at the junction between the Aravalli Craton in
the north and the Dharwar Craton to the south. It is also a part of the Narmada–Son
belt. The Lunavada Group occupies parts of Sabarkantha and Panchmahal districts
in Gujarat (Fig. 1a). Gupta and Mukherjee (1938) first mapped the area. The study
area extends from 23° 01 25.51 N to 73° 37 50.3 E and from 23° 23 11.62 N
to 73° 45 36.4 E. It comes within the Survey of India’s topo-sheet No. 46 E/12.
The best way to reach the study area is by the Highway NH 48 that connects Delhi
and Mumbai. Another way to reach the study area is by train, first from Vadodara to
Godhra, and then from Godhra to Lunavada.
Mahi and Panam are the two major rivers in the study area. The area comprises
peneplains to the north of Lunavada and toward the south and east is an intensely
folded and faulted terrain. The peneplains consist of softer rocks such as chlorite
schists and mica schists, which are easily eroded and covered by soil; whereas hard
resistant quartzites associated with these metapelites forms the fold limbs (Fig. 1b,
also see Fig. 2). Structural geology and tectonics of the area have been worked out by a
few workers so far, e.g., Gupta and Mukherjee (1938), Iqbaluddin (1997), Mamtani
(1999) and Joshi (2013). Recently, Mukherjee et al. (2020) published numerous
structural field photographs from the Lunavada area.
This entire sequence is complexly folded at least thrice and was subsequently
intruded by the Godhra Granite (Mamtani et al. 2000). There are very well-defined
complex deformation events, which can be seen regionally in the satellite image
(Fig. 3). The southern part of the study area around Lunavada, Santrampur and
further south is characterized by regional-scale superposed folds. The northern part
of the study area shows tight folds and close-spaced axial planar fractures (Fig. 4).
A sinistral shear zone (Mamtani et al 1999) exists in the northern part of the study
area with a mean trend of N50° E–S 230° W (Fig. 5).

2 Structural Detail Deduced from Google Earth Image

Polyclinal fold: These folds show axial surfaces with variable or contrasting closure
directions (Fig. 6).
Second-order folds: M-shaped (Fig. 7a) and Z-shaped folds (Fig. 7b) are documented
easily after zooming in the Google Earth image.
Curved axial trace of the fold: This is interpreted from the image (Fig. 8). Curved axial
trace of the fold would mean that the folded rock underwent at least two generations
of compression.
Structural Analyses of the Lunavada–Santrampur Area (Gujarat, India) … 623

Fig. 1 a Part of the


lithostratigraphic map of the
SAMB around Lunavada
(after Gupta et al. 1980). L =
Lunavada, G = Godra, V =
Vadodara and S =
Santrampur. b Geological
map of the study area.
Schists of different
metamorphic grades
(chlorite, biotite and garnet
biotite schists) are shown by
different symbols (after
Mamtani et al. 2001)
624 G. H. Chauhan et al.

Fig. 2 Sentinel-2 image displaying intercalation of quartzite with metapelites generated using band
ratio of band 4(red) by band 2(blue) after applying ATCOR

Fig. 3 Sentinel-2 band combination (12-11-2) displaying the study area


Structural Analyses of the Lunavada–Santrampur Area (Gujarat, India) … 625

Fig. 4 Google Earth Image of the study area

Fig. 5 ASTER GDEM (90m) Hillshade azimuth of 315° and solar elevation of 30° showing sinistral
shear zone
626 G. H. Chauhan et al.

Fig. 6 Google Earth image of polyclinal fold (inside box)

3 Data Analyses Using Remote Sensing Images

Use of remote sensing images in structural geological data is a well-established


procedure (e.g., Misra et al. 2014; Babar et al. 2017; Misra and Mukherjee 2017;
Vanik et al. 2018; Dasgupta and Mukherjee 2019). Meer et al. (2014) presented the
potential of Sentinel-2 imagery (10 m resolution) in structural interpretation. Such an
imagery with ASTER GDEM (90 m resolution) was used for structural trendlines and
lineaments interpretation. QGIS (V.2.18.24) was the open-source software used for
processing and mapping the data during the analysis. Image processing techniques
like atmospheric correction and band rationing were used. Fill-nodata algorithm was
used for processing of DEMs to make it error free. Geoprocessing tool, Topographic
openness, was used to understand the surface concavities and convexities.
The Sentinel-2 images were atmospherically corrected by applying the Dark
Object Subtraction-1(DOS-1) method. A bandset containing the blue band (band
2) with a 10-m resolution, and two short wave infrared bands (band 11 and 12) with
resolution of 20 m were created to visualize the area (Fig. 3). To differentiate the
quartzite from metapelites, a band ratio was created by ratioing the red band (band 4)
with the blue (band 2). Quartzites appeared creamish whereas the metapelites purple
(Fig. 4).
Hillshades of ASTER DEM were generated with azimuth to highlight all the
smaller linear features. In order to identify linear topographic features from the
DEM, seven shaded relief images were generated. The first shaded relief image
Structural Analyses of the Lunavada–Santrampur Area (Gujarat, India) … 627

Fig. 7 a Google Earth image of the second-order M-type fold. b Google Earth image of the
second-order Z-type fold
628 G. H. Chauhan et al.

Fig. 8 Google Earth image of curved axial trace

Fig. 9 Hillshade derived from ASTER GDEM with azimuth of 45° and solar elevation of 45°
Structural Analyses of the Lunavada–Santrampur Area (Gujarat, India) … 629

Fig. 10 Hillshade derived from ASTER GDEM with azimuth of 90° and solar elevation of 45°

created had a solar azimuth (sun angle) of 45° (Fig. 9) and a solar elevation of 45°.
The other six shaded relief images were created with six contrasting illumination
directions 90° (Fig. 10), 135° (Fig. 11), 180° (Fig. 12), 225° (Fig. 13), 270° (Fig. 14)
and 315° (Fig. 15). The second step is to merge the two DEMs which best represents
the surface, i.e., 135° and 45° (Fig. 16).
Topographic openness is a tool used to understand the openness of the topography
(Elmahdy 2010). It describes the degree of dominance or enclosure of a point relative
to the surrounding terrain, in eight different directions within a given radial distance.
Measured above the surface, a positive openness emphasizes convex features in the
landscape. Measured below the surface, a negative openness emphasizes concave
features in the landscape. Positive openness (Fig. 17) reflects the surface upward
and/or ridge (footwall of fault). On the other hand, a negative openness (Fig. 18)
reflects surface downward and/or channel (fault zone).
The topography which looked more or less even was better visualized after using
the Topographic Openness algorithm. The Positive openness map highlighted the
concave up features like the fold, it’s deformed limbs and other elevated parts present
in the area; however, the Negative openness map emphasized the convex landform
present below the concave up features and its surrounding. A depression is seen
to pop-up in the negative openness map between the Mahi river channel and the
northern limb of the fold in the northwest direction indicating deeper crust.
630 G. H. Chauhan et al.

Fig. 11 Hillshade derived from ASTER GDEM with azimuth of 135° and solar elevation of 45°

Fig. 12 Hillshade derived from ASTER GDEM with azimuth of 180° and solar elevation of 45°
Structural Analyses of the Lunavada–Santrampur Area (Gujarat, India) … 631

Fig. 13 Hillshade derived from ASTER GDEM with azimuth of 225° and solar elevation of 45°

Fig. 14 Hillshade derived from ASTER GDEM with azimuth of 270° and solar elevation of 45°
632 G. H. Chauhan et al.

Fig. 15 Hillshade derived from ASTER GDEM with azimuth of 315° and solar elevation of 45°

Fig. 16 Merging of two hillshades with azimuth 315° and 45° and solar elevation of 45°
Structural Analyses of the Lunavada–Santrampur Area (Gujarat, India) … 633

Fig. 17 Positive openness map (enhancing surface upward)

Fig. 18 Negative openness map (enhancing surface downward)


634 G. H. Chauhan et al.

Fig. 19 Lineaments extracted by merging two hillshades with azimuth of 315° and 45° and solar
elevation of 45°

4 Analysis of Lineaments

Automatic technique of lineament identification was avoided as many non-


meaningful linear may crop-up due to illumination, topography, shadow, etc. Two
DEMs (Figs. 9 and 15), with azimuths 315° and 45° and a solar elevation of 45°, were
merged to trace the lineaments. A total of 92 lineaments were identified (Figs. 19 and
20) and plotted on the rose diagram (Fig. 21) with the help of Rose.net (V.0.10.0.0,
year: 2012) software. The Rose diagram clearly shows that NW and NE are the two
dominant trends.
Structural Analyses of the Lunavada–Santrampur Area (Gujarat, India) … 635

Fig. 20 Map displaying lineaments extracted from ASTER GDEM

Fig. 21 Rose diagram


showing distribution of
lineaments in the study area
636 G. H. Chauhan et al.

5 Conclusions

Study of Google Earth images reveals eye-catching polyclinal folds, second-order


folds and superposed folds from Santarampur area (Gujarat, India). Image analyses
gives us improved idea about the topography as well as the dominant trend of natural
lineaments in the area.

Acknowledgements The amazing detail of structures in remote sensing images from Santarampur
was pointed out by Satardu Bhattacharya (Space Application Center) to SM. Thanks to Springer
(Marion Schneider, Alexis Vizcainoand the proofreading team) for assistance and Aditya Joshi (MS
University Baroda) for interaction. Subhobroto Mazumder (ONGC Dehradun) provided detail useful
comments on this article. Thanks to Marion Schneider, Annett Buettener, Boopalan Renu, Alexis
Vizcaino, Doerthe Mennecke-Buehler and the proofreading team (Springer). Dutta and Mukherjee
(2021) summarize this work.

References

Babar, Md., Kaplay, R. D., Mukherjee, S., & Kulkarni, P. S. (2017). Evidences of deformation of
dykes from Central Deccan Volcanic Province, Aurangabad, Maharashtra, India. In S. Mukherjee,
A. A. Misra, G. Calvès, & M. Nemčok (Eds.), Tectonics of the Deccan Large Igneous Province
(Vol. 445, pp. 337–353). London: Geological Society of London Special Publications.
Dasgupta, S., & Mukherjee, S. (2019). Remote sensing in lineament identification: Examples from
western India. In A. Billi & A. Fagereng (Eds.), Problems and solutions in structural geology and
tectonics. Developments in Structural Geology and Tectonics Book Series (Vol. 5, pp. 205–221).
Amsterdam: Elsevier. ISSN: 2542-9000. ISBN: 9780128140482.
Dutta, D., & Mukherjee, S. (2021). Introduction to Structural Geology and Tectonics Field
Guidebook—Volume 1. In S. Mukherjee (Ed.), Structural Geology and Tectonics Field Guide-
book—Volume 1. Switzerland: Springer Nature Switzerland AG. Cham. pp. xi-xvi. ISBN:
978-3-030-60142-3.
Elmahdy, S. (2010). Topographic openness algorithm for characterizing geologic fractures of Kuala
Lumpur limestone bedrock using DEM. Journal of Geomatics, 4, 63–65.
Gupta, B. C., & Mukherjee, P. N. (1938). Geology of Gujarat and southern Rajputana. Recruitment
of the Geological Survey of India, 73(2), 103–208.
Iqbaluddin, S. U. D., & Javed, A. (1997). Geomorphology and landscape evolution of bharatpur
district, rajasthan. Journal of the Indian Society of Remote Sensing, 25(3), 177–186.
Joshi, A., Limaye, M. A., & Deota, B. S. (2013). A model representing successive deformational
events of Ankalwasynform, Lunavada Group, Gujarat. Gondwana Geological Magazine, 28, 1–4.
Mamtani, M. A., Greiling, R. O., Karanth, R. V., & Merh, S. S. (1999). Orogenic deformation and
its relationship to AMS fabric—An example from the Southern Margin of the Aravalli Mountain
Belt, India. Memoir Geological Society of India, 44, 9–24.
Mamtani, M. A., Karanth, R. V., Merh, S. S., & Greiling, R. O. (2000). Tectonic evolution of the
southern part of Aravalli Mountain Belt and its Environs: possible causes and time constraints.
Gondwana Research, 3, 175–187.
Mamtani et al. (2001). Time relationship between metamorphism and deformation in Proterozoic
rocks of the Lunavada region, Southern Aravalli Mountain Belt (India) - a microstructural study.
Journal of Asian Earth Sciences, 19(1), 195–205.
Meer, F. D., Werff, H. M. A., & Ruitenbeek, F. J. A. (2014). Potential of ESA’s Sentinel-2 for
geological applications. Remote Sensing Environment, 148, 124–133.
Structural Analyses of the Lunavada–Santrampur Area (Gujarat, India) … 637

Misra, A. A., Bhattacharya, G., Mukherjee, S., & Bose, N. (2014). Near N-S paleo-extension in the
western Deccan region in India: Does it link strike-slip tectonics with India-Seychelles rifting?
International Journal of Earth Sciences, 103, 1645–1680.
Misra, A. A., & Mukherjee, S. (2017). Dyke-brittle shear relationships in the Western Deccan
Strike Slip Zone around Mumbai (Maharashtra, India). In S. Mukherjee, A. A. Misra, G. Calvès,
M. Nemčok (Eds.), Tectonics of the Deccan Large Igneous Province (Vol. 445, pp. 269–295).
London: Geological Society of London Special Publications.
Mukherjee, S., Bose, N., Ghosh, R., Dutta, D., Misra, A. A., Kumar, M., Dasgupta, S., et al. (2020).
Structural geological atlas. Berlin: Springer. ISBN: 978-981-13-9825-4.
Vanik, N., Shaikh, H., Mukherjee, S., Maurya, D. M., & Chamyal, L. S. (2018). Post-Deccan trap
stress reorientation under transpression: Evidence from fault slip analyses from SW Saurashtra,
western India. Journal of Geodynamics, 121, 9–19.
Fundamentals of Lithostructural
Mapping: Example from the SW Part
of the Proterozoic Bhima Basin,
Karnataka, India: A Note on Dharwarian
Crustal Evolution

Sukanta Goswami, Shivam Shrivastava, Suman Das,


and Purnajit Bhattacharjee

Abstract Lithostructural mapping in any type of geological terrains like the younger
Himalayan mountain or the older Archaean Cratons has always been an immensely
helpful tool. There are some challenges for geologists in mapping certain geolog-
ical terrain. Although the use of remote sensing, GIS and other feasible tech-
nology is enhancing day by day, accurately produced field survey data cannot be
challenged and considered as most reliable and fundamental. The fundamentals of
projection techniques, toposheet preparation, components and working principles for
brunton compass, magnetic declination effects, Global Positioning System (GPS),
litho contact tracing, outcrop correlation by extrapolation and interpolation, etc. are
of prime significance for field geologist. Though there are a number of publica-
tions available on these aspects, systematic examples with sketches, data tables and
outcrop photos with specific case studies remained pending. In this chapter, compre-
hensive descriptions with a different approach on a few less commonly available but
useful combination of techniques for lithostructural mapping are demonstrated with
reference to a study area along with the SW basement of Proterozoic Bhima basin.
Evolution of Archaean crust is also explained with a brief description.

1 Introduction

The basic understanding and techniques of geological mapping are well documented
in several literatures (e.g., Barnes and Lisle 2004; Lisle 2004; McClay 2007; Kruhl
2017; Billi and Fagereng 2019). This contribution gives some upgraded guidelines
on lithostructural mapping, which are not covered yet in such a perspective and can
be helpful for the field geologists. Presently, some ground rules are defined so that

S. Goswami (B) · S. Shrivastava · P. Bhattacharjee


Atomic Minerals Directorate for Exploration and Research, Bangalore 560072, India
e-mail: sukantagoswami.amd@gov.in
S. Das
Geological Survey of India, Central Head Quarters, Kolkata 700016, India

© Springer Nature Switzerland AG 2021 639


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_22
640 S. Goswami et al.

accuracy and precision can be maintained in mapping techniques. Therefore, the


map cannot vary for the same area irrespective of different workers at different times
and can be reproduced any time. Understanding and representation of rock types and
structural aspects together in the map may give quite valuable information on sub-
surface geology and especially helpful for exploration of natural economic resources.
Field observation may provide significant insight into the deeper crust and to some
extent even the upper mantle. Lithostructural mapping can be used to represent a
complete depth of erosion and exhumation (up to even the uppermost mantle of
the geological past at places) from the now exposed outcrops. Therefore, proper
lithostructural mapping can be helpful in understanding a thorough picture of the
structural variation associated with certain deformation, which in turn can indicate
past tectonics. For the present purpose thorough ideas on different coordinate systems
and projection types are required in advance. It is assumed that the basic concepts
(e.g., the recognition criteria of rock types and structures, what to measure and what
to describe, how to collect, analyze and interpret the field data, etc.) are known to
all.
In this chapter, we start the discussion with some useful principles which are used
practically in the field during mapping. There are certain aspects described in detail
in subsequent paragraphs viz., 1. terrain-specific rock types (whether sedimentary or
igneous or metamorphic), 2. morphology of different geological structures, and their
interrelationship, 3. primary and/or secondary structural fabrics, and 4. manifestation
of all features on the map. Finally, an example is shown from the SW part of the
Proterozoic Bhima basin (Karnataka, India). The main aim in this present context
is to understand deformation history of the Eastern Dharwar Craton (EDC) and its
implications on Archean crustal evolution in India.

2 Prerequisite of Lithostructural Mapping

Besides the field items (hammer, chisel, compass, tape, hand lens, notebook, scale,
protractor, pencil, pen, eraser, knife, streak plate, GPS, camera, acid bottle, magnet
and water bottle, etc.), the most essential object is topographic base map or the
toposheet (i.e., a type of map characterized by the detail and quantitative represen-
tation of relief, using contour lines with specific contour interval, scale, geographic
direction and index). Basic understanding of some principles and techniques is a
vital point in lithostructural mapping. The terminologies like scale of toposheet and
map, coordinate system, projection, etc. should be known to all.
Using different projection techniques (viz. azimuthal, conical and cylindrical),
information on the surface of the spherical Earth is possible to be transferred onto a
flat piece of paper (Figs. 1a–c). This imaginary piece of paper can be regarded as a
projected map, which shows no distortion where it touches the Earth. Away from the
touching points, distortions increase, and thus, irrespective of projection techniques
middle of a map usually touches the Earth. Therefore, it is understood that azimuthal
Fundamentals of Lithostructural Mapping: Example … 641

Distortion increases with distance


from the centre (solid red circle)
b

Distortion increases away from median line (Red line)

Distortion increases away from median line (Red line)

Fig. 1 Three types of projection techniques a Azimuthal projections showing longitude lines
fanning out from the centre and latitude lines as concentric circles. b Conical projection showing
the shapes of land masses near mid-latitude are close to the true shape. c Cylindrical projection
showing that any line drawn on the map is true direction

projection is best for polar areas, conical projection is best for mid-latitudes (~20° –
60° North and South). For the equatorial areas like southern Indian terrain cylin-
drical projection is used as per the national mapping policy, 2005 (Survey of India
2005). In this context, the mapping format can be based on angle or distance. A
Universal Transverse Mercator (UTM) system provides a constant distance relation-
ship anywhere on the map. In the angular coordinate systems, the distance covered
by a degree of latitude at the equator remains the same but distance covered by a
degree of longitude differs towards the poles. The present mapping work is based
on cylindrical projection map with angular geographic coordinate system, in which
degree-minute-second (D/M/S) format is used. The coordinates can also be reported
in degree decimal (DD.DDDD) format, in which minutes and seconds are represented
as a fraction of one degree.
The Earth’s surface is not uniform but varies from mountains to plateaus and
plains. The elevation and depressions in the Earth’s surface are called the physical
642 S. Goswami et al.

features or relief features of the Earth. The map used to show these topographic relief
features is called a relief map. All the topographical maps in India are prepared by
the central agency called Survey of India, under the Department of Science and Tech-
nology. The production of new updated topographical maps is not being generated
after implementation of National Mapping Policy since 19th May 2005. Toposheets
are available on 1:250,000; 1:50,000 and 1:25,000 scales as Open Series Maps (OSM)
for the use of general public/civilians for supporting development activities in the
country. These are based on WGS-84 Datum and UTM Projection and do not contain
grid and classified information. They can be obtained from all map sale centre and
other authorized agents deputed all over the country. Defence Series Maps (DSM)
contain significant strategic and classified information so accurately that they belong
to restricted category. Most of the international borders with hilly disputed lands are
kept under such type. DSM is prepared on 1:250,000; 1:50,000 and 1:25,000 scales
for the use of defence forces of India for supporting national security requirements.
The Survey of India is only authorized for preparation and printing of DSM. For sale
and distribution, the responsibilities are assigned to ADGMS (GSGS), Ministry of
Defence.
The fundamental principles of indexing and cartographic design is simple and
to identify specific toposheet of a particular area, the international base map with a
scale of 1: 1,000,000 is (covering land areas only from 4 to 40° N Latitude and 44 to
124° E Longitude) divided into 136 segments with 4° latitude × 4° longitude areas.
The Indian territory comes within segment numbers ranging from 40 to 92 (Fig. 2a).
For further detailed representation, each segment is divided into 16 parts with 1°
latitude × 1° longitude areas (with 1:250,000 scale). These are also called degree
sheets, which again subdivided into 16 subcells of 30 latitude × 30 longitude (with
1:100, 000 scale). Again 15 latitude × 15 longitude segments are numbered from
1 to 16 in a columned manner (with 1:50,000 scale). The 15 × 15 toposheets are
divided into 4 sheets, each of 7.5 and are numbered as NW, NE, SW and SE. The
schematic depiction of the present study area around 56D/11 toposheet is shown for
better visualization (Fig. 2b).

3 Objectives and Principles

The distribution of various lithologic units and geological structures of different times
indicate specific events. The lithostructural map can give insight into the relationships
among three fundamental aspects, i.e., time, space and events. The occurrence of
different rock units suggests a specified province. For example, the greenstone belts
and granitoids cannot be expected inside the Proterozoic Bhima basin until and unless
there is a major deformation. Systematic lithostructural mapping can put constrain on
age, time stratigraphic position and interrelationships among different rock types and
tectonic events, which in turn provide clues to interpret crustal evolution. Prior to the
discovery of radiometric dating geologists used relative dating methods to understand
the possible ages of rocks. Radiometric dating provided an absolute dating method
Fundamentals of Lithostructural Mapping: Example … 643

760E; Toposheet no. 56 800E;


a 68∞E 72∞ 76∞ 80∞ 84∞ 88∞ 92∞ 96∞E b 200N (1:1,000,000) 200N
32
42

37 51 60 A E I M
36∞N

33 38
B F J N
43 90
52

32∞
61
81
C G K O

34 39 91
D H L P
82
44 53
62 71 760E; 800E;
28∞ 77
160N 160N
35 Toposheet no. 56D 770E;
760E;
92 170N (1:250,000) 170N
45
40 54
72 78 83

1 9 13
63

24∞
5

41 84
93
2 6 10 14
46 55
73

64 79

20∞ 3 7 11 15
A E I M

47 B
F J N 4 8 12 16
56 74 85
65
C G K
16∞
O
760E; 770E;
D H L P 160N 160N
Toposheet no. 56D/11
76030’E; (1:50,000) 76045’E;
57
86 16030’N 16030’N
48 66

12∞

NW NE
49 67 87
58

8∞N
SW SE
0 200km 400km 600km

50 59 88 76030’E; 76045’E;
80∞ 84∞
72∞E 76∞ 68 88∞ 92∞E 16015’N 16015’N

Fig. 2 a Basic concept of toposheet formulation over outline map of India (modified after Survey
of India, 1996). b Division principles of toposheets into smaller scales

since the early twentieth century. However, such an expansive and time-consuming
method is not always possible practically when episodic deformation signatures
are clearly distinguishable. So, lithostructural mapping in field itself can give huge
information on relative ages readily.
A brief recap of the following fundamental laws/principles are very helpful in this
regard with present examples.
The Law of Superposition in undeformed sedimentary terrain is always useful.
This law states that in a tectonically undisturbed terrain, the overlying sedimentary
rock layer is always younger than the one beneath it and also older than the one
above it. For the Bhima Group of sedimentary rocks in the study area, the principle
is very well applicable with older conglomerate and younger limestone (Fig. 3a).
The later disturbances like bioturbation, folding and/or faulting can cause ambiguity
(Fig. 3b, c). However, this principle allows sedimentary layers to be viewed as a
form of vertical timeline. The regular order of occurrence of fossils in rock layers,
i.e., the Law of faunal succession is another basic law. This law is not of much
use in our Archaean to Meso Proterozoic terrain in the present context. Although
644 S. Goswami et al.

a Young
b
Limestone

Shale
Sandstone
v
Conglomerate

Basement Dyke/vein
complex Erosion surface

Greenstone/schist belt Old

c e
d F1
Lava flow
F2 Granitoids
Erosion surface

Xenolith

Fig. 3 a Cartoon showing younging direction towards top for the undeformed sedimentary beds.
The limestone unit is youngest and conglomerate is oldest sedimentary unit. Basement complex
indicate greenstone xenolith within granite and dyke/vein cut across granite. Hence successive
younger rocks can be noted as greenstone-granite-vein/dyke. b Diagram showing a folded sequence
where law of superposition is not applicable. c Diagram showing a faulted sequence where law of
superposition is not applicable. d Principle of cross-cutting relationship explains that F 1 is older
than F 2 fault. e Principle of inclusion explains that xenolith is older than host granitoids

the occurrences of microfossils are reported (e.g., Suresh and Sundara Raju 1983),
it is not possible to recognize them with naked eyes. The principle of Uniformi-
tarianism states that the geologic processes observed in operation that modify the
Earth’s crust at present have worked in the same way over geologic time. This prin-
ciple states "present is the key to the past" (Scott 1963). The past geological history
can be explained by presently ongoing observations of geological events. The prin-
ciple of intrusive relationships concerns cross-cutting intrusions. In geology, when
an igneous intrusion (e.g., dyke or vein) cuts across a formation of sedimentary or
igneous/metamorphic rock, it can be determined that the igneous intrusion is younger
than the sedimentary or igneous/metamorphic rock. Therefore, if a dyke/vein is found
to cut across granitoid but not the cover sedimentary rocks, then dyke/vein can be
considered as pre-sedimentary but post granite emplacement event (Fig. 3a). There
may be a number of different types of intrusions including batholiths, plutons, stocks,
laccoliths, lopoliths, sills, dykes and veins. The principle of cross-cutting relation-
ships can be explained in the vicinity of faults. In a sequence, the formations that
are cut by fault are older and those that are not cut must be younger than the fault.
This law is possible to demonstrate in field, where the sedimentary sequence is over-
lain by Deccan lava flow. The topmost lava is not affected by fault but displacement
of older sedimentary rocks indicates post sedimentation but pre-volcanism faulting
(Fig. 3c). Finding the key bed/marker bed may be helpful to determine if the fault is
a normal or reverse fault. In Fig. 3c, shale and sandstone beds act as marker horizons
to characterize the fault as a normal dip-slip fault. From the relative displacements
of different generations of faults (viz. F 1 , F 2, etc.), it is even possible to determine
Fundamentals of Lithostructural Mapping: Example … 645

relative age of faults (Fig. 3d). The principle of inclusions and components states
that the inclusions must be older than the formation that contains them. Presence of
metabasalt xenoliths of older greenstone belt patches is significant in this context
(Fig. 3a, e). In sedimentary rocks, it is common for gravel from an older formation
to be ripped up and included in a newer layer. A similar situation with igneous rocks
occurs when xenoliths are found. These foreign bodies are picked up as magma
or lava flows and are incorporated, later to cool in the matrix. As a result, xeno-
liths are older than the rock which contains them. Another principle regarding the
original horizontality is useful to visualize that the deposition of sediments occurs
as essentially horizontal beds and usually steep dipping beds indicate some sort of
deformation and tilting. Therefore in the case of dipping beds, the down-dip direc-
tion is indicative of successive younger beds inside the basin (Fig. 4a). The principle
of lateral continuity is regarding the lateral persistence of sedimentary layers in all
directions. This can be useful in old sedimentary terrain where part of the rock record
is separated by a valley or other erosional feature, can be assumed to be originally
continuous. The undeformed thick shale beds in the study area often occur as isolated
flat horizontal topped hillocks like the mesa and butte (Fig. 4b). The soil-covered
areas are essentially mapped from the elevation data and lateral correlation of such
remnant hillocks by the use of this principle of lateral continuity.
Another additional aspect is the decorum of fieldwork and mapping. The field-
based work should be done in a responsible manner as per the geological fieldwork
code published by the Geologists’ Association and Finkl (1988).

a Younger beds Younger beds

Grain size reduction

b Surface of erosion Younger beds

Grain size reduction

Fig. 4 Principle of horizontality and lateral continuity for sedimentary beds indicate outcrops
a Before and b After erosion
646 S. Goswami et al.

4 Geology

The position finding on maps is the first important primary task for field geologists,
who must locate themselves to better than 1 mm level accuracy of their correct posi-
tion on the base map and toposheet. Although the GPS can be used for such purposes,
there are some circumstances when GPS signals may be unavailable in deep valleys
or dense forest-covered areas. Thus a prior knowledge of broad regional geolog-
ical setup is always helpful. Therefore, pacing, compass bearing and back-bearing
become useful for locating the position on the base map. The prior information on
fundamental aspects like field behaviours and terrain geology, accessibility, suitable
season for a visit, mode of transport and other logistic support may save time and
help in smooth conduction of the work. We have presented a brief summary of the
regional geology of the terrain followed by local geology in which lithostructural
mapping is carried out. This preliminary idea on broad geological setting in cratonic
level can give the reader an easy to visualize terrain set up.
The Dharwar Craton (DC) with about 350,000 km2 area forms a part of the south
Indian block in Karnataka and Andhra Pradesh states of the Indian shield. Central
Indian Tectonic Zone (CITZ) is separating DC from the north Indian block. The
southern margin of DC is characterized by late Archean charnockitic zone partially
rejuvenated in the Neoproterozoic (Bhaskar Rao et al. 2003; Ghosh et al. 2004). The
DC is subdivided into two blocks (Fig. 5) viz., Eastern Dharwar Craton (EDC) and
Western Dharwar Craton (WDC) with Chitradurga Shear Zone (CSZ) in between
(Swami Nath and Ramakrishnan 1981; Naqvi and Rogers 1987; Ramakrishnan and
Vaidyanadhan 2008). The lithology, metamorphic grade, magmatism and volcano-
sedimentary environment are not exactly similar in EDC and WDC. According to
the established stratigraphy of WDC (after Swami Nath and Ramakrishnan 1981)
~3.4–3.3 Ga old Gorur Gneiss along with ancient ~3.0–3.2 Ga old Supracrustals

N
170
Mapping
area

Chitradurga shear zone

160
Deccan lava

WDC EDC Proterozic basins

Cuddapah basin
Younger granitoids
150
Greenstone belts

50km Gneiss and migmatites

740 760 780

Fig. 5 Regional geological map of Dharwar Craton (Modified after GSI 1994)
Fundamentals of Lithostructural Mapping: Example … 647

of Sargur Group form the basement in the WDC (Swami Nath and Ramakrishnan,
1990; Bhaskar Rao et al. 1991). The Peninsular Gneissic Complex (PGC) of ~3.3–
3 Ga tonalitic–trondhjemitic–granodioritic (TTG) rocks (Peucat et al. 1993; Chad-
wick et al. 1997, 2000; Jayananda et al. 2000) show intrusive and tectonic contact
with older rocks and deformed angular unconformity contact with younger Dharwar
Supergroup. The Late Archaean greenstone belt rocks (Dharwar Supergroup/schist
belt) were deposited during 2.9–2.6 Ga (Taylor et al. 1984; Nutman et al. 1996;
Kumar et al. 1996; Jayananda et al. 2008). The younger potassic granitoid plutons
are ranging in age from 2.7 to 2.5 Ga (Balakrishnan et al. 1990; Friend and Nutman
1991; Nutman et al. 1996; Jayananda et al. 2006).
Unlike WDC, stratigraphy of EDC show greenstone/schist belts equivalent to
Dharwar Supergroup (~2.6–2.8 Ga), interspersed with different granitoid bodies
of ~2.5–2.6 Ga (Dey 2013; Dey et al. 2014, 2017; Jayananda et al. 2000, 2018).
The linear and irregular series of sub-parallel supracrustal greenstone belts in the
northern part exhibit NNW-SSE trend. The present study area is situated in the
EDC where distinctive stratigraphy or marker horizons are not well established.
The Closepet granite (CG) in the western part of EDC is intruded parallel to the
CSZ. EDC comprises several elongated plutons, which are broadly parallel to the
greenstone/schist belts (Narayanaswami 1970). Chadwick et al. (2000, 2007) called
these plutons as Dharwar Batholith (DB). The U–Pb dating of zircon (SHRIMP)
reveals the emplacement age of the DB ranges from 2.7 to 2.5 Ga (Friend and
Nutman 1991; Krogstad et al. 1995; Nutman et al. 1996; Nutman and Ehlers 1998).
Jayananda et al. (2000) have suggested that the plutonic intrusion of K-rich granite
is indicative of Late Archaean juvenile crustal accretion of the EDC. The western
boundary of the EDC is defined by a 200 km wide lithologic transitional zone from
the PGC of WDC to the CG of EDC (Ramakrishnan and Vaidyanadhan 2008). The
CG of 2513 ± 5 Ma and equivalent plutonic emplacements indicate a widespread
Neoarchaean phase of plutonism which is considered to mark as the stabilization age
for the WDC and EDC (Friend and Nutman 1991; Mojzsis et al. 2003).
Platformal Proterozoic sedimentation is observed in Cuddapah, Bima and Kaladgi
basins of DC. However, EDC forms the basement of Bhima and Cuddapah basins only
(Goswami 2019). Along the SW margin of Bhima basin, the lithostructural mapping
gives further detail information on local geology of the area, which is a broadly E–W
rectangular block located in the northern part of the Karnataka state (Fig. 5). This
study area is well connected by south-central railway from Bangalore. The nearest rail
station is Yadgir located about 50 km NE of the study area. Geologically, the Bhima
basin is a linear slightly sinuous NE-SW trending Neoproterozoic basin covering
an area of about 5200 sq km (Fig. 6). Stratigraphy and sedimentation history of the
basin is discussed in several literatures (e.g., Mishra et al. 1987, Kale and Peshwa
1995; Jayaprakash 1999). However, field-based lithostructural analysis is not been
carried out so far in the SW sector due to lack of surface exposures and soil covering
which made the work more challenging.
648 S. Goswami et al.

F F
GEOLOGICAL MAP OF BHIMA BASIN F
Chincholi
5 0 5 10 15Km

17°
F
15' N
Kagna river Bashirabad
W E
F F Sedam
Farhatabad Devantegnur
S

KasturapalleF
F Wadi
F
17°'
Ramtirth
F
F

F Fracture zone
F

LATERITE
INDEX
16°
45' DECCAN TRAP
Kurlagere F F
F Gogi F CHERT BEDS (INTER - TRAPPEANS)
PURPLE SHALE(HARWAL Fm.)
LIMESTONE( KATAMDEVARHALLI Fm.)
SAND STONE/ SHALE(HALKAL Fm.)
LIMESTONE(SHAHABAD Fm)
SHALE
(RABANPALLI Fm)
ARENITE
16°
30' Wajhal DYKE
F
F
YOUNGER GRANITE
YOUNGER SCHIST BELT
F Thirth TintiniF PENINSULAR GNEISS
OLDER SCHIST BELT (SARGUR)
F
F
FAULT
SURFACE RADIOACTIVITY
16°
15' 76°15' 76°30' 76°45' 77° 77°15' 77°30'

Fig. 6 Geological map of Bhima basin. (modified after Kale et al. 1990)

5 Lithostructural Mapping

The mapping area comprises denudational landforms with elevation ranges from
400 to 550 m. The basement igneous complex consists of older patches of Archaean
greenstone belt rocks and two types of plutonic occur as elevated hillocks and irreg-
ular ridges. The younger Late Proterozoic sedimentary rocks are mostly occurring
as horizontal flatbeds which are often discontinuous laterally due to the presence
of intense deformation features like fault zones and erosion. Geological survey of
India has been surveying over the area since a long with different objectives of
mineral exploration. The earlier workers of GSI (e.g., V. R. Venkoba Rao, Field
Session (FS) 1954–55; Srinivasa Rao, FS 1965–66; B.K. Ranganath, FS 1971–72;
S. Narasimha, FS 1971–72; P. Shanmugam and S.K. Hans, 1983–84; D. Sundara-
vanan and T. Krishnappa 1988–89) have provided significant information on geology
including rock types, structures and economic materials. The resources like building
materials including limestone and granite slabs, minerals like spodumene, sulphides
of Cu, Au, Tin, Ruby, etc. are reported earlier.
Therefore, with this background information the mapping works are started with
an ultimate objective specific to analyze tectonic history with reference to the
Dharwarian crustal evolution.
Fundamentals of Lithostructural Mapping: Example … 649

5.1 Methodology

The initial traverse planning is a crucial part of lithostructural mapping. After demar-
cating the study area as a rectangular block systematic grid can be made as per the
terrain configuration and specific requirements based on the scale of study. Presently
1:50,000 scale mapping over about 500 sq km areas may be suitable for square grid
pattern, where traverse spacing and observation intervals within each traverse are
similar. However, often rectangular grid pattern may also be obtained. Practically, it
is almost impossible to maintain a systematic grid of study points as per the planning
due to some logistic constraints and other factors like topographic and anthropogenic
factors, water bodies, steep scarp/slopes, mine sites, highways, wildlife sanctuaries,
etc. (Fig. 7). Therefore, to save time and energy it is better to take the advantage
of Geologic Traverse Generator (GTG), a GIS program for delineation of traverses
strategically for field geologists (Reeves 2015). Although proposed traverse paths
and a series of "must visit" sites are automatically generated, the ultimate decision
should be taken logically by field geologists. After reaching all the locations the
details of elevation must be cross-checked by GPS as well as toposheet and then
all possible field data including lithology and structures must be documented. In
this way after connecting all the location point data, a contour map with optimum
elevation interval (presently 20 m) may be recreated with rock types and structural
data. Such types of maps are of immense help because it is possible to understand
the thickness of litho units instantly and vertical displacement can also be visual-
ized easily. Therefore, the method of traverse planning, contouring, recognition of
rock types and structural data, tracing lithounit across and along traverse direction,
correlation of lithostructural data plots are fundamental base upon which the entire
interpretation and analysis are dependent.
Figure 8 shows the lithostructural map of the studied area where it can be seen that
the eastern part is more structurally disturbed and occupied by basement granitoids
compared to the western part where horizontal undeformed cover sedimentary rocks
of Bhima basin is exposed. Thus, a further detailed study is carried out especially in

Lithocontact (marked Traverse 2


T1 Traverse 1
with red colour)
following the Traverse 3
contour 320m 340m
440m
420m
280m 300m
400m
380m 260m
440m
420m
W.S. 240m
220m
380m 400m 200m
360m 380m
360m
340m
200m
280m

Fig. 7 A hypothetical contour map with traverse plan according to topographic and logistic
constrains. Green band indicate vertical mafic dyke with straight intersection line with topog-
raphy. T1 indicates true thickness of horizontally bedded unit. W. S. is for wildlife sanctuary. Three
traverse lines with accessible traverse points are shown for example
650

0Km 5Km 10Km 15Km 20Km


76°20'E 76°25'E
N
76°30'E 76°35' 76°40'
16°30' 16°30'

Kaladevanahalli
Hebbal Hebbal
Chennur Khurd
Y Buzurg
Wajhal Benakanahalli Feebly developed
50° Gneissic foliation
Devapur 10° 10°
F F 60° L
Talikota 70° 5° Slickenlines
Malleswar F 10°
20° 15°

Devatakala
540m contour
60°
F Chennur 520m contour
70° 70°
Banhatti Hunasagi 10° Tanda
F 10°
500m contour
10°
G 480m contour
70° 5°
Bandepanahalli 10° 70° 460m contour
70°
G 80°
F G 80° 5°
50° 440m contour
F G Wajhal Tanda 70° Bahirpur
F 420m contour
70° 5°
S 80°
Tanda 400m contour
F 20° Siddapur Deccan trap
F Limestone
F F F Shale
Arenite
10°
Basement
F Bahiraur Granitoids
F F Srinivasapur G
X
F G Plunging
Mafic dyke Fault 10° 10°
Kangonhalli antiform
F F Brecciation Shear 16°25'
16°25' 76°35' zone 76°40'
76°20'E

Fig. 8 Contoured lithostructural map along the SW basement of Bhima basin. (This work)
S. Goswami et al.
Fundamentals of Lithostructural Mapping: Example … 651

the eastern sector (Fig. 9) and a cross-section is made to understand the effects of
faulting.

5.2 Interpretation

In this work, the map is prepared on the basis of field observations with a number
of structural data. Besides the field outcrops of different lithounits and their struc-
tural attributes care is also taken on RL data. Thus the map is made in a contoured
base to avoid any scope of misinterpretation on field disposition. For simplicity, the
contour intervals (20 m) are shown in different colours for better visualization of the
effects of topography on structural correlation (elevation range from 400 to 540 m).
The mapped area (N16° 25 to N16° 30 latitude and E76° 15 to E76° 40 longitude)
shows the general undeformed horizontal sedimentary rocks, which are disturbed
and tilted near the fault zones only. The topography is broadly peneplained in nature
with an average elevation of 450 m in general with a gentle slope towards east.
The contours suggest about 520 m elevation near Talikota in the western extreme
and 400 m in the eastern extreme near Benakanahalli area. However, unlike irreg-
ular rough topography, the flat horizontal sedimentary rocks of Bhima basin often
show erosional remnants over basement granitoids as mesa and butte (Fig. 10).
Isolated granitic hillocks of the basement complex are often affected by brittle-
ductile shearing. In the eastern part, dominantly granitoids of different generations
with patches of greenstone/schist belt along with mafic dykes and quartz-pegmatite
veins are common. The NE-SW shear zone is often observed to affect the basement
complex. Basic dyke emplacement and associated calcification in granite is also
evidenced near-fault zone. The central part of the mapped area is affected by E–
W faulting. Wajhal-Hunasagi area exhibits northern downthrown block with basinal
sediments and southern upthrown side dominated by basement complex. The western
part of the map is elevated topographically and covered by basaltic lava of Deccan
Trap over sediments. Isolated block faults are often observed and along such faults,
granitoid exposes as a pop-up block. In the eastern sector, general undeformed sedi-
mentary rocks comprising arenite and shales of Rabanpalli Formation followed by
Shahabad Limestone are horizontal. Therefore the occurrence of younger rocks can
be noted at higher elevation compared to older lithounit. However, there are certain
locations where younger limestone beds occur at relatively lower elevation compared
to older arenites and shales. Since the beds are horizontal the presence of fault must
be the cause of such discrepancies.
0 Km 5 Km 10 Km
652

N
76°30' 76°35' 76°40'
16°30'
L L S
L L S
L L L
L L S S
L L L S G
S S
S G
L L Kaladevanahalli
Hebbal Buzurg Hebbal Khurd G
L L Chennur
G
L Y Wajhal L G G G Benakanahalli Feebly developed
50° Gneissic foliation
G Devapur G 10° 10°
G
60° L L G
L G 70° 5° Slickenlines
G G 10°
15°
L20° G
G G Devatakala
G G
G 60°
A G Chennur Tanda G
70°
520m contour
70°
Hunasagi 10°
G G 10°
500m contour
G 10°
A 70°
480m contour
G G 5°
10° 70° G 460m contour
G G
G 80° 70°
50° 80° 5° 440m contour
G G Wajhal Tanda 70° Bahirpur
S A 420m contour
G 70° 5° G
A G 400m contour
80° 20°
Tanda G
Siddapur D Deccan trap
G A
G G L Limestone
S G
D S G G S Shale
G A
A A Arenite
10° G
G Basement
G A
A Bahiraur Granitoids
A Srinivasapur G G
G G
X G G
A A Mafic dyke Fault 10° 10° Plunging
S Kangonhalli G
G Brecciation Shear zone antiform
16°25' 16°25'
76°30' 76°35' 76°40'

Strike slip
D fault S/A Shear Shear zone
S S Wajhal
500m Srinivasapur A A A/G zone and dyke Strike slip
A A Tanda Wajhal
Wajhal Fault
G G fault village
G G G G G G G G
400m G G G L L L L L L L L L L L
G G G S S S L L
G G G A A A S S S S S S
G G G A A A A A A A
G G G G G G G G G G
G G G G G G
300m
X Y

Fig. 9 Lithostructural map of the eastern sector along Benakanahalli-Wajhal-Devapur tract showing the eastern tip of Wajhal fault in the central part of the map
near Chennur. The N–S cross-section along X–Y can be seen to understand the effects of faulting. (This work)
S. Goswami et al.
Fundamentals of Lithostructural Mapping: Example … 653

Fig. 10 Horizontal beds of Rabanpalli Shale of Bhima basin form mesa and butte

6 Outcome

Based on the lithological and structural attributes of the study area a local strati-
graphic column is made (Table 1). The wide range of rock types of different ages
and significant variation in structural fabrics are described below.

6.1 Rock Types

6.1.1 Basement Complex

Patches of Archaean Greenstone Belt (Mangalur Schist Belt)

At about 20 km ENE of Wajhal village, the southern continuation of the mangalur


schist belt is observed with patchy outcrops that are cross-cut by younger granitic
plutons. Near Bonal village typical outcrops of greenstone belt are examined. The
rocks are mostly dark mafic in nature with grain size ranging from gabbroic to
basaltic. The well-developed foliation planes are mostly striking NNW–SSE with

Table 1 Local stratigraphy


Rock type Age
of the study area
Soil, laterite, boulders Recent
Deccan trap Upper Cretaceous to lower
Eocene
Bhima group of sedimentary Meso-Neoproterozoic
rocks
Mafic dykes Palaeoproterozoic
Closepet granite equivalents Neoarchaean
Peninsular gneiss equivalents Mesoarchaean
Greenstone/schist belts Mesoarchaean
654 S. Goswami et al.

Fig. 11 Outcrop photo of typical metamorphosed schist belt patches. a Apophyses of younger
granite in the metabasics. b Pillow breccia c. Pillow structures preserved within schist belt rock.
d Spilitic nature of meta pillow basalt

a dip amount ranging from sub-vertical to 70° due WSW. Low-grade greenschist
facies metamorphism can be suspected from the presence of chlorite, biotite flakes.
The superposed hook-shaped folding at some outcrops is noted. This greenstone
belt rock unit is the oldest rock type in the area. Apart from these there are several
patches of schist belt rocks occur as xenoliths within younger granitoids as well as
gneissic and migmatitic rocks near Benakanahalli and east of Chennur tanda. Pillow
breccia or, hyaloclastite, spilites, granitic apophyses and veins are very common
in the preserved schist belt rock patches (Fig. 11a–d). The basement complex of
the Bhima basin is characterized by oldest greenstone belts occurring as patches of
schistose meta basic rocks within the later emplaced gneissic-migmatitic complex
and younger phases of granitoids and pegmatites. The granitoids are observed to
show two different characteristics.
Fundamentals of Lithostructural Mapping: Example … 655

Fig. 12 a Typical outcrop of TTG occurs in south of Malleshwar and Talikota. The outcrops are very
much scanty and mostly soil-covered area with faulted contact between TTG and Shahbad limestone
is preserved. b The zoomed view of the preserved fault plane along which TTG and Shahbad
limestone come in contact due to faulting. c Typical migmatitized gneiss of PGC equivalent TTG.
d Zoomed view of TTG with dominant biotite, hornblende rich melonosome and quartzo-feldspathic
leucosome rich portions form net structure

Older TTG of Peninsular Gneiss Equivalent Rocks

The first group of granitoids is medium to coarse-grained and greyish in appear-


ance with relatively more dark minerals like biotite and hornblende. The flat eroded
peneplaned outcrops are often seen below soil cover. Feebly develop NNW–SSE to
N–S gneissic foliation with aligned dark and light minerals are notable. However,
the outcrops of such weakly lineated and foliated granitoid are limited and better
exposed in the NE beyond the study area. However, the areas like Banekenahalli,
Hunasagi, south of Malleshwar and Talikota area exhibit some outcrops of typical
dark TTG of PGC equivalents (Fig. 12a, b). These granitoid vary from dark to light
grey in colour medium grain size and compositionally range from granite to gran-
odiorite with plagioclase feldspar, quartz, biotite and hornblende. These are younger
compared to greenstone belt rocks. The older granitoids are often migmatitized and
feebly developed gneissic foliation and colour banding are common (Fig. 12c, d). This
may be broadly categorized as tonalite-trondjhemite-granodiorite (TTG). However,
further details can be described from petrographic studies under a microscope. The
656 S. Goswami et al.

geochemical characterization of the granitoids will also be carried out from XRF and
REE data.

Younger Granite-Syenite-Pegmatite Rocks

In the study area, the most significant and very common feature in the basement
complex is younger plutonic emplacements cross-cutting older plutons (Fig. 13a,
b). The older granitoids are affected by younger potash feldspar rich pink to brick
red granitoids which are very coarse-grained and emplaced by the forceful intrusion
in the form of vein, injection, plug, etc. Based on the emplacement set up, these
younger granitoids can be categorized into three sub-types as alkali feldspar granite
to quartz syenite plug, pegmatitic veins with graphic texture and pink granite plutons.
Altogether, these three sub-types represents a particular time range possibly formed
after partial melting of older granitoids of TTG composition. These younger intrusive
rocks are mostly equivalent in terms of their emplacement timings and occur as either
injection or along wide fracure zones. These younger granite-syenite-pegmatites
are more common compared to the older granitoids and mostly occur as hillocks
(Fig. 14a). The coarse brick red coloured granitoids are mainly alkali feldspar granite
which is holocrystalline with pheneritic texture (Fig. 14b). Such alkali feldspar gran-
ites are occurring together with other alkali-rich syenite and granitic rocks. Very
coarse-grained pink syenites (Fig. 15a) comprises of > 80% feldspar (mostly micro-
cline and plagioclase) with minor quartz (10–15%) are commonly found along with
red granite. Often pegmatitic appearance is observed near the margin of the syenite
bodies where large interlocking crystals (>3 cm) of quartz and feldspar are common
(Fig. 15b). The intergrowth texture between quartz and alkali feldspar is characterized
as graphic texture in pegmatites (Fig. 16a, b). Presence of this feature is indicative
of simultaneous crystallization of quartz and feldspar.
Apart from these the basement complex is affected by later formed quartz reef/vein
and mafic dykes. These linear intrusive bodies are youngest feature of basement do
not cut basinal sedimentary rocks.

6.1.2 Basinal Sedimentary Rocks

Field observations suggest that general undeformed horizontal to sub-horizontal sedi-


mentary rocks are affected only near the fault zones. The rock units are indicative
of a typical fining upward transgressive sequence comprising basal conglomerate-
grit-arenite-shale sequence of Rabanpalli Formation followed by Shahabad Lime-
stone. Beds are generally horizontal and disturbed near-fault zones where dip amount
ranging from 80° to 60° towards north. The dip amount of the beds gradually die out
away from fault zone. At about 50 m from fault zone dip become 15° and about
70 m from fault zone beds become horizontal. The lithological descriptions in the
following paragraphs are made for the purpose to understand the environment and
associated conditions of sedimentation.
Fundamentals of Lithostructural Mapping: Example … 657

Fig. 13 a Younger brick red coarse alkali feldspar granite emplacement across older granite with
pegmatitic contact near Chinnur Tanda. b Pegmatite vein along fracture in older granite near Wajhal
Tanda

Conglomerate-Grit-Arenite Facies

These facies comprises around 30–50% volume of detrital clasts of granule (~4 mm)
to cobble (~150 mm) size at the bottom part (Fig. 17). The lowermost conglom-
erate lying above basement complex (e.g., granitoids + greenstone belt rocks) is
characterized by matrix-supported polymictic clasts, which are mostly granitic and
quartz as well as cherty in nature. The clasts are mostly moderate to poorly rounded
658 S. Goswami et al.

Fig. 14 a Brick red coarse alkali feldspar granite emplacement cutting older granite and E–W meta
basics near Wajhal Tanda. b Brick red alkali feldspar granite outcrop near Wajhal Tanda

with medium degree of sphericity. Such a poorly sorted conglomerate-grit unit


shows dominantly quartzo-feldspathic matrix indicating low degree of textural as
well as mineralogical maturity hence less transportation. The maximum thickness
of conglomerate-grit unit is ~2 m. The graded nature of this unit is noted from
decreasing order of grain size. Such fining upward transgressive sequence gradually
transformed into arenitic horizon which is about 5 m thick brownish coloured iron
rich sandstone with poorly developed primary sedimentary structures.
Fundamentals of Lithostructural Mapping: Example … 659

Fig. 15 a Coarse pink syenite outcrop near Chinnur Tanda. b Feldspar-quartz interlocking in
pagmatitic syenite body near Chinnur Tanda
660 S. Goswami et al.

Fig. 16 a Hand specimen of graphic texture. b Outcrop of graphic granite-pegmatite

Purple to Red Shale Facies

The shale facies is lying above arenite unit with conformable relation and gradational
transition indicating a relatively deeper quiet environment. The thickness of these
facies is about 110 m as measured from the elevation of an undeformed horizontally
Fundamentals of Lithostructural Mapping: Example … 661

Fig. 17 Fining upward sequence in the section view of conglomerate-grit-arenite facies

bedded flat top hillock occurring as "Mesa". The overlying Shahbad limestone occurs
as cap over shale along flat hilltop. The clay minerals cannot be identified under even
a hand lens. The colour of the shale beds is often exhibit localized variety ranging
from grey spots to greenish-grey tint. However, dominantly purple to red colours
are characteristic of these facies (Fig. 18a, b) possibly due to the occurrence of the
ferric iron (Fe+3 ) oxide mineral hematite (Fe2 O3 ). This is indicative of oxidizing
conditions at the time of hematite formation. Reduction sots are also common as
grey spots within this shale due to possible presence of decomposing organic matter.
Organic matter can reduce the iron and enabling it to be removed in solution, which
results in the removal of the red colouration.
662 S. Goswami et al.

Fig. 18 a Purple/red shale of Rabanpalli Formation and overlying Shahbad limestone. b Possible
organic content and oxidation state of iron (blue shaded region) on the basis of colour. Plot after
Potter et al. (1980)

Fig. 19 a Brecciated Shahbad limestone and chert nodules in Yadhalli. b Dipping limestone beds
with chert nodules near Banahatti

Limestone Facies with Chert Nodules

Milky white to grey limestone of Shahbad Formation is lying conformably above the
shale of Rabanpalli Formation as capping along top of the hillocks. These limestone
beds are a featureless undeformed horizontal unit with occassional chert nodules
formed during diagenesis. Stylolites are common with sharp-pointed zigzag patterns
to box-shaped broad undulations. Near the fault zones, plumose joints are common
in steeper limestone beds along with calcite veins and other conjugates as well as
extensional joints along with strike joints (Fig. 19a, b).

6.1.3 Deccan Lava Flow

In the study area, basaltic lava flows of the Deccan trap occur above Rabanpalli
Formation in the south of Wajhal (Kiriani Gudda) at elevation about 540 m in the
Fundamentals of Lithostructural Mapping: Example … 663

Fig. 20 a Spheroidal weathering in basalt and b. vesicular basalt with amygdaloidals and different
degrees of alteration and oxydation. c Zoomed in view of vesicles and amygdales

southern block of Wajhal fault. In the northern block trap basalt occupy most of the
areas and covering the Shahbad limestone. The flow beds are very well recognized
with characteristic compositional and textural variations. Spheroidal weathering is
common (Fig. 20a) with vesicular and amygdaloidal textures and different types
of alteration with colour banding (Fig. 20b). Yellow to rusty-coloured altered and
weathered basaltic soil contain the mineraloid limonite [∼FeO(OH) · nH2 O], and
some brown coloured minerals goethite (FeOOH).

6.2 Structure

The map (Fig. 8) suggests general undeformed horizontal sedimentary rocks, which
are disturbed and tilted near the fault zones only. In the south of Wajhal Tanda (Fig. 9)
the gritty sandstone occurs at 515 m RL and granites at 500 m RL. However, the
limestone beds occur at about 420 m RL near Wajhal village, which is located about
2.5 km north of Wajhal Tanda. This is indicating about a northern downthrown block
movement of a broadly E–W fault in between two locations (i.e., Wajhal Tanda and
Wajhal). Similarly, the western side of Wajhal Tanda showed the occurrence of gritty
sandstone at about 420 m RL. The clastics of the Rabanpalli Formation are a tran-
gressive fining upward sequence with conglomerate-grit-arenite-shale sequence. The
offset of mafic dykes is indicating a strike-slip movements which is also supported
by the sub-horizontal to low angle pitch (5–15° ) slickenlines on E–W to 110° –290°
fracture planes (Fig. 21a). Therefore parallel right step over pattern of strike-slip
movement lead to pop-up block movement. This is also supported by sub-horizontal
unloading joints (Goswami et al. 2019a) and brick red alkali feldspar granite block
intrusions (Fig. 21b, c). Similar type of block faulting is observed in the western sector
near Bandepanahalli village (Fig. 8), where older granitic basement got uplifted and
come in the same elevation with younger Rabanpalli Shale and Shahbad Limestone
beds. There is a narrow E–W uplifted block along which granite and mafic dyke are
apparently observed to cut through the basinal sediments. The trend of older pre-
sedimentary dyke and post-sedimentary block fault is the same, i.e., E–W. Thus in
this area near Bandepanahalli the dyke is apparently looks like post-sedimentary in
nature. The wajhal-Hunasagi fault is a broadly curvilinear ESE–WSW fault which
664 S. Goswami et al.

Fig. 21 a Slickenside plane developed along 110° /80° => 200° fault plane with epidotization give
greenish tint. Slickenlines due to fault movement varies in plunge from general horizontal to rarely
25° and strike-slip component is dominant regionally. b Unloading joint development in granitoids.
c Pop-up block faulting due to sinistral right stepover pattern of strike-slip fault movements

is ending near east of Chennur village with characteristic tip line faulting (N–S)
associated with it to compensate for the stress distribution. Due to presence of such
tip line faulting the granite/limestone occurrence is noted in the same elevation. The
limestone beds are horizontal in general except near the fault zone, where the steeper
northerly dip (ranging from 30° to 80° ) of limestone beds are noticed near SE of
Wajhal village, NNE of Hunasagi, N of Banahatti (Fig. 22a, b). Doubly plunging
antiforms are also noteworthy near Hunasagi (Fig. 23a, b). The higher dip near-fault
zone is certainly due to fault dragging effects. The mapped area represents several
sub-parallel E–W fault segments instead of a single fault and strike-slip fault arrays
often lead to development of pop-up block movement. Therefore, Wajhal-Talikota
tract is characterized by a number of major E–W faults with apparently northern
downthrown block movements. The western part of the study area is mostly covered
with Deccan lava and the fault might have passed below the lava flow. Minute obser-
vation can give clue towards a fact that there is a progressive southerly offset of
the E–W fault segments as one proceed from east to west of the study area. Further,
there is a new feature seen in the field to demarcate fault even in flat peneplaned areas
without proper outcrops. There is development of lime kankar (Fig. 24) wherever
the fault is observed to cut through limestone. This structural guide is cross-checked
several times in field and based on the occurrences of such kankar at some soil-
covered areas the fault has been traced and extrapolated. The confirmed outcrops of
fault within limestone granite contact also exhibit the same features. Details of origin
of such kankar are explained by Mukherjee et al (2017) in this context.

6.2.1 Structural Data

The geological map with plotted structural data showed variation in deformation
mechanism with time. Table 2 represents a structural datasheet to maintain a system-
atic practice of data utilization. Field observations reveal that the basement rocks
are mostly deformed in a ductile manner and brittle features are developed at later
events during basin formation possibly during rifting and dyke emplacements. Mostly
undeformed horizontal basin-fill sedimentary rocks represent tectonically stable span
Fundamentals of Lithostructural Mapping: Example … 665

Fig. 22 a Steeply dipping limestone beds near-fault zone. b Generalised horizontal limestone beds
in undeformed areas away from fault zone
666 S. Goswami et al.

Fig. 23 a Doubly plunging antiform shows preserved portion of western part of closure. b Easterly
directed closure

Fig. 24 Lime Kankar along


the fault affecting the
limestone

of time. Further reactivation of the faults might have taken place after sediment
depostion and diagenesis specifically during upliftment and exhumation. This latest
deformation event has affected the sedimentary rocks to some extent locally. Thus,
the sedimentary rocks show steeply dipping beds, small scale localized folds, and
displacements near the fault zone.

6.2.2 Ductile Deformation Structures

Ductile features are identified from typical changes in shape without developing
sharp discrete fractures. NNW-SSE foliation development in greenstone belt patches
and gneissic rocks is related to the oldest phase sinistral sharing (Fig. 25a, b). Later,
the shear zone and associated folds and boudins reveal a progressive dextral sharing
Table 2 Structural data sheet
Area Lat/Long T.S. No. Structural data Rock type Remarks
Bonal 16° 30 56D/10 NW-SE Schist belt 140° /70° => 230° Gabbro to basalt Mangalur greenstone belt
31.6"N and later intruded by Granitoid
76° 39 53.2"E granitoid

Wajhal Tanda 56D/11 Fracture 90° –270° , vertical Mafic dyke Cut across all the granitoids
16° 27 18.2"N
76° 32 58"E
Wajhal Tanda 56D/11 N45E dextral 0° –180° , vertical Medium grained En echelon fracture formed
16° 27 18.2"N shearing N60E en echelon grey Granitoids due to NE–SW dextral
76° 32 58"E sharing

Hunsagi north 56D/11 Vertical joint sets Horizontal N–S and N30E joint sets
16° 29 8.7"N limestone with N15E acute bisector
76° 30 37.6"E (σ 1)
Fundamentals of Lithostructural Mapping: Example …

Horizontal bed

Hebbal kurd 56D/11 Fracture zone N45E vertical Brecciated Brittle nature of fracture
16° 28 5.9"N fracture zone of fracture granite with quartz vein/reef
76° 37 0.7"E about 50 m width with quartz reef

Wajhal Tanda 56D/11 Fracture set 40° /65° =>130° Medium grained Radioactivity up to 0.05
16° 27 22.3"N grey granite mR/hr along fracture trend
76° 32 53.2"E

(continued)
667
Table 2 (continued)
668

Area Lat/Long T.S. No. Structural data Rock type Remarks


Wajhal Tanda 56D/11 Fault plane 80° /65° =>170° Medium grained Diagonal slip Fault cut
16° 27 22.3"N Plunging Pitch 15° =>80° grey granite older fractures along N–S
76° 32 53.2"E slickenlines NNE joint set and NNW-SSE trends

Wajhal Tanda 56D/11 Cross cutting sets N–S vertical Coarse alkali Broadly E–W fault is
16° 27 22.3"N of fault and 80° /70° =>230° feldspar granite younger than NNE set and
76° 32 53.2"E fractures 30° /40° =>120° and E–W trend N–S fracture is youngest set
pegmatite

Wajhal 56D/11 Flat limestone No remarkable deformation Shahbad Mostly soil covered area
16° 27 18.2"N Formation
76° 32 58"E
Horizontal bed
Wajhal Tanda 56D/11 Flat grit bed No remarkable deformation Rabanpalle Patchy outcrop surrounded
16° 26 43.8"N Formation by granitoids
76° 33 1.5"E
Horizontal bed
North of 56D/7 Steep beds with Bedding Steeply dipping Rabanpalle/ Shahbad Fm
Banahatti 95° /60° =>5° limestone/shale Contact zone affected by
16° 28 23.5"N Fracture sets Chert nodules, wajhal Fault
76° 27 59.6"E 150° /30° =>240° stylolites and
35° /60° =>125° strike joint
95° /30° =>185° observed
(continued)
S. Goswami et al.
Table 2 (continued)
Area Lat/Long T.S. No. Structural data Rock type Remarks
Yadahalli 56D/10 Brecciated chert Limestone and Chert nodules are broken
16° 31 41.2"N nodules in brecciated chert into pieces like chips of
76° 32 27.9"E limestone nodules biscuit
Horizontal bed
affected by
brecciation
West of wajhal 56D/11 Dipping grit beds Bed 55° /15° =>325° Rabanpalli Fining upward sequence of
Tanda with 2 sets of Fracture sets arenitic conglomerate-grit-arenite
16° 25 48.2"N fracture 30° /60° =>120° formation
76° 31 34.8"E 120° /70° =>210°

SW of wajhal 56D/11 Horizontal bed, Fracture sets Rabanpalli Horizontal sandstone beds
Tanda Three sets of 15° /70° =>285° Formation
16° 25 30.8"N fracture 75° /70° =>345° middle part
76° 32 3.3"E 150° /vertical
Fundamentals of Lithostructural Mapping: Example …

Chinnur Tanda 56D/11 Shared rock with 60° /60° =>250° Brick red to flesh Often pegmatitic with
16° 27 35.9"N two sets of joint 120° /sub-vertical red coarse alkali crystal size above 3 cm
76° 33 48.8"E feldspar granite
to quartz syenite

Chinnur Tanda 56D/11 Shared rock with E–W shear fracture Sheared foliated About 200 m wide
16° 27 35.9"N two sets of joint bounded by N60E quartz syenite to deformation zone with
76° 33 48.8"E dextral share zone alkalifeldspar small scale parallel ENE
granite share zone

(continued)
669
Table 2 (continued)
670

Area Lat/Long T.S. No. Structural data Rock type Remarks


Chinnur Tanda 56D/11 Shared rock with E–W shear fracture Sheared foliated 100° –280° and N50E joint
16° 27 35.9"N two sets of joint bounded by N60E quartz syenite to sets with N75E acute
76° 33 48.8"E dextral share zone alkalifeldspar bisector (σ1)
granite
Chinnur Tanda 56D/11 Shared rock with N–S and N60E shear Sheared foliated Swinging of foliation
16° 27 35.9"N two sets of joint fracture near hinge quartz syenite to planes near hinge areas
76° 33’48.8"E and folding area alkalifeldspar
Plunge of fold axis granite
15° =>210°

Chinnur Tanda 56D/11 Shared rock with Boudinaged fold with Sheared foliated Quartz veins are folded
16° 27’35.9"N two sets of joint N15E axis shows quartz syenite to with thick hinge and thin
76° 33 48.8"E and folding dextral sharing alkalifeldspar limbs
granite

Chinnur Tanda 56D/11 Pink syenite with 60° /sub vertical Sheared foliated Uranium anomaly up to
16° 27 35.9"N N60E joint quartz syenite to 0.07mR/hr for 5 m × 1.5m
76° 33 48.8"E alkalifeldspar dimension
granite
S. Goswami et al.
Fundamentals of Lithostructural Mapping: Example … 671

Fig. 25 a σ type porphyroclast of biotite fish showing sinistral sense of sharing. b Pure biotite rich
layer of migmatite aligned along NNW–SSE
672 S. Goswami et al.

Fig. 26 Boudinaged fold in younger alkali feldspar granite associated with dextral sharing; near
Chennur Tanda. a Sketch; b Outcrop photo; c Schematic diagram shows progressive stages of
deformation

along ENE–WSW (Fig. 26a–c). Net-like migmatite indicative of metatexite, devel-


oped with two or more systematic parallel sets of leucosome/melanosome. The inter-
section of different sets of neosome exhibit a net-like pattern (Fig. 12c, d). These
are formed due to the presence of different sets of extensional shear bands and the
anatectic melt could have migrated from host into the shear bands. There are rootless
folds associated with shear zone where the partial melting and flow of quartz to the
lower stress areas like fold hinge provided complex structures (Fig. 27). The minor
folds within granitoid is tricky job to identify due to very low competence contrast
developed from minor composition contrast and shear plane formation (Fig. 28).
However, there is signature of two different events of sharing as discussed above.

6.2.3 Brittle Deformation Structures

The brittle shearing in the form of fault, mafic dyke and quartz reef emplace-
ment along fracture zone and intense veination and brecciation within the basement
complex (Fig. 29) possibly occurs during initiation of the Bhima basin formation.
All of these phenomenon and features are associated with dominantly extensional
tectonics. Post sedimentation tectonics are also brittle in nature, dominantly occur as
fault and associated drag folding at places. Bedding planes are tilted in the vicinity
Fundamentals of Lithostructural Mapping: Example … 673

Fig. 27 ENE–WSW shear zone and progressive dextral sharing lead to development of thick hinge
thin limb type threefolds, the quartz flows to hinge area and detached limbs produced rootless fold
segments
674 S. Goswami et al.

Fig. 28 Minor folds within granitoids formed due to shearing

of fault zone. At places strike-slip faulting can be evidenced from apparent displace-
ments of mafic dykes and horizontal slickenside lineation (Fig. 21a). Since most
of the soil-covered areas are difficult to interpret the outcrops are only taken into
consideration for structural analysis and wherever possible projection is made.

7 Discussion

Now, the question is "Why should we prepare and study the lithostructural map?"
The answer may be given from different viewpoints including energy and mineral
resource finding, developing geologic hazard protection mechanisms, social and
cultural development and also as a part of research programme. The lithostructural
mapping can also give clues to unravel the crustal evolution, which is one of the
prime field of research nowadays. After considering all the field studies and different
aspects like geomorphology, structure, stratigraphy, deformation events and elevation
contour and map-section analysis following interpretation can be drawn:
1. There is not a single discontinuous E–W fault from Wajhal to Talikota tract, but
a series of sub-parallel fault segments which forms a zone of the fault system.
2. The segments of E–W faults are often associated with block faulting related to
right step over pattern of strike-slip faulting.
Fundamentals of Lithostructural Mapping: Example … 675

Fig. 29 Brittle fracture zone with a NE–SW trending brecciated ridge with intense quartz veination,
brecciation and hematitization

3. E–W faults are not interconnected everywhere but reactivation leads to develop-
ment of doubly plunging folds nearby major E–W trend.
4. Therefore, the upthrown and downthrown blocks are recognized first from RL
data and contact relationships of different lithounits across the fault.
676 S. Goswami et al.

Fig. 30 Granitoids showing apparently developed smooth planes due to E–W faulting

5. Structural data suggest the initial normal faulting lead to tilting of beds. The
broadly E–W fault passes through limestone, shale and granite and develops a
typical planer fabric with northerly dipping planes (dip amount near-fault is 60°
to 70° ). The granitoids are even affected by this faulting and developed apparent
fault plane (Fig. 30).
6. The drag folding and associated doubly plunging folds (Fig. 23a, b) are associated
with reactivated movement in reverse sense. This is supported by other areas like
Chikkalur, where the similar setting is evidenced in the form of well preserved
high angle reverse faults and doubly plunging folds related to Wadi fault. In
fact, the high angle reverse fault cannot be formed in a single compressive stress
event as per Coulomb’s fracture criterion and Anderson theory. Thus, the initially
formed normal fault plane gave a smooth pathway for easier reverse movement
and associated folding.
7. It can be concluded that the faults are reactivated reverse fault with diagonal slip.

8 Crustal Evolution

The most fundamental concept for the geologists is that although the lithostructural
mapping gives ideas on the present-day Earth system (i.e., ~4.6 Ga old), proper
interpretation can give ideas on the evolution of the Earth to its present condition. It
is now understood that the evolution of the dynamic Earth will be continuing to change
its properties. Since, the internal layering of the Earth is characterized on the basis of
Fundamentals of Lithostructural Mapping: Example … 677

different chemical composition (i.e., continental and oceanic crust, mantle and core)
and physical/mechanical properties (i.e., lithosphere, asthenosphere, mesosphere,
outer core and inner core), the nomenclature also varies. The present field-based
works are restricted to the continental crustal surface, where systematic studies on the
structure and rock types give insight into sub-surface dynamics and magma evolution.
Thus, a crustal study in turn gives valuable indirect information on mantle properties.
The crustal evolution is a complex process and there is an interplay of various factors
including formation of proto crust followed by its destruction and renewal. Different
events during Archaean and later time in geological past led to crustal dichotomy,
i.e., the contrast in composition and nature. Complexity in the cratonic rocks is
testimony of significant information on Precambrian crustal evolution. During this
initial phase of the Earth some unique features are developed which can give clues to
understand corresponding events responsible behind their formation (Table 3). Some
of the important examples of oldest crustal signature with its evolution are shown
below:
• So far the oldest zircon (4.4 Ga) age is found in Western Australia’s Jack Hills
(Wilde et al. 2001).
• Oldest rock of 4.03 Ga Acasta gneiss is located in slave craton, Canada (Bowring
and Williams 1999).
• Oldest rock in India is of 3.7 Ga (OMG) Singhbhum Craton (Ghosh et al. 1996;
Mishra et al. 1999).
• Oldest greenstone belt in India is 3.51-Ga Southern Iron Ore Group, Singhbhum
Craton (Mukhopadhyay et al. 2008).
• Gorur gneiss (3.4 Ga) is the oldest rock of Dharwar Craton (Beckinsale et al.
1980, 1982).
• Earliest signs of life on Earth are recorded from the Isua greenstone belt
western Greenland (3.8 Ga) (based on stable carbon isotope ratio) (Mojzsis et al.
1996; Abigail 2016; Roland 2016).
• The oldest known fossil is domical stromatolite (3.5 Ga) of Western Australia,
Dresser Formation, Warrawoona Group, Pilbara Supergroup (Walter et al.

Table 3 Geochronologic event sequence with Dharwarian crustal evolution


Eon Events
Hadean (4.6–3.8 Ga) Initial molten Earth, dominant green house gases, unstable crust,
extraterrestrial bombardment, volcanism
Archaean (3.8–2.5 Ga) Hydrosphere, green house gas reduction, thin crust, greenstone
belts, algoma type BIF, cyano-bacteria, partial melting, primitive
TTG, shield, craton
Proterozoic (2.5–0.54 Ga) Modern plate formation, LIP, underplating,
differentiation-fractionation, potassic granite, thick crust, mobile
belt, deformation-metamorphism, supercontinent and drifting,
surficial processes, oxygenation-ozone, eukaryotic creatures, red
beds, superior type BIF, wide basin
678 S. Goswami et al.

1980; Buick et al. 1981, 1995; Groves et al. 1981; Buick 2001; Van Kranen-
donk et al. 2003, 2004)
• Oldest fossil in Dharwar Craton is stromatolites of Vanivilas Formation,
Chidradurga Group, Dharwar Supergroup (2.6 Ga) (Srinivasan et al. 1989, 1997).
• Recently dated oldest crustal material in India is 4.24–4.03 Ga xenocrystic zircons
from OMTG of the Champua area, Singhbhum Craton (Chaudhuri et al. 2018).
• Oldest paleosol in India is the pyrophyllite-bearing paleosol (3.29–3.02 Ga)
Keonjhar, Singhbhum Craton (Mukhopadhyay et al. 2014).
The inconsistency in rock record is explained by different orogeny and associ-
ated deformation events to interrupt the crustal development. Clearly distinguish-
able ~3 Ga TTG of Peninsular Gneissic Complex (PGC) equivalents were derived
from partial melting of older amphibolite supracrustals possibly in the presence
of water derived from hydrous minerals (Rapp and Watson 1995). However, with
gradual cooling of the Earth during Archaean, water vapour got condensed to form
hydrosphere. There is another opinion on the origin of hydrosphere by asteroid
impact (Daly and Schultz 2018). Anyhow, carbon dioxide (CO2 ) content started
decreasing by precipitation of carbonates at the bottom of global ocean. The pres-
ence of dolomites at places in EDC greenstone belts supports the same (Goswami
et al. 2016, 2017). The air was mostly nitrogen-rich at that time and BIF horizons
are limited in dimension due to limited oxygen. It is natural that the Earth surface
became solid but, the interior was still hot and liquid. The initial thin crust could
not produce wide and deep ocean basins. In fact, major fault could break apart the
crust of that time. The Algoma-type BIFs in the EDC greenstone belts were formed
in fault-bounded narrow basin over older crust. Minor detrital sedimentary rocks in
these sequences are texturally and mineralogically immature, indicating negligible
transportation. Mostly chemogenic sediments like carbonates, BIF and chert at the
top are indicative of significant event of oxygen origin. The growth of EDC in penin-
sular shield started with frequent volcanism which produced elements like Fe, Si,
etc. Archaean ocean contained much volcanic-derived ferrous iron (Fe2+ ), which got
deposited as hematite (Fe2 O3 ) in BIFs. The oxygen that combined the ferrous iron
was possibly provided as a waste product of cyanobacterial metabolism. These initial
volcanic islands were the only land surface with evidence of oxidation and precipita-
tion of BIF. The solar radiation might have produced initial organic compounds from
methane and ammonia by photodissociation. Presence of many types of filamentous
microfossils of 3.45 Ga in the cherts and evaporites of the Pilbara region suggest
that photosynthesis had begun to release oxygen into the atmosphere by that time.
The presence of fossil molecules in the cell walls of 2.5-billion-year-old blue-green
algae (cyano-bacteria) establishes the existence of rare oxygen-producing organisms
by that period. Therefore the cyanobacteria or blue-green algae is responsible for
Algoma type BIF by oxidation of iron. In this way removal of greenhouse gases like
CO2 and the gradual cooling of the atmosphere leads to some changes in atmospheric
conditions. Sedimentary rocks in greenstone belts of EDC give valuable information.
At present, most carbonates and oxidized red soils are being deposited within 30° ,
Fundamentals of Lithostructural Mapping: Example … 679

phosphorites within 45° and evaporites within 50° latitudes. The tillites may repre-
sent glaciation or polar latitude. Therefore the sedimentary rocks like carbonates,
BIF and sandstones of EDC greenstone sequence suggest continental drifting by
plate movements within limit.
It is expected that the initially developed crust was too thin to tolerate much defor-
mation and/or extraterrestrial impacts. Thus outpouring of mantle-derived ultramafic
lava along the broken crust led to the development of oldest greenstone belts that could
not survive due to sinking and partial melting. This repetitive process of lava erup-
tion over initial thin crust, its cooling-sinking-partial melting and associated primary
differentiation is responsible for the development of ultramafic-mafic-TTG-granite-
felsic volcanic rock assemblage in the oldest part of EDC. TTG to granite develop-
ment is of course a Late Archaean event when enough crustal thickness (~15–20 km)
was achieved by the above described repetitive vertical accretion process.
Till the end of Meso Archaean, perhaps greenstone volcanic complex, mafic
plutonics, and TTG were the dominant igneous rocks of EDC. During this time
the crust achieved sufficient thickness and it was not so easy anymore to break the
crust into pieces. Thus, mantle could not be directly exposed to the surface but
reside below the crust. Upper part of such mantle become part of the lithosphere and
got enough residence time for further differentiation to form different mechanical
layering with characteristic physico-chemical properties. The temperature difference
among semisolid layers led to generation of two types of convection cells i.e., lower
mantle to upper mantle and mesosphere to asthenosphere (Montelli et al. 2004). Thus
mantle plume activity and upwelling of mantle caused extension and stretching of
lithosphere to develop intense fractures/cracks across the crust, which became the
pathway of dyke swarms. Thus, instead of direct breaking the thicker crust exhibited
dyke swarms and large igneous provinces. Further, extension led to development of
normal faults and initiation of rifting to create larger and deeper basins. The Cuddapah
basin and its basement complex is a classic example of this entire phenomenon in
EDC.
The array of normal faulting followed by rifting and at last creation of divergent
plate boundary by breaking down the plate due to convective motions of mantle mate-
rials is a sequential event of later crustal evolution. This convection is responsible for
intraplate hot spots as well as the plate motions around the Earth surface and there-
fore opening and closing of ocean basins, shifting of continents, and development
of mountains. In the EDC such plate tectonics started possibly by Late Archaean
Kenoran orogeny (Aspler et al. 2001) and suturing of EDC and WDC took place
along with juvenile Closepet granite formation. Continental crust development was
accelerated after formation of lithospheric plates during the late Archaean. Since our
discussion is restricted till Archaean, later tectonic history is beyond the scope of
this chapter.

Acknowledgements Thanks to Soumyajit Mukherjee (IIT Bombay) for an invitation, internal


review, and handling this article. We are indebted to the director and all the colleagues in the Atomic
Minerals Directorate (AMD) for support. Thanks to Marion Schneider, Annett Buettener, Boopalan
680 S. Goswami et al.

Renu, Alexis Vizcaino, Doerthe Mennecke-Buehler, and the proofreading team (Springer). Dutta
and Mukherjee (2021) summarize this article.

References

Abigail, A. (2016). Geology: Evidence of life in Earth’s oldest rocks. News and Views 31 August.
Nature.
Aspler, L. B., Wisotzek, I. E., Chiarenzelli, J. R., Losonczy, M. F., Cousens, B. L., McNicoll, V.
J., et al. (2001). Paleoproterozoic intracratonic basin processes, from breakup of Kenorland to
assembly of Laurentia: Hurwitz Basin, Nunavut. Canadian Sedimentary Geology, 287, 141–142.
Balakrishnan, S., Hanson, G. N., & Rajamani, V. (1990). Pb and Nd isotope constraints on the origin
of high Mg and tholeiitic amphibolites Kolar Schist Belt, South India. Contributions Mineralogy
Petrology, 107, 279–292.
Barnes, J. W., & Lisle, R. J. (2004). In Basic geological mapping (4th ed.). The Atrium, Southern
Gate, Chichester, West Sussex PO19 8SQ, England:Wiley. ISBN 0-470-84986-X.
Beckinsale, R. D., Drury, S. A., & Holt, R. W. (1980). 3360 M. yr. old gneisses from the south
Indian Craton. Nature (London), 283, 469–470.
Beckinsale, R. D., Reeves-Smeith, G., Gale, N. H., Holt, R. W., & Thompson, B. (1982). Rb-
Sr and Pb-Pb whole rock isochron ages and REE data for the Archaean gneiss and granites,
Karnataka State, South India; Indo-Us Workshop on the Precambrian of South India (Vol. 35–36).
Hyderabad, India: National Geophysical Research Institute abs.
Bhaskar Rao, Y. J., Naha, K., Srinivasan, R., & Gopalan, K. (1991). Geology, geochemistry and
geochronology of the Archaean Peninsular Gneiss around Gorur, Hassan District, Karnataka.
Indian Journal of Earth System Science, 100(4), 399–412.
Bhaskar Rao, Y. J., Janardhan, A. S., Vijaya Kumar, T., Narayana, B. L., Dayal, A. M., Taylor, P.
N., et al. (2003). Sm-Nd model ages and Rb-Sr isotopic systematics of charnockites and gneisses
across the cauvery shear zone, southern India: Implications for the Archaean-Neoproterozoic
terrane boundary in southern Granulite Terrain. Geology Society of India Memories, 50, 297–317.
Billi, A., & Fagereng, A. (2019). Problems and solutions in structural geology and tectonics. In
Developments in structural geology and tectonics series. Series Editor: Soumyajit Mukherjee.
Elsevier ISBN: 978-0-12-814048-2.
Bowring, S. A., & Williams, I. S. (1999). Priscoan (4.00–4.03 Ga) orthogneisses from northwestern
Canada. Contributions to Mineralogy and Petrology, 134, 3–16.
Buick, R., Dunlop, J. S. R., & Groves, D. I. (1981). Stromatolite recognition in ancient rocks: an
appraisal of irregular laminated structures in an early Archean chert–barite unit from North Pole,
Western Australia. Alcheringa, 5, 161–181.
Buick, R., Groves, D. I., & Dunlop, J. S. R. (1995). A biological origin of described stromatolites
older than 3.2 Ga: Comment. Geology, 23, 191.
Buick, R. (2001). Life in the Archean. In D. E. G. Briggs & P. R Crowther (Eds.), Paleobiology II.
blackwell science (pp. 13–21), London, UK:Oxford Press.
Chadwick, B., Vasudev, V. N., & Hegde, G. V. (1997). The Dharwar craton, southern India, and its
Late Archaean plate tectonic setting: current interpretations and controversies. Indian Academy
of Science (Earth and Planetary Science) Proceedings, 106(4), 1–10.
Chadwick, B., Vasudevan, V. N., & Hegde, G. V. (2000). The Dharwar craton, southern India,
interpreted as the result of late Archaean oblique convergence. Precambrian Research, 99, 91–
111.
Chadwick, B., Vasudev, V. N., Hegde, G. V., & Nutman, A. P. (2007). Structure and SHRIMP U/Pb
zircon ages of granites adjacent to the Chitradurga schist belt: Implications for Neoarchaean
convergence in the Dharwar craton, southern India. Journal Geological Society of India, 69,
5–24.
Fundamentals of Lithostructural Mapping: Example … 681

Chaudhuri, T., Wan, Y., Mazumder, R., Ma, M., & Liu, D. (2018). Evidence of Enriched, Hadean
Mantle Reservoir from 4.2–4.0 Ga zircon xenocrysts from Paleoarchean TTGs of the Singhbhum
Craton, Eastern India. Scientific Reports, 8, 7069. https://doi.org/10.1038/s41598-018-25494-6
Daly, R. T., & Schultz, P. H. (2018). The delivery of water by impacts from planetary accretion to
present. Science Advances, 4(4): eaar2632. https://doi.org/10.1126/sciadv.aar2632
Dutta, D, & Mukherjee, S. (2021). Introduction to Structural Geology and Tectonics Field Guide-
book—Volume 1. In S. Mukherjee (Ed.), Structural Geology and Tectonics Field Guidebook—
Volume 1. Switzerland AG: Springer. Cham. pp. xi-xvi. ISBN: 978-3-030-60142-3.
Dey, S. (2013). Evolution of Archaean crust in the Dharwar craton: The Nd isotope record.
Precambrian Research, 227, 227–246. https://doi.org/10.1016/j.precamres.2009.07.002.
Dey, S., Nandy, J., Choudhary, A. K., Liu, Y., & Zong, K. (2014). Origin and evolution of granitoids
associated with the Kadiri greenstone belt, eastern Dharwar craton: A history of orogenic to
anorogenic magmatism. Precambrian Research, 246, 64–90.
Dey, S., Halla, J., Kurhila, M., Nandi, J., Heilimo, E., & Pal, S. (2017). Geochronology of Neoar-
chaean granitoids of the NW eastern Dharwar craton: Implications for crust formation. Geological
Society London Special Publications, 449(1), 89–121.
Finkl C. W. (1988) Geological fieldwork, codes. In General geology. Encyclopedia of earth science.
Boston, MA:Springer
Friend, C. R. L., & Nutman, A. P. (1991). SHRIMP U-Pb geochronology of the Closepet granite
and Peninsular gneisses, Karnataka, South of India. Journal of the Geological Society of India,
38, 357–368.
GSI reports of various field session (e.g., V. R. Venkoba Rao, FS 1954–55; Srinivasa Rao, FS 1965–
66; B.K. Ranganath, FS 1971–72; S. Narasimha, FS 1971–72; P. Shanmugam and S.K. Hans,
1983–84; D. Sundaravanan and T. Krishnappa, 1988-89).
Geological Survey of India (1994). Geological map of Dharwar Craton, Vasundhara Project.
Ghosh, D., Sarkar, S. N., Saha, A. K., & Ray, S. L. (1996). New insights on the early
Archaean crustal evolution in eastern India: Re-evaluation of lead-lead, samarium-neodymium
and rubidium-strontium geochronology. Indian Minerals, 50, 175–188.
Ghosh, J. G., deWit, M. J., & Zartman, R. E. (2004). Age and tectonic evolution of Neoproterozoic
ductile shear zones in the Southern Granulite Terrain of India, with implications for Gondwana
studies. Tectonics, 23 (TC3006). https://doi.org/10.1029/2002tc001444
Goswami, S., Sivasubramaniam, R., Bhagat, S., Kumar, S., & Sarbajna, C. (2016). Algoma type BIF
and associated submarine volcano-sedimentary sequence in Ramagiri granite-greenstone terrain,
Andhra Pradesh India. Journal of Applied Geochemistry, 18(2), 155–169.
Goswami, S., Upadhayay, P. K., Bhattacharjee, P., & Murugan, M. G. (2017). Tectonic setting
of the Kadiri schist belt, Andhra Pradesh. Indian Acta Geologica Sinica-english edition, 91(6),
1992–2006.
Goswami, S., Maurya, V. K., Tiwari, R. P., Swain, S. & Verma, M. B. (2019a). Structural analysis
of T. Sundupalle greenstone belt and surrounding granitoids, Andhra Pradesh, India. Arabian
Journal of Geosciences. https://doi.org/10.1007/s12517-019-4793-2
Goswami, S. (2019). Structural attributes of schist-migmatite-gneiss complex of the East Dharwar
Craton, India (abstract). National seminar on strategic mineral exploration for sustainable
development: Emerging trends and challenges, 5th–6th April, Bangalore.
Groves, D. I., Dunlop, J. S. R., & Buick, R. (1981). An early habitat of life. Scientific American,
245, 64–73.
Jayananda, M., Moyen, J. F., Martin, H., Peucat, J. J., Auvray, B., & Mahabaleshwar, B. (2000). Late
Archean (2550–2520 Ma) juvenile magmatism in the eastern Dhar-war craton, southern India:
Constraints from geochronology, Nd–Sr isotopes andwhole rock geochemistry. Precambrian
Research, 99, 225–254.
Jayananda, M., Chardon, D., Peucat, J. J., & Capdevila, R. (2006). 2.61 Ga potassic granites and
crustal reworking in the western Dharwar craton, southern India: Tectonic, geochronologic and
geochemical constraints. Precambrian Research, 150, 1–26.
682 S. Goswami et al.

Jayananda, M., Kano, T., Peucat, J. J., & Capdevila, R. (2008). 3.35 Gakomatiite volcanism in
the western Dharwar craton Southern India: Con-straints from Nd isotopes and whole rock
geochemistry. Precambrian Research, 162, 160–179. https://doi.org/10.1016/J.precamres.2007.
07.010.
Jayananda, M., Santosh, M., & Aadhiseshan, K. R. (2018). Formation of Archean (3600–2500 Ma)
continental crust in the Dharwar Craton, southern India. Earth-Science Reviews, 181. https://doi.
org/10.1016/j.earscirev.2018.03.013.
Jayaprakash, A. V. (1999). Evolutionary history of bhima basin. In V. S. Kale & d R. H. Sawkar
(Eds.), Field workshop on ‘integrated evaluation of the Kaladgi and Bhima basins (pp. 23–28).
Bangalore: Geological Society of India, Abstract volume (compilers).
Kale, V. S., Mudholkar, A., Phansalkar, V., & Peshwa, V. V. (1990). Stratigraphy of the bhima group.
Journal of the Palaeontological Society of India, 35, 91–103.
Kale, V. S., & Peshwa, V. V. (1995). Bhima Basin (p. 142). Bangalore: Geological Society of India.
Krogstad, E. J., Hanson, G. N., & Rajamani, V. (1995). Sources of continental magmatism adjacent
to the late Archaean Kolar Suture zone, South India: Distinct isotopic and elemental signature of
two late Archaean magmatic series. Contributions to Mineralogy and Petrology, 122, 159–173.
Kruhl, J. H. (2017). In Drawing geological structures. Wiley, Inc. Series: Geological field guide;
7180. ISBN 9781119387237
Kumar, A., Bhaskar Rao, Y. J., Sivaraman, T. V., & Gopalan, K. (1996). Sm-Nd ages of Archaean
metavolcanics of the Dharwar craton, South India. Precarnb. Research, 80, 206–215.
Lisle, R. J. (2004). In Geological structures and maps: A practical guide (3rd ed.). Elsevier
Butterworth-Heinemann. ISBN 0 7506 5780 4
McClay, K. R. (2007). In The mapping of geological structures. Wiley. ISBN-13 978-0471-93243-7
(P/B)
Mishra, R. N., Jayaprakash, A. V., Hans, S. K. & Sundaram, V. (1987). Bhima group of upper
proterozoic-A stratigraphicpuzzle. Memoirs Geological Society of India, 6, 227–237.
Mishra, S., Deomurari, M. P., Wiedenbeck, M., Goswami, J. N., Ray, S., & Saha, A. K. (1999).
207Pb/206Pb zircon ages and the evolution of the Singhbhum Craton, eastern India: An ion
microprobe study. Precambrian Research, 93, 139–151.
Mojzsis, S. J., Arrhenius, G., Mc Keegan, K. D., Harrison, T. M., Nutman, A. P., & Friend, C.
R. (1996). Evidence for life on earth before 3800 million years ago. Nature, 384(6604), 55–59.
https://doi.org/10.1038/384055a0.hdl:2060/19980037618.PMID8900275.
Montelli, R., Nolet, G., Dahlen, F. A., Masters, G., Engdahl, E. R., & Hung, S. H. (2004). Finite-
frequency tomography reveals a variety of plumes in the mantle. Science, 303(56), 338–343.
Mukherjee, A., Bhattacharjee, P., Natarajan, V., Bhatt, A. K., Zakaulla, S., & Rai, A. K. (2017). Lime
kankar as surface signature of concealed dykes-A guide to borehole planning for uranium explo-
ration in Rachakuntapalle-Giddankivaripalle area, southern Cuddapah basin, Andhra Pradesh,
India. Journal of Geosciences Research, 1, 107–113.
Mukhopadhyay, J., Beukes, N. J., Armstrong, R. A. Zimmermann, U., Ghosh, G., & Medda, R.
A. (2008). Dating the oldest greenstone in India: A 3.51-Ga precise U-Pb SHRIMP zircon age
for Dacitic Lava of the southern iron ore group, Singhbhum Craton. Journal Geology 116(5),
449–461.
Mukhopadhyay, J., Crowley, Q., Ghosh, S., Ghosh, G., Chakrabarti, K., Misra, B., et al. (2014).
Oxygenation of the Archean atmosphere: New paleosol constraints from eastern India. Geology,
43(6). https://doi.org/10.1130/g36091.1.
Mojzsis, S. J., Devaraju, T. C., & Newton, R. C. (2003). Ion microprobe U-Pb determinations
on zircon from the late Archaean granulite facies transition zone of southern India. Journal of
Geology, 111, 407–425.
Naqvi, S. M., & Rogers, J. J. W. (1987). Precambrian geology of India (p. 223). New York: Oxford
University Press.
Narayanaswamy, S. (1970). Tectonic setting and manifestation of upper mantle in the Precambrian
rocks of south India (pp. 377–403). NGRI, Hyderabad: Proceedings Symposium on the Upper
Mantle Project.
Fundamentals of Lithostructural Mapping: Example … 683

Nutman, A. P., Chadwick, B., Krishna Rao, B., & Vasudev, V. N. (1996). SHRIMP U-Pb zircon
ages of acid volcanic rocks in the Chitradurga and Sandur Groups and granites adjacent to Sandur
schist belt. Journal of the Geological Society of India, 47, 153–161.
Nutman, A. P., & Ehlers, K. (1998). Evidence for multiple Palaeoproterozoic thermal events and
magmatism adjacent to the broken hill Pb–Zn–Ag orebody, Australia. Precambrian Research,
90, 203–238.
Peueat, J. J., Mahabaleshwar, B., & Jayananda, M. (1993). Age of younger tonalitic magmatism
and granulitic metamorphism in the South India transition zone (Krishnagiri area): Comparison
with older Peninsular gneisses from the GorurHassan area. Jounal Metamorphic Geology, 11,
879–888.
Potter, P. E., Maynard, J. B., & Pryor, W. A. (1980). Sedimentology of shale (p. 306). New York:
Springer.
Ramakrishnan, M., & Vaidyanadhan, R. (2008). Geology of India (Vol. 1, p. 556). India, Bangalore:
Geological Society.
Rapp, R. P., & Watson, E. B. (1995). Dehydration melting of metabasalt at 8–32 kbar: Implications
for continental growth and crust-mantle recycling. Journal of Petrology, 36(4), 891–931. https://
doi.org/10.1093/petrology/36.4.891.
Reeves, R. (2015). Deriving traverse paths for scientific fieldwork with multicriteria evaluation
and path modeling in a geographic information system (Master’s Thesis). University of Southern
California.
Roland, P. (2016). Wavy geenland rock features ‘are oldest fossils. BBC news Retrieved 31st August
2016.
Scott, G. H. (1963). Uniformitarianism, the uniformity of nature, and paleoecology. New Zealand
Journal of Geology and Geophysics, 6(4), 510–527. https://doi.org/10.1080/00288306.1963.104
20063. ISSN 0028-8306
Srinivasan, R., Shukla, M., Naqvi, S. M., Yadav, V. K., Venkatachala, B. S., Uday Raj, B., et al.
(1989). Archaean stromatolites from the Chitradurga schist belt, Dharwar Craton. South Indian
Precambrian Research, 43(3), 239–250.
Srinivasan, R., Pantulu, G. V. C., & Gopalan, K. (1997). Rare earth element geochemistry and Rb-Sr
geochronology of Archaean stromatolitic cherts of the Dharwar craton, south India. Proceedings
Indian Acadamy Sciences (Earth Planetary Science), 106(4), 369–37
Suresh, R., & Sundara Raju, T. P. (1983). Problematic Chuaria, from the Bhima basin. South India.
Precambrian Research, 23(1), 79–85.
Survey of India (2005, May 19). National Mapping Policy. http://www.surveyofindia.gov.in/pages/
display/246-publications
Swami Nath, J., & Ramakrishnan, M. (1981). The early Precambrian supracrustals of southern
Karnataka. Geological Survey of India Memoirs, 112, 350.
Swami Nath, J., & Ramakrishnan, M. (1990). Early Precambrian supracrustals of southern
Karnataka. A present classification and correlation. Geological Society India Memoirs, 19,
145–163.
Taylor, P. N., Chadwick, B., Moorbath, S., Ramakrishnan, M., & Viswanathan, M. N. (1984).
Petrography, chemistry and isotopic ages of Peninsular gneisses, Dharwar acid volcanic rocks
and Chitradurga granite with special reference to the Archaean evolution of the Karnataka craton.
Precambrian Research, 23, 349–379.
Van Kranendonk, M. J., Webb, G. E., & Kamber, B. S. (2003). Geological and trace element evidence
for a marine sedimentary environment of deposition and biogenicity of 3.45 Ga stromatolitic
carbonates in the Pilbara Craton, and support for a reducing Archaean ocean. Geobiology, 1,
91–108.
van Kranendonk, M. J., Collins, W. J., Hickma, N. A. H., & Pawley, M. J. (2004). Critical tests
of vertical versus horizontal tectonic models for the Archaean East Pilbara Granite-Greenstone
Terrane, Pilbara Craton Western Australia. Precambrian Research, 131, 173–211.
Walter, M. R., Buick, R., & Dunlop, J. S. R. (1980). Stromatolites, 3400–3500 Myr old from the
North Pole area, Western Australia. Nature, 284, 443–445. https://doi.org/10.1038/284443a0.
684 S. Goswami et al.

Wilde, S. A., Valley, J. W., Peck, W. H., & Graham, C. M. (2001). Evidence from detrital zircons
for the existence of continental crust and oceans on the Earth 4.4 Gyr ago. Nature, 49(6817),
175–178. https://doi.org/10.1038/35051550. PMID 11196637.
A 3D Photogrammetric Approach
in Mapping Meso-Scale Folds and Shears
in Structurally Controlled Syngenetic
Mn-Mineralised Zones of Shivrajpur
Region, Eastern Gujarat, India

Aditya U. Joshi

Abstract The term mapping means an illustration of an entity/entities present within


a same/different domains to acknowledge its association with each other. This repre-
sentation in the past few decades has given a broad spectrum in terms of its utility and
applicability. The present article offers a simplified text to apply an unorthodox way
of mapping the terrain using a 3D photogrammetric approach. This technique has
emerged though less than a decade has an ability to solve many geological problems
concerned with outcrop accessibility as well as its interpretations. Here, this method
has been used in an abandoned mine section to map the folded outcrop enriched with
manganese mineralisation from the Shivrajpur Formation, Champaner Group, a part
of Upper Aravalli Supergroup. The former segment of the chapter depicts the overall
framework of rocks that belong to the area under investigation and analogous struc-
tures developed within them, while the latter explicit the use of 3D photogrammetry
in meso-scale mapping as well as its interpretations.

1 Introduction

Conventional mapping techniques decode the regional-scale structures. At times,


these methods have been very useful in understanding the dynamics of plate move-
ments. For example, it would have been really difficult to understand style of collision
between the Indian and the Eurasian plates, unless mapping the entire mountain range,
which possess the thrust geometries. Alternatively, any structurally deformed meta-
morphosed terrain may require a justified structural evolution in order to answer
the varieties of structures present within that region. Although these answers are
acquired based on the structural map prepared on large-scales, one would require
supplementary evidences using field photographs and/or microstructural data. In
addition, working on both remote sensing and field structural geology together to
answer common research question has been a powerful method (Misra et al. 2014;

A. U. Joshi (B)
Department of Geology, Faculty of Science, The Maharaja Sayajirao University of Baroda,
Vadodara, Gujarat 390002, India
e-mail: adityaujoshi@gmail.com

© Springer Nature Switzerland AG 2021 685


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_23
686 A. U. Joshi

Dasgupta and Mukherjee 2017, 2019). However, this task becomes a bit tricky in
case of working in less deformed and/or low-grade metamorphites due to its simple
deformation and less complexity. In other words, during the preparation of large-scale
maps, some of the small-scale structures are suppressed, which actually holds a lot
more potential to witness the structural complexity of the entire region. The present
chapter tries to address this problem by providing a solution using 3D photogram-
metric approach. The area selected is an abandoned mine consisting of eye-catching
meso-scale tight isoclinal to open folds of various orders. Furthermore, the present
region is a classical example of structurally controlled Mn mineralisation developed
syngenetically in the rocks of the Shivrajpur Formation of the Champaner Group,
Aravalli Supergroup. The area is located 70 km NE of Vadodara city and lies at the
eastern margin of the Gujarat state (Fig. 1a). The abandoned mine section is popular
in the type locality as “Bhat Mines” situated NE of Shivrajpur town.
The mining history of this area suggests that the mine was initiated by The
Shivrajpur Syndicate Limited, registered under M/s Click Nixon (at Bombay), in
the year 1906, which was responsible to extract through open cast/underground
mining methods and to export high-quality manganese to England (Rasul 1965).
The company progressed till the year 1969, and thereafter, it was handed over to the
Gujarat Mineral Development Corporation (GMDC) Ltd. The GMDC took active
role in Mn production until they faced an obstacle in the early 90s. In the year
1990, the area was declared as Jhambughoda Wildlife Sanctuary, and on account
of uncompromising environmental norms, the Mn mine located near Shivrajpur and
many other regions within the terrain got abandoned. Currently, the corporation is
engaged in old manganese dumps removal in and around the abandoned working
sites.
The Bhat mines can be traced through two entry points. (i) After Bhat Nala Bridge,
2 km interior towards Juni Bhat village; (ii) Another entry through vehicle can be
reached through Shivrajpur—Gujarat State Road Transport Company (GSRTC), old
Bus stand crossing, a road leading towards Hatnimata or Sarasuva. Along this present
road, an old gate heading towards Bamankuva mines routes towards the backside of
the Juni Bhat village (Fig. 1b).
The detail of the overall topography, lithostratigraphy, sedimentation history,
tectonics and metamorphism pertaining to the Champaner Group has been described
in Chap. 18 of this book. Hence, author restricts himself, in the first half by explaining
the nature of Mn ores present within the Champaner region, followed by an in-depth
description of structural features available in the abandoned mine section. The other
half of the chapter deals with the steps for using freely available software for 3D
photogrammetry. This will help to know the method of data acquisition and incorpo-
ration, in order to reconstruct a fold outcrop in 3D. Students can make its optimum
use in their dissertation projects to elucidate structural attributes of folds and will be
helpful to draw noteworthy interpretations from a single outcrop. As in the present
case, the curved axial plane of a folded outcrop and associated brittle shear indicator
can be used to interpret the nature of Mn accumulation along the fold hinges as well
as following the plunge directions.
A 3D Photogrammetric Approach in Mapping Meso-Scale … 687

Fig. 1 Location map of the GMDC abandoned mine section located NE of Shivrajpur, Eastern
Gujarat

2 Nature of Mn Ores Presents Within the Champaner


Region

The Champaner region of eastern Gujarat consists of the highest volume of Mn


deposits found after Central India. The ores found in the present region are mainly
pyrolusite, psilomelane, braunite and winchite (Fermor 1909; Roy 1958). In the case
688 A. U. Joshi

of the lower Aravallis, majority of the groups were studied in great detail by several
workers due to their structure and associated mineralisation (Blanford 1869). The
Champaner Group came into the limelight on account of its manganese deposits,
based on which it was correlated in terms of lithology with the Dharwar rocks of the
South India (Fermor 1909). In his book, “The Manganese ore deposits of India”, he
enumerated the mode of occurrence and the origin of Mn found within the Cham-
paner region. Based on origin, these ores are mainly syngenetic, formed on account
of the accumulation during sediment transport of the Shivrajpur Formation of the
Champaner Group. Another variety of Mn ore, winchite (Mn-rich hornblende), is
developed epigenetically, due to the granitic intrusion, and is mostly restricted to
the skarn rocks of the Champaner region (Joshi and Limaye 2018). Fermor further
classified the ores developed on account of sedimentation and limited metamorphism
as primary ores, whereas those formed due to granitic fluid as the secondary ores.
Beer (1919), in his “Notes on rocks from Pavagad to Dohad”, jotted an important
observation that the Mn ores bodies found at Shivrajpur are structurally controlled
and occur as twisted lenses engulfed in deformed quartzites and slates.

3 Lithological Setting at Bhat Mines

The rocks present within the abandoned mine area belong to the Shivrajpur Formation
of the Champaner Group (Gupta et al. 1997). The major rock types are thick band
of Mn-rich quartzite intercalated with thin bands of Mn-bearing phyllites. There
are in all three mines excavated from the Champaner region, viz. 1. Shivrajpur, 2.
Bamankuva and 3. Pani mines, which consist of a total length of the host >14 km.
The area selected as a type section in this chapter consists of ~2 km in length and
represents a part of Shivrajpur/Bhat mines (Fig. 2). The Bhat mine consists of two
low mounds, one of Mn rich and the other of Mn-devoid host buried under thin soil
cover. The northern part of the Bhat mine is characterised by Mn-rich host and has
been excavated by driving inclined shafts at the elevated regions, while the same
host is extracted by drawing open pits in the valley regions. The inclined shafts are
3–4 m wide, 8–12 m deep, excavated at 15–20° along N275°–285° (Fig. 3). The
open pits located in the valley regions are irregular in shape and vary in size. The
smallest pit can be encountered along the eastern side of the abandoned mine near
Juni Bhat village (Fig. 4), which consist 10 × 05 m (L × W) (longest dimension
measured within the pit), whereas the largest pit has dimensions up to 280 × 118 m.
The southern part of the mine consists of Mn-devoid host or very less amount of Mn
content within these rocks. Though being less productive with regard to metalliferous
deposit, this portion of the mine shows decent evidences in terms of structure and
helps in drawing interpretations through its structural complexity. The entire area has
a large amount of excavated Mn-host dump, mainly restricted to the areas adjacent
to open pits and inclined shafts.
The quartzites of the mine area are fine to medium grained, massive, carbon black
in colour with the presence of botryoidal Mn nodules (Fig. 5). This rock shows the
A 3D Photogrammetric Approach in Mapping Meso-Scale … 689

Fig. 2 Geological map of the abandoned mine section

presence of excellent box work (guides to ore for Mn deposit) (Fig. 6). Occasionally,
the quartzitic bands show limonitic alterations on account of the Fe content (Fig. 7).
The phyllitic band which occurs as intercalations within the quartzites is quartz
dominant, massive, fine grained, buff brown to steel grey in colour and display
prominent primary sedimentary structure such as compositional banding. The rock
also shows good preservation of meso-scale secondary structures in the form of folds
and lineations (Fig. 8).
690 A. U. Joshi

Fig. 3 Field photograph of the inclined shaft excavated along the highlands of the abandoned mine

4 Structural Setting at Bhat Mines

Structurally, the Mn-bearing hosts of the Bhat mines represent two phases of defor-
mation, viz. D1 and D2 , to develop F1 and F2 plunging folds having ESE–WNW and
E–W trends, respectively (Joshi et al. 2018; Joshi 2019a, 2019b; Joshi and Limaye
2020; Mukherjee et al. 2020). The F1 are second-order tight folds with interlimb
angles <25°. These folds are well exposed along the valley regions having trend of
the plunging fold axis towards N280°/20° (Fig. 9). F1 fold hinges are also charac-
terised by the development of fold mullions having similar trend (Fig. 10). There
also exists top-to-down (top-to-N280°) shears in vertical section, at low angle with
the axial planes of the F1 folds that shifted the F1 hinges (Fig. 11). These shears are
represented by well-developed tensional gashes present along the southern limbs of
F1 folds and the same climbs along the F1 fold hinge portions (Fig. 12). The F2 folds
are mainly restricted to the ridges or the highlands and act as an overburden to F1
folds (Fig. 13). F2 folds show the first-order open-type folds with a plunge of 21°
towards N270–275° (Fig. 14). In addition, the transition between F1 and F2 shows
development of fold vergence to develop variety of Z, M and S-folds of lower orders
(Figs. 15 and 16).
A 3D Photogrammetric Approach in Mapping Meso-Scale … 691

Fig. 4 Field photograph of the open pit (punctured Mn-rich quartzite anticline) for the extraction
of metalliferous deposit

5 3D Photogrammetry of Meso-Scale Folds and Shears

In order to reconstruct the meso-scale fold and associated brittle shears in 3D, an
appropriate outcrop depicting all possible structural features has been selected for the
3D photogrammetry. This segment addresses various stages of the 3D photogram-
metric survey: (i) device used and software selection along with their technicalities;
(ii) field operations to guide data acquisition and software operations to understand
the data processing up to 3D model generation.

6 Equipments and Software

Basic handheld camera or cell phone camera with at least 12 megapixel resolution
gives better results. In the present case, iPhone 11 device having 12 MP camera
resolution with f/2.2 aperture for a depth of field equivalent has been used to take the
images. KAISER BAAS XS3 Gimbal is used to click the images at steady condition
(Fig. 17). In order to reconstruct the entire outcrop in 3D using the field images,
Agisoft Metashape Standard (1-month trial version) software is used that can be easily
downloaded through the following website https://www.agisoft.com/. The download
692 A. U. Joshi

Fig. 5 Botryoidal nature of manganese nodules within the host rock

section of the website provides software in two modes; Mode 1 is the standard mode,
which is can be worked out using simple cameras or even with the help of cell phone
cameras. Mode 2 is the professional version with a good accessibility to process the
images taken through drones and helps in geo-referencing a project. Also, one can
generate detail digital elevation model (DEM) through the professional version. The
description to follow discusses the standard mode software, which is easy to handle
and can be used for outcrop-scale interpretations.

7 Field and Software Operations

Obeying the basic rule of photogrammetry, all images needs to be taken at 60%
overlap, at each and every possible angle (Fig. 18). It will be beneficial to snap the
outcrop in the circular flight mode, as if a mapper is taking images with the help of
manual drone. In addition, few images are required to be taken at zoomed in mode,
in order to appreciate the depth dimension of the outcrop.
After downloading the software and installing it within the system, click on start a
free 30-day trial and press ok (Fig. 19a). Once the software is ready to perform the task
go to file/save as, this will help to save the initial file of the project/give desired name to
the project (Fig. 19b) and then go to workflow (available at the top of the taskbar)/add
A 3D Photogrammetric Approach in Mapping Meso-Scale … 693

Fig. 6 Field photograph of box work, guides to ore for manganese

Fig. 7 Limonitic alterations of Mn-rich quartzite bands due to the presence of Fe in the host
694 A. U. Joshi

Fig. 8 Field photograph of Mn-bearing phyllites found intercalated with thick quartzitic bands

photos. Once the photos are added, one can see the number of snaps selected for 3D
construction located at the lower window to the work area (Fig. 19c). The next step is
to align photos (click on workflow/align photos). A small window will appear which
will ask to set the accuracy. Select the accuracy level high and press ok (Fig. 19d).
This process will take more time depending on the speed of the processor and align
photos to appropriate angles. Proper alignment can be achieved by taking appropriate
images in the field. Improper overlaps may lead to problems in the alignment, due to
which 3D reconstruction might be inappropriate. Based on the data generated, it can
be seen in the description section that a total of 46 images were used; out of which
the software has generated 50,194 point cloud data (Fig. 19e). The workspace area
consists of a lot of points, which also includes the noise of the surrounding region. The
next step is to remove noise using “Free-Form Selection tool” (Fig. 19f) by simply
selecting the noise by dragging the cursor and deleting it. (Fig. 19g). Post-deleting
the noise, there exist 34,323 points as actual point data cloud required for the project.
Once all noise is removed, then click on workflow/Build Dense Cloud (Fig. 19h),
select quality as high and press ok (Fig. 19i). When the processing is completed,
click on Model/View Mode/Dense Cloud (Fig. 19j).Comparing both the point, data
set reveals significant addition of points from 34,330 to 10,699,634 (Fig. 19k, l). In
addition, one can view the nature of camera angles set during imaging the outcrop
with the help of show camera mode available at the top within the task bar (Fig. 19m).
In order to give even texture to the entire outcrop, the next step is to build mesh, an
A 3D Photogrammetric Approach in Mapping Meso-Scale … 695

Fig. 9 Tight isoclinal plunging F1 fold, marker standing in the broken F1 hinge

option within the workflow (Fig. 19n). This will allow developing arbitrary 3D surface
of the entire outcrop. Once the mesh has been created, the unnecessary holes can be
filled up by clicking on Tools/Mesh/Close Holes, to 100% and press ok (Fig. 19o, p).
Lastly, the reconstructed model can be exported through File/Export/Export Model
and save it in 3D PDF mode (Fig. 19q, r). The final 3D PDF model is possible to
edit in various software, in order to draw or highlight important components of the
generated model (Fig. 19s). A simple flowchart describing all important steps of the
software operations is given in (Fig. 20).

8 Outcrop Mapping Results

The folded outcrop of Mn-rich quartzite selected for the 3D reconstruction is ~4.2
ft in height from the soil + Mn host dump covered ground, 3–0.8 ft wide and ~14.2
ft long, measured manually (Fig. 21a). The outcrop is broken from few places due
to the want of Mn ore present within; however, the entire structure is in situ and
display majority of the features that lead to probable interpretations with reference
to its genesis. The fold is oriented E–W with its two limbs on the either side, i.e. N
and S, respectively. The southern limb is characterised by the presence of tensional
gashes, a brittle shear sense indicator, which climbs opposite to the plunge direction
696 A. U. Joshi

Fig. 10 Field photograph of fold mullions at the F1 fold hinge

to the hinge portions of the fold (Fig. 21b). Northern limb is devoid of any shear
sense indicators and merely project fractures along with minor limonitic alterations.
The fold closes upwards with its upper layer slightly shifted towards the northern
direction as compared to its lower layers. The fold axis of the upper layer trends
N270°, while the preceding layers in the N280° direction. The overall plunge amount
and interlimb angle of the fold are 20° and 23°, respectively. The axial plane of the
shifted hinge has a dip of ~80° towards N180°, and the normal hinge has a vertical
axial plane. The dip isogons of the upper layers converge towards the inner arc side,
which is tighter than the outer arc of the fold. The overall trace of the hinge line
is curved. The tensional gashes observed along the southern limb are composed of
quartz and provide top-to-down (top-to-N280°) shear direction along vertical section.
A 3D Photogrammetric Approach in Mapping Meso-Scale … 697

Fig. 11 Field photograph showing de-shifting of the F1 hinge (observed along the central lower
portion of the photograph). It needs to be noted that the hinge line is curved in opposite to plunge
direction

9 Outcrop Interpretations/Conclusions

Based on the above structural attributes, qualitative description of the fold geometry
can be given as follows: (i) based on the sense of curvature (Bailey et al. 1939), it can
be treated as antiform; (ii) on the basis of the plunge of the fold axis (Fleuty 1964),
the fold is gently plunging fold; (iii) orientation of the axial plane (Willis and Willis
1929) suggests that the fold is upright to inclined in nature; (iv) to that of the dip of
axial plane (Fleuty 1964), it is a steeply inclined to upright fold; (v) on the basis of
the interlimb angle (Fleuty 1964), it can be termed as tight isoclinal fold; (vi) the dip
isogons method (Ramsay 1967) suggests that the fold belongs to Class 1B; (vii) the
shifting of the hinge at the upper layers of the fold and overall curved hinge line is
on account of the top-to-down (top-to-N280°) shear. This shear is probably syn-D2
and responsible for the Mn mineralised material to accumulate along the shifted fold
hinges and also towards the plunge direction.
698 A. U. Joshi

Fig. 12 Evidences of tensional gashes along the southern limb of F1 folds and same climbing along
the hinges at the up plunge direction

Fig. 13 Panoramic view of Bhat mines showing F2 folds as overburden to F1 folds


A 3D Photogrammetric Approach in Mapping Meso-Scale … 699

Fig. 14 Field photograph showing first order F2 open fold

Fig. 15 Transition between F1 and F2 showed by fold vergence. Clockwise rotation has resulted
into “Z”-shaped fold
700 A. U. Joshi

Fig. 16 Lower-order “M”-shaped fold developed at the core portions of preceding layers

Fig. 17 Photograph of the devices used in the project. iPhone 11 cell phone for photography
mounted on KAISER BAAS XS3 Gimbal for clicking stable pictures
A 3D Photogrammetric Approach in Mapping Meso-Scale … 701

Fig. 18 Field photograph of the outcrop selected for the project. In order to map the outcrop images
with 60% overlap have been clicked from all directions in the circular flight mode manner
702 A. U. Joshi

Fig. 19 a Steps to work out the software: Step to select free 30-day trail. b Saving initial file of
the project. c Added photos in the software for 3D outcrop reconstruction. d Align photos step by
selecting accuracy level to high. e 50,194 point cloud data representation in the left window generated
by 46 images. f “Free-Form Selection tool” used in removal of noise. g Selection procedure of noise
point cloud data (dragging the cursor over unwanted points added on account of background noise
and press delete). h Procedure to build dense cloud data, after noise removal. i Selection of quality
as high in order to complete the task in stipulated time (Note. Selection of highest quality will take
a lot of time to develop dense cloud data and not advisable for low configured systems). j Processed
dense cloud data can be viewed in the view mode by clicking on dense cloud. k Comparison of
point data cloud and l dense data cloud, drastic increments in points from 34,330 to 10,699,634.
m Image capturing positions and angles note that there is 50–60% overlap in all images; in some
cases, it is up to 80%. n Build Mesh step to develop arbitrary 3D surface of the entire outcrop.
o Step to fill unnecessary open spaces within the reconstructed outcrop by selecting the close holes
p option to 100%. q, r Procedure to export the reconstructed model to the 3D PDF mode. s Final
output of the 3D PDF of the reconstructed model
A 3D Photogrammetric Approach in Mapping Meso-Scale … 703

Fig. 19 (continued)
704 A. U. Joshi

Fig. 20 Simplified flowchart depicting various stages of software operations


A 3D Photogrammetric Approach in Mapping Meso-Scale … 705

Fig. 21 a, b Image of a 3D virtual outcrop used for demonstrating the structural attributes

Acknowledgements The author is highly indebted to the device sponsors (Gimble support received
from) Mrs. Soniya and Mr. Omkar Joshi, New Zealand. The author is grateful to the Chief Conser-
vator of Forest, Government of Gujarat, to give necessary permission to work within the limits
of the sanctuary area. Valuable suggestions from Prof. M. A. Limaye (MSU, Baroda) are duly
acknowledged. AUJ is thankful to Dr. Soumyajit Mukherjee (IIT, Bombay) and an anonymous
reviewer for providing valuable comments for the improvement of the article. Thanks to Marion
Schneider, Annett Buettener, Boopalan Renu, Alexis Vizcaino, Doerthe Mennecke-Buehler and the
proofreading team (Springer). Dutta and Mukherjee (2021) encapsulate this work.

References

Bailey, E. B., Weir, J., & McCallien, W. J. (1939). Introduction to geology. London: Macmillan.
Beer, E. J. (1919). Notes on rocks from Pavagad to Dohad. Transactions on Mining and Geological
Institute of India, 13, 73–127.
Blanford, W. T. (1869). On the geology of the Taptee and lower Nerbudda valley and some adjoining
districts. Memoirs of the Geological Survey of India, 6(3), 163–207.
706 A. U. Joshi

Dasgupta, S., & Mukherjee, S. (2017). Brittle shear tectonics in a narrow continental rift:
Asymmetric non-volcanic Barmer basin (Rajasthan, India). The Journal of Geology, 125,
561–591.
Dasgupta, S., & Mukherjee, S. (2019). Remote sensing in lineament identification: Examples from
western India. In A. Billi, & A. Fagereng (Eds.) Problems and solutions in structural geology and
tectonics. Developments in structural geology and tectonics book series. (Vol. 5, pp. 205–221)
Series Editor: S. Mukherjee. Elsevier. ISSN: 2542-9000. ISBN: 9780128140482.
Dutta, D., Mukherjee, S. (2021). Introduction to Structural Geology and Tectonics Field Guide-
book—Volume 1. In S. Mukherjee (Ed.) Structural Geology and Tectonics Field Guidebook—
Volume 1. Springer Nature Switzerland AG. Cham. pp. xi-xvi. ISBN: 978-3-030-60142-3.
Fermor, L. L. (1909). The Manganese ore deposits of India. Memoirs of the Geological Survey of
India, 37.
Fleuty, M. J. (1964). The description of folds. Proceedings of the Geologists’ Association, 75,
461–492.
Gupta, S. N., Arora, Y. K., Mathur, R. K., Iqbaluddin, Prasad, B., Sahai, T. N., & Sharma, S. B.
(1997). The Precambrian geology of the Aravalli region, southern Rajasthan and NE Gujarat.
Memoirs of the Geological Survey of India, 123, 1–262.
Joshi, A. U. (2019a). Fold interference patterns in Meso-Proterozoic Champaner fold belt (CFB)
Gujarat, western India. Journal of Earth System Science. https://doi.org/10.1007/s12040-019-
1075-z.
Joshi, A. U. (2019b) Structural evolution of Precambrian rocks of Champaner Group, Gujarat,
western India. Unpublished Ph.D. Thesis, The Maharaja Sayajirao University of Baroda, pp. 1–
190.
Joshi, A. U., & Limaye, M. A. (2018). Rootless calc-silicate folds in granite: An implication towards
syn to post plutonic emplacement. Journal of Earth System Science, 127(5):ID67.
Joshi, A. U., & Limaye, M. A. (2020) Anisotropy of magnetic susceptibility (AMS) studies on
quartzites of champaner group. In T. K. Biswal, S. K. Ray, & B. Grasemann (Series Editors)
Upper Aravallis: An implication to decode regional tectonics of southern Aravalli mountain belt
(SAMB), Gujarat, Western India; Special Volume on Society of Earth Scientist Series, under the
heading of Structural Geometry of Mobile Belts of the Indian Subcontinent. ISSN: 2194–9204.
ISBN: 978–3–030–40592–2. pp.199–211. https://doi.org/10.1007/978-3-030-40593-9.
Joshi, A. U., Sant, D. A., Parvez, I. A., Rangarajan, G., Limaye, M. A., Mukherjee, S., et al. (2018).
Sub-surface profiling of granite pluton using microtremor method: Southern Aravalli, Gujarat,
India. International Journal of Earth Sciences, 107, 191–201.
Misra, A. A., Bhattacharya, G., Mukherjee, S., & Bose, N. (2014). Near N-S paleo-extension in the
western Deccan region in India: Does it link strike-slip tectonics with India-Seychelles rifting?
International Journal of Earth Sciences, 103, 1645–1680.
Mukherjee, S., Bose, N., Ghosh, R., Dutta, D., Misra, A., Kumar, M., et al. (2020). Structural
geological atlas (p. 623), Springer.
Ramsay, J. G. (1967). Folding and fracturing of rocks. New York: McGraw-Hill.
Rasul, S. H. (1965). The manganese ore of Shivrajpur. Economic Geology, 60, 149–162.
Roy, B. C. (1958). The manganese-ore deposits of PanchMahals and Baroda districts, Bombay
State. Geology Survey of India Bulletin Series A, (15).
Willis, B., & Willis, R. (1929). Geologic structures (2nd ed., p. 518). New York: Mc-Graw-Hill
Book Co.
Vein Geometry Around Bhuj (Gujarat,
India)

Mohammad Walid Omid, Soumyajit Mukherjee, and Sudipta Dasgupta

Abstract The Bhuj area (Kutch, Gujarat, India) is a good and rather unexplored
place in terms of structural geology. This is despite the Kutch region has been
studied in detail for palaeontology, sedimentology and stratigraphy. We report three
morphologies of vein sets of quartz within the sandstone country rock. The area
shows a good possibility of conducting fieldwork with B.Sc. students.

1 Introduction

Veins are joins that are filled up by secondary minerals and can form by fluid over-
pressure see Passchier and Trouw (2005), Bons (2010), Bons et al. (2012), Mukherjee
(2014a, b, 2015), Fossen (2016). Mukherjee et al. (2019) and Ghosh and Mukherjee
(2020) as introductory reading materials on vein geometries.
Figure 1 is a recent summary of tectonic elements of the Kutch region.

2 Veins

Three types of quartz vein patterns within sandstone are noteworthy: (i) X-pattern
or acute angle-obtuse angle intersection pattern (Figs. 2, 3, 4 and 5), (ii) high-angle
intersection pattern (Figs. 6 and 7), and (iii) vein displaced/folded against or near
another vein (Figs. 8 and 9).

M. W. Omid · S. Mukherjee (B) · S. Dasgupta


Department of Earth Sciences, Indian Institute of Technology Bombay, Powai, Mumbai,
Maharashtra 400076, India
e-mail: smukherjee@iitb.ac.in; soumyajitm@gmail.com

© Springer Nature Switzerland AG 2021 707


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_24
708 M. W. Omid et al.

Fig. 1 The Kachchh Rift Basin, reproduced from Shaikh et al. (2020). The study area Bhuj is
shown. White inward-pointed double arrows: SHMax, black dotted 1584 lines represent structural
contours with 1000 feet interval, Black dashed lines: Bouger Gravity Anomaly contours in mGal,
Red stars: major seismicity, IBF: Island Belt Fault, GDF: Gora Dungar Fault, GF: Gedi Fault, SWF:
South Wagad Fault, KMF: Kachchh Mainland Fault, VGKNFS: Vigodi-Gugriana-Khirasra-Netra
Fault System, KHF: Katrol Hill Fault, NKF: North Kathiawar Fault. IBU: Island Belt Uplift, DU:
Desalpar Uplift, WU: Wagad Uplift, KMU: Kachchh Mainland Uplift, NHRFZ: Northern Hill
Range Fault Zone, KHRFZ: Katrol Hill Range Fault Zone. Red, blue and yellow double arrows:
relative plate motion at convergent, divergent and transform-fault boundaries respectively
Vein Geometry Around Bhuj (Gujarat, India) 709

Fig. 2 “X-pattern” of sub-vertical veins in horizontal outcrop of sandstone along Kodhki Road.
a Strike of vein-1: 85°. Strike of vein-2: 10°. b Strike of vein-1: 315°, Strike of vein-2: 20°. c Strike
of vein-1: 330°, Strike of vein-2: 25°. d Strike of vein-1: 355°, Strike of vein-2: 40°

Fig. 3 “X-pattern” of steeply-dipping to sub-vertical veins in horizontal outcrop of sandstone along


Kodhki Road. a Strike of vein-1: 355°. Strike of vein-2: 30°. Dip: 70°. b Strike of vein-1: 345°,
Strike of vein-2: 35°. Dip 82°. c Strike of vein-1: 330°, Strike of vein-2: 25°. d Strike of vein-1:
225°. Dip: 80°. Strike of vein-2: 170°. Dip: 85°
710 M. W. Omid et al.

Fig. 3 (continued)

Fig. 4 “X-pattern” of steeply-dipping to sub-vertical veins in horizontal outcrop of sandstone along


Kodhki Road. a Strike of vein-1: 40°. Dip: 86°. Strike of vein-2: 330°. Dip: 88°. b Strike of vein-1:
340°, Strike of vein-2: 27°. c Strike of vein-1: 55°. Dip: 86°. Strike of vein-2: 335°. Dip 88°. d Strike
of vein-1: 350°. Dip: 86°. Strike of vein-2: 40°. Dip: 82°
Vein Geometry Around Bhuj (Gujarat, India) 711

Fig. 5 “X-pattern” of steeply-dipping to sub-vertical veins in a horizontal outcrop of sandstone


along Kodhki Road. a Strike of vein-1: 50°. Strike of vein-2: 10°. b Strike of vein-1: 50°, Strike
of vein-2: 10°. c Strike of vein-1: 230°. Strike of vein-2: 335°. d Strike of vein-1: 30°. Strike of
vein-2: 340°

Fig. 6 High-angle intersection pattern of veins a horizontal outcrop of sandstone along Kodhki
Road. a Strike of vein-1: 292°. Strike of vein-2: 10°. b Strike of vein-1: 255°, Strike of vein-2: 15°.
Dip of both the veins are 90°. c Strike of vein-1: 230°. Strike of vein-2: 335°. d Strike of vein-1:
30°. Strike of vein-2: 340°
712 M. W. Omid et al.

Fig. 6 (continued)

Fig. 7 High-angle intersection pattern of veins a horizontal outcrop of sandstone along Kodhki
Road. a Strike of vein-1: 310°. Dip: 90°. Strike of vein-2: 200°. Dip: 84°. b Strike of vein-1: 320°.
Dip: 87°. Strike of vein-2: 40°. Dip: 88°. c Strike of vein-1: 340°. Dip: 82°. Strike of vein-2: 80°.
Dip: 85°
Vein Geometry Around Bhuj (Gujarat, India) 713

Fig. 8 Vein displaced/folded against or near another vein. Horizontal outcrop of sandstone along
Kodhki Road. a Strike of vein-1: 210°. b Strike of vein-1: 330°. c Strike of vein-1: 210°. d Strike
of vein-1: 140°

Fig. 9 Vein displaced/folded against or near another vein. Horizontal outcrop of sandstone along
Kodhki Road. a Strike of vein-1: 160°. Dip: 80°. Dip direction: 70°. b Strike of vein-1: 160°. Dip:
85°. c Strike of vein-1: 25°. Dip: 70°
714 M. W. Omid et al.

Acknowledgements The fieldwork for MWO and SM was funded by IIT Bombay. Assistance by
Dhawale Mayur (IISER Pune), and Aashu Pawar and Nidhi Lohani (Panjab University) during field-
work is thanked. Mohamedharoon A. Shaikh (MS University Baroda) rearranged figures. Thanks to
Marion Schneider, Annett Buettener, Boopalan Renu, Alexis Vizcaino, Doerthe Mennecke-Buehler
and the proofreading team (Springer). Dutta and Mukherjee (2021) summarize this work.

References

Bons, P. D., Elburg, M. A., & Gomez-Rivas, E. (2012). A review of the formation of tectonic veins
and their microstructures. Journal of Structural Geology, 43, 33–62.
Dutta, D., Mukherjee, S. (2021). Introduction to Structural Geology and Tectonics Field Guide-
book—Volume 1. In S. Mukherjee (Ed.) Structural Geology and Tectonics Field Guidebook—
Volume 1. Switzerland AG, Springer Nature. Cham. pp. xi-xvi. ISBN: 978-3-030-60142-3.
Fossen, H. (2016). Structural geology. Cambridge University Press.
Ghosh, R., Mukherjee, S. (2020). Structures of lesser/greater Himalaya in and around an out-of-
sequence thrust in the Chaura-Sarahan area (Himachal Pradesh, India). In S. Mukherjee (Ed.)
Structural Geology and Tectonics Field Guidebook - Volume 1. Switzerland AG, Springer Nature.
Mukherjee, S. (2014). Review of flanking structures in meso- and micro-scales. Geological
Magazine, 151, 957–974.
Mukherjee, S. (2014b). Atlas of shear zone structures in Meso-scale (pp. 1–124). Cham, Springer
Geology. ISBN 978-3-319-0088-6.
Mukherjee, S. (2015). Atlas of structural geology. Amsterdam: Elsevier. ISBN: 978-0-12-420152-1.
Mukherjee, S., Bose, N., Ghosh, R., Dutta, D., Misra, A.A., Kumar, M., et al. (2019). Structural
geological atlas. Springer. ISBN: 978-981-13-9825-4.
Shaikh, M., Maurya, D.M., Mukherjee, S., Vanik, N., Padmalal, A., & Chamyal L. (2020). Tectonic
evolution of the intra-uplift Vigodi-Gugriana-Khirasra-Netra fault system in the seismically active
Kachchh rift basin, India: Implications for the western continental margin of the Indian plate.
Journal of Structural Geology.
Oriented Rock Samples for Detailed
Structural Analysis

Krzysztof Gaidzik and Jerzy Żaba

Abstract A significant part of the detailed structural analysis can be done in the
field. However, in many cases, especially in studies on ductile shear zones, further
in-depth microstructural analysis is needed. For this purpose, correct collection of
oriented rock samples with various tectonic structures is crucial for legitimate struc-
tural analysis. Samples collected in the field with known spatial orientation permit
the determination of various geometric and kinematic characteristics of the observed
structures. Oriented samples collected from field also permit to note and interpret
new, previously omitted and/or invisible structures that have not been noticed in
the field. This article presents techniques useful for sample collection in the field,
coordinate systems applicable in the structural studies, and methods to re-orientate
samples, together with ways to select thin section orientation. These are essential
steps toward detail structural analysis.

Keywords Oriented rock sample · Structural analysis · Field method ·


Re-orientation · Thin section

1 Introduction

Detailed structural studies of various tectonic features in different scales include


three essential components. These are as follows: (1) geometric, (2) kinematic, and (3)
dynamic analysis, together with statistical calculations. All of those mentioned above
are based on geometric-kinematic observations in the field related to the morphology,
spatial orientation, and mutual superposition of defined structures, as well as in strain
analysis. A large part of tectonic structures can be identified and initially described
in the field itself. However, in many cases, further in-depth study, both meso- and
microstructural, together with laboratory analysis, are needed (e.g., Kozłowski et al.
1986; Hansen 1990; McClay 2013; Gaidzik and Żaba 2017; Dutta and Mukherjee

K. Gaidzik (B) · J. Żaba


Institute of Earth Sciences, University of Silesia, B˛edzińska 60, 41-200 Sosnowiec, Poland
e-mail: krzysztof.gaidzik@us.edu.pl

© Springer Nature Switzerland AG 2021 715


S. Mukherjee (ed.), Structural Geology and Tectonics Field Guidebook—Volume 1,
Springer Geology, https://doi.org/10.1007/978-3-030-60143-0_25
716 K. Gaidzik and J. Żaba

2019, 2021; Żaba and Gaidzik 2019). For this purpose, proper collection of rock
samples with various tectonic structures is crucial for legitimate structural analysis.
Oriented samples allow determining the actual spatial orientation of these newly
found structures. It can be done using various techniques and devices, e.g., triaxial
re-orientator (Kozłowski et al. 1986). Such tools enable giving a spatial orientation
to sample in the laboratory analogous to how these were exposed in the outcrop.
This, in turn, makes it possible to measure the spatial orientation of all the structures
and their parameters, as if we had measured them in-field exposure. Apart from these
advantages, oriented rock samples can also be used to produce oriented thin sections
for in-depth microstructural and petrostructural analyses (e.g., Knopf and Ingerson
1938; Passchier and Trouw 2005).
This chapter focuses on the accurate collection of oriented rock samples in the
field for further detailed laboratory micro- and mesostructural analyses.

2 Samples Collection in the Field

Preferably, the directions of strike and dip should be marked with the indelible marker
on a specific surface within the sample (Figs. 1 and 2). According to the convention,
the strike is marked with a long straight horizontal line, whereas dip, as a short
line perpendicular to the strike pointing out the direction of the plane inclination
(Figs. 1 and 2e). The surfaces for the marking can be bedding planes, foliation
surfaces, fractures, joints, fault planes, cleavages, etc. However, primary surfaces,
such as bedding or foliation, are recommended, as these are the easiest to sample
and usually provide the most useful kinematic analysis (Hansen 1990). The most
important requirements include a generally smooth surface with a known origin. If
possible, the orientation of any line features visible in the field (e.g., a- or b-lineation,

Fig. 1 Location of oriented samples within the outcrop: a the measured surface is seen from
the top—normal position, b the measured surface is seen from the bottom—reverse (overturned)
position
Oriented Rock Samples for Detailed Structural Analysis 717

Fig. 2 Examples of samples collected in the field with the signs of strike and dip marked before
these were hammered off. a Sample in the outcrop before taking any measurements, geological
compass for the scale, b measurements of strike and dip of the foliation plane; the measured surface
is seen from the top—normal position, c sample with the sign of foliation strike and dip, d sample
with the sign of foliation strike and dip and trend and plunge of b-lineation, e sample in the outcrop
before taking any measurements, and f sample with the sign of strike and dip; the measured surface
is seen from the bottom—reverse (overturned) position
718 K. Gaidzik and J. Żaba

Fig. 3 Final preparation of the sample before storage: a oriented sample with all the necessary
measurements marked and accompanied by digital notes of the site and collected measurements
and b the plastic bag’s preferable layout used for sampling with all the necessary information and
measurements

slickenlines, mineral lineations, intersection lines, etc.) should also be marked with
an arrow pointing out the direction of the plunge (Fig. 2f). The measured 3D angles
of recorded planes and lines should be checked for consistency and marked on the
sample itself (using the indelible marker) and noted in the field notebook (Fig. 3). Of
crucial importance is to record how the measurements have been taken: (a) measured
surface is seen from the top—normal position and (b) measured surface is seen from
the bottom—reverse (overturned) position, as this would significantly impact the
forthcoming interpretation (Kozłowski et al. 1986; Figs. 1 and 2). In a particular case
of a vertical surface, the facing direction should be noted as well.
After this preparatory stage, a geological hammer, a chisel, and eye protection
are needed to take the sample from the outcrop. It is a good practice to select a
relatively small sample that can be easily removed but contain all the structures
of interest. When succeeded, lightly tap off the sharp edges to round them so that
the sample does not split the plastic bag or any other material used for its storage.
The next step is to secure and record the sample correctly. The sample should be
labeled with a simple acronym name and number, not confused with other specimens
(Fig. 2f). The most recommended idea is the first letter from the site (or first 2–3
letters depending on the sites and how many samples with different locations, but
with similar names are planned to be collected) with the consecutive numbering
either chronological, according to one direction (e.g., from north to south, etc.).
The label should be recorded in the rock sample itself, the bag especially prepared
for sample collection, and the notebook (Fig. 3). The labeling of the plastic bag
containing the oriented rock sample should include the following information: name
of the sample, date of collection, lithology, the tectonic structure of interest, spatial
orientation measurements (strike, dip, trend, and plunge according to the preferred
convention), locality, and name of the geologist who collected the sample (Fig. 3).
Alternately, these information can be jotted down in the note book itself, accompanied
Oriented Rock Samples for Detailed Structural Analysis 719

Fig. 4 Internal coordinate


system defined individually
for each oriented rock
sample

by a geological sketch of the outcrop together with meso- to macroscopic structures


observed and their relationship to the sample. The photographic documentation of
the outcrop before and after taking the sample is also needed.

3 Coordinate System

Before any further analysis, especially if oriented thin sections are planned, the
internal coordinate system should be defined and marked on the sample. The prefer-
able convention is to use the lower case “abc” system (Fig. 4). Both “a” and “b” axes
lie on the foliation surface, whereas the “c” axis is perpendicular. The “b” axis is
always parallel to the b-lineation, such as the axes of minor folds, aligned elongate
minerals on foliation surface parallel to minor folds, etc. The “a” axis, on the other
hand, displays the direction of tectonic transport, is perpendicular to the “b” axis,
and can be expressed by elongation lineation, mineral lineation indicated by aligned
streaks on foliation surface, slip lineation, groove, or striations on foliation surface,
or flow lineation in the direction of flow on foliation surface.
However, uncomplicated in theory, applying these rules in actual examples are not
invariably straightforward due to the potential lack of planar and/or tectonic struc-
tures. In such cases, different solutions can be applied depending on the particular
sample and the purpose of the study (Table 1).

4 Triaxial Re-orientator

An advantage of oriented samples over non-oriented ones is the feasibility to measure


structures omitted or not visible in the field. Commonly, due to the lack of time,
weather conditions, outcrop exposure toward the sunlight (shade, shadows), vegeta-
tion, visibility of some small structures, the wetness of rocks, etc., some structures
720 K. Gaidzik and J. Żaba

Table 1 Examples of potential puzzle situations and possible solutions in determining the oriented
rock samples’ internal coordinate system
Problem Solution
1 structural surface, 0 lineation Any direction on the foliation can be treated and marked as
the “b” axis, “a” axis perpendicular to it, and “c” axis
perpendicular to the foliation surface
2 structural surfaces, 0 lineation The intersection line can be treated as the “b” axis, while the
better-defined surface as the ab plane
>2 structural surfaces, 0 lineation The “b” axis is marked as the intersection between the
principal and the second foliation surfaces, and the “a” axis
is on the main foliation surface, perpendicular to the “b” axis
0 structural surfaces, 1 lineation Lineation is the “b” axis, at the right angle in any direction is
the “a” axis, and at the right angle to both is the “c” axis
1 structural surface, 2 lineations If one of them is b-lineation, this one defines the “b” axis, and
if both are b-lineations, the better-defined marks the “b” axis
0 structural surface, 0 lineation Any direction defines the “b” axis, and the other directions
are perpendicular

are not observed until detailed laboratory examination of collected samples. Gener-
ally, these are but are not limited to different types of lineations (Vanik et al. 2018;
Shaikh et al. 2020) or small-scale cleavage planes. In such cases, the application of
special devices constructed of antimagnetic materials, such as triaxial re-orientator
(Fig. 5a), allows to orient the sample in space exactly how it was in the outcrop,
thus authorizing further structural measurements (Kozłowski et al. 1986). It can be
done by adequately manipulating the knobs and movable joints with the geolog-
ical compass’s assistance (Figs. 5 and 6). When the oriented sample is set as in
the outcrop, the missing measurements can be taken the same as those in the field
(Figs. 5g, h and 6b).

5 Thin Sections

One of the main aims of such studies is to recognize the sense of movement in ductile
shear zones for further stress analysis, and the correct orientation of the thin section
in space is essential. Thus, the best practice is to cut thin sections perpendicular to
axes “a” and “b” of the internal coordinate system (see Sect. 2; Fig. 7). The most
commonly used in the microstructural and petrostructural studies are thin sections
in ac and bc planes (Fig. 7). If the sample has been collected for kinematic analysis,
then the preferred surface should parallel the bc plane. However, in some cases,
the analysis of strain and crystallographic fabrics is needed in the perpendicular
direction as well, i.e., ac plane (Hansen 1990). The ab planes are rarely used for thin
section purposes (Kozłowski et al. 1986). Before cutting the sample, the desirable
thin section’s exact position should be marked using a marker or pen. After the thin
Oriented Rock Samples for Detailed Structural Analysis 721

Fig. 5 Triaxial structural re-orientator (a) and the main stages of its application: b oriented sample
installed in the triaxial re-orientator, c and d re-orientation of the sample using the knobs and
movable joints with the assistance of the geological compass, e and f sample re-oriented to the
in situ position ready for further structural measurements, g strike and dip measurement of axial
cleavage unseen in the field, and h trend and plunge measurement of fold axis omitted in the field
722 K. Gaidzik and J. Żaba

Fig. 6 a Re-orientation of the sample taken from the reverse position using the triaxial structural
re-orientator with the assistance of the geological compass and b trend and plunge measurement of
slickelines unfeasible in the field

Fig. 7 3D model presents cutting thin sections from oriented rock samples with the internal
coordinate system

section is cut, its orientation toward coordinate system axes should be marked for
ease of interpretation.

Acknowledgements Soumyajit Mukherjee invited to submit and reviewed this article. Dutta and
Mukherjee (2021) summarized this work.
Oriented Rock Samples for Detailed Structural Analysis 723

References

Dutta, D., & Mukherjee, S. (2019). Opposite shear senses: Geneses, global occurrences, numerical
simulations and a case study from the Indian Western Himalaya. Journal of Structural Geology,
126, 357–392.
Dutta, D., & Mukherjee S. (2021). Extrusion kinematics of UHP terrane in a collisional orogen:
EBSD and microstructure-based approach from the Tso Morari Crystallines (Ladakh Himalaya).
Tectonophysics 880, 228641.
Dutta, D., & Mukherjee, S. (2021). Introduction to Structural Geology and Tectonics Field Guide-
book—Volume 1. In S.Mukherjee (Ed.), Structural geology and tectonics field guidebook,
SwitzerlandAG: Springer Nature. Cham. pp. xi-xvi. ISBN: 978-3-030-60142-3.
Gaidzik, K., & Żaba, J. (2017). Comparison of fold deformation metamorphic cover of the
Karkonoszegranitoids. Przegl˛adGeologiczny, 65(12), 1548–1554.
Hansen, V. L. (1990). Collection and preparation of thin sections of oriented samples. Journal of
Geological Education, 38(4), 294–297.
Knopf, E. B., & Ingerson, F. E. (1938). Structural Petrology. Geol. Soc. Amer. Mem., 6, 1–270.
Kozłowski, K., Żaba, J., & Fediuk, F. (1986). Petrology of metamorphic rocks (pp. 383–390).
Skrypty Uniwersytetu Śl˛askiego.
McClay, K. R. (2013). The mapping of geological structures (168p). New York: Wiley.
Passchier C. W., & Trouw R. A. (2005). Microtectonics. Berlin: Springer Science & Business Media.
Shaikh, M., Maurya, D. M., Mukherjee, S., Vanik, N., Padmalal, A., & Chamyal, L. (2020). Tectonic
evolution of the intra-uplift Vigodi-Gugriana-Khirasra-Netra Fault System in the seismically
active Kachchh Rift Basin, India: Implications for the western continental margin of the Indian
plate. Journal of Structural Geology, 140, 104124.
Vanik, N., Shaikh, H., Mukherjee, S., Maurya, D. M., & Chamyal, L. S. (2018). Post-Deccan trap
stress reorientation under transpression: Evidence from fault slip analyses from SW Saurashtra,
western India. Journal of Geodynamics, 121, 9–19.
Żaba, J., & Gaidzik, K. (2019). Variscan stress regime rotation: Insights from the analysis of
kink folds in the northern margin of the Bohemian Massif, South Poland. Comptes rendus—
Geoscience, 351, 395–405. https://doi.org/10.1016/j.crte.2019.04.003

You might also like