Download as pdf or txt
Download as pdf or txt
You are on page 1of 201

Geophysical Monograph Series

Geophysical Monograph Series


196 Extreme Events and Natural Hazards: The Complexity 222 Magnetosphere‐Ionosphere Coupling in the Solar System
Perspective A. Surjalal Sharma, Armin Bunde, Vijay P. Dimri, Charles R. Chappell, Robert W. Schunk, Peter M. Banks,
and Daniel N. Baker (Eds.) James L. Burch, and Richard M. Thorne (Eds.)
197 Auroral Phenomenology and Magnetospheric Processes: 223 Natural Hazard Uncertainty Assessment: Modeling and
Earth and Other Planets Andreas Keiling, Eric Donovan, Decision Support Karin Riley, Peter Webley, and Matthew
Fran Bagenal, and Tomas Karlsson (Eds.) Thompson (Eds.)
198 Climates, Landscapes, and Civilizations Liviu Giosan, Dorian Q. 224 Hydrodynamics of Time‐Periodic Groundwater Flow:
Fuller, Kathleen Nicoll, Rowan K. Flad, and Peter D. Clift (Eds.) Diffusion Waves in Porous Media Joe S. Depner and Todd
199 Dynamics of the Earth’s Radiation Belts and Inner C. Rasmussen (Auth.)
Magnetosphere Danny Summers, Ian R. Mann, Daniel N. 225 Active Global Seismology Ibrahim Cemen and
Baker, and Michael Schulz (Eds.) Yucel Yilmaz (Eds.)
200 Lagrangian Modeling of the Atmosphere John Lin (Ed.) 226 Climate Extremes Simon Wang (Ed.)
201 Modeling the Ionosphere‐Thermosphere Jospeh D. Huba, 227 Fault Zone Dynamic Processes Marion Thomas (Ed.)
Robert W. Schunk, and George V. Khazanov (Eds.) 228 Flood Damage Survey and Assessment: New Insights from
202 The Mediterranean Sea: Temporal Variability and Spatial Research and Practice Daniela Molinari, Scira Menoni, and
Patterns Gian Luca Eusebi Borzelli, Miroslav Gacic, Piero Francesco Ballio (Eds.)
Lionello, and Paola Malanotte‐Rizzoli (Eds.) 229 Water‐Energy‐Food Nexus – Principles and Practices
203 Future Earth – Advancing Civic Understanding of the P. Abdul Salam, Sangam Shrestha, Vishnu Prasad Pandey,
Anthropocene Diana Dalbotten, Gillian Roehrig, and and Anil K Anal (Eds.)
Patrick Hamilton (Eds.) 230 Dawn–Dusk Asymmetries in Planetary Plasma
204 The Galápagos: A Natural Laboratory for the Earth Environments Stein Haaland, Andrei Rounov, and Colin
Sciences Karen S. Harpp, Eric Mittelstaedt, Noemi Forsyth (Eds.)
d’Ozouville, and David W. Graham (Eds.) 231 Bioenergy and Land Use Change Zhangcai Qin, Umakant
205 Modeling Atmospheric and Oceanic Flows: Insightsfrom Mishra, and Astley Hastings (Eds.)
Laboratory Experiments and Numerical Simulations 232 Microstructural Geochronology: Planetary Records Down
Thomas von Larcher and Paul D. Williams (Eds.) to Atom Scale Desmond Moser, Fernando Corfu, James
206 Remote Sensing of the Terrestrial Water Cycle Venkat Darling, Steven Reddy, and Kimberly Tait (Eds.)
Lakshmi (Ed.) 233 Global Flood Hazard: Applications in Modeling, Mapping
207 Magnetotails in the Solar System Andreas Keiling, Caitriona and Forecasting Guy Schumann, Paul D. Bates, Giuseppe T.
Jackman, and Peter Delamere (Eds.) Aronica, and Heiko Apel (Eds.)
208 Hawaiian Volcanoes: From Source to Surface Rebecca Carey, 234 Pre‐Earthquake Processes: A Multidisciplinary Approach to
Valerie Cayol, Michael Poland, and Dominique Weis (Eds.) Earthquake Prediction Studies Dimitar Ouzounov, Sergey
209 Sea Ice: Physics, Mechanics, and Remote Sensing Pulinets, Katsumi Hattori, and Patrick Taylor (Eds.)
Mohammed Shokr and Nirmal Sinha (Eds.) 235 Electric Currents in Geospace and Beyond Andreas Keiling,
210 Fluid Dynamics in Complex Fractured‐Porous Systems Octav Marghitu, and Michael Wheatland (Eds.)
Boris Faybishenko, Sally M. Benson, and John E. Gale (Eds.) 236 Quantifying Uncertainty in Subsurface Systems
211 Subduction Dynamics: From Mantle Flow to Mega Celine Scheidt, Lewis Li, and Jef Caers (Eds.)
Disasters Gabriele Morra, David A. Yuen, Scott King, Sang 237 Petroleum Engineering Moshood Sanni (Ed.)
Mook Lee, and Seth Stein (Eds.) 238 Geological Carbon Storage: Subsurface Seals and
212 The Early Earth: Accretion and Differentiation James Badro Caprock Integrity Stephanie Vialle, Jonathan Ajo‐Franklin,
and Michael Walter (Eds.) and J. William Carey (Eds.)
213 Global Vegetation Dynamics: Concepts and Applications in 239 Lithospheric Discontinuities Huaiyu Yuan and Barbara
the MC1 Model Dominique Bachelet and David Turner (Eds.) Romanowicz (Eds.)
214 Extreme Events: Observations, Modeling and Economics 240 Chemostratigraphy Across Major Chronological Eras
Mario Chavez, Michael Ghil, and Jaime Urrutia‐Fucugauchi (Eds.) Alcides N.Sial, Claudio Gaucher, Muthuvairavasamy
215 Auroral Dynamics and Space Weather Yongliang Zhang Ramkumar, and Valderez Pinto Ferreira (Eds.)
and Larry Paxton (Eds.) 241 Mathematical Geoenergy:Discovery, Depletion,
216 Low‐Frequency Waves in Space Plasmas Andreas Keiling, and Renewal Paul Pukite, Dennis Coyne, and Daniel
Dong‐Hun Lee, and Valery Nakariakov (Eds.) Challou (Eds.)
217 Deep Earth: Physics and Chemistry of the Lower Mantle 242 Ore Deposits: Origin, Exploration, and Exploitation
and Core Hidenori Terasaki and Rebecca A. Fischer (Eds.) Sophie Decree and Laurence Robb (Eds.)
218 Integrated Imaging of the Earth: Theory and Applications 243 Kuroshio Current: Physical, Biogeochemical and
Max Moorkamp, Peter G. Lelievre, Niklas Linde, and Amir Ecosystem Dynamics Takeyoshi Nagai, Hiroaki Saito,
Khan (Eds.) Koji Suzuki, and Motomitsu Takahashi (Eds.)
219 Plate Boundaries and Natural Hazards Joao Duarte and 244 Geomagnetically Induced Currents from the Sun to the
Wouter Schellart (Eds.) Power Grid Jennifer L. Gannon, Andrei Swidinsky, and
220 Ionospheric Space Weather: Longitude and Hemispheric Zhonghua Xu (Eds.)
Dependences and Lower Atmosphere Forcing Timothy 245 Shale: Subsurface Science and Engineering Thomas
Fuller‐Rowell, Endawoke Yizengaw, Patricia H. Doherty, Dewers, Jason Heath, and Marcelo Sánchez (Eds.)
and Sunanda Basu (Eds.) 246 Submarine Landslides: Subaqueous Mass Transport
221 Terrestrial Water Cycle and Climate Change Natural and Deposits from Outcrop to Seismic Profiles Kei Ogata,
Human‐Induced Impacts Qiuhong Tang and Taikan Oki (Eds.) Andrea Festa, and Gian Andrea Pini (Eds.)
Geophysical Monograph 247

Iceland
Tectonics, Volcanics,
and Glacial Features
Tamie J. Jovanelly

This Work is a co‐publication of the American Geophysical Union and John Wiley and Sons, Inc.
This Work is a co‐publication between the American Geophysical Union and John Wiley & Sons, Inc.

This edition first published 2020 by John Wiley & Sons, Inc., 111 River Street, Hoboken, NJ 07030, USA and the
American Geophysical Union, 2000 Florida Avenue, N.W., Washington, D.C. 20009

© 2020 the American Geophysical Union

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in any
form or by any means, electronic, mechanical, photocopying, recording, or otherwise, except as permitted by law. Advice
on how to obtain permission to reuse material from this title is available at http://www.wiley.com/go/permissions

Published under the aegis of the AGU Publications Committee


Brooks Hanson, Executive Vice President, Science
Carol Frost, Chair, Publications Committee
For details about the American Geophysical Union visit us at www.agu.org.

Wiley Global Headquarters


111 River Street, Hoboken, NJ 07030, USA

For details of our global editorial offices, customer services, and more information about Wiley products visit us at
www.wiley.com.

Limit of Liability/Disclaimer of Warranty


While the publisher and authors have used their best efforts in preparing this work, they make no representations
or warranties with respect to the accuracy or completeness of the contents of this work and specifically disclaim all
warranties, including without limitation any implied warranties of merchantability or fitness for a particular purpose.
No warranty may be created or extended by sales representatives, written sales materials, or promotional statements for
this work. The fact that an organization, website, or product is referred to in this work as a citation and/or potential source
of further information does not mean that the publisher and authors endorse the information or services the organization,
website, or product may provide or recommendations it may make. This work is sold with the understanding that the
publisher is not engaged in rendering professional services. The advice and strategies contained herein may not be suitable
for your situation. You should consult with a specialist where appropriate. Neither the publisher nor authors shall be liable
for any loss of profit or any other commercial damages, including but not limited to special, incidental, consequential, or
other damages. Further, readers should be aware that websites listed in this work may have changed or disappeared
between when this work was written and when it is read.

Library of Congress Cataloging‐in‐Publication data is available.

Hardback: 9781119427094

Cover Design: Wiley


Cover Image: © Tamie Jovanelly

Set in 10/12pt Times New Roman by SPi Global, Pondicherry, India

Printed in the United States of America

10 9 8 7 6 5 4 3 2 1
This book would have never been imagined if it were not for my wonderful husband,
Joe(y) Cook, whose dreams and love are as far reaching as an Askja ash cloud.
And believe me, that is really far.
CONTENTS

Preface...................................................................................................................................................ix

Introduction1
1. The Geologic Framework of Iceland������������������������������������������������������������������������������������������������������ 3

Part I Tectonics 5
2. Overview of Tectonics in Iceland............................................................................................................ 7

3. Tectonics of the Reykjanes Peninsula and Southwestern Region........................................................... 15

4. Tectonics of the South and Southeastern Regions................................................................................. 27

5. Tectonics of the Northeastern Region.................................................................................................. 31

6. Tectonics of the Western Region.......................................................................................................... 39

Part II Volcanics 43
7. Overview of Volcanics in Iceland......................................................................................................... 45

8. Volcanics of the Reykjanes Peninsula and Southwestern Region.......................................................... 61

9. Volcanics of the South and Southeastern Regions................................................................................ 65

10. Volcanics of the North and Northeastern Regions................................................................................ 87

11. Volcanics of the Western Region........................................................................................................ 107

Part III Glacial Features 115


12. Overview of Glacial Features in Iceland............................................................................................. 117

13. Glacial Features of the Reykjanes Peninsula and Southwestern Region.............................................. 137

14. Glacial Features of the South and Southeastern Regions.................................................................... 145

15. Glacial Features of the Northern and Western Regions...................................................................... 159

Glossary���������������������������������������������������������������������������������������������������������������������������������������������������� 173

References������������������������������������������������������������������������������������������������������������������������������������������������ 177

Index��������������������������������������������������������������������������������������������������������������������������������������������������������� 199

vii
PREFACE

My first venture to Iceland was in 2006 when it was glossary, and GPS coordinates for most locations for
still off the radar of most tourists. With the keys to a easy reference.
Toyota Yaris and a series of paper road maps, my older The support I had in writing this book was truly
brother Jim and I circled the island in 11 days. Although endless. Encouragement came in the form of multiple
we were perpetually lost, we had found a pristine bouquets of flowers and sugar‐free Red Bulls delivered
landscape with amazing views, incredible geology, and to my office by husband (Joe Cook), positive reviewer
no road signs. I was hooked. Over the course of the feedback (Dr. Kent Murray and Dr. Sheila Seaman),
next decade I would visit nearly every summer, bring- and advice from colleagues (Dr. Ed Harvey and wife
ing with me fortunate undergraduate students who Carol Rogers, and Dr. Mary Anne Holmes). Substantial
could keep up with the hiking, as well as my desire to project contributions came from Nathan Mennen who
explore everything about the island—including the prepared all the figures for the book, Emily Larrimore
sampling of the dreadful rotten shark cuisine. My who wrote some of the interesting displayed boxes you
preparation for teaching Physical Geology and will find within the text, and Amanda Tomlinson who
Advanced Geological Field Studies courses in Iceland formatted references and glossary terms. I also need to
became the inspiration to write this book. thank the many Berry College Geology “Home Team”
As a geologist, I hope to not only capture the island’s students who traveled with me to Iceland: Mallory
natural beauty, but also to enhance it through detailed Paulk, Maggie Midkiff‐Maddrey, Russell Maddrey,
descriptions that link the relationships between Matthew Bentley, Emma Cook, Emily Larrimore,
­structure, process, and time to the island’s evolution. Carley Carder, Amanda Tomlinson, JT Keiffer,
I reviewed innumerable peer‐reviewed scientific papers Timothy Wooley, and Andrew Elgin. Undoubtedly,
in order to deliver the reader with the most up‐to‐ they were the guinea pigs for this book and provided
date research on interesting, and sometimes debated, insight into the content it should contain.
geological theories regarding an island being split in I cannot believe how much fun I had writing this
half due to plate tectonic motion. This text is not just book and I want to thank John Wiley & Sons, Inc.
intended for academics, but also for novice geologists Publishers and the American Geophysical Union for
who want to understand the magnificent scenery at a providing me with this opportunity. I enjoyed the
deeper level. To encourage this, I provide background whole process: manuscript collecting, reading, learning,
introductions and figures that offer information on and the solitude of the writing process. For me, every
foundational geological concepts. Additionally, the day was a chance to study more about the country
book has been intentionally organized for travelers to and science that I love so dearly. With that stated,
use, by highlighting Iceland’s most popular destina- I realize that I am standing on the shoulders of the
tions and putting each region into a contextual per- foundational Icelandic geologists that came before me:
spective. More specifically, the book is organized into Helgi Björnsson, Páll Einarsson, Agust Gudmundsson,
three main sections: tectonics (Chapters 2–6), volca- Guðrún Larsen, Kristjan Sæmundsson, Oddur
nics (Chapters 7–11), and glacial features (Chapters Sigurðsson, and Thor Thordarsen. Each of these life-
12–15). The book can be read from cover to cover, or it long explorers have written books and documents that
can be utilized as a travel guide by traversing the island I encourage the reader to look at first hand for more
from the capital city Reykjavik counter clockwise. For complete understanding. During the 15 months or so
the latter use, the reader can refer to the introductions that it took to write the manuscript I often reflected
to each Part (Chapters 2, 7, 12) followed by chapters back on my Advisor, Dr. Sheri Fritz, at the University
organized into four cardinal quadrants describing the of Nebraska‐Lincoln who worked tirelessly on manu-
southwest (Chapters 3, 8, 13), southeast (Chapters 4, script writing out of a labor of love, or so it seemed.
9, 14), northeast (Chapters 5 and 10), and northwest Her commitment to science has always inspired me to
(Chapters 6, 11, 15). The book provides an index, work harder.

ix
Introduction
1
The Geologic Framework of Iceland

The island of Iceland, with its northern tip just and perpendicular transform faults with a (to the east)
61 km south of the Arctic Circle, has a long construc- parallel offset ridge (Figure 1.1) [Einarsson, 2008].
tive history that started 130 million years ago during Prior to the current tectonic setting, where the
the last Pangea cycle. Spreading of new ocean floor at Kolbeinsey and Reykjanes ridges form the spreading
mid‐ocean ridges to separate continents following this centers, the extinct Aegir Ridge to the east, which ran
breakup and the onset of a large mantle plume (radius parallel to these systems [Kristjánsson, 1979; Weir
about 300 km) beneath Greenland are thought to have et al., 2001; Tronnes, 2002], was important in the
led to excessive mantle upwelling [Wolfe et al., 1997; initiation of the North Atlantic Ocean during the
Holbrook et al., 2001; Rickers et al., 2013]. These events Eocene (about 56 to 34 million years ago) as rifting
coincide with dates of continental flood basalts and and seafloor spreading began separation of Greenland
mid‐Cretaceous volcanism along the Arctic Mendelev from Norway. At 24 million years ago (Late Oligocene
Ridge, Alpha Ridge, and Ellesmere Island [Lawver and to Early Miocene), as the overall size and temperature
Müller, 1994; Johnston and Thorkelsen, 2000; of the mantle plume continued to dissipate, the
Sigmundsson, 2006]. Uplift accompanying igneous Reykjanes–Kolbeinsey plate boundary was centered
intrusions in northwest Europe, Greenland, and over the hot spot (Figure 1.1) [Fitton et al., 1997;
Canada between 64 and 52 million years ago immedi- Kodaira et al., 1998; Holbrook et al., 2001]. Since then,
ately initiated passive volcanic margins, thereby setting the main Reykjanes and Kolbeinsey ridges have moved
the stage for magmatic upwelling and island creation 240 km to the northwest so that the plume is now
[Saunders et al., 2007]. located under Iceland’s largest ice cap, Vatnajökull.
During the past 60 million years, the overall northwest Consequently, this is also where the crust is thickest
migration of the North American plate carrying (~40 km; Sigmundsson, 2006]. The position of the hot
Greenland and the southeast migration of the Eurasia spot has been established through a combination of
Plate have determined the position of the Iceland hot earthquake data [Oskarsson et al., 1985; Einarsson,
spot. The process of rifting has separated the two major 1991; Weisenberger, 2010], seismic crustal structure
plates, Eurasia and North America (Figure 1.1), with the and tomography data [Flòvenz and Gunnarson, 1991;
Mid‐Atlantic Ridge forming a divergent plate boundary Foulger et al., 2006], and seismic reflection and refrac-
between them. Seafloor spreading is occurring at approx- tion data [Holbrook et al., 2001].
imately 2 cm per year, or 20 km per million years [Sella The plume‐origin hypothesis suggests that volcanism
et al., 2002; Geirsson et al., 2006]. Two Icelandic was initiated by ascending mantle‐derived magma
microplates (or blocks, i.e., Hreppar in the south and from beneath thick continental lithosphere and sub-
Tjörnes in the north) have formed at the intersection sequently from beneath oceanic lithosphere as rifting
of (to the west) the Reykjanes and Kolbeinsey ridges continued and the ocean basin grew [Sigvaldason, 1974a].

Iceland: Tectonics,Volcanics, and Glacial Features, Geophysical Monograph 247, First Edition. Tamie J. Jovanelly.
© 2020 American Geophysical Union. Published 2020 by John Wiley & Sons, Inc.

3
4 The Geologic Framework of Iceland

Figure 1.1 Tectonic context of Iceland. At present Iceland is divided by the Kolbeinsey Ridge in the north
and the Rekjanes Ridge in the south. The yellow circles show the position of the mantle plume from
50 million years ago to present. [Modified from Fitton et al. [1997]; design credit Nathan Mennen.]

Alternative hypotheses that consider mechanisms for Most of the 350,000 km2 basaltic plateau making up
large magma generation include excess magmatism Iceland lies below sea level, with about 30% of the
from melting of mantle and/or recycled ocean‐crust island being above sea level, up to a maximum relief of
material [Foulger, 2006] and a rift model whereby the 2110 m above the ocean surface [Gudmundsson, 2000].
development of a North Atlantic spreading center is The submarine shelf surrounding the island ranges
solely reliant on plate‐tectonic mechanisms and not hot‐ 50–200 km wide and gently slopes to depths of 400 m
spot development [Ellis and Stoker, 2014]. [Thordarson and Larsen, 2007].
Part I: Tectonics
2
Overview of Tectonics in Iceland

2.1. PRESENT TECTONIC SETTING by identifying distinct sedimentary horizons containing


plant remains between lava formations.
The present tectonic setting of Iceland is driven by the Of these zones, that in the east has been the most active
continued spreading of the Mid‐Atlantic Ridge (MAR); during the past 2–3 million years (Myr). The Eastern
specifically, the Kolberinsey Ridge (KR) in the north and Volcanic Zone, and numerous other past rift‐jump struc-
the Reykjanes Ridge (RR) in the south (Figure 1.1). These tures, were mapped using subaerial lava‐flow bodies (e.g.,
subaerial expressions of the MAR in Iceland are charac- dipping of Tertiary basalt strata) that could be identified
terized by various seismically and volcanically active on surface geologic maps through various aged folded
centers often referred to as neovolcanic zones [Einarsson, basalts (Figure 2.1; Böðvarsson and Walker, 1964;
1991]. Three major neovolcanic zones are recognized Jóhannesson and Sæmundsson, 2009; Hjartarson et al.,
where the main processes are normal faulting and volcanic 2017]. The South Iceland Seismic Zone, an area charac-
fissuring: North Volcanic Zone, West Volcanic Zone, and terized by high earthquake activity, accommodates
East Volcanic Zone (Figure 2.1). These neovolcanic zones the offset between the East and West Volcanic Zones
are bounded by perpendicular transform faults that [Stefánsson et al., 2006; Einarsson, 1991]. The East
connect the RR and KR with a parallel offset ridge axis, Volcanic Zone intersects the North Volcanic Zone at the
with the Mid‐Iceland Belt (MIB) forming a triple junction triple junction beneath Vatnajökull, formed by the MIB
beneath Vatnajökull [Sigmundsson, 2006]. (Figure 2.1). The North Volcanic Zone connects to the
The island has undergone dynamic change through a (KR) via the Tjörnes Fracture Zone. Here, the term
series of rift jumps that first began 24 million years ago “fracture zone” describes the zone that connects the
(Ma) in northern Iceland, which initiated the first rift parallel off‐set ridge axis to the KR segment of the MAR.
zone [Harðarson et al., 1997; Hjartarson et al., 2017]. Transform plate movement along the Tjörnes Fracture
Magmatic upwelling through rift jumping is a prominent Zone has resulted in another major center of seismicity
process in the evolution of Iceland [Hjartarson et al., and deformation. The Snæfellsnes Volcanic Belt reacti-
2017]. As described by Mittelstaedt et al. [2008) rift vated at 2 Ma and is moving to the southeast, whereas the
jumps are induced by magmatic heating from an off‐axis southern part of the East Volcanic Zone is currently
hot spot (at present, under Vatnajökull), which results in propagating to the southwest. The Reykjanes Volcanic
a change in the location of the ridge axis. The magma Belt in southwest Iceland is the subaerial expression of
produced by the hot spot thins the lithospheric crust the RR and connects to the West Volcanic Zone.
thereby initiating new rifting to form a new ridge axis. In
Iceland, this process is combined with east and west diver- 2.2. BACKGROUND GEOLOGY
gence of two continental plates, resulting in the rift axes
becoming less active as they move away (e.g., relocate) Effusive volcanism during the Tertiary from seafloor
from the hot spot intensity. Denk et al. [2011] recognizes spreading in the North Atlantic region began to build
unconformities that accompany rift jumps in Iceland up a massive basalt plateau (estimated 350,000 km2)

Iceland: Tectonics,Volcanics, and Glacial Features, Geophysical Monograph 247, First Edition. Tamie J. Jovanelly.
© 2020 American Geophysical Union. Published 2020 by John Wiley & Sons, Inc.

7
8 TECTONICS

Figure 2.1 Iceland’s volcanic zones, associated plate boundaries, and general geologic age of bedrock. KR,
Kolbeinsey Ridge; RR, Reykjanes Ridge; EVZ, East Volcanic Zone; WVZ, West Volcanic Zone; NVZ, North
Volcanic Zone; SISZ, South Iceland Seismic Zone; MIB, Mid‐Iceland Volcanic Belt; TFZ, Tjörnes Fracture
Zone; ÖVB, Öræfi Volcanic Flank; RVB, Reykjanes Volcanic Belt; SVB, Snæfellsnes Volcanic Belt. [Adapted
from Sæmundsson [1979]; design credit Nathan Mennen.]

[Sæmundsson, 1979]. Lavas of similar composition or poorly lithified beds such as till, glaciofluvial
have been found in northwest Britain, Faroe Islands, deposits, marine and fluvial sediments, as well as
and Greenland, which help confirm plate movement soils (Table 2.1; Gardner, 1885; Walker, 1960;
and provide documentation of the scale of this depo- Sæmundsson, 1979]. Due to its abundance (covering
sitional event [Roberts and Hunter, 1979]. The overall about half the total area of Iceland) and its extensive
age of the island exposed above sea level is geologi- exposure in the east and west, there is even an
cally young, with the oldest rocks found to the east Icelandic term for the dark basalt, blágrýtismyndun
and west (14–16 Ma; Moorbath et al., 1968; [Sæmundsson, 1979].
McDougall et al., 1984), whereas rocks in the The Tertiary Basalt Formation is composed mainly
northern region may be only 12 Ma [Sæmundsson, of basaltic lava flows (>83%) comprising tholeiite
1986]. Walker [1960] completed the first published petrology (typical of continental plateau basalts and
lithological account of the Tertiary units on Iceland. mid‐ocean ridges), olivine (typical of ocean basins),
Iceland is divided geologically into three main and porphyritic basalts (representative of intrusive
groups: Tertiary Basalt Formation (Upper Tertiary), and extrusive processes) (Thórarinsson et al., 1959;
Grey Basalt Formation (Upper Pliocene to Lower Klein and Langmuir, 1987; Shorttle and Maclennan,
Pleistocene), Mòberg Formation (Upper Pleistocene), 2011]. Although dominated by basalt, rhyolitic lavas
and the Upper Pleistocene and Holocene unconsolidated (8%), andesitic lavas (3%), and interbasaltic beds
OVERVIEW OF TECTONICS IN ICELAND 9

Table 2.1 Geologic timescale and associated major climate events in Iceland

Era Period Epoch Age Stage Sub‐stage Formation Major events


0–2.5 ka Late
Bog Period (sub‐Atlantic)
2.5–5 ka Late
Holocene Birch Period (sub‐Boreal)
5–7.2 ka Early Bog Period (Atlantic)

Upper Pleistocene Formation


7.2–9.3 ka Early Birch Period (Boreal)
9.3–10 ka Pre‐Boreal Ice Age glaciers melt
10–11 ka Younger Cooling in northern
Dryas hemisphere; glaciers
grow
Late Pleistocene

11–12 ka Weichselian Allerød Warmer climate


Quaternary

12–20 ka Older Icelandic ice sheets


Dryas quickly retreat
20–110 ka Eurasian ice sheet at
maximum; last glacial
stage
115–130 ka Eemian Last interglacial stage
Pleistocene Pleistocene
Middle

130–300 ka Saale Glacial stage


CENOZOIC

300–700 ka

0.7–2.5 Ma Start of full‐scale


Plio‐Pleistocene
glaciations
Early

Formation

2.5–3.3 Ma Pacific Ocean fauna arrive


Pliocene

in Iceland. Bering Strait


opens
3.3–7 Ma Climate begins to cool
Warm, temperate climate
Miocene

Teriatery Basalt Formation


Late

7–12 Ma
Teriatery

Miocene

12–18 Ma
Middle

18–25 Ma Origination of Iceland


Miocene
Early

Note. ka, thousand years ago; Ma, million years ago. Modified from Thordarson and Höskuldsson [2014]; design credit Nathan
Mennen.
10 TECTONICS

c­omposed of tephra and sediment (6%) also can be basalt piles. Most feeder dikes have north to south ori-
found [Einarsson, 1994]. Over time, vesicles and entation, their mean thickness is about 3 m, and widths
fractures present in rock can become infilled post‐dep- range from 1 to 10 m [Sæmundsson, 1986]. Textbook
ositionally with minerals such as quartz, jasper, chalce- examples of dike swarms found in the eastern fjords
dony, calcite, and zeolites. Large calcite crystals can make up between 3 and 15% of the rock outcrop.
sometimes found are referred to as “Iceland spar” Thordarson [2012] describes the dike swarms as subsur-
[Einarsson, 1960]. Other than interbasaltic red‐bed face components of fissure swarms in active volcanic
clays, sedimentary rocks (<10% of the Tertiary systems, whereby the swarms are linked to localized
succession) and fossils are rare (Box 2.1). For example, deposits of andesite–rhyolite lava and tephra marking
only 50 genera or species of plants have been docu- the location of extinct central volcanoes; 40 extinct
mented on the Tjörnes Peninsula [Einarsson, 1994; volcanic systems have been identified in the Tertiary
Thordarson and Höskuldsson, 2014]. succession.
Most of the Tertiary landscape‐altering events Deposition of the Tertiary basalts ended at 2.5 Ma
resulted from fissure eruptions, although the presence coinciding with the onset of an extensive glacial period
of central or shield volcanoes can be spectulated [Eiríksson, 2008]. The crustal response to the weight of
through the numerous feeder dikes that cross cut the ice overburden was translated through normal faulting

Box 2.1 The Miocene and Pliocene fossils of the Tjörnes Peninsula

Iceland’s oldest rocks date back to 14–16 Ma several sea‐level fluctuations [Thoroddsen, 1892].
­(during the Middle Miocene), some of which in The Tjörnes beds are the only notable pre‐
northern Iceland host a fossil record [Grímsson and Quaternary marine deposits in the country, and
Símonarson, 2008]. Iceland is not a particularly they include intertidal, littoral, and subtidal strata
fossiliferous locality because only approximately deposits representing a high‐energy marine deposi­
5% of its bedrock comprises sedimentary rock tional system [Field et al., 2017]. In total the Tjörnes
[Ólafsdóttir and Dowling, 2014]. The majority of beds are 500 m thick and can be separated into
sedimentary bedrock is found on the Tjörnes three biozones based on the dominant mollusc
Peninsula, which subsequently has the most exten­ species: Lower Pliocene Tapes Zone, Upper
sive fossil record of Iceland. Documenting the fos­ Pliocene Mactra Zone, and Serripe Zone. Many of
sils of Iceland is important because Iceland’s the mollusc species found in both the Tapes and
position between the Arctic and Boreal regions, Mactra Zones are either extinct in modern times or
and its position between North America, Greenland, largely migrated to live in warmer waters than that
and Europe/Asia provides insight into the biogeo­ presently surrounding Iceland.
graphic and evolutionary history of the fossilized The Tapes Zone comprises thin lignites alternating
organisms found there, and the climatic history of with marine deposits rich in molluscs such as
the region [Grímsson and Símonarson, 2008; Arctica islandica and species of Cardium and
Wappler et al., 2014]. The fossiliferous strata have Mytilus. These species indicate that the Tapes Zone
a relative abundance of invertebrates and vegeta­ was deposited in a shallow‐marine environment,
tion, but occurrences of vertebrates are uncommon. alternating with subaerial lignite deposition. The
The Tjörnes Formation in northeast Iceland dates Mactra Zone also comprises fossiliferous marine
back to the Early Pliocene, is 1200 m in thickness, deposits alternating with subaerial lignites. In this
and comprises four lithological units: Kaldakvisl zone the now extinct Mactra species can be found
lavas, Tjörnes beds, Höskuldsvík lavas, and the alongside extant species of Glycimeris. The youn­
Bredavík Group [Field et al., 2017]. The Tjörnes gest zone, the Serripes, accounts for just over half of
beds are particularly fossiliferous and are com­ the thickness of the Tjörnes beds. The Serripes Zone
posed primarily of marine sediments with abun­ comprises predominantly marine sediments but
dant shell and gastropod fossils, interbedded with unlike the Tapes and Mactra Zones interbedded lig­
fluvial and lacustrine sediments, and also lignite nite deposits are very infrequent and very thin, and
deposits, indicating that the area has experienced occur near the top of the zone. Mollusca preferring
OVERVIEW OF TECTONICS IN ICELAND 11

a cold climate, such as Serripes gronlandicus and small deer were found in a red interbasaltic sand­
species of the genera Neptunea and Macoma, stone bed of the Burstarfell Formation in
begin appearing in the fossil record pointing this Vopnafjörður, slightly southeast of the Tjörnes
zone, but overall the sea temperature was still Peninsula. The finds comprised fragmented
higher than it is at present [Thoroddsen, 1892]. shoulder and scapula bones, and were of Pliocene
The fossil floral record of Iceland dates back to age, dated between 3.5 and 3 Ma. This discovery
the Middle Miocene, and there is a dramatic indicates that it was possible Iceland supported
difference between older and younger flora char­ assemblages of small herbivorous mammals that
acteristics, due to the gradual cooling of Iceland’s were the ancestors of those that became isolated
climate beginning at 12 Ma. This environmental on Iceland when it became an island in the Early to
change caused thermophilus plants (plants that Middle Miocene or Late Oligocene [Grímsson and
thrive in warmer climatic conditions) to become Símonarson, 2008]. As recently as 2011, the first
extinct, and cold tolerant plants to become more marine vertebrate fossil was found in a cliff face of
dominant [Grímsson and Símonarson, 2008]. the Tjörnes Peninsula. A partial right whale skull
Further pollen analysis carried out on the lignite was found in the Mactra Zone of the Tjörnes beds.
deposits found in the Mactra Zone shows that the Age estimates on the skull are varying, with the
region supported coniferous forests in the middle minimum age being 3.4 Ma and the maximum age
Pliocene. Pollen from fir, spruce, pine, and larch being 4.63 Ma. This find indicates that the Tjörnes
can be found in the Mactra Zone, as well as broad­ Formation has valuable potential for further dis­
leaf species such as oak, beech, hazel, and holly, covery of marine vertebrate fossils, which are so
along with alder, birch, and willow. These tree rare in the fossil record of Iceland. Further discov­
species represent a warmer climate during the mid­ eries such as this could also develop understanding
dle Pliocene than Iceland’s present climate. of communities and evolution of paleomarine
Vertebrate fossils are virtually unknown in ­vertebrate organisms at high Arctic latitudes [Field
Iceland’s fossil record. In the 1990s, bones of a et al., 2017].

with throws between 10 and 20 m and main fault Although superficially similar in appearance to the
strikes generally trending N–E or N–SW. The Tertiary Basalt, the Grey Basalt is characterized as
characteristic dip of the Tertiary Basalt Formation is being lighter in color and coarser in texture
between 5° and 15°, but the direction of dip is variable [Kjartansson, 1960]. The Móberg Formation is
depending on region (Figure 2.1; Einarsson, 1960; described as a palogonite tuff and breccia that is a
Sigmundssson, 2006). Measurements of Tertiary Basalt result of hydration and alteration processes
Formation dips identify folds in the basalt that relate [Kjartansson, 1960].
to the rift‐jump complexes previously described. In the Pliocene Epoch at about 5 Ma climate began
Conformable boundaries between the Tertiary and to cool, and by Early Pleistocene times (2.5 Ma) a
Quaternary basalt deposits in Iceland are not obvious major ice cap had formed, which by Late Pleistocene
[Einarrson, 1994]. Dating of changes in polarity of times (26 ka) covered the entire island of Iceland,
Earth’s magnetic field, however, have made it possible lasting until the present Holocene Interglacial
to delineate Tertiary and Quaternary basalts, as the (Figure 2.2). In the Weichselian Glacial stage, b
­ etween
Quaternary Period spans the present Brunhes normal the Eemian and Holocene interglacials, five glacial
polarity epoch and the previous Matuyana predomi- events took take place [Thordarson and Höskuldsson,
nately reversed polarity epoch [Kristjansson and 2014], culminating in the Last Glacial Maximum,
McDougall, 1982]. The top of the penultimate normal when ice likely extended well beyond the current
polarity epoch, the Gauss, marks the end of the shoreline. Early reports of a glacial moraine found
Tertiary Period at approximately 2.6 Ma. 130 km offshore [Ólafsdóttir, 1975; Norðahl et al.,
There are two notable Plio‐Pleistocene depositional 2008] were ­ followed by other geomorphological
sequences that ultimately make up about 25% of studies that determined the ice extent was only
exposed rock at the surface: Grey Basalt (3.0–0.7 Ma) 10–20 km on the coastal shelf [Hjort et al., 1985].
and Móberg Formations (0.7 Ma to 10,000 ka). Regardless of the offshore position of the ice margin,
Figure 2.2 Growth of ice sheets in comparison to landmass on Iceland over the Tertiary and Quaternary
Periods. [Modified from Thordarson and Höskuldsson [2014]; design credit Nathan Mennen.]

Figure 2.3 Development of a möberg ridge. Strike and dip symbols identify depression within the surface of
the ice sheet: (a) subglacial eruption forms; (b) pillow lava cone or pillow ridge develops; (c) möberg ridge or
cone result post‐eruption. [Figure modified from Thordarson and Larsen [2007]; design credit Nathan Mennen.]
OVERVIEW OF TECTONICS IN ICELAND 13

Photo 1 An example of a möberg ridge in southwestern Iceland. [Courtesy of Tamie Jovanelly.]

Iceland’s landscape was influenced dominantly by the dominate the landscape in the neovolcanic zones.
expansion and retreat of Pleistocene ice sheets. The Moreover, the substantial meltwaters from Weichselian
resulting Pliocene–Quaternary geomorphologic evo- ice sheets 1 km thick produced abundant fluvial
lution includes features such as the mörberg ridges deposits and spectacular erosional landscape features
(Figure 2.3) and tabletop mountains (Photo 1), which such as canyons.
3
Tectonics of the Reykjanes Peninsula and Southwestern Region

A major rift jump from the Snæfellsnes Peninsula high‐temperature geothermal system, and clusters of fis-
(Figure 2.1) to the Reykjanes Peninsula occurred sure swarms [Clifton et al., 2002]. Related seismicity of
approximately 6–7 Ma to initiative spreading there the southwest area aligns with the fissure swarms, result-
[Sæmundsson, 1979; Jóhannesson, 1980]. It is on the ing in shallow (2–5 km) earthquakes that are typically
Reykjanes Peninsula where three systems merge: the M = 4 or less [Angelier et al., 2008; Iceland Meteorological
West Volcanic Zone (WVZ), the Reykjanes Volcanic Office, 2018), although, two major earthquakes with
Belt (RVB), and the oblique‐trending RR, i.e. the MAR magnitudes between 7 and 8 were reported in 1784 and
[Keiding et al., 2006]. The unique steepness of the RR 1896. Periods of activity were recorded in 1929–1935,
as it meets the Reykjanes Peninsula makes it one of the 1967–1973, and 2000, indicating a return period of ~30
only places in the world where a mid‐ocean ridge is years [Einarsson, 2008]. The latest occurring largest
visible on land. This change in dip has been attributed earthquakes (M > 5) have been linked to strike‐slip fault-
to faster spreading rates, as confirmed by topographic ing [Árnadóttir et al., 2004] and crustal stretching
and seismic studies [Einarsson, 2001]. Although there [Keiding et al., 2009]. Currently, it is common for the
has been a gradual decline in activity of the RVB and earthquake swarms to begin in the west, which then trig-
WVZ, since between 2 and 3 Ma when the rift started gers aftershock sequences eastward along the fault
migrating east [Keiding et al., 2008; Sæmundsson, [Keiding et al., 2009] that are likely tied to bookshelf
1974], small‐scale earthquake activity (M < 5) persists faulting (Figure 3.2) [Einarsson, 2008]. Bookshelf fault-
today in a narrow band (~4 km wide) along the plate ing occurs when stress is released in transform fault
boundary. zones during strike‐slip earthquakes. In turn, the release
There are six major rift areas (denoted as fissure in pressure causes the blocks between the fault zones to
swarms) in the southwest area of the Reykjanes rotate.
Peninsula that align with the present NE–SW active rift The fissure swarms (Photo 2) on the western part
zone: Reykjanes, Eldvörp–Svartsengi, Fagradalsfjall, of the Reykjanes Peninsula, which reflect extensional
Krýsuvík, Brennisteinsfjöll, and Hengill–Langjökull and normal faulting on northeast‐striking planes
(Figure 3.1). These active rifts are arranged en echelon [Einarsson, 1991], were formed beneath Pleistocene
along the RVB, making it very different from the other glaciers (i.e., hyaloclastite ridges) or are postglacial
volcanic zones identified in Figure 2.1. Hence, the posi- eruptive features. In contrast, the fissures on the east-
tion of the active rifts in the RVB relate to the oblique ern side are predominantly right‐lateral strike‐slip
movement of the plate, which results in more shearing faults that crosscut the plate boundary in north–south
movement and less spreading perpendicular to the plate patterns [Einarsson, 1991]. As described by Jakobsson
axis [Weir et al., 2001; Sigmundsson, 2006]. Each of the et al. [1978] and Sæmundsson [1979], these eruptive fis-
six rift areas in the RVB has its own magma supply, sures are spaced on average approximately 5 km apart.

Iceland: Tectonics,Volcanics, and Glacial Features, Geophysical Monograph 247, First Edition. Tamie J. Jovanelly.
© 2020 American Geophysical Union. Published 2020 by John Wiley & Sons, Inc.

15
16 TECTONICS

Figure 3.1 Active volcanic zones on the Reykjanes Peninsula. The main zone begins with the Reykjanes
Volcanic Belt (RVB) and ends with the Western Volcanic Belt (WVB). Six fissure systems trend perpendicular
to the main zone. [Modified from Jóhannesson and Sæmundsson [1998]; design credit Nathan Mennen.]

The landscape of the Reykjanes Peninsula is consid- 2007]. Reservoir temperatures as high as 300°C at 200–
ered to be geologically young, with about half the area 500 m below surface have been reported from borehole
covered by Holocene lava fields. Specifically, the topog- drilling [Ármannsson and Kristmannsdottir, 1992;
raphy is molded by 16 eruption events that occurred Bjarnason, 2000], which corresponds to hydrothermal
between 875 and 1340 CE from five volcanoes mineral alteration as temperatures increase with depth
[Thordarson and Höskuldsson, 2014]. The rocks exposed (e.g., smectite–zeolite, chlorite, chlorite–epidote;
in the southwest area are not older than 3.2 Ma and Gísalson, 1973; Mawejje, 2007]. Steep hyaloclastic
most are Early Pleistocene in age (Figure 2.1). The ridges that helped form the closed‐basin system for the
material of most mountains in the area is hyaloclastite NE–SW orientation of Lake Kleifarvatn (63.9263,
formed subglacially. The Reykjanes Peninsula dates to −21.9744) dominate Krýsuvík Valley [Morales, 1992].
the Bruhnes normal polarity magnetic epoch Lake Kleifarvatn (9.7 km2) is the deepest lake in the
(<0.8 Ma; Middle to Late Pleistocene), which corre- southwest region and is among the deepest lakes in
sponds to the age of the RR segment of the MAR. Iceland (107 m) (Photo 3). To the east Lake Kleifarvatn
is surrounded by tabletop ridges Geitahlid and
3.1. KRÝSUVÍK VALLEY AND LAKE Kistufell, and to the west a subglacial hyaloclastic
KLEIFARVATN ridge (Sveifluhals). These features are evidence of both
effusive and explosive subglacial eruptions [Imsland,
The Krýsuvík Valley is one of five high‐temperature 1973; Jónsson, 1978]. In 2001, three earthquakes of
geothermal areas explored for potential resource M = 4.5 occurred in the nearby South Iceland Seismic
management on the Reykjanes Peninsula [Mawejje, Zone (SISZ), however, seismicity was not detected at
Tectonics of the Reykjanes Peninsula and Southwestern Region  17

Figure 3.2 Bookshelf faulting occurs when rift segments are offset from one another and the motion bet-
ween segments is accommodated by transform faults that are oriented orthogonally to the main rift axis.
Concurrently, this lateral motion induces a clockwise rotation of the faults. [Modified from Angelier et al.
[2008]; design credit Nathan Mennen.]

Lake Kleifarvatn despite the appearance of a fissure 3.2. BLUE LAGOON


400 m long and 0.3 m wide that lowered the lake sur- AND GEOTHERMAL ENERGY
face by ~4 m [Thordarson, 2012]. The lake is fed by
groundwater and precipitation and has since refilled. The Blue Lagoon (63.8804, −22.4495) is one of
Near Lake Kleifarvatn is an area that was once a research Iceland’s most visited sites with more than 500,000
borehole drill site, called Seltún (63.8970, −22.0548). people visiting this man‐made spa yearly (Photo 5)
From a boardwalk it is easy to view sulfur deposits, [Icelandic Tourism Bureau, 2018]. The 5000 m2 lagoon
mudpots, and fumaroles. Also nearby are eight explosion is the result of the wastewater output from Svartsengi
craters of maar volcanic eruptions. The largest geothermal plant that has been in operation since 1976
crater, Graenvatn (63.8848, −22.0536) is 300 m in [Gudmundsóttir et al., 2010]. The geothermal water
diameter and 44 m deep (Photo 4; Thordarson and under Svartsengi is richer in minerals than typical geo-
Höskuldsson, 2014]. thermal areas due to the salted ocean water, which
Photo 2 An example of a fissure swarm in southwestern Iceland. [Courtesy of Russell Maddrey.]

Photo 3 Lake Kleifarvatn is the deepest lake in southwestern Iceland and fills the bottom of Krýsuvík Valley.
[Courtesy of Tamie Jovanelly.]
Photo 4 Grænvatn, near Lake Kleifarvatn, is an example of a maar volcanic eruption. [Courtesy of Russell
Maddrey.]

Photo 5 The Blue Lagoon geothermal spa is a popular tourist destination for rest and relaxation. [Courtesy
of Russell Maddrey.]
20 TECTONICS

seeps through the porous lava and combines with the horst and graben sequence. The only other place in
hot groundwater. Because the water in the Svartsengi Iceland where the MAR can be seen is in Þingvellir
Geothermal Power Plant has such high mineral content National Park (see section 3.4). Here, the movement
(e.g., 70 ppm Cl, conductivity 300 μS cm−1) it has been between the North American and Eurasian Plates
deemed too corrosive to be used directly as a hot water can be clearly seen in the cracks or faults that traverse
source for housing [Thorolfsson, 2005]. Consequently, the region. Also from this location, you can see
the power plant uses the 240°C hot water to heat up móberg hills, which are composed of pillow lavas
fresh groundwater, which in turn is pumped into homes made of hydroclastic tephra and breccia that line the
on Reykjanes Peninsula. The characteristic milky blue rift valley as it approaches a lighthouse in the dis-
color of the Blue Lagoon results from the high silica tance. The other pertinent surface features include a
content extracted in the process of cleaning the water historic Stampar fissure eruption from 1226 CE over-
before it is used in energy production. Since the actual lying a 2000‐year‐old fissure eruption lava, and an
mineral‐rich hot water is not used directly, whatever is 8000‐year‐old picritic lava shield with hydroclastic
not used for heating the groundwater is pumped ridges from the last glacial episode (12 ka; Franzson
directly into the lava fields surrounding the power et al., 2002].
plant, creating the Blue Lagoon [Thorolfsson, 2005]. Reykjanesviti is close to a geothermal power plant
The pumped water cools during the production pro- and museum (63.8664, −22.6997) that is open to the
cess and by the time it reaches the bathing area it is public during normal business hours. This power
approximately 38°C [Blue Lagoon, personal commu- plant has been operational since the late 1960s and is
nication]. As a result of continued hot‐water produc- adjacent to visible steam vents and solfataras. An
tion at Svartsengi power plant, ever more water is exploration borehole drilled in 2002 to a depth of
pumped into the lava fields, which means the lagoon 2054 m (temperature of 320°C at the base of the
keeps growing a few centimeters every year borehole) revealed subsurface hydrothermal alter-
[Gudmundsóttir et al., 2010]. ation below 500 m depth. These data are of interest
The lava fields of Svartsengi were formed in a because the progressive alteration zonation is the
volcanic eruption in 1226 CE. Due to the young age, opposite signature to that found in the Svarstsengi
the lava field has high porosity and permeability. and Eldvorp geothermal fields (i.e., Blue Lagoon),
Hydrogen sulfide that is released during the process which are also on the Reykjanes Peninsula [Franzson,
of creating geothermal energy can corrode pipelines 2000; Franzson et al., 2002].
if not combined with ambient air. For this reason,
ambient air is used to clean the system and to lower 3.4. ÞINGVELLIR AND HENGILL
the temperatures to below 150°C, creating the soft,
slimy, ground that lines the bottom of the lagoon The divergent plate boundary seen at Reykjanesviti
[Thorolfsson, 2005]. This same silica mud is used to trends 70°NW along the whole Reykjanes Peninsula
produce the skin‐care products for which the Blue [Einarsson, 2008] until it joins with the WVZ and SISZ
Lagoon is known. at the Hengill triple junction (Figure 2.1) within the
Hengill–Langjökull volcanic system [Einarsson, 2008].
3.3. MID‐ATLANTIC RIDGE AT REYKJANESVITI The WVZ has been the main site for crustal spreading
in southern Iceland for the past 6–7 Myr at 2 cm year−1
The mid‐ocean ridge system is the longest and the and is currently spreading at a rate of 0.8 cm year−1
most extensive chain of mountains on Earth [Perlt et al., 2008]. The Hengill–Langjökull system is
(65,000 km), but being located underwater, more than the easternmost of a series of six closely spaced
90% of this mountain range remains hidden from basaltic fissure systems that cut diagonally across the
view [Toomey, 2012; Wright et al., 2012]. Reykjanesviti Reykjanes Peninsula (Figure 3.1) and consists of a
(63.8155, −22.7040) is one of the few places on Earth series of NE–SW‐trending fissure vents, crater rows,
where a segment of the MAR occurs subaerially as a and small shield volcanoes occupying a strongly
surface expression of the submerged RR (Figure 1.1) faulted graben that has been mapped recently by
in the North Atlantic Ocean. The RR is spreading at Friese [2008].
an average rate of 2.5 cm year−1 [DeMets et al., 1994]. The location known as Þingvellir National Park
The MAR is geologically significant because it marks (64.2553, −21.1322) presents a rare example of an
the boundary where the North American and exceptionally well‐exposed rift valley that marks the
Eurasian tectonic plates meet and separate forming a crest of the MAR and the boundary between the
Tectonics of the Reykjanes Peninsula and Southwestern Region  21

Photo 6 Þingvellir National Park provides a rare example of an exceptionally well‐exposed rift valley that
marks the crest of the boundary between the North American and Eurasian tectonic plates. A braided river
occupies the subsided graben. [Courtesy of Russell Maddrey.]

North American and Eurasian tectonic plates (Photo 6). that marks the western boundary of the North
The rift valley narrows from about 20 km in the American plate and can be walked along in the park.
northeast to about 10 km in the southwest [Thordarson This feature is 7.7 km in length, 64 m at its widest
and Höskuldsson, 2014]. The extension appears as point, and has maximum throws of 40 m [Gudmundsson,
nearly parallel fissures and down‐thrown fault blocks 1987]. Its equivalent across the graben, marking the
running along the length of the valley [Friese, 2008]. eastern boundary of the Eurasian plate is Hrafnagjá
On the west (North American) side of the valley, the (64.2572, −21.1107), which is 11 km long, 68 m wide
blocks step down toward the east, while the situation and has a maximum throw of 30 m [Gudmundsson,
is reversed on the opposite (European) side. The 1987]. The Þingvellir faults are believed to be the sur-
graben floor (4.7 km wide) slopes to the southwest face expressions of deeply rooted normal faults. The
toward Lake Þingvallavatn (>100 m above sea level) numerous fissures encountered on the valley floor are
and continues to subside at about 0.1 cm year−1 with a of similar origin and are perpendicular to the surface
total subsidence of 40 m for the past 9000 years [Islam [Ingolfsson, 2008].
et al., 2016]. Rifting within the graben occurs episodi- The low‐lying bottom of the graben drains a braided
cally, with the last major earthquake activity river system into Lake Þingvallavatn, which is the larg-
­occurring in 1789 CE when the graben floor subsided est natural lake in Iceland (84 km2). The deepest part
1–2 m. Additionally, the valley walls (horsts) continue of the lake measures 114 m, which means that its floor
to move apart at a rate of about 0.7 cm year−1, totaling is below sea level. Bedrock, sediments, and faults were
70 m of horizontal extension over the past 9000 years. mapped using seismic reflection profiles and sono-
Almannagjá is a dramatic tensional fracture (Photo 7) graphs [Thors, 1992; Thordarson and Höskuldsson,
22 TECTONICS

Photo 7 Almannagjá is a dramatic tensional fracture in Þingvellir National Park. [Courtesy of Russell
Maddrey.]

2014]. The Þingvellir lava (Þingvallahraun) covers the eruption event lasting 50–100 years. A number of lava
northern part of Lake Þingvallavatn and the graben flows from this eruption entered the southern reaches of
floor north and east of the lake. It originated in a the graben, covering a 200 km2 area with an estimated
major fissure eruption to the southeast of Hrafnabjörg, volume of 17 km3 [Thordarson and Höskuldsson, 2014].
around 9100 years ago. The many single flows of this Although the area’s geology is renowned, Þingvellir
lava are best exposed in the fault scarp of Almannagjá may be best known to Icelanders as the initial site of
where numerous sheets of individual lava lobes were the commonwealth general assembly, AlÞing. It was
stacked as the eruption progressed [Ingolfsson, 2008]. established in 930 CE and is still in operation today,
From this sequence of flows it was determined that the making it the oldest working parliament in Europe.
oldest part of the lake basin was formed during the Sessions were held at this site until 1798, when, it is
Weichselian glaciation (Table 2.1), thereby resulting in thought, multiple volcanic and tectonic events lead to
the formation of subglacial hyaloclastites. Postglacial the meetings being moved to the capital city Reykjavík
lavas (namely Þingvallahraun, Nesjahraun) flowed [Thordarson and Höskuldsson, 2014].
into the basin (~30 km3) to modify its shape [Thors, The area southeast of Lake Þingvallavatn contains
1992], thereby damming outlets and thus maintaining the Hengill, an active central volcano (64.1833,
the lake level today. −21.3333; 803 m), which is associated with the ­present
Volcanoes surround the Þingvellir graben and illus- accretionary zone. Also in this volcanic complex is the
trate the connection between rifting and volcanism. extinct Grensdalur central volcano, which is associated
Mount Skjaldbreiður (1060 m) is a shield volcano seen with an accretionary zone that became inactive at
to the north of the Þingvellir graben. Like much of the about 0.7 Ma when the spreading center migrated
region, it was also formed 9000 years ago in a single west. Between 1993 and 1998 a major volcano‐tectonic
Tectonics of the Reykjanes Peninsula and Southwestern Region  23

Photo 8 Hellisheiđi Geothermal Plant harnesses subsurface steam and water to generate electricity for
Reykjavík. The plant is at the base of Krafla. [Courtesy of Tamie Jovanelly.]

episode took place that produced thousands of earth- Iceland’s most renowned geysers, the Stóri Geysir
quakes up to M = 5.5. These events were driven by an (64.3104, −20.3024), can be found near the town of
intrusion at 7 km depth. A uniform uplift has been Haukadalur. Historic accounts of the Stóri Geysir, or
observed (2 cm year−1), while the seismic records show as it is better known the “Great Geysir,” reveal that it
episodic activity associated with strike‐slip faulting as has been erupting since 1294 CE. Sæmundsson [1979]
well as normal faulting. determined that the heat source for the area’s geysers
The high‐temperature Hengill geothermal fields came from shallow igneous intrusions, although dur-
are the second largest in Iceland [Pálmason et al., ing the past 10,000 years there has not been any
1985] and the Nesjavellir and Hellisheiði geo- volcanic activity in the area. Pasvanoglu et al. [1998]
thermal fields (Photo 8; Box 3.1) located southwest calculated mineral saturation indexes for the waters in
of Hengill are major producers of geothermal the geothermal park and concluded that the water
energy for Reykjavík. from Stóri Geysir is in near equilibrium at 220–240°C.
Determining the chronology of siliceous sinter deposits
3.5. STÓRI GEYSIR AND GULFOSS associated with hot springs and geysers is difficult due
to the lack of reliable silica dating methods [Jones
The first documented use of the Icelandic word et al., 2007]. Jones et al. [2007], however, was able to
“geysir” was in 1597 by Björn Jónsson a farmer from determine the history of Stóri Geysir by using inter-
the town of Skaðsá [Nielsen, 1937]. The word “geysir” mixed tephra deposits of the 3.5 m thick depositional
can be translated to mean gusher or “the one who apron. The tephrochronology was determined from
rages” and is related to the Icelandic verb “gjósa,” palagonite, a post‐depositional alteration product
which means to erupt [Torfason, 1985, p. 3]. One of from the interaction of water and tephra derived from
24 TECTONICS

Box 3.1 Hellisheiði Geothermal Power Station


Emily Larrimore
Iceland’s incredible geothermal capacity and had increased by nearly four times [Ragnarsson,
numerous hydrothermal reservoirs give the small 2015]. Today, the Hellisheiði Power Station is a
island amazing potential to be completely self‐ cogeneration plant, meaning that the plant pro-
sustaining and independent from fossil fuels. duces both electricity and heat from one source of
Although the population growth of Iceland fuel [Ragnarsson, 2015].
remains at 1%, increased tourism has put stress An innovative and adaptive element of the
on the country’s need for energy production. For Hellisheiði Power Station is that it was built in
example, from 2010 to 2015 tourism increased modular units, such that the power station has the
20% per year, resulting in 1 million foreign visi- ability to expand as the demand for geothermal
tors exploring the island’s beauty in 2016. The energy generation from industry and municipal
expansion of tourism has been met recently with use increases [Gunnlaugsson and Gislason, 2005].
the need to expand its use of natural resources for The overall energy production process at Hellisheiði
energy production. In 2008, about 25% of the occurs in three stages: (a) the collection and
electricity generated in Iceland was from geothermal processing of steam, (b) electricity generation,
sources, and overall 82% of the primary energy and (c) heating of cold water for space heating.
used in Iceland is from renewable domestic First, during the collection and processing stage,
sources [Harðarson, 2014]. steam and water are extracted from drillholes to
Owing to its close proximity to the capital city be taken to a unit called the separation station. In
Reykjavík, geologists began to further explore the the separation station, the steam and water are
Hengill geothermal area (Nesjaveillir, Hellisheiði, separated, then transported to the steam turbines
and Hveragerdi), located in the middle of the in the main powerhouse. Each individual steam
Western Volcanic Zone (Figure 2.1) [Zakharova turbine has a rated capacity of 40 MWe, and the
and Spichak, 2012]. The Hengill geothermal area turbines are used in the second stage to generate
has about 600 surface manifestations of hydro- electricity from the steam. In the third stage, the
thermal activity, such as hot springs, geysers, and steam is used to preheat the cold water in a con-
steam jets, making it a very attractive region for denser. In tube heat exchangers, final heating of
geothermal energy utilization and development the separated geothermal water occurs. One of
[Zakharova and Spichak, 2012]. Additionally, the the finals steps in the production process is to
presence of hydrothermal reservoirs connected by pass the heated geothermal water through a deaer-
both horizontal and vertical fissures allows for ator to release excess dissolved oxygen and other
easy movement of hot fluids from depths of 1–2 gases through boiling. This is done because water
km [Gunnlaugsson and Gislason, 2005]. The saturated with dissolved oxygen has a corrosive
temperature of this geothermal resource varies effect on steel, which would complicate the trans-
between 200°C and approximately 320°C portation process [Gunnlaugsson and Gislason,
[Harðarson, 2014]. 2005]. To transport the energy created at
The Hellisheiði Power Station (Photo 8; 64.0371, Hellisheiði, the power plant is connected to a
−21.4012) was built by Reykjavík Energy and 220 kV overhead power line that is then extended
began producing electrical energy in 2006 with to a grid. The power plant also has a buried water
an initial production capacity of 90 MWe [Lund transmission pipeline 18 km long that connects
et al., 2008]. (MWe, electricity output of a power the plant to Reykjavík’s water distribution system
plant in megawatts. The electricity output of a [Gunnlaugsson and Gislason, 2005].
power plant is equal to the thermal overall power The hydrothermal reserves of the Hveragerði
multiplied by the efficiency of the plant.) In 2010, geothermal area are mainly used for greenhouse
the Hellisheiði Power Station began producing heating by residents of the region, but Nesjaveillir
hot water for district heating in Reykjavík, where also has a smaller scale electrical power station
about one‐third of the population of Iceland lives. that utilizes other nearby hydrothermal reservoirs
As of 2011, the total installed capacity of the plant [Zakharova and Spichak, 2012].
Tectonics of the Reykjanes Peninsula and Southwestern Region  25

Photo 9 Strokkur (The Churn) erupts every 10 min and has a column height of 30 m. [Courtesy of Tamie
Jovanelly.]

the nearby Katla, Hekla, and Veiðivötn volcanoes. It in the geothermal park, however, erupts every 10 min
was found that Stóri developed in three main phases: and has a column height of 30 m (Photo 9). Pasvanoglu
10,000–4000 years ago hot spring waters formed the et al. [1998] determined that the total flow rate in this
geyser; 3000–900 years ago was a period of weathering geothermal field has decreased over time from 14 L s−1
and little sinter deposition; 900–800 years ago saw the in 1967 to 9 L s−1 in 1994. It is inconclusive as to
rebirth of the geyser [Jones et al., 2007]. Stóri’s eruption whether drilling in the area is related to these changes
recurrence interval and intensity has fluctuated since or not. A short hike from Strokkur to the west is
its origin; its annual growth also varies between 0.2 Laugarfjall Mountain, which provides a panoramic
and 1.4 kg year−1 m−2 [Tobler et al., 2008]. In 1845, it view overlooking the geothermal park.
was recorded that the maximum column height rose to Gulfoss (64.310l4, −20.3024), which translates to
170 m. This occurred just before a 53‐year period of “Golden Waterfall,” is Iceland’s most popular water-
dormancy ended by a major eruption in 1896 when fall (Photo 10; Iceland Tourism Bureau, 2018]. The
column height was described at 60 m. A man‐made meltwater of the Langjökull glacier feeds the Hvítá
channel was dug around the rim of the geyser in 1935 River, which follows NNE-trending faults before it
to promote faster recirculation of surface water to the enters the 2.5 km gorge to form a cataract‐segmented
central vent, which initially worked, but the channel waterfall. The two‐tiered waterfall drops a total of
quickly became clogged with silica deposition and the 32 m (first tier, 11 m; second‐tier, 21 m). The depth of
geyser stopped erupting again. An earthquake in 2000 the canyon at the southern end (called Gullfossgjúfur)
reinvigorated the geyser and it has since been erupting has a maximum depth of 70 m. There are two com-
intermittently. Strokkur (The Churn), another geyser peting ideas about the origins of the canyon. One is
26 TECTONICS

Photo 10 Gulfoss is one of Europe’s largest waterfalls. This picture depicts the second tier drop of 21 m.
[Courtesy of Russell Maddrey.]

that a catastrophic jökulhlaup flood of meltwater dur- causes differential weathering whereby the softer
ing the Younger Dryas (11.7—10.5 ka) glacial retreat sedimentary unit quickly erodes under mechanical
­
formed the canyon. Another, which may be more plau- weathering processes induced by the presence of water.
sible, has to do with the varied lithology of the area. With a rate of erosion at 25 cm year−1, the canyon
The stratigraphy is such that thick sedimentary units would have formed during the Younger Dryas
(<10 m) are capped by thin lava flows (~1–3 m). This (10,700 ka) stadial.
4
Tectonics of the South and Southeastern Regions

The SISZ trends east–west between two major ments across southern Iceland indicate that the
volcanic rift zones of southern Iceland, namely the East extension between the EVZ and the WVZ is latitude
Volcanic Zone (EVZ) and the WVZ at the Reykjanes dependent. For example, as described by Islam et al.
Peninsula. The SISZ is a transform fault zone that is [2016], in the northern region the EVZ accommodates
70–100 km long and 20 km wide and is considered 1.98 cm year−1 of extension, whereby the WVZ accom-
important for its seismotectonic activity (Figure 2.1) modates 0.26 cm year−1. In the southern region the
[Stefánsson and Halldórsson, 1988]. The Mid‐Iceland EVZ accommodates 1.38 cm year−1 and WVZ
Belt (MIB) to the north is a subparallel transform fault 0.7 cm year−1. Surface deformation zones include sev-
that also joins the EVZ and the WVZ. The block that eral volcanic systems in the EVZ with their associated
sits in between the MIB and SISZ is referred to as the fissure swarms as illustrated by Hekla, Katla (including
Hreppar microplate and has been described only Eldgjá fissure), Grímsvötn (including Lakagígar fis-
recently [LaFemina et al., 2005; Einarsson, 2008]. sure), and Bárðarbunga (including the Vatnaöldur,
Rifting during the past 1100 years has been mostly Veiðivötn, and Tröllagígar eruptive fissures; see Figure
in the EVZ, but seems to have been switching between 7.2). A notable rifting episode occurred during the
the (older) WVZ and the (younger) EVZ on a time- Laki eruption (1783–1784 CE; 64.0647, −18.2261),
scale of thousands to tens of thousands of years which produced about 16 km3 of erupted material and
[Einarsson, 2008]. Although it is agreed that the SISZ formed the Lakagígar eruptive fissures [Jónsson et al.,
is associated with a rift jump, it is still inconclusive as 1997]. The most recent rifting event occurred in 2014–
to whether the jump occurs by continual rift propaga- 2015 at the Bárðarbunga volcanic system (63.6411,
tion or by alternating mechanisms that allow both the −17.5281), which involved a dike intrusion 45 km long
EVZ and WVZ to be active [Sigmundsson et al., 1995]. that formed over a 2‐week period [Hjartardóttir et al.,
The EVZ is a southward propagating accretionary 2016a]. The total dike volume was about 0.5 km3
plate boundary (2–5 cm year−1 from the presumed hot [Sigmundsson et al., 2015; Gudmundsson et al., 2016].
spot under Vatnajökull) that started forming 2 Ma A series of north–south surface fractures have been
[Einarsson, 1991, 2008]. Over the past 10,000 years the mapped in the SISZ to indicate a tectonic process
EVZ has accounted for 85% of the divergence called bookshelf faulting (Figure 3.2) [Einarsson et al.,
[Sigmundsson et al., 1995; Islam and Stufkell, 2017] 1981]. Here, a series of faults are aligned perpendic-
and represents an interrifting phase. This interrifting ular to the main transform motion of the SISZ and are
phase is dominated by steady spreading that is in good spaced about 1 km from the adjacent fault [Einarsson,
agreement with the global plate‐motion models devel- 2008]. Specifically, the overall direction of movement
oped by DeMets et al. [2010] and Pedersen et al. [2009]. north of the SISZ is to the west, whereas south of the
During the Holocene, however, the rates of spreading SISZ it is to the east. Since the movement north of
along the EVZ and WVZ have not been equal the fault is occurring faster, the transform motion in
[LaFemina et al., 2005]. Recent results obtained by the SISZ is dominantly left‐lateral therefore the
analysis of Global Positioning System (GPS) measure- rotation of the blocks appears counterclockwise
­

Iceland: Tectonics,Volcanics, and Glacial Features, Geophysical Monograph 247, First Edition. Tamie J. Jovanelly.
© 2020 American Geophysical Union. Published 2020 by John Wiley & Sons, Inc.

27
28 TECTONICS

[Sigmundsson et al., 1995]. Modeling of GPS data by level by ~4 m [Clifton et al., 2003]. Following these
LaFemina et al. [2005] supports rotational movements, events, on 29 May 2008 an earthquake doublet occurred
like bookshelf faulting, such as are found on the on the western margin of the SISZ, 2 km south of the
Hreppar microplate. This evidence best supports the town of Hveragerði (64.0261, −21.1786) that resulted in
propagating rift hypothesis previously mentioned a Mw = 6.1 event [Decriem et al., 2010]. According to
[Einarsson, 2008]. geodetic data ­ analyzed by Decriem et al. [2010],
Some of the largest earthquakes (Mw = 6.0–8.0) in his- continued seismic activity along the SISZ is likely as
toric time have occurred in the SISZ and are usually they determined that this sequence (2000–2008) released
related to movement along this transform boundary that only about half as much of the stored energy compared
produces bookshelf faults [Bonafede et al., 2007]. to the 1896–1912 CE event.
Sigmundsson et al. [1995] determined that to be in step
with relative motion of the North American and 4.1. FAULTS AND FAULT COMPLEXES
Eurasian plates that the average slip rate of the faults OF THE TERTIARY BASALTS
within the SISZ must be about 0.05–0.5 cm year−1 per
fault. Coincidently, due to rift‐jump mechanisms, the The Tertiary plateau basalts of Iceland have been
accumulation of brittle crust over time, and a closer grouped with similar rocks in northwest Britain,
overall proximity to the hot spot, earthquakes closer to Faroes Islands, and Greenland that are also linked to
the EVZ tend to be larger for any given episode and effusive fissure volcanism from central volcanoes
gradually dissipate moving westward [Einarsson, 1991, ­towering 300–1,000 m above the broad and flat‐lying
2008; Trønnes, 2002]. Historic earthquakes along the lava plains [Roaldset, 1983; Thordarson, 2012]. The
SISZ have interesting patterns of seismicity character- collective thickness of the Tertiary Basalt Formation is
ized by sequences that last up to 30 years; a complete estimated to be 10,000 m, although the measured
cycle is ~140 years [Stefánsson and Halldorsson, 1988]. thickness is never much more than 3000 m due to lat-
Earthquake events have been recorded since 1103 CE, eral spreading during the cooling phase (Table 4.1)
however, the times between 1732–1734, 1784, and [Einarsson, 1991]. In eastern Iceland, the stratigraphic
1896 CE are particularly well documented due to prop- thickness accumulated from 700 lava flows over 11 Myr
erty damage and loss of lives [Einarsson et al. 1981; (13–2 Ma) [Sæmundson, 1979]; indicating an average
Stefánsson and Halldórsson, 1988]. accumulation rate of 700 m 106 year−1 [Thordarson and
After an 88‐year cessation of seismic activity, it is Höskuldsson, 2014]. Mapping of the plateau basalts
thought that a new earthquake sequence (30‐year cycle)
began as part of the movement initiated by the 17 June
Table 4.1 Lithology, percentage, and thickness
2000 earthquake (Mw = 6.5) that occurred in the east of Icelandic rock types
central SISZ [Bonafede et al., 2007; Decriem et al., 2010].
This event was followed up by another Mw = 6.5 earth- Lithology Thickness (m) %
quake on 21June 2000 that was ~20 km west of the first Tholeiitic basalt 2134 48
one. Field assessment by Pederson et al. [2001] showed Olivine basalt 1036 23
that there was 2.4 m of dextral displacement on near‐ Porphyritic basalt 549 12
vertical faults. These events triggered multiple large‐ Andesitic lavas 122 3
scale aftershocks (Mw >5.0) on the Reykjanes Peninsula Rhyolitic lavas 335 8
approximately 80 km from the epicenter [Clifton et al., Sedimentary and pyroclastic 274 6
Total thickness 4450 100
2003]. As described in section 3.1, this earthquake
activity caused the destabilization of normal faults Note. Adapted from Jakobsson et al. [2008]; design credit
under Lake Kleifarvatn that resulted in a fall of water Nathan Mennen.

Figure 4.1 West to east outcrop section along the south side of Reydarfjördur starting at Langjökull.
[Adapted from Thórarinsson et al. [1959]; design credit Nathan Mennen.]
Tectonics of the South and Southeastern Regions  29

indicates 25 km of eroded basalt from edges of the pla- Throughout southeast Iceland the regular structure
teau, resulting in low‐lying topography adjacent to sea of the Tertiary lava pile is commonly interrupted by
level (Figure 4.1), as can be seen in Reydarfjördur paleofissure swarms (5–15 km wide) of dikes that
(65.0354, −14.2160) [Einarsson, 1991]. crosscut the basalts. This is an example of the corifting

Photo 11 (a) Road to Egilsstađir showing dipping Tertiary Basalt Formation in outcrop. [Courtesy of Russell
Maddrey.] (b) A close‐up image of Tertiary Basalt Formation stacking commonly seen in the eastern fjords.
[Courtesy of Tamie Jovanelly.]
30 TECTONICS

Figure 4.2 An example of a geological cross‐section in eastern Iceland showing regional tilting with
­associated rock‐type descriptions by location. [Modified from Sigmundsson [2006]; design credit Nathan
Mennen.]

phase that follows the interrifting phase, which is often strata to 5–10° near sea level [Sigmundsson, 2006] with
observed at active divergent plate boundaries. During the overall direction of dip toward the central and
the corifting phase, dike injections and eruptions southern parts of the island or, more specifically,
intrude the upper crust, releasing stresses and exten- ­dipping toward the volcanic zones from which they
sional strains that are accumulated during the interrift- originated (Figure 2.1) [Sæmundsson, 1967]. The
ing phases [Pederson et al., 2009; Islam and Sturkell, development of synclinal structures centered on rift
2017]. For example, in Reydarfjördur (65.0354, areas is a flexural response to increases in loading.
−14.2160), Walker [1960] mapped some 500 dikes in a Consequently, anticlinal structures also formed during
40 km long section with a thickness of 1525 m, which rift relocation in the SISZ, as well as in association
he calculated corresponded to an expansion of ~4%. with major unconformities (Figure 2.1) [Sæmundsson,
As described by Thordarson [2012] the dike swarms are 1967]. Prior to the tilting of the basalts by tectonic
associated with andesite–rhyolite lava and tephra movement, vertical uplift caused some 600 m of
deposits that likely represent the location of extinct ­displacement [Thórarinsson et al., 1959]. Textbook
central volcanoes: 40 have been identified in the Tertiary examples of the Tertiary basalts can be seen en route
succession. to Egilsstaðir as Highway 1 ascends along the lake
A characteristic feature of the Tertiary Basalt Lagarfljót (65.2180, −14.4728; Photo 11). A side trip
Formation is the regional tilting of the lava pile. In on Highway 953 from Mjóifjörður (65.2000, −13.8000)
general, the dip can vary from near zero in the upper will find other well‐known exposures (Figure 4.2).
5
Tectonics of the Northeastern Region

The EVZ meets the North Volcanic Zone (NVZ) in overlap of their respective fissure swarms, which tend
conjunction with the MIB beneath Vatnajökull where to spread in episodic events. This occurs because
a mantle plume (hot spot) presently resides [Darbyshire fractures within the fissure swarms are activated dur-
et al., 2000], to form a triple junction (Figure 5.1). The ing rifting events, when magma intrudes the fissure
current offset between the spreading axes of the NVZ swarms to form dikes or even fissure eruptions
and WVZ and the overall divergent plate boundary [Sæmundsson and Karson, 2006]. The density of fis-
defined by the RR and the KR is 100–150 km sures within the swarms is greatest within 20–30 km of
(Figures 1.1 and 2.1) [Björnsson et al., 1977; Angelier the central volcano, while the distal areas of the fissure
et al., 1997]. The KR is connected to the island via the swarm are mostly characterized by noneruptive
transform Tjörnes Fracture Zone (TFZ) that joins fractures. Historically documented rifting events
the NVZ. The NVZ is simpler than other parts of the within the NVZ have occurred in the Krafla volcanic
MAR in Iceland as here the plate boundary has only system (1724–1729 and 1975–1984 CE), Askja
one oblique branch, with the main zone being active volcanic system (1875–1876 CE), and the Þeistareykir
for the past 6–7 Myr and for the past 2 Myr propa- volcanic system (1618 and 1885 CE) [Sigurðsson and
gating northwards [Trønnes, 2002]. The long‐term Sparks, 1978; Magnúsdóttir and Brandsdóttir, 2011;
local seafloor‐spreading rate is measured at Hartley and Thordarson, 2013]. As noted by Einarsson
1.8 cm year−1 [DeMets et al., 1994] and the simplicity [2008], although all share characteristics of a volcanic
of the system lends itself well to the study of the system, they also each have distinguishable features.
rheology of an interrifting phase showing strain
­ For example, the Þeistareykir central volcano does not
accumulation over an area much wider than the have a developed caldera indicating the absence of
­associated vertical deformation [Pedersen et al., 2016]. shallow crustal magma chambers. In this volcanic
A crustal thickness of 30–40 km in the axial rift zone system, olivine–tholeiitic lava shields are more
of northeast Iceland has been estimated by magneto- numerous than elsewhere in the axial rift zone. Next,
telluric measurements, whereas thickness decreases the Kraftla system has a caldera, but it has been sepa-
(20–30 km) with age towards the older Tertiary areas rated by a fissure swarm, causing the infilling of the
to the east and west of the NVZ, and with increasing depression [Brandsdöttir and Einarsson, 1979; Johnsen
distance from the hot spot [Björnsson, 1985]. The main et al., 1980; Hauksson, 1983; Tryggvason, 1994]. In
tectonic feature of the NVZ is fissure swarms, although turn, the Askja system is dominated by the activity of
strike‐slip faults, normal faults, and hyaloclastite a central volcano, despite having one of the longer
ridges are also present. The fissure swarms are mainly ­fissure complexes in the group.
situated in the center of the 200 km long and 50 km In summary, the NVZ plate boundary connects the
wide NVZ, which comprises five volcanic systems each central volcanoes found in the NVZ; the NVZ is
having a central volcano: Þeistareykir, Krafla, oblique to the individual fissure swarms in this area.
Fremrinámar, Askja, and Kverkfjöll (Figure 5.2). Björnsson [1985] suggested that this might indicate that
These systems are arranged en echelon with some the fissure swarms in northeast Iceland are created in

Iceland: Tectonics,Volcanics, and Glacial Features, Geophysical Monograph 247, First Edition. Tamie J. Jovanelly.
© 2020 American Geophysical Union. Published 2020 by John Wiley & Sons, Inc.

31
32 TECTONICS

Figure 5.1 Crustal accretion, relocation, and propagation of the Icelandic rift zones in the past 12 Myr.
Notable features include: at 8 Ma, spreading along the Snæfellsnes and Skagi rift zones; at 6–4 Ma, the
incipient propagation and mature development of the West Volcanic Zone (WVZ) and North Volcanic Zone
(NVZ) after new rift initiation at about 7 Ma; at 2 Ma to present, the southward propagation of the East
Volcanic Zone (EVZ), initiated at about 3 Ma. [Modified from Trønnes [2002]; design credit Nathan
Mennen.]

the upper brittle crust mainly by tensile crack the 35 km of onland exposure (i.e., in a straight line
formation. The central volcanoes of these volcanic from Tjörnes to Dalvík) characterized by zones of
systems are areas of volcanic silica rocks and geo- crushed rocks associated with localized steeply dip-
thermal energy. Upper Pleistocene rocks in this ping Pleistocene lava flows; dips between 14° and 36°
­divergent zone, however, often produce pliable lavas are found in coastal areas closest to the junction between
that rarely produce distinct edifices, or if they do ini- the Húsavík–Flatey Fault and the KR [Rögnvaldsson
tially, they are prone to post‐depositional subsidence et al., 1998; Angelier et al., 2000; Bergerat and Angelier,
[Einarsson, 1991]. 2000; Garcia et al., 2003]. Additionally, in this area
extensive sets of mineral veins occur, as well as
5.1. FLATEYJARSKAGI AND TJÖRNES numerous minor faults that strike subparallel to the
PENINSULAS TFZ [Gudmundsson et al., 1993; Angelier et al., 2000].
As the TFZ is oriented obliquely to the spreading
The Flateyjarskagi (65.9977, −17.9857) and Tjörnes direction, both right‐lateral and extensional faulting
(66.1500, −17.1500) peninsulas are crosscut by the are present [Angelier et al., 2000]. The TFZ has been in
TFZ to create a tectonically complicated area that is existence for 8 Ma and is the result of a rift jump that
characterized by moderate seismic activity (M < 5.0), initiated spreading in the NVZ [Garcia et al., 2003].
crustal extension, and transform faulting. The TFZ Consequently, this rift jump in the northern region of
extends 1–3 km offshore, providing the potential for Iceland is thought to be considerably older (~6 Ma)
tsunami events in this area, as evidenced by the tsu- than that of the SISZ, which results in more mature
nami warning initiated on 26 March 2017 from an off- transform processes and complex stress patterns
shore quake. Distinctive morphological features mark (Figure 3.2) [Bergerat and Angelier, 2008].
Tectonics of the Northeastern Region  33

Figure 5.2 Five active fissure belts found in the Northern Volcanic Zone on the Tjörnes Peninsula. [Design
credit Nathan Mennen.]

The TFZ comprises three subparallel northwest‐ right‐lateral HFF intersects the normal faults of the
trending seismic zones: Grímsey Oblique Rift (GOR), Þeistareykir fissure swarm (Figure 5.2) [Magnúsdóttir
Húsavík–Flatey Fault (HFF), and Dalvík Zone (DZ). and Brandsdóttir, 2011]. The offshore component of
Historically the HFF has been considered the prin- the HFF forms a valley 5–10 km wide and 3–4 km deep
ciple fault because it connects with the KR offshore that meets the Kolbeinsey sea stack (66.6700, −18.5000)
(Figure 2.1) [Einarsson, 1991; Bergerat and Angelier, [Gudmundsson, 2000]. The fault can be seen from the
2008; Stefánsson et al., 2008; Metzger et al., 2012]; top of Mount Húsavíkurfjall (66.0500, −17.3000)
however, the GOR was formed ~2 Ma and currently coming on shore in the bay at Húsavík. Total fault dis-
takes up most of the tectonic movements within the placement is some 60 km over 6 Myr [Sæmundsson,
TFZ [Sigmundsson, 2006; Magnúsdóttir et al., 2015]. 1974; Thordarson and Höskuldsson, 2014]. Both exten-
Over the past decade, the submerged northern portion sional and compressional stress fields associated with
of the GOR, specifically near the island Grímsey, the TFZ have several effects on local tectonic processes
which is 40 km off the northern coast, has had fre- and the topography in Húsavík. Extensional forces
quent earthquake swarms with durations that span have resulted in the presence of two pull‐apart (sag)
days or weeks [Iceland Metorological Office, 2018]. basins developed in low‐lying areas [Botnsvatn,
The maximum magnitude of such earthquakes exceeds 66.0058, −172753; Höskuldsvatn, 65.9850, −17.1983).
5.0, with faulting occurring as right‐lateral strike‐slip. In contrast, compressional processes have lead to
The HFF occurs on land at the town of Húsavík vertical uplift of Pleistocene strata and dip‐slip
(66.0450, −17.3383) and continues eastward as the behavior on the HFF. Up to 200 m high fault scarps
34 TECTONICS

(e.g., the Skjólbrekka slope just north of Húsavík) on Mývatn. Pleistocene basalt ridges toward the west
the Tjörnes Peninsula indicate a significant dip‐slip and möberg ridges (Blafjall and Sellandafjall) sur-
component causing vertical displacement of up to round the shallow lake basin to the east. Numerous
1400 m on land and 1100 m offshore [Gudmundsson postglacial crater rows, open fissures, and normal
et al., 1993; Rögnvaldsson et al., 1998]. Four major faults (vertical d­ isplacement ranging from 10 cm to
earthquakes have occurred on the HFF during the 3 m) are found on the east side of the lake. A swarm
past 200 years. In 1755, an earthquake with an esti- of fault scarps west of Mývatn, running from Mount
mated M = 7.0 occurred in Skjálfandi Bay and in 1838 Vindbelgur (65.6248, −17.0672) northwards through
a M = 6.5 event occurred near the western end of the Lake Sandvatn (64.4260, −20.1824), are of early
bay. The last major earthquake (M 6.5) on the eastern postglacial age. The faults and eruption fissures in
part of the fault occurred in 1872 and caused extensive the Mývatn area are organized in a NNE–SSW orien-
damage and surface ruptures in the town of Húsavík. tation and are part of a much longer fracture zone
Most of the present‐day seismicity occurs offshore on extending from Axarfjördur (66.0017, −175122) to
the northwest part of the HFF. An analysis of the Sellandafjall (65.4163, −17.0411) [Thordarson and
earthquake swarms along both the GOR and HFF Höskuldsson, 2014].
zones found the majority (i.e., 90%) of earthquake The area may be best known geologically for the
focal depths were shallower than 10 km [Rögnvaldsson Mývatn Fires of 1724–1729 CE when, 800 years after
et al., 1998]. Additionally, recent research suggests that humans first settled in the area [Olafsson, 1979], a
the HFF currently has the potential for a M 6.8 earth- basaltic fissure eruption produced 0.45 km3 of basaltic
quake [Metzger et al., 2011, 2012]. This is a concern lava covering 30 km2 from a discontinuous fissure
because Húsavík has over 2000 inhabitants. 11 km long [Granvold, 1984]. As described by Björnsson
The DZ forms the southernmost part of the TFZ. et al. [1978] the volcanic and seismic activity associated
Although there are no obvious fault scarps on the sur- with this event was intense for short periods (days)
face representing the DZ there has been some modern and then followed by longer periods of quiescence.
seismic activity. Most notably, in 1963 a M = 7.0 earth- Sæmundsson [1978] has evaluated historic Icelandic
quake occurred in the westernmost region of the DZ documentation of the Mývatn Fires based on geolog-
near the mouth of the Skagafjörður River. In 1934 the ical mapping of the area [Thoroddsen, 1925]. From
town of Dalvík experienced a M = 6.2 event (estimate these accounts it was determined that most of the lava
based on historic accounts). The DZ, however, may was erupted during the last 2 years of the episode.
best be known for its morphological features, such as During a major eruption on 30 June 1729, a lava flow
river canyons (e.g., Jokulsarglijufur National Park). traveled 15 km from Leirhnjúkur (65.7200, −16.7939)
toward the town of Reykjahíð (65.6449, −16.9059),
5.2. LAKE MÝVATN AND MÝVATN FIRES where it covered four farms but was diverted around
a church that is still standing today [Myvtan Visitor
The area of Mývatn (65.6039, −16.9761) is part of Center, personal communication]. Accounts of the
the Krafla Fissure (Figure 5.2) with bedrock of first 3 years, from 17 May 1724, describe intense
mostly basalt and hyaloclastites [Thórarinsson, 1979]. ­earthquake swarms, fault movements, and elevation
The original lake basin was first created around changes that took place in a rapid series of events,
38 ka, ­during what is referred to as the Lúdent Cycle, including the hydromagmatic eruption to form the
with an effusive eruption of Ketildyngja (65.4293, Viti crater (65.04697, −16.72612) that will be discussed
−16.6544) that dammed outflow from melting gla- in subsequent chapters. Also on 17 May 1724, fourteen
ciers to create a basin [Olafsson, 1979]. The lava fields rootless cone craters were formed at the surface
between Reykjahlið and Vogar Farm were produced (Photos 12 and 13).
during this time from an eruption at Jarðbaðshólar
(65.6031, −17.0047). The current morphology of 5.3. KRAFLA FIRES
Lake Mývatn (4.2 m deep, 37 km2) was determined by
a large fissure eruption ~23 ka. The lava flowed down The Krafla Fissure (Figure 5.2; ~80 km long and
the Laxárdalur Valley (65.6362, −17.1635) to the 4–10 km wide) contains the 200,000‐year‐old central
coastal plain of Adaldalur (65.8936, −17.3848) and volcano, Krafla (65.7286, −16.7478), and extends from
then to the Arctic Ocean, a distance of 50 km from the TFZ in the north at Axarfjördur to south of Lake
Photo 12 Rootless cones at Mývatn. [Courtesy of Tamie Jovanelly.]

Photo 13 Side profile of a rootless cone at Lake Mývatn. [Courtesy of Tamie Jovanelly.]
36 TECTONICS

Table 5.1 Krafla Fires eruptions (1975–1984) pattern and position of the fissures has raised the
Eruption Date Duration (days)
possibility of there being another magma chamber
south of the Krafla caldera. This is particularly
1 20–21 December1975 2 important because of the potential expansion of geo-
2 27–29 April 1977 3
thermal fields. Currently, the Krafla Geothermal
3 8–9 September 1977 2
Energy facility operates in the Namafjall area, about
4 16 March 1980 1
5 10–18 July 1980 9 10 km south of the Krafla central volcano.
6 18–23 October 1980 6 Additionally, the use of satellite imagery combined
7 30 January to 4 February 1981 6 with aerial photographs, digital elevation models, and
8 18–23 November 1981 6 other measurement techniques, has provided intense
9 4–18 September 1984 15 ground surveying of this event. Using this technology,
Hollingsworth et al. [2013] quantified that the total
Note. Design credit Nathan Mennen.
volume of magma added to the upper crust was 1.2–
2.1×109 m3 and from comparisons with ground
fracturing observations confirmed the overall episodic
Mývatn (65.6031, −17.0047). After 250 years of inac- pattern of rifting events.
tivity of the Mývatn Fires, the Krafla Fissure opened
to start an episode termed the Krafla Fires (1975–1984 CE) 5.4. DETIFOSS, SELFOSS, AND HAFRAGILSFOSS
[Björnsson et al., 1978; Opheim and Gudmundsson, WATERFALLS
1989]. From 1975 to 1984 CE there were nine mag-
matic eruptions and 20 rifting events that occurred At the headwaters of the Jökulsárgljúfur canyon,
within 7 km of the central volcano, with each event the Jökulsá á Fjöllum river is deeply incised into the
lasting a maximum of 5 days (Table 5.1) [Tryggvason, surrounding terrain, with the drop in elevation
­
1984, 1986]. ­occurring at three large vertical waterfalls, all within a
The first event on 20 December 1975 began with 5‐km‐long reach [Baynes et al., 2015]: Selfoss (13 m
earthquake swarms (M = 4.0) around the Krafla high), Dettifoss (54 m high; 65.8147, −16.3846; Photo
­caldera that resulted in ground fissuring and small 14), and Hafragilsfoss (20 m high). The Jökulsá á
lava eruptions north of the hyaloclastite ridge, Fjöllum river is the second longest river in Iceland
Leirhnjúkur (65.420, −16.480). Soon after, the first with a length over 200 km and experiences large
of many rifting events would follow in active pulses summer discharge flows reaching 1500 m3 s−1 [Wells,
lasting a few months at a time. At the end of each 2016]. The bedrock is composed mostly of columnar
pulse, the magma reservoir estimated to be 3 km below basalt, with several subhorizontal lava flows stacked
the caldera would gradually inflate at a rate of one on top of the other [Baynes et al., 2015;
7–10 mm day−1, with magma infilling the reservoir at Figure 5.3). The unique Icelandite, a type of andesite
a rate of 5 m3 s−1 [Björnsson et al., 1979]. Once the rock that is iron rich, can also be found intermittently.
­reservoir met a critical threshold the pattern would The paleoposition of the river can be suggested by
begin to repeat itself, with earthquakes, vertical ground three strath terrace surfaces found at different eleva-
movement (<2 m), and sometimes volcanic eruptions tions throughout the canyon.
[Einarsson, 1978; Tryggvason, 1980]. Notable rifting The initiation of this canyon is suspected to be the
events include 10 July 1980 when a 5–6 km2 lava field result of extreme flood events, called jökulhlaups, that
was produced, followed by a 5‐day rifting event from caused catastrophic landscape changes after the end of
13 to 23 November 1980 that erupted another 17.5 km2 the last glaciation [Sæmundsson, 1973; Eliasson, 1977;
of lava at the surface. Baynes et al., 2015]. This is supported by the correla-
With instrumental monitoring in place during the tion of exposed surface ages to three periods of intense
Krafla Fires, scientists were able to closely observe canyon cutting about 9, 5, and 2 ka during which mul-
the event. Extensive mapping by Hjartardóttir et al. tiple large knickpoints retreated large distances
[2012] illustrated uneven distribution patterns of the (> 2 km) during Holocene glacial melting events
fissure eruptions during the 1975–1984 episode. (Photo 14; Baynes et al., 2015].
Specifically, they noticed that although a cluster of Additionally, around 9 ka two volcanic fissures
fissures was within ~7 km of the caldera, there was erupted in the area: Rauðhólar and Hljóðaklettar. The
also a cluster in the southern region of the Krafla Rauðhólar fissure is thought to contain the largest
Fissure up to 70 km away. The irregularity in the crater row on Earth (nearly 150 craters) and was
Tectonics of the Northeastern Region  37

Photo 14 Strath terraces and knickpoints at Detifoss. [Courtesy of Russell Maddrey.]

Figure 5.3 A lithological cross‐section at Detifoss. [Adapted from Thórarinsson [1959]; design credit
Nathan Mennen.]

­ issected by the Jökulsá á Fjöllum river, which cut


d (65.8294, −16.4136). Both the Rauðhólar and
through the feeder channel of one of the volcanic cra- Hljóðaklettar fissures produced large lava fields,
ters. This opened a unique cross‐sectional view of the although catastrophic fluvial flooding events have
magma channel in the canyon’s east wall near Hafragil eroded most of the deposits.
6
Tectonics of the Western Region

The 80 km long and 10 km wide Snæfellsness to become convergent, thereby developing areas of
Peninsula is about 1 h drive north of Reykjavík. The uplift (mountains) that now divide the p ­ eninsula in
interior mountain chain provides picturesque views of half. Three volcanic zones formed during this time that
the 700–1000 m tall peaks of the Snæfellsjökull strato- have similar characteristics to those seen in the active
volcano (Photo 15). The Snæfellsness Peninsula is NVZ today. The volcanic systems of the SVB,
­distinctly different from other parts of the island in Snæfellsjökull (64.8060, −23.7768), Lýsuskarð
that the Snæfellsnes Volcanic Belt (SVB) is orientated (64.8520, −23.2075), and Ljosufjöll (64.9145,
east–west, being almost perpendicular to the main −22.5803), are arranged en echleon and trend WNW
zones of rifting in Iceland (Figure 6.1). Prior to the (i.e., mimicking paleoplate movement). Despite the
establishment of the WVZ, the rift zone at Snæfellsnes mountain range and volcanic systems, the reported
Peninsula extended northwards to Skagi (64.3665, crustal thickness on the peninsula is between 15 and
−20.5832) and was the main spreading center from 20 km, which is the thinnest crust on Iceland, with the
about 15 Ma to 7 Ma when the hot spot was offshore Mohorovičić discontinuity depth beneath Snæfellsnes
(Figure 1.1). At this time the Snæfellsnes–Skagi Rift being another 4 km deeper toward the mantle [Allen
Zone linked directly to the KR [Martin et al., 2011]. et al., 2002b; Du et al., 2002]. Lavas of the three
The center of the zone is marked by a synclinal struc- volcanic systems lie unconformably above a 1.2 Ma
ture in the lava [Jóhannesson, 1980] (Figures 2.1 and sedimentary horizon that can be viewed in Búlandshofði
6.2). Evidence for an older rift zone is found at the (64.9429, −23.4889). The three volcanic systems are
extreme northwest of Iceland where a 14.9 Ma uncon- interesting tectonically because they comprise alkali
formity separates lava dipping towards the younger olivine basalts and transitional alkali basalts that are
axis of the Snæfellsnes–Skagi Rift Zone, incidentally typical of flank zone deposition, in contrast to the pre-
separating the older lava flows [Harðarson et al., 1997]. dominantly tholeiitic flood basalts in the center of the
The Snæfellsnes–Vatnsnes Rift Zone emerged some island [Jakobsson, 1972; White and McKenzie, 1989;
8 Ma when the mantle plume created a spreading axis. Jakobsson et al, 2008; see Figure 7.1 below]. Although
This now ancient rift was active until 6 Ma when the the SVB was very active during the Holocene (25 effu-
spreading axis migrated towards the WNW at sive and two explosive eruptions) none have occurred
0.3 cm year−1. As the active spreading axis moved away, during human settlement or in the past 1100 years,
by 3 Ma the mantle plume changed the position of the which is why many consider the system to be extinct.
axis to form a new rift (namely the WVZ) closer to its The Snæfellsness Peninsula is regarded today as an
center (Figure 1.1). By 1 Ma, the edge of the mantle aseismic region, although six earthquakes (M < 5.0)
plume was still in proximity of the Snæfellsness occurred from 1912 to 1962 [Sigurðsson, 1970] and
Peninsula. Continued plate movement caused the microseismicity (M < 1.0) has been monitored
divergent plate boundary that crosscut the peninsula underneath the southern and northeastern flanks of

Iceland: Tectonics,Volcanics, and Glacial Features, Geophysical Monograph 247, First Edition. Tamie J. Jovanelly.
© 2020 American Geophysical Union. Published 2020 by John Wiley & Sons, Inc.

39
40 TECTONICS

Photo 15 The largest peak on the Snæfellsness Peninsula is the stratovolcano, Snæfellsjökull. [Courtesy of
Tamie Jovanelly.]

Snæfellsjökull volcano that originated from depths and young toward the center axis (Figure 6.2). It
of 8–15 km. This seismic activity likely reflects should be borne in mind, however, that most of the
movement of fluids within the ductile crust [Fuchs Tertiary basalts on the peninsula are buried beneath
et al., 2013]. The peninsula continues to be monitored Plio‐Pleistocene–Holocene deposits (Figure 6.1). To
by the Iceland Meteorological Office. the north, toward the Vestfirðir Peninsula, and to the
south, along part of the Mýrar region, Tertiary basalts
6.1. SNÆFELLSNES SYNCLINE can be seen in age sequence, i.e., becoming older
AND BORGARFJÖRÐUR ANTICLINE toward the flanks of the syncline.
In Borganes (Figure 6.1; 64.5609, −21.9010) evidence
The axis of the Snæfellsnes Syncline (Figure 2.1) of the Borgarfjorður Anticline can be seen. Opposite to
reflects the active spreading center of the paleo‐ a syncline structure, the oldest rocks (here 13 Ma) occur
Snæfellsnes–Vatnsnes Rift Zone that began rupturing at the axis of the anticline (Figure 6.2) and become pro-
some 6 Ma, hence, the youngest of the Tertiary Basalt gressively younger toward the flanks: ~12 Ma in the
Formation exposed on the Snæfellsness Peninsula are WVZ to the south and 6–8 Ma towards Hítardalur
the same age (Figure 6.1). In accordance with the prin- (64.8279, −22.0651) in the north (Figures 2.2 and 6.1).
ciples of syncline formation, these young basalts are It is likely that the syncline and anticline structures
found along its center, which stretches obliquely from formed post‐rift jump as the plate movement trajectory
the north (Alftafjörður) to the south coast (Garðar). changed to WNW thereby creating convergence and
On both sides of the syncline, the strata will dip toward closure of the Snæfellsnes–Vatnsnes Rift Zone.
Tectonics of the Western Region  41

Figure 6.1 Location map of Snæfellsness Peninsula with associated ages and rock type. The Snæfellsnes
Volcanic Belt (SVB) is illustrated. The strike and dip symbols highlight the Snæfellsnes Syncline and
Borgarfjöd ̄ur Anticline structures. [Design credit Nathan Mennen.]

Within the rock record across the peninsula the


Tindar horizon (8 Ma) experiences almost continuous
deposition with few discontinuities, whereas the
Hreðavatn horizon records notable discontinuities and age
differences between the Tertiary basalts. At Hreðavatn,
for example, the age difference is about 5 Ma (12 Ma
on bottom, 6.5 Ma on top) between distinctive events.
Identifying this unconformity lead to development
of the rift‐jump hypothesis with the extinction of
the Snæfellsnes–Vatnsnes Rift Zone. Accordingly, the
Tertiary basalts are observed to be older toward the
WVZ (away from the rift axis), whereas ages at Mount
Figure 6.2 Schematic diagrams of anticline and syncline Hafnafjall and Hreðavatn are the same as they are the
structures. [Design credit Nathan Mennen.] same distance from the plate divergence axis.
Part II: Volcanics
7
Overview of Volcanics in Iceland

7.1. VOLCANIC SETTING and pyroxenes from a tholeiitic basalt magma. The
evolution of the Fe‐rich magma would then cause the
The dynamic landscapes of Iceland are the result of iron to become depleted in successive liquids, forcing
active spreading of the MAR and the presence of a the magmatic composition toward rhyolite. The other
hot spot that is currently located under Iceland’s larg- is that dehydration melting of tholeiitic basalts due to
est glacier, Vatnajökull. The high volumes of magma high heat flow from hot‐spot proximity may also pro-
production compared with other parts of the MAR vide a mechanism for the generation of Icelandic rhyo-
results in crustal thickness averaging about 30 km lite (Figure 7.1) [Sigvaldason, 1974a; Thy et al., 1990].
beneath the hot spot. This active magmatism continues Over the past 11 centuries, ejections of basalt, andesite,
to increase crustal capacity by contributing intrusions dacite, and rhyolite magmas have resulted in varying
(dikes, primarily), surface eruptions, and addition by eruption styles, including: effusive eruptions producing
underplating. The island’s neovolcanics are dictated by high‐magnitude lava floods; mixed Strombolian to
those areas associated with crustal spreading (volcanic Plinian eruptions; explosive phreatomagmatic and
rift zones) as well as volcanic eruptions (Figure 2.1). magmatic eruptions spanning the volcanic explosively
The predominant type of basalts found on Iceland are index (VEI) from low (0) to high (8) intensity
tholeiitic basalts, which are unusual in that they tend [Thordarson and Larsen, 2007] (Table 7.1).
to show higher concentrations of discordant trace ele- There is some discrepancy in the total number of
ments that classify them as enriched mid‐ocean ridge active and dormant volcanoes on the island, for two
basalts (EMORB) [Nelson, 2003]. Different from most reasons. First, the North American Plate continues to
MOR scenarios, tholeiitic basalts transition to alkali move westward as are the volcanic zones (namely, the
basalts within the rift zones (Figure 7.1) [Sigurdsson, NVZ, EVZ, WVZ) (Figures 1.1 and 2.1). Herein, a
1966; Jakobsson, 1972a, 1979a, 1979b]. This atypical volcanic zone is defined as comprising active central
amount of silica‐rich deposits (approximately 10–12%) volcanoes and extensional fissures, or “fissure swarms,”
is found at both active and inactive (Tertiary age) which are associated with normal faulting, geothermal
central volcanoes compared with other islands built on activity, and the possibility of a post‐eruption caldera.
oceanic crust [Gunnarson et al., 1998]. It was initially Second, although monogenetic (i.e., single eruption)
thought that the silica was introduced by the melting basaltic volcanoes are by definition extinct, there are
of continental crust beneath Iceland. Two viewpoints many on Iceland that remain in zones considered vol-
regarding silica magma production have emerged, canically active [Walker, 1999]. Most scholars agree,
however. One is fractional crystallization of primitive however, that there are currently 33 active volcanoes
basalts or the partial melting of hydrothermally altered on the island of Iceland (Figure 7.2), and these repre-
basaltic crust suites [Schattel et al., 2014]. As explained sent nearly all volcano types and eruption styles known
by Nelson [2003], Fe‐enrichment trends would be on Earth [Thórarinsson and Sæmundsson, 1979;
expected by the removal of early crystallizing olivines Thórarinsson, 1981]. The volcano types are unevenly

Iceland: Tectonics,Volcanics, and Glacial Features, Geophysical Monograph 247, First Edition. Tamie J. Jovanelly.
© 2020 American Geophysical Union. Published 2020 by John Wiley & Sons, Inc.

45
46 VOLCANICS

Figure 7.1 Tholeiitic basalt phase diagram showing changes in alkalinity as it corresponds to Iceland’s rift
zones. [Design credit Nathan Mennen.]

Table 7.1 The volcanic explosivity index (VEI)


VEI Eruption description Tephra volume (km3) Cloud column height (km) Eruption type Duration (h)
1 Gentle 10,000 (m )3
0.1–1 Hawaiian/strombolian <1
2 Explosive <0.01 1–5 Strombolian/vulcanian <1–6
3 Severe 0.01–0.1 3–15 Plinian/vulcanian <1–12
4 Cataclysmic 0.1–1 10–25 Plinian/vulcanian 1–>12
5 Paroxysmal 1–10 >25 Plinian/ultraplinian 6–>12
6 Colossal 10–100 Plinian/ultraplinian >12
7 Supercolossal 100–1000 Plinian/ultraplinian >12
8 Megacolossal >1000 Ultraplinian >12
Note. Modified from Newhall and Self [1982]; design credit Nathan Mennen.

distributed between volcanic zones: three within the Figure 7.2), which are all located in the EVZ. In
RVB, nine in the WVZ, two in the MIB, five in the addition, there are approximately 150 dormant
NVZ, eight in the EVZ, three in the SVB, and three in Icelandic volcanoes, i.e., volcanoes that have been
intraplate volcanic belts [Jóhannesson and Sæmundsson, inactive for a minimum of 10,000 years.
1998] (Figure 7.2). According to Thordarson and The resulting volcanic systems are variable: twelve
Larsen [2007], 80% of the verified historic eruptions are comprised of a fissure swarm and a central vol-
have taken place within the four most active volcanic cano, seven of a central volcano, nine of a fissure
systems (Grímsvötn, Bárdarbunga, Hekla, and Katla; swarm and a central domain, and two with a single
OVERVIEW OF VOLCANICS IN ICELAND 47

Figure 7.2 Location map of active and extinct volcanoes on Iceland. Active volcanoes are listed by number.
EVZ, East Volcanic Zone; WVZ, West Volcanic Zone, NVZ, North Volcanic Zone; SVB, Snæfellsnes Volcanic
Belt. [Modified from Thordarson and Larsen [20067; design credit Nathan Mennen.]

central dome [Thordarson and Larsen, 2007]. This mixed. There has been recent discussion surrounding
delineation is based on recurrence of eruptions hap- the current classification of effusive eruptions in
pening on the same vent system and further suggests Iceland as either monogenetic or polygenetic. Sheath
the separation of central volcanoes (polygenetic, i.e., and Canon‐Tapia [2015] noted that fissure eruptions
built from multiple eruptions, type) from basalt vol- can have either single or multiple feeder‐dike injec-
canoes (monogenetic eruption type). Exploration of tions. They argue that these flood basalt events should
the geologic record through tephrochronology and be termed accordingly as “diffuse volcanic fields”
detailed mapping indicates that some 205 historic instead of by the volcano classification type.
events have occurred, which averages 20–25 eruptions There have been approximately 13 eruptions since
per century [Larsen and Eiríksson, 2008a]. Of these 205 the Víkings settled Iceland in 874 CE, with some events
events, 192 represent distinct eruptions and 13 are clas- causing dramatic lifestyle and landscape changes, as
sified as long‐term (one month to several years) fissure discussed in Box 7.1 and the following chapters. Hekla
eruption events often termed “fires” (e.g., Table 5.1). is among the most famous volcanoes on the island as it
Of the eruptions identified, approximately 78% are has erupted 18 times since 1104 CE, with the last time
classified as explosive, 8% as effusive, and 13% as being in 2000 CE. Also in the EVZ, Katla has erupted
48 VOLCANICS

Box 7.1 Living with Volcanoes


Emily Larrimore
Globally, almost 500 million people live within attacks, bronchitis, and chest tightness, and can
exposure range of a volcano that has been active worsen preexisting conditions such as chronic bron-
in recorded history [Hansell and Oppenheimer, chitis and advanced heart problems [Horwell and
2004]. To those who experience an eruption Baxter, 2006]. Studies including historical and con-
living within this range can cause negative acute temporary eruptions have found that volcanogenic
and chronic health effects, as well as negative gases have accounted for <1–4% of fatalities caused
psychological impacts. In Iceland, multiple by eruptions, however, this is likely an underesti-
volcanic eruptions can occur within an inhabitant’s mate of total deaths caused by volcanic gases as
lifetime so it is important that the effects of erup- these studies do not fully account for overall volcanic
tions are studied thoroughly. degassing not associated with eruptions [Hansell
Being that Iceland is endowed with many unique and Oppenheimer, 2004]. Poisoning from volcano-
volcanic features that can become destructive fairly genic H2S emissions can cause unconsciousness
quickly, it is important to study the impact that and death within minutes of exposure to concentra-
natural disasters have on the population of tions of >700–1000 ppm [Hansell and Oppenheimer,
Iceland. Hlodversdottir et al. [2016] completed a 2004]. Population studies of chronic low‐level
population‐based study on the effect that the 2010 exposure show dose‐dependent trends for increases
Eyjafjallajökull eruption had on the physical and in diseases of the nervous system [Hansell and
mental health of 2066 exposed children who lived Oppenheimer, 2004]. Furthermore, CO2 released
near the volcano. This study found that adverse during an eruption with a concentration 30%
physical and mental health issues such as increased can cause unconsciousness in 1 min. Prolonged
likelihood of respiratory symptoms, increased exposure can cause seizures and lead to death
anxiety, increased frequency of headaches, and [Hansell and Oppenheimer, 2004]. Emissions of
increased frequency of sleep disturbances were SO2 can cause short‐term effects such as eye irrita-
exhibited by children living in exposed areas tion and severe airway constriction in asthmatics
6–9 months after the eruption [Hlodversdottir [Horwell and Baxter, 2006]. Epidemiological and
et al., 2016]. This study also found children who clinical studies of the health effects of volcanic gases
had experienced damage to their homes or other are still lacking, and future studies are required to
possessions were also more likely to experience properly gauge the full scale and intensity that vol-
anxiety and depression. The researchers in this canogenic emissions have on human health. This is
study were surprised to find that in 2013, three especially true as tourists are becoming increasingly
years after the eruption of Eyjafjallajökull, the chil- fascinated with volcanic areas such as Iceland.
dren in this study were still experiencing adverse Icelandic eruptions also have great effects on
symptoms. One reason for this is that children are continental Europe. In 1783–1784 a series of erup-
an extremely vulnerable group during a natural tions of the Laki volcano lasting 8 months resulted
disaster, and they require developmentally appro- in the death of around a quarter of the population
priate intervention and preventative strategies of Iceland, and created a low‐lying “haze” that
in order to avoid severe psychological trauma caused wide‐sweeping effects and damage
[Hlodversdottir et al., 2016]. throughout Europe, especially in the United
In terms of physical health effects, epidemiolog- Kingdom [Sonnek et al., 2017]. Volcanic emissions
ical studies have found a strong relationship ­between caused significant crop failures in Europe, resulting
both acute and chronic exposure to particulate in increased mortality, as well as negative health
matter and mortality [Ostro, 1984; Pope and effects including eye irritation, headaches, and
Dockery, 2006; Brook et al., 2010]. The morphology breathing difficulties [Sonnek et al., 2017]. At pre-
of volcanic ash is similar to that of asbestos in that it sent, volcanic emissions have the potential to work
is a respiratory hazard caused by insoluble particles in combination with anthropogenic air pollution to
in the form of fibers [Horwell and Baxter, 2006]. create remarkably harmful air quality [Hansell and
After heavy ash fall, short‐term health problems Oppenheimer, 2004]. In addition, gases emitted
from inhaling particulate matter can include asthma during volcanic eruptions can be very damaging to
OVERVIEW OF VOLCANICS IN ICELAND 49

technological infrastructure fabricated with copper, geysers. Geotourism is advantageous for Icelandic
aluminum, nickel, tin, and silver [Sonnek et al., society as it can promote conservation and sustain-
2017]. Volcanic gasses such as H2S, SO2, HCl, and ability, educate the general public on potential
HF can cause exceptional corrosive damage to risks and hazards, and can benefit local economies
telephones, computers, circuit boards, and other and encourage rural development [Ólafsdóttir and
telecommunication equipment, potentially result- Dowling, 2014].
ing in failure [Sonnek et al., 2017]. Volcanic eruptions obviously cause negative
Although sometimes destructive, Iceland’s effects on society that can result in financial loss,
geology offers it extraordinary economic opportu- environmental degradation, and even fatalities, but
nities in the form of tourism. Geotourism, a the unique volcanic nature of Iceland has also
relatively new trend within tourism, promotes provided many economic, energy, and tourist
travel destinations that are unique in their landscape opportunities. Further research and development
and geodiversity. Geoparks, like the Katla Geopark, of models to predict the negative impacts and chal-
have become popular among travelers for the lenges caused by future eruptions are of vital
promise of getting up close and personal with importance so that the benefits of Iceland’s geo-
geologic features such as volcanoes, glaciers, and thermal activity can outweigh its costs.

20 times since settlement. Since ~1100 CE, it is sug- the volcano’s cone‐shape structure, or edifice, is built
gested that Iceland is responsible for more than one‐ by the symmetrical accumulation of lava or pyroclastic
third of all new lava exposed at the surface, with material around this central vent system through repeated
magma productivity >5 km3 per century [Thordarson eruptions (i.e., polygenetic eruption type), which may or
and Larsen, 2007; Thordarson and Höskuldsson, 2008]. may not be bimodal in magmatic composition. They
This total includes the Earth’s largest eruption of the commonly have a stratified appearance with alternating
past 200 years, as measured by volume, at Bárðarbunga lava flows, airfall tephra, pyroclastic flows, volcanic
in 2014–2015. This eruption lasted six months and lahars, and/or debris flows (Figure 7.4). The three largest
produced 85 km2 of lava, which equates to the size of stratovolcanoes, Öræfajökull (64.5400, −23.8071;
Manhattan. The largest tholeiitic eruption since 2100 m), Eyjafjallajökull (63.631, −19.6083; 1667 m), and
1100 CE occurred over a 50‐day duration at Laki Snæfellsjökull (64.8400, −23.9017; 1446 m), are currently
(1783–1784 CE) producing 370 km2 of lava. in volcanic belts where little or no crustal spreading
occurs (Figure 7.2, Photo 15).
7.2. VOLCANIC MORPHOLOGY Although there are numerous calderas existing today in
Iceland, the most recent formation of a caldera occurred
The morphology of Iceland’s volcanoes varies widely after the Bárðarbunga (64.6411, −17.5281) eruption in
in size and shape, and there is variation in overall life 2015. A caldera forms by the collapse of a volcano into
spans ranging from 0.5 to 1.5 Myr [Jakobsson et al., itself to form a large caldron‐like depression in its place.
1978; Sæmundsson, 1978, 1979; Jakobsson, 1979a] The depression forms after the rapid evacuation of a
(Figure 7.3). Undoubtedly, Iceland is unique in the diver- magma chamber that has been left without structural
sity of volcanic features represented. Moreover, the com- support for the crust above the magma chamber.
plexities of extensional rift tectonics and the continued
presence of a hot spot leads to interesting discussion of 7.2.2. Shield Volcanoes and Associated Features
volcano formation and structure. Although there are
currently 33 active volcanoes on the island, there are also Skjaldbreiður (64.4092, −20.7525) (Photo 16) is a
more than 50 extinct central volcanic features represent- large‐volume (>1 km3) basalt lava shield volcano with
ing 0.78–0.15 Ma (Figure 7.2). a summit crater [Rossi, 1996]. Shield volcanoes in
Iceland can be distinguished from stratovolcanoes by
7.2.1. Stratovolcanoes and Associated Features their heights being typically about one‐twentieth of
their widths. Additionally, the lower slopes are often
Many of Iceland’s volcanoes contain (or once contained) gentle (~3°), but the middle slopes become steeper
a central vent underlying the summit crater, which would (~10°) and then flatten at the summit. This gives shield
classify them as stratovolcanoes or composite cones. Here, volcanoes a flank morphology that is convex in an
50 VOLCANICS

Figure 7.3 Volcano morphology and time line, with relative sizes, shapes, and life spans of various types
of volcanoes. The volcano images are vertically exaggerated two times and the craters and calderas are
exaggerated four times. The approximate erosional expectances of the volcanic structures are listed in
years. [Modified from Decker and Decker [2005]; design credit Nathan Mennen.]

Figure 7.4 Basic volcano morphology. [Design credit Nathan Mennen.]

upward direction. Their overall broad shape results until it separates from the flow. At this time, the gen-
from the extrusion of low‐viscosity mafic lava with low eral mass behind it moves forward. The top of the flow
gaseous content that spreads outward from the sum- cools quickly while the molten magma underneath
mit area. The resulting surface structures will be deter- cools slowly as it is sheltered from contact with the
mined by magma discharge rate and the steepness of atmosphere; a jagged appearance results. In contrast,
the slope over which the lava flows. “Aa” lava flows gentle slopes and lower discharge events form smooth
develop when there is a high discharge rate and steep undulating or ropy masses called pahoehoe that are
slopes (Photo 17). The partially solidified front of the typically <1 m thick (Photo 18). Pahoehoe flows move
flow steepens due to the mass of flowing lava behind it in sheets, or by the advance of the lava toe. Increasing
Photo 16 Pahoehoe field in front of Skjaldbreid ̄ur Shield Volcano located on the Reykjanes Peninsula.
[imageBROKER/ Alamy Stock Photo.]

Photo 17 An example of an aa lava flow. [Courtesy of Tamie Jovanelly.]


52 VOLCANICS

Photo 18 An example of a pahoehoe lava flow. [Courtesy of Tamie Jovanelly.]

viscosity on the part of the lava or shear stress ­ resent. Sometimes the age of the scoria cone can
p
­introduced in the topography can change a pahoehoe be discerned by shape of the edifice whereby newer
flow into an aa flow, but the reverse never occurs. Lava deposits will have slopes up to 35°, and older eroded
tubes are also common features associated with shield cones typically have gentler slopes (15–20°). Unlike
volcanism. By definition, lava tubes are cave‐like strato or shield volcanoes, scoria cones have straight
volcanic features formed by the hardening of the sides and very large summit craters with respect to
outer shell of the tube. These structures help further their relatively small edifices. The red‐oxidized inte-
the transmission of lava because the tube insulates the riors reveal that the scoria cone was breached during
lava moving within. Lofthellir lava cave (65.6317, its formation.
−17.4536) on Snæfellsnes Peninsula provides a text- Volcanic structures often associated with Iceland
book example. Effusive eruptions can also produce are linear‐type volcanoes or volcanic fissures. Hekla
spatter rings and scoria cone volcanoes (Photo 19) that (63.9923, −19.6658; 1491 m) is such an example as it is
represent short (days to months) and small (<0.1 km3) a volcanic ridge that is elongated along the strike of its
to medium (0.1–1 km3) mafic events associated with main eruptive fissure (Photo 20). Fissure eruptions that
shield volcano systems. The height is generally less did not centralize can leave a monogenetic crater row
than 300 m and can occur as discrete volcanoes on whereby multiple craters indicate multiple single‐event
basaltic lava fields, or as a parasitic cone generated by eruptions (Photo 21). The crater rows are more common
flank eruptions on shield volcanoes or stratovolcanoes. in Iceland than in any other volcanic district; the largest
Scoria cones are composed almost wholly of ejected ones have an area of ~50 km2 and craters may be
basaltic tephra that is commonly lapilli size, although counted in hundreds or even thousands [Thórarinsson
bomb‐size fragments and lava spatter may also be et al., 1959]. In size, they range from 1–2 m high and
Photo 19 An example of a spatter cone. [Courtesy of Tamie Jovanelly.]

Photo 20 The Hekla volcano formed by an elongated fissure eruption found along its ridge crest. [Courtesy
of Tamie Jovanelly.]
54 VOLCANICS

Photo 21 An example of a rootless cone. [Courtesy of Tamie Jovanelly.]

2–3 m in diameter to more than 30 m high and with a conducted by Licciardi et al. [2007], which indicated
base of 300–400 m, such as at Mývatn. The failure of 12 of 13 subglacial eruptions had a final eruptive
an eruption to centralize may be due to weak compres- phase during the last deglaciation; in a­ccordance
sive stress or because the magma ascent is high due to with the hypothesis that melt production in Iceland
low viscosity magma [Németh and Smith, 2017]. is enhanced by isostatic readjustment through
creation of new conduits transporting magma to the
7.3. SUBGLACIAL ERUPTIONS surface. Ice confinement and changes to meltwater
drainage are among features that distinguish subgla-
The terms subglacial and intraglacial classify cial volcanism from submarine and subaqueous
volcanic eruptions by the landforms created by such eruptions [Gudmundsson, 2004; Smellie, 2000, 2006].
events. Specifically, subglacial events apply only to As noted above, an important feature of volcanism in
those processes occurring under ice cover without Iceland is the interaction of magma with ice and water
direct contact with the atmosphere. The term as it was mostly ice‐covered during the Pleistocene, and
­intraglacial is more general since it also applies to today 10% of the country is still covered, including
eruptions and volcanoes that have broken through some of the most active volcanoes. Ten of the pres-
the glacier [Jakobbsson and Gudmundsson, 2008]. The ently active central volcanoes have been shaped by
interaction of magma with the presence of ice and subglacial eruptions (i.e., Grímsvötn, Bárdarbunga,
­
water is an important part of the volcanic processes Mýrdalsjökull, Eyjafjallajökull, Hengill, Tindfjöll,
in Iceland that affects both explosivity and mineral Hofsjökull, Tungnafellsjökull, Askja, and Kverkfjöll).
texture. For example, the average eruption rates at Phreatomagmatic basaltic eruptions (i.e. Surtseyan‐
the end of the last glacial period in Iceland (~12 ka) style) are common in Iceland and include subglacial,
were 100 times higher than during either glaciated or subaerial, and submarine eruptions from monogenetic
unglaciated conditions [Maclennan et al., 2002]. This central vent and fissure eruptions, as well as eruptions
was also confirmed by cosmogenic 3He ­chronology from polygenetic central volcanoes [Thordarson and
OVERVIEW OF VOLCANICS IN ICELAND 55

Figure 7.5 Comparison between tindar (top) and tuya (bottom) subglacial structures. [Modified from
Jakobsson and Johnson [2012]; design credit Nathan Mennen.]

Larsen, 2007]. These eruptions differ from phreatic tuyas) or möberg ridges (also known as tindars)
eruptions, which are steam‐driven explosions that (Figure 7.5). To avoid confusion, Russell et al. [2014]
occur when water beneath the ground or on the surface defines a tuya as a positive‐relief volcano having a
is heated by contact with magma, lava, hot rocks, or morphology that results from ice confinement during
new volcanic deposition (e.g., tephra and pyroclastic‐ eruption and comprising a set of lithofacies that reflect
flow deposits). During both types of eruption, how- direct interaction between magma and ice/meltwater.
ever, the intense heat of the magma (upwards of 1170°C More specifically, Russell et al. [2014] separates these
for basalts) may cause water to boil and flash to steam, glaciovolcanic features into four morphological cate-
thereby generating an explosion of steam, water, ash, gories: flat‐topped, conical, linear, or complex. Using
blocks, and bombs. The products from phreatomag- this as a guide, tindars are generally found in semiregu-
matic eruptions are often discerned because they con- lar linear rows having an abrupt peak and a 2:1 ratio
tain juvenile clasts, have low vesicule content, and are (length vs. width) [Jakobsson and Gudmundsson, 2008].
considered to be better sorted and finer grained [Heiken Tuyas also have a broad lava cap that is missing from
and Wohletz, 1985]. tindars.
Lateral subglacial fissure eruptions lead to the
7.3.1. Subglacial Features formation of elongated möberg sheets consisting of
hyaloclastite lava, pillow lava (Photo 22), and/or
During the Pleistocene, the majority of Iceland was columnar basalts (Figure 7.6). These sheets can flow
covered by an ice sheet (Figure 2.2) resulting in subgla- considerable distances (up to 35 km) under the glacier
cial eruptions and initial explosive eruptive phases due with widths about 2 km and thickness of 150 m [Walker
to magma–water interactions as seen at Grímsvötn and Blake, 1966]. Hyaloclastite lava is hydrated tuff‐like
(64.4203, −17.2100) and Katla (63.6397, −19.1550). breccia rich in volcanic glass and has the appearance of
Subglacial eruptions from a central vent can lead to angular flat fragments sized between a few millimeters
the formation of tabletop mountains (also known as to a few centimeters. This fragmentation occurs by the
56 VOLCANICS

Photo 22 Pillow lava basalts are often covered by moss in southwestern Iceland. [Courtesy of Tamie
Jovanelly.]

force of the volcanic explosion or by the thermal shock different angles of cooling fronts. Thus, the division of
during rapid cooling. Subglacial‐eruption pillow basalts colonnade and entablature is the result of slow cooling
are formed if the ice to water pressure is adequately from the base upward and rapid cooling from the top
high. (This process is comparable to the dry formation downward (Figure 7.7).
of pahoehoe lava lobes known as inflation or endoge- Localized phreatic eruptions can cause the formation
nous growth). Under low‐pressure conditions, magma of rootless cones when lava (1100°C or greater) crosses
will immediately cool due to the rapid heat transfer over a wet surface, such as a swamp or a lake. Rootless
from the ice to the magma resulting in glassy mineral cones are often confused with crater rows as the tephra
textures, like obsidian [Gudmundsson, 2005]. A structure from the single‐event explosion builds up crater‐like to
that forms from the solidification of (mainly) basaltic give the appearance of a volcanic crater. Additionally,
magmas under the presence of water is columnar basalt. they are dissimilar in that rootless cones are not
When the lava cools, contraction occurs at centers that connected to a subsurface magmatic conduit system.
are equally spaced allowing a hexagonal fracture pattern
to develop [Spry, 1962] (Figure 7.6). The fracture 7.4. TEPHROCHRONOLOGY
pattern that forms at the cooling surface will tend to be
propagated down the lava as it cools, forming long, Tephrochronology is the technique of using tephra
geometric columns perpendicular to the surface flow of ash from individual volcanic eruption events to
water (on top of the lava) that can range in size within a ­determine a chronological order for the purposes of
deposit. A shift in colonnade‐type columns (e.g., those reconstructing geologic time, paleoenvironments, or
columns that are straight) to entablature (e.g., those col- archaeological events. The origins of tephrochronology
umns that are irregular) may be due to changes in water began in Iceland where it was first used in the 17th
flow direction or intensity, whereby flow changes create century to describe soil sections [Thórarinsson, 1981c;
OVERVIEW OF VOLCANICS IN ICELAND 57

Figure 7.6 Hexagonal columnar basalt. [Photo courtesy of Russell Maddrey; design credit Nathan Mennen.]

Figure 7.7 Entabular versus colonnade columnar basalt. [Modified from Spry [1962]; photo courtesy of
Russell Maddrey; design credit Nathan Mennen.]

Davies, 2015]. Centuries later, in the 1930s, an Icelandic [Sigurðsson, 1983]. Undoubtedly this tool, which
doctoral student, Sigurður Thórarinsson, began to started out as being a relative dating technique, has
make relationships between pollen and tephra dispersal developed into an important absolute technique where
patterns [Sigurðsson, 1983; Davies, 2015]. In his firm dates can be determined and tied to the geochem-
­dissertation, he coined the term “tephrochronology” ical signatures of specific volcanoes, thereby expanding
58 VOLCANICS

its use as a means to determine the distal extent of subject in the field of volcanology. Although some of
volcanic ash and volume. Isopach maps are common Iceland’s volcanoes are well known for repetitive cycles,
representations of tephra distribution that are based on they offer only suggestions into eruption frequency,
land surveying (for recent events) or investigation of and offer little usefulness in terms of volcanic hazard
soil profiles (historic events). The overall dispersal of mitigation or preparation strategies. For example,
tephra is influenced by environmental factors that Hekla’s 10‐year eruption pattern began in 1970, how-
include explosion type and duration, plume height, pre- ever, it has remained dormant since its last explosion
vailing wind direction, and strength at time of eruption in 1990. Other volcanoes, like Katla, have been
[Óladóttir et al., 2011]. extremely violent before historic time for which only
Although it is not an exact science, tephrochronology the tephra record can give some clues of dispersal pat-
can aid in determining the volume of ejected material. terns, volume, etc. Repose time can be linked to explo-
Volumes of tephra are commonly published as: bulk sivity whereby the longer the pause between eruptions
volume or compacted volume, freshly fallen or uncom- the larger and more intense will the next eruption likely
pacted volume, and dense rock equivalent (DRE). be. Some have suggested that monitoring the duration
Additionally, in areas where tephra has fallen offshore of precursory activity, like the infilling of a shallow
two volumes may be listed: that on land and that on magma reservoir for example, could help to predict
land and sea at sea. For events not recorded during his- volcanic eruptions [Passarelli and Brodsky, 2012].
toric times estimating the volumes can provide an Physical parameters, such as these, involve the run‐up
indicator of explosivity that can be correlated to the time of magma movement. Rubin [1995] suggests that
VEI (Table 7.1). Many tephrochronology scientists con- the silica content of the magma is ultimately the driver,
sider the Grímsvötn 2004 event to be one of the best however, as it influences the viscosity, and therefore
field recorded data sets for representing accuracy in vol- the run‐up time. Here, run‐up time is described as the
umetric definitions of tephra fallout [Larsen Larsen and time elapsed from the initiation of observed magmatic
Eiríksson, 2008b]. This success is largely because of the unrest to the onset of the magmatic eruption [Passarelli
combination of ground, aerial, and satellite observa- and Brodsky, 2012]. Although a positive correlation
tions, and the ability to sample relatively undisturbed between repose and run‐up time has been found for all
material on Vatnajökull ice cap using state‐of‐the art classes of magma composition volcanoes evaluated
equipment [Oddsson et al., 2007; Jude‐Eton et al., 2012]. (34 in total), it is agreed that precursory activity is both
Larsen and Eiríksson [2008a], however, note that only a hard to establish and is largely influenced by the tec-
small portion of the historic eruptions occurring under tonic setting. The question posed by Thelen et al.
Vatnajökull produced enough tephra to be dispersed [2010] still remains, what earthquakes or parameters
beyond the ~8000 km2 ice cap. are important to interpret an impending eruption? For
In Iceland, tephrochronology included in soils can now, there is some agreement that repose time and
date back to 9000 ka. Between 600 and 700 tephra seismic energy released are essential into forecasting
layers are known in soil, sediment, and ice [Larsen and eruptions [Thelen et al., 2010].
Eiríksson, 2008a] with more than 500 basaltic in com-
position, and approximately 100 that are primary sili- 7.6. VOLCANIC EXPLOSIVITY INDEX
ceous in composition [Óladóttir et al., 2008, 2011].
One of Iceland’s most active volcanoes, Hekla, pro- Newhall and Self [1982] proposed the VEI as a means
duced three unique silica tephra deposits (referred to to categorize recent and paleo events for the purpose of
as H3‐1050 BC, H4‐2250 BC, H5‐5100 BC) that are making comparisons based on the explosive character
often used as marker beds for tephrochronology of the of an eruption, utilizing a logarithmic scale of 0–8
island. These tephra layers are found in soil profiles on (lowest to highest) to depict magnitude. The criteria
90% of the island making widespread correlation pos- included in the VEI are listed in order of decreasing
sible. Additionally, because of their distinct silica‐rich, reliability: volume of ejecta, column height, qualitative
white to yellow color, and laminar bedding, they description of event (e.g., gentle, explosive, colossal),
become easily identifiable. classification (e.g., Hawaiian type, Vulcanian type),
duration (hours of continual blast), tropospheric injec-
7.5. VOLCANIC REPOSE tion, and stratospheric injection. For example, a value
of 0 can represent a nonexplosive event that produces
Understanding the repose of all types of volcanoes <10,000 m3 of tephra and a value of 8 can represent a
has been a highly important and much investigated cataclysmic event that produces 1.0 × 1012 m3 tephra
OVERVIEW OF VOLCANICS IN ICELAND 59

(Table 7.1). Therefore, the VEI for either modern or Newhall and Self [1982] estimated VEIs for some
paleo events can be approximated by eruption magni- 8000 eruptions, including those occurring on Iceland,
tude (volume) or intensity (mass eruption rate). using data published in Simkin et al. [1981]. Also noted
Newhall and Self [1982] acknowledged that there are was a requirement that whenever an explosion was
limitations to the VEI. In particular the scale does reported without a description, a correction could be
not include atmospheric or geographic (i.e. latitude applied by adding a VEI value that accounts for paleo
proximity) data, nor does it attempt to classify injection events as a proxy for missing data. In addition a
type. Subsequently it as been acknowledged that non- correction could be applied if the historic description
explosive eruptions producing large pyroclastic flows lent itself to an event likely being larger or smaller than
may receive a higher VEI than necessary and that using reported (e.g., reporting VEI 3 instead of 2). Newhall
DRE to report bulk volume may be difficult for com- and Self [1982] compiled a list of the largest explosive
parative purposes [Newhall et al., 2018]. Additionally, eruptions in the world since 1500 CE with VEI ≥ 4;
Houghton et al. [2013] observed challenges in applying Icelandic volcanoes represented 1% of those eruptions.
the VEI to phreatic events that sometimes only produce Today, volcanologists extensively use the VEI for cre-
small eruptive volumes of tephra that are locally dis- ating hazard maps and to determine tephra dispersal
persed [Houghton et al., 2013]. forecasting [Houghton et al., 2013].
8
Volcanics of the Reykjanes Peninsula and Southwestern Region

Southwestern Iceland and the submarine northern and is considered the largest known monogenetic
RR comprise a complicated rift system composed of v­olcano in Iceland [Jakobbson and Johnson, 2012].
overlapping spreading centers that include: the RR, Of the 258 intraglacial volcanoes identified in the
the Reykjanes Peninsula (comprising the RVB), and WVZ, 101 are consistent with tuya formation in that
the WVZ (Figures 1.1, 2.1, and 3.1). The 170-km‐long their lavas were deposited above water level. Jakobbson
WVZ (Figure 2.1) was the main focal point of crustal and Johnson [2012] suspected that this might indicate
spreading in south Iceland at 6–7 Ma, but volcanic deposition in lakes within the ice sheet. This is consis-
activity reduced during Pleistocene and postglacial tent with the proposal of Eason et al. [2015], who agree
times, especially compared with the RVB. The main that these features formed at the onset of deglaciation
decline in volcanic intensity of the WVZ in Pleistocene when ice sheets melted rapidly, providing the opportunity
times resulted from both the development of the SISZ for uplift of the underlying mantle, which in turn influ-
(Figure 2.1) and the reduction in tectonic activity with enced the melting rates in the upper mantle. Although
decreased rates of spreading. The last volcanic eruption abundant, it is likely that the tindars and tuyas in the
in the WVZ occurred 1000–1300 CE, which correlates WVZ correspond to low‐volume Holocene eruptions
to activity on the Snæfellsnes Peninsula. Conversely, from volcanic fissures and volcanic lava shields; the
postglacial volcanic activity has been dynamic on the average WVZ volume is 0.11 km3 compared with
Reykjanes Peninsula. Today, however, in the WVZ 1.11 km3 lava shields on the Reykjanes Peninsula
there is still an abundance of evidence for crustal [Andrew and Gudmundsson, 2007]. It is therefore sug-
spreading (faulting) and hydrothermal activity, which gested that the morphology and structure of the intra-
includes geysers, mudpots, and steam vents. glacial volcanoes in this region were largely dependent
on tectonic control (i.e., fissure length and width) and
8.1. VOLCANISM OF THE WESTERN availability of magma at the time of eruption. In the
VOLCANIC ZONE WVZ, a relatively short eruption fissure produced
small tindar when little magma was available, and a
A relatively large part of the WVZ is covered by the small tuya or eventually a larger tuya when more
ice cap, Langjökull. For this reason, the landscape of the magma was erupted and the eruption lasted longer; a
WVZ is characterized by Late Pleistocene (0.01–0.78 Ma) long eruption fissure produced a tindar [Andrew and
intraglacial volcanism that has been influenced by Gudmundsson, 2007].
the presence of an ice sheet 500–2000 m thick in the Evidence of explosive tuya features, like the
interior. The largest of these intraglacial volcanic Kerlingarfjöll central volcano (64.6275, −19.2569), are
features being the tuya called Eiríksjökull (64.7697, also found in the WVZ [Stevenson et al., 2010]. This
−20.4020), which has an estimated volume up to 48 km3 Quaternary rhyolite outcrop rises 1488 m above a

Iceland: Tectonics,Volcanics, and Glacial Features, Geophysical Monograph 247, First Edition. Tamie J. Jovanelly.
© 2020 American Geophysical Union. Published 2020 by John Wiley & Sons, Inc.

61
62 VOLCANICS

g­ lacial outwash plain and is the second largest deposit and small shield volcanoes like the Reykjanes Volcano
(200 km2) in Iceland resulting from at least 21 different (63.8500, −22.5660) with a summit at 140 m. The
eruptions over the past 275 kyr (from approximately Reykjanes volcanic system is the westernmost of a
68 to 345 ka) [Flude et al., 2010]. Holocene basalt lava four closely spaced fissure system that extends diago-
flows cap the tuya formations in the northeast of the nally across the Reykjanes Peninsula. Holocene (e.g.,
complex suggesting that Kerlingarfjoll was ice covered Hellnahraun, Kapelluhraun) and modern eruptions
for most of its volcanic history [Gronvold, 1972]. This (e.g., Reykjanes Fires) that have occurred at several
central volcano has been of particular interest in the locations on the NE–SW trending fissure system
discussion and understanding of continental crust cover most of the volcanic system, but rocks span
formation, wherein studying silicic material in oceanic the age of 3.2 Ma and become progressively younger
crust provides insights into relevant magmatic to the south. Three shield volcanoes (Hrútagjá and
processes [Flude et al., 2010]. Moreover, the use of reli- Þráinskjöldur, 63.9568, −21.9842; Sandfellshæð
able absolute dating techniques like K–Ar or Ar/Ar 63.8710, −22.5905) were formed during the Pleistocene–
methods on these deposits has contributed to providing Holocene transition that are important to the overall
a chronological framework for explosive volcanism in shaping of the peninsula some 9000–14,000 ka. With
Iceland; for example, in determining that repose cumulative volumes of lava nearing 15 km3, these
periods for rhyolite volcanism are on the order of tens flows are the major factor responsible for raising the
of thousands of years, thus making it possible for peninsula above sea level. More specifically, pahoe-
Kerlingarfjöll to erupt again. Additionally, the detailed hoe lavas from the half shield volcanoes, Hrútagjá and
chronology of the ryholitic deposits has revealed that Þráinskjöldur, spread radially northwards to form 20 km
the explosive activity spans the last four glacial periods. of coastline between Straumsvík and Vogararstapi.
The ages are not precise enough to assign an eruption Iceland’s largest tumuli result from Hrútagjá lava
to a particular glacial or interglacial stage, but some flows. The main crater is a shallow depression sitting
eruptions may have taken place in Northern on top of a lava rise. In the culmination phase of the
Hemisphere interglacial periods [Flude et al., 2010]. eruption (6000 CE), a solid crust of lava around the
The volcanic features of the WVZ landscape makes crater was inflated by magma, resulting in extensive
for ideal comparisons between three main eruptive rise. Later, when the magma escaped through chan-
environments: effusive deep‐water eruption producing nels, a subsided block of lava was left behind, marked
pillow lavas; explosive shallow‐water eruption producing by deep fissures on the margins [Sigurgeirsson, 2010].
hyaloclastites; and an effusive subaerial capping lava In contrast, Þráinskjöldur is heavily dissected by
eruption [Eason et al., 2015]. Often these stages coin- fissures and faults from rifting in the Reykjanes
cide with, or are followed by, dike and irregular minor volcanic system. Sandfellshæð provides a textbook
intrusions. The Middle to Late Pleistocene intraglacial example of a shield volcano with its gently sloping
volcanic rocks of the WVZ belong to the palagonitic sides (<3°) that rise, to 90 m above sea level and its
Möberg Formation, which comprises all strata formed lava covers 120 km2. The summit crater is well
during the Brunhes geomagnetic epoch to the end of defined: 450 m wide and 20 m deep. Forming during
the Pleistocene [Jakobsson and Gudmundsson, 2008]. glacial time (14,000 ka), Sandfellshæð is older than
The resulting rock types found in the WVZ are often the other two volcanoes previously mentioned and
picrites, olivine tholeiites, and tholeiite basalts, with formed when sea level was about 30 m lower than
only a small portion (~0.5 vol%) of the magma emerg- present.
ing as intermediate or silica rich [Jakobsson and The Younger Stampar cone row (63.9155, −22.3650)
Johnson, 2012]. that was formed during the Reykjanes Fires (1210–
1240 CE) is the youngest volcanic formation on the
8.2. VOLCANISM OF THE REYKJANES Reykjanes Peninsula. This 4.5-km‐long formation
PENINSULA covers the northern tip of the peninsula and can be
recognized from a top of the Mid Ocean Ridge.
In the southwest of the Reykjanes Peninsula, there Additionally, toward the interior of the peninsula
are six major volcanic regions that align with the modern‐ the Arnasetur (63.9056, −22.4106) monogenetic
day NE–SW active rift zone: Reykjanes, Eldvörp– volcanic fissure with scoria and spatter craters
Svartsengi, Fagradalsfjall, Krýsuvík, Brennisteinsfjöll, formed in an eruption that was also part of the epi-
and Hengill–Langjökull (Figure 3.1). These regions sode. The resulting fissure is 2 km in length and the
consist of lava fields, postglacial basaltic crater rows, lava formed covers 20 km2.
Volcanics of the Reykjanes Peninsula and Southwestern Region  63

8.3. ESJA MOUNTAIN AND EXTINCT spreading of the WVZ. Due to erosion, the gabbro
VOLCANIC SYSTEMS (KOLLAFJÖRÐUR intrusions that once were the root of the central
AND STARDALUR) ­volcanoes can be seen at the base of Mount Esja.
Fissure ­
eruptions fed by dikes followed earlier
Esja is a mountain complex of three central volcanic activity [Forslund and Gudmundsson, 1991]
­ olcanoes that dramatically dominates the northern
v and can be seen along the northern slopes of
skyline of Reykjavík, and is a popular hiking spot Eyararfjall (64.3191, −21.7261) and Lokufjall
as it is only 10 km from the city center. The summit (64.2818, −21.8257).
is called Hábunga and sits 914 m above the low‐
lying area. Esja was created in the late Pliocene, 8.4. RAUÐHÓLAR ROOTLESS
when _during warm periods lava flowed and in the CONE COMPLEX
cold periods ridges of tuff were built up under
the glacier. The oldest part of the mountain range The term Rauðhólar translates to “red hills” in
(dating to 3.2 Ma) can be explored at the base of the Icelandic and has been coined as a geologic term to
western slope and the younger rhyolitic cap deposits describe rootless cone features. Although there are
dating to 1.8 Ma can be seen on the eastern side, other Rauðhólar chain complexes in Iceland (namely,
called Móskarðshnúka (64.2431, −21.5278). The in Vesturdalur near Mývtán), some of the easiest to
1650 m total thickness of the Esja succession was visit are near Reykjavík in a town that bears the
formed by the extinct Stardalur and Kollafjörður name of the feature (64.0986, −21.7797). This local-
volcanic ­ systems that originated with the initial ized phreatic eruption spanning 1.2 km2 in the RVB

Photo 23 Mount Hengill, southwest Iceland. [Courtesy of Tamie Jovanelly.]


64 VOLCANICS

created a cluster of phreatic pseuodcraters that 8.5. HENGILL AND ÞINGVELLIR


formed when the Leitahraun lava flowed over a wet-
land 5200 ka. The Leitahraun lava was emitted from The Hengill–Langjökull fissure swarm is located in
a shield volcano located on the eastern side of the the southern part of the WVZ. It extends at least 30 km
Bláfjöll (63.9842, −21.6583) mountain complex. The north of the Mount Hengill central volcano (64.0856,
main pathway of the lava was toward the south to −21.3133; Photo 23) and covers an area of 100 km2.
the coast near Þorlákshöfn, however, the lava also This active volcano (805 m in height) last experienced
flowed towards the west, crossing wetlands at an eruption around 1000 CE, which is said to have
Rauðhólar into Lake Elliðavatn, then along the occurred during a meeting of the Icelandic parliament
channel of River Elliðaár and into the sea at at Þingvellir. Mount Hengill has been volcanically
Elliðavogur in Reykjavík. The Rauðhólar pseu- active since 0.8 Ma and has produced four postglacial
docraters in Elliðaárhraun are part of Reykjavík’s (< 10,000 ka) lavas [Sæmundsson, 1967]. The volcano is
nature reserve of Heiðmórk that was created in 1950 mostly composed of hyaloclastites, but contains pillow
to preserve these unique features. Originally, there lavas, pillow breccias, as well as andesite, dacite, and
were over 80 craters, but the gravel from them was rhyolite. The extensive Nesjavellir and Hellisheiði high‐
used for road and airport construction during World temperature geothermal area encompasses Mount
War II. Now, as a result, some interior structures of Hengill, which is the major producer of geothermal
the pseudocraters can be examined closely. Other energy for Reykjavík (Box 3.1). The area experiences
rootless cones can be observed west of Kirkjubæjar­ continuous background small‐magnitude earthquake
klaustur in Álftaver (63.5261, −18.3603) and at activity that correlates spatially with surface heat loss.
Landbrotshólar near Katla, where the largest More specifically, between 1994 and 2007 about 100,000
(50 km2) complex in Iceland can be found. The age earthquakes were recorded, reaching intensity up to
and origin of the lava at Landbrotshólar has been M = 5.0 [Jakobsdottir, 2008]. Crustal doming with a
debated, but its origin is believed to be the volcanic maximum uplift rate of 2 cm year−1 coincided with the
eruption of Eldgja in 934 CE. seismic activity, until collapse in 2000.
9
Volcanics of the South and Southeastern Regions

The EVZ and NVZ neovolcanic zones connect fallout deposits: a 1947 tephra has been identified in
beneath Vatnajökull. The EVZ is thought to have western Russia and an 1845 tephra has been identified
begun its formation ~2–3 Ma within pre‐existing crust in the northern islands of the British Isles [Dugmore
[Foulger and Natland, 2003]. The EVZ connects to the et al., 1995; Wastegard et al., 2000].
WVZ with two strike‐slip faults, the MIB and the SISZ The origin of these silica magmas within the EVZ
(Figure 2.1). Due to the proximity of the hot spot continues to be highly debated. The two main view-
under Vatnajökull, and the faster rate of spreading points regarding silica magma production include (a)
along the Eurasian plate margin, the EVZ represents fractional crystallization of primitive basalts and (b)
the most active volcanism on the island during historic partial melting of hydrothermally altered basaltic
times (i.e., post‐874 CE to present; Figure 9.1). The crust [Schattel et al., 2014]; however, Scerrisdottir
EVZ neovolcanic zone is responsible for about 77% of [2007] proposed a combination of these circumstances.
historic eruptions and ~60% of the erupted magma Schattel et al. [2014] compared rift and off‐rift vol-
volume. According to Thordarson and Larsen [2007] canoes in the NVZ and EVZ (i.e., Askja vs. Hekla) in
the Grímsvötn volcanic system alone accounts for order to investigate the physicochemical conditions of
about 38% of the verified historic eruptions, while the silica magma origin with respect to possible crustal
three other systems collectively represent ~39% of the assimilation. Although the controversy continues,
remaining events (Bárdarbunga, ~14%; Hekla, ~13%; Schattel et al. [2014] concluded that magmas behave
Katla, ~12%). The other three systems on the EVZ differently at rift and off‐rift localities. For example,
that have also been active in historic time are the authors suggest that the rhyolitic magmas at Hekla
Vestmannæyjar, Eyjafjöll, and Torfajökull, which originated deeper in the crust (≥4 km), under lower
account for 3% of the events [Larsen et al., 1998; temperatures, but with higher water content compared
Larsen, 2002]. Additionally, central volcanoes in the with those at Askja.
EVZ mentioned above, as well as Eyjafjallajökull, have The type of activity represented within the EVZ varies
over the past 11 centuries produced a minimum of 24 depending on the proximity (north or south) to the
silica‐rich tephra eruptions [Larsen et al., 1999]. junction of the SISZ, which has resulted in uneven dis-
Isopach maps based on the identification of tephra tribution of central volcanoes (Figures 2.1 and 7.2);
layers by Larsen et al. [1999], with contributions from notable clusters occur near the southern tip of the
Thórarinsson [1954, 1967], have determined variable EVZ and in eastern central Iceland, which is for
directions, distal extent, and volume of tephra deposi- the most part covered by the western part of the
tion for these explosive events (Figure 9.2). The esti- Vatnajökull ice cap (Figure 9.1). The extensive ice
mated volumes of the 24 eruptions vary widely from cover and limited exposure of rocks has resulted in
10 to <0.01 km3 [Larsen et al., 1999] (Figure 9.3, much less detailed knowledge of the geology of this
Table 9.1). Two Hekla events produced the furthest area compared to other parts of the EVZ [Gudmundsson

Iceland: Tectonics,Volcanics, and Glacial Features, Geophysical Monograph 247, First Edition. Tamie J. Jovanelly.
© 2020 American Geophysical Union. Published 2020 by John Wiley & Sons, Inc.

65
66 VOLCANICS

Figure 9.1 Active fissure, volcanic systems, and central volcanoes found in southern and central Iceland.
[Modified from Gudmundsson and Högnaddóttír ̄ [2007]; design credit Nathan Mennen.]

and Högnadóttir, 2007]. Bedrock topography of the of a Bouguer anomaly map that revealed anomalies
western part of Vatnajökull was mapped in consider- associated with central volcanoes and fissure swarms
able detail in the 1980s and 1990s using radio echo in the area [Gudmundsson and Högnadóttir, 2007].
soundings [Thórarinsson, 1974; Thórarinsson and A change of eruptive styles and compositional prod-
Sæmundsson, 1979; Bjórnsson and Einarsson, 1990; ucts along the EVZ is also evident [Jakobbson, 1972a;
Larsen et al., 1998]. This was followed by production Meyer et al., 1985]. North of the SISZ it is common to
Volcanics of the South and Southeastern Regions  67

Figure 9.2 Geology of the Hekla region showing distribution of historical lava flows. [Modified from
Höskuldsson et al. [2007]; design credit Nathan Mennen.]

see rifting structures, such as fissure swarms (e.g., Laki, Olaðottír et al. [2011] noted that in prehistoric time
Eldgjá) and hyaloclastite ridges, that represent past the EVZ experienced a lag of 1–3 kyr between peak
glaciation and isostatic rebound situations. To the activity for volcanoes directly above the mantle plume
south of the SISZ large central ­ volcanoes (e.g., compared with volcanoes located in the nonrifting
Grímsvötn) that illustrate high‐frequency but low‐ part of the EVZ or closer to the island boundary. From
volume eruptions dominate the landscape. Iron‐rich this time–space relationship the authors drew the
basalts characterize the central EVZ (e.g., Veidivötn, conclusion that a significant increase in future volca-
Grímsvötn), whereas furthest south, at the rift tip in nism can be expected there, following increased levels
Vestmannæyjar, alkali basalts dominate. In the interior, of volcanism above the plume.
FeTi tholeiites can be found at Hekla (Figure 7.1). Volcanic activity in the EVZ has influenced airline
Each volcanic center has distinctive geochemical char- travel multiple times in the past decade (e.g.,
acteristics that require differences in source compo- Eyjafjallajökull, 2010; Grímsvötn, 2004; Hekla,
nents and their proportions at distances of 10–20 km. 2000). The Eyjafjallajökull eruption in 2010 resulted
The large shield volcanoes that populate the WVZ and in the cancellation of 100,000 flights and affected the
NVZ are absent in the EVZ [Sinton et al., 2005]. travel of about 10 million passengers costing some
68 VOLCANICS

9.1. GRÍMSVÖTN AND BÁRÐARBUNGA

The subglacial Vatnajökull volcanoes are amongst


the most active in the country in historic time
[Thordarson and Larsen, 2007]. Together with the inter-
play of associated fissure swarms, the central vol-
canoes of Grímsvötn, Bárðarbunga, and Kverkfjökull
produce explosive basaltic eruptive products because
of magma interacting with glacial meltwater. Grímsvötn
and Bárðarbunga are closer to the hot‐spot plume
compared with Kverkfjökull (Figure 9.1).
The deposits of the Vatnajökull volcanoes are
challenging to investigate due to presence of glacial
ice and erosional, distribution, and preservation
processes. The eruptive histories of three partly sub-
glacial volcanic systems (Grímsvötn, Bárðarbunga,
and Kverkfjökull) have been analyzed using tephra,
soil profiles, and written records [Thórarinsson, 1950;
Larsen et al., 1998; Thordarson and Larsen, 2007;
Larsen and Eiríksson, 2008a]. However, some of the
Figure 9.3 Directions of ash dispersal from volcanic erup- written accounts have been found to be incorrect
tions at Hekla. Numbers correlate to events listed in
regarding the activity of Kverkfjökull and only a
Table 9.1. [Design credit Nathan Mennen.]
fraction of the historic eruptions at volcanoes
beneath Vatnajökull have produced tephra that was
deposited and preserved in soils outside the ice cap
and in the barren highland areas [Larsen and
$50 billion in lost Gross Domestic Product [Oxford Eiríksson, 2008]. Even the recent small eruptions of
Economics, 2010] (Figure 9.4). Casadevall [1994] out- Grímsvötn in 1983 and 2004, for example, had
lined the impacts of volcanic eruptions on aircraft limited tephra dispersal and deposited none outside
operations: decreased visibility, exterior abrasion the 8000 km2 ice‐cap.
to the airplane, temporary engine failure, and A key question addressed by Olaðóttir et al. [2011]
permanent engine damage. Collection of baseline is whether these volcanoes have always been as active
data from historic volcanic eruptions in the EVZ as observed during the last millennium or has their
therefore is important for predicting potential haz- activity exhibited periodic or other systematic long‐
ards to transtlantic flight paths and European air term changes? Through this study, the authors deter-
travel. More specifically, determining seasonal wind mined the estimated eruption frequency of each
patterns, average eruption column height, ash par- volcanic system: Grímsvötn ~7 eruptions in 100 years;
ticle size distribution, and ash volume are crucial Bárðarbunga ~5 eruptions in 100 years; Kverkfjöll
variables to be included in computer models. Using 0–3 eruptions in 100 years, but showing periodic
a minimum plume height of 12 km and other data activity with repose times of >1000 years [Olaðóttir
from past EVZ eruptions, Leadbetter and Hort [2011] et al., 2011]. A “Mid‐Holocene low” was also discov-
examined the probability and timing of volcanic ash ered as all three volcanic systems experienced lulls in
reaching European airspace. Results from the com- activity from 5 to 2 ka. This change in activity is
puter model determined that all countries in Europe attributable to decreases in magma generation from
will receive some volume of volcanic ash within 24 h the mantle plume rather than to ice‐load/glacial‐thick-
following an eruption; countries closer to Iceland, ness relationships.
like the United Kingdom and those in northwest Grímsvötn (64.417, −17.333) has been one of the
Europe have a higher risk of ash concentrations most active volcanoes in Iceland since 1200 CE,
exceeding a 5% threshold. The countries having the including six eruptions between 1934 and 2011. The
highest risk of deposition within a 4‐day period volcano covers an area of ~62 km3. The bedrock topog-
include Scotland and those in Scandinavia [Leadbetter raphy of Grímsvötn central volcano has been mapped
and Hort, 2011]. by radio echo sounding [Björnsson and Einarsson,
Volcanics of the South and Southeastern Regions  69

Table 9.1 Tephrochronology of Hekla with associated volume, repose period, and eruption date. The numbers listed
correspond to those on Figure 9.3
Dormant Tephra volume
Number Date CE years (km3) VEI Notes
1 1104 >300 ? ? A highly destructive eruption that ejected its tephra over
the entire northern half of Iceland
2 1158 53 >0.2 4 Formed the Efrahvolshraun lava flow
3 1206 47 >0.15 4
4 1222 15 ? ?
5 1300 77 >0.5 4 Second largest tephra erution; formed the Selsundshraun
lava flow
6 1341 40 ? ? Flourine from the ejected tephra poisoned livestock
7 1389 47 >0.2 4 Formed the Nordurhraun lava flow
8 1510 120 >0.75 4
9 1597 86 ? ?
10 1636 39 ? ?
11 1693 56 ? ? Produced upwards of 60,000 m3 s−1 of tephra which was
eventually carried as far as Norway
12 1766 72 1.3 5 Hekla’s largest volcanic eruption recorded; 3–5 cm of
tephra deposited; flooding caused from melted ice and
snow at Hekla’s peaks
13 1845 77 0.63 4 This 7 month long eruption caused widespread death
among local livestock and wildlife
14 1947 102 0.8 4
15 1970 16 0.2 4
16 1980 10 0.12 4 This eruption spread along Hekla’s 6.9 km fissure,
depositing upwards of 20 cm of tephra
17 1991 9.5 0.15 4
18 2000 9 0.11 4 Hekla’s most recent volcanic eruption produced an ash
plume 10 km high, which traveled over 300 km from
the eruption site
Note. Modified from Larsen et al. [1999], Oladóttir et al. [2011], and Thórarinsson [1970]; design credit Nathan Mennen.
VEI, volcanic explosivity index.

1990] revealing a composite caldera that Gudmundsson the overlying ice, and meltwater accumulates in a sub-
and Milsom [1997] divided into three regions. A glacial lake within the caldera until the surrounding ice
combined magnetic and seismic reflection survey of is breached. When this happens, water escapes to cause
the main caldera indicated that the caldera floor is a jökulhlaup in the River Skeiðará, after having trav-
made of volcanic clastic sediments, lava flows and sills eled ~50 km beneath the Skeiðarárjökull outlet glacier.
[Gudmundsson, 1992]. The Grímsvötn caldera lake Jökulhlaups occur there every 1–10 years and last from
(10–12 km wide, 200–300 m deep) is largely confined days to weeks, each time releasing 0.4–4 km3 of water
to the main caldera and is overlain by a floating [Björnsson, 2003].
ice‐shelf 240–260 m thick. Direct surface expressions A devastating jökulhlaup occurred in November
of hydrothermal activity are visible at its highest peak, 1996 following an eruption of Grímsvötn, however, no
Grímsfjall (1722 m), on the southern caldera rim, in lives were lost due to detailed monitoring (e.g., hydro-
the form of fumaroles, steam outlets, and ice caves. logical, seismic, geodetic) carried out by the IMO. The
Most of the phreatomagmatic eruptions persist for flooding event itself did not occur until 13 days after a
days to weeks. Volcanic eruptions at Grímvötn fissure eruption occurred between the Bárðarbunga
often coincide with jökulhlaups when the water level volcano and Lake Grímsvötn, but indicators from cal-
in the caldera lake rises to a critical height of 1425– dera subsidence and subsurface monitoring of water
1450 m a.s.l. Geothermal activity continuously melts flow allowed for proper evacuation and the closing of
70 VOLCANICS

Figure 9.4 The aviation zones shut down (red) or restricted (orange) by the 2010 eruption of Eyjafjallajökull.
[Deltafalcon/Wikimedia Commons; Public Domain.]

the Ring Road (at Hringvegur) prior to the arrival of All historic eruptions at Grímsvötn took place
the flood wave. The velocity of the glacial River from the ice‐covered central volcano, with exception
Skeiðará peaked at 55,000 m3 s−1, or more than five of the 1783–1784 Laki fissure eruption [Thórarinsson,
times that of a “normal” megaflood. Ultimately, a sec- 1974; Gudmundsson and Björnsson, 1991]. The most
tion of road across the Skeiðará sandur was washed recent eruption occurred on 21 May 2011 when a
away by the intensity of the flood event (Photo 24). 20–25 km ash cloud was accompanied by multiple
Smaller jökulhlaups have occurred since the 1996 event earthquakes [Petersen et al., 2012]. The ash cloud
(e.g., 2004), when appropriate warnings have also been from this eruption was classified as 10 times larger
issued by the IMO. To provide a historic perspective, than the 2004 eruption and produced an order of
from 1600 to 1934 about one jökulhlaup occurred magnitude larger volume of magma (0.2–0.3 km3
every decade with a discharge of about 5–7 km3 of DRE) [Sigmarsson et al., 2013; Jude‐Eton et al., 2012].
water; post‐1938 two or three jökulhlaups would occur Although flight travel was interrupted in northern
every decade with correspondingly smaller discharges Europe, it was much less widespread than the 2010
between 0.5 and 3.5 km3 [Björnsson, 1992; Björnsson disruption after the Eyjafjallajökull eruption.
and Gudmundsson, 1993]. Another notable event occurred on 28 December
Volcanics of the South and Southeastern Regions  71

Photo 24 The remnants of a bridge wiped out along the Ring Road (Route 1) by a jökulhlaup flooding event
in 1996 linked to the eruption of Grímsvötn. This is known as the Skeiđará Bridge Monument. [Courtesy of
Russell Maddrey.]

1998 when a week‐long eruption took place but did 9.2. KATLA AND EYJAFJALLAJÖKULL
not trigger a glacial outburst.
Bárðarbunga is the second most active volcano in Katla (63.633, −19.083), which translates to “kettle”
historic time (with Hekla a close third) and second is a gentle sloping (<4°) polygenetic bimodal volcano
tallest peak in Iceland (2009 m). Bárðarbunga is part that sits within the rift of the EVZ (Photo 25). The
of a volcanic system that is approximately 200 km long 30 km diameter central volcano is within a 110 km
and 25 km wide. The caldera is about 80 km2, up to long volcanic system that narrows to <1 km in the
10 km wide, and 700 m deep. The crater is hidden by northeast; the Eyjafjallajökull volcano is also within
850 m of glacial ice. this system. The summit is 1480 m a.s.l., where the cal-
Although the eruption frequency is less than that dera is covered by a 590 km2 glacier called Mýrdalsjökull
of Grímsvötn, activity on the two systems often that has an approximate thickness of 230 m [Oladöttir
behaved concurrently [Larsen et al., 1998; Sigmarsson et al., 2008]. Radio echo measurements completed by
et al., 2000]. Also similar to Grímsvötn, the Björnsson et al. [2000] determined that the caldera was
Bárðarbunga central volcano has been capped with 650–700 m deep and seismic profiling completed by
ice historically, but approximately 70% of the overall Gudmundsson et al. [1994] revealed a shallow (~3 km)
system is ice free [Sæmundsson, 1978]. Despite this, magma chamber below the surface. Subsequently, the
there have only been 3 of 23 confirmed eruptions radio echo measurements also revealed a complex
along the ice‐free fissure swarm [Larsen, 1984; drainage system that influences the direction of glacial
Thordarson and Larsen, 2007]. meltwater discharge during jokhulhaups at Katla.
72 VOLCANICS

Photo 25 The gently sloping Katla volcano in the Eastern Volcanic Zone. [Courtesy of Emily Larrimore.]

Thordarson and Larsen [2007] consider Katla to be the in particular with a high frequency of events occurring
most productive volcanic system in Iceland in terms of at 3500–5500 BCE and 8000–9500 BCE [Oladöttir
magma volume production, considering that the 934– et al., 2005]. Larsen et al. [2001], however, estimate that
940 CE Eldgjá Fissure Fires (63.633, −19.083) event the amount of material produced by an individual
resulted in a 700 km2 lava field. event is typically <1 km3. In the vicinity around
The explosive eruption history of Katla dates back Mýrdalsjökull, the Katla tephra layers have specific
to the Pleistocene (~10,000 BCE), since when a characteristics that help discern them from those
minimum of 300 subglacial events have likely occurred deposited by Hekla. Specifically, most of the tephra
from the central volcano [Larsen, 2000], averaging six layers show distinctive bedding planes due to intermittent
eruptions per century and an average repose time of deposition and shifting wind strength and direction
25 years [Óladóttir et al, 2005] (Figure 9.5, Table 9.2). during the eruption [Larsen, 2010]. Additionally,
Of these 300 eruptions, 208 tephra layers are found in Larsen [2010] notes two distinct differentiation patterns
the stratigraphy of the eastern flank with 190 classified within some deposits. The K1357 deposit, for example,
as prehistoric and 18 as historic [Oladöttir et al., 2005, contains fragmented fine‐grained tephra that coarsens
2008]. The vast majority of the tephra layers have min- upward, which may indicate the influence of meltwater
eralogy that indicates basaltic origins, with two periods during the initial eruption phase. The reverse sequence
Volcanics of the South and Southeastern Regions  73

is found at K1755 suggesting a lower water‐to‐magma


ratio. Additionally, there have been at least 12 small
to moderate volume (<0.01–0.27 km3, calculated as
uncompacted tephra) silica eruptions (trachydacitic
to dacitic) occurring between 5500 and 1700 BCE
[Larsen et al., 2001; Óladóttir et al., 2008]. In teph-
rochronology these deposits are often referred to as
SILK, with a silica content between 63 and 67%
[Larsen et al., 2001]. SILK deposits, as described by
Larsen [2010], are yellow to green in appearance and
sometimes ­contain needle‐like glass shards. As expected,
comparisons of the distal fallout mapping of SILK
versus basalt tephra units indicate less dispersion
related to their brief Plinian eruptions (Figure 9.5).
Oladöttir et al. [2008] suggest that changes in the min-
eral composition of the tephra over time are the result
of cyclic behavior of Katla’s subsurface magma con-
duit system. Notably, the changes indicate that when
the “plumbing system” evolves from a simple chamber
to one containing complex sills and dikes, the eruption
frequency will increase. Consequently, Katla is cur-
rently in a period of low eruption frequency and
Figure 9.5 Directions of ash dispersal from volcanic
simple magma development.
eruptions at Katla. Numbers correlate to events listed in
Table 9.2. The initial eruption, ~10,000 BCE, is linked to a
Plinian eruption that dispersed 10 km3 of rhyolitic
tephra (with lesser amounts of basaltic‐icelandite and
icelandite) and airborne ash some 1300 km away from
the source [Thordarson and Höskuldsson, 2008]. This
event is tied to a pyroclastic flow deposit referred to as
Table 9.2 Tephrochronology of Katla with associated the Sólheimar ignimbrite [Tomlinson et al., 2012]. The
volume, date, dormancy period, and volcanic explosivity Sólheimar ignimbrite has special significance for
index (VEI) value. The numbers listed correspond to those ­several reasons. First, this unit represents one of the
on Figure 9.5 largest eruptions from Katla producing >6 km3 of
pyroclastic material as measured from proximal
Dormant Tephra
Number Date CE years volume (km3) VEI
deposits [Lacasse et al., 2007], and likely represents the
source magma chamber. Next, the ignimbrite is often
1 920 >300 0.27 4 considered the source of the bimodal Vedde Ash
2 1262 342 0.48 4
deposit [Lacasse et al., 1995] that dates to the Younger
3 1357 95 0.20 4
4 1416 59 ? ?
Dryas Stadial in northern Europe (9823–10,365 BCE)
5 1440 24 ? ? [Birks et al., 1996]. For this reason, the Vedde Ash
6 1500 60 0.50 4 (in Nordic countries) or Skógar tephra (in Iceland)
7 1580 80 ? ? has been used to correlate tephrochronology from
8 1612 32 0.04 3 ­terrestrial cores in Iceland and northern Europe (e.g.,
9 1625 13 0.50 4 Blockley et al., 2007], to ice cores in Greenland
10 1660 35 0.26 4 (e.g., Grönvold et al., 1995], and marine cores from the
11 1721 61 0.33 4 North Atlantic (e.g., Jennings et al., 2002]. An inter-
12 1755 34 0.80 4 esting feature of this tephra is that it contains a mix-
13 1823 68 0.05–0.1 3 ture of both acidic and basaltic glass types that could
14 1860 37 0.05–0.1 3 indicate one transitional eruption, two eruptions, or
15 1918 58 0.5–1 4
a Plinian eruption followed by fissure deposition
Note. Design credit Nathan Mennen. [Norðahl and Hafliðason, 1992; Lacasse et al., 1995].
74 VOLCANICS

Again, although this tephra is often used as a example, reported estimates of the fallout tephra from
correlative tool for the last glacial termination it the1918 event exceeded 50,000 km2 on land with depo-
remains questionable. A petrogenesis evaluation of the sition reaching the Faroe Islands on the first day of the
ignimbrite and the ash by Tomlinson et al. [2012] was eruption [Larsen, 2010].
not able to confirm correlation between the ignimbrite Like tephra, the deposition that occurs via flooding
and the Vedde Ash, although they suggest that per- is an important part of the geologic record for Katla.
haps the Vedde Ash soon followed the initial eruption. Moreover, the fast‐moving and voluminous jökulhlaups
The first eruption to occur in the Katla volcanic produced at Katla are considered the most destructive
system after the Víking settlement in 874 CE was in Iceland [Tómasson, 2002; Russell et al., 2010];
60 years later in 934 CE. Since then, Katla has erupted sometimes even being referred to as “Katlahlaups”
20 times with a recurrence interval between 13 and (Figure 9.6). Regardless of eruption size, volcanic
95 years [Thordarson and Larsen, 2007]. All the historic activity under Mýrdalsjökull creates meltwater mixed
eruptions have been subglacial phreatomagmatic erup- with ice blocks and volcanic debris. As discussed by
tions generating basaltic tephra deposits ranging in Larsen [2010], since 1179 CE all documented Katla
volume from <0.1 km3 to >0.5 km3, with the most jökulhlaups have originated at the caldera and move
notable and widespread being 1625 CE, 1755 CE, and through a system called “Kötlujökull Pass” that leads
1918 CE [Thórarinsson, 1981c] (Figure 9.5). For to the coastline at the Mýrdalssandur. The Kötlujökull

Figure 9.6 Jökulhlaup flooding events since 1612 CE from Katla onto the Mýrdalur sandur plain. [Modified
from Larsen [2010]; design credit Nathan Mennen.]
Volcanics of the South and Southeastern Regions  75

Pass is a 15 km route that can be on top of, within, or migrating antidunes and the development of stationary
below the Kötlujökull glacial lobe to the east of Katla chute and pool structures may be associated with
(Figure 9.6). Through historic description, it appears localized hydraulic jumps.
that there is little warning before the flooding event as Low‐levels of seismic activity have been monitored
it can occur before, or just after, the first appearance of in the vicinity of the Katla volcanic zone for the past
an ash column. Flood deposits have been traced back 50 years. From these data two distinct seismic areas
to 1612 CE (Figure 9.6) and every jökulhlaup since have been determined: a location referred to as
1660 has extended the coastline, including the 1918 Goðabunga rise in the west and the caldera itself
eruption [Larsen, 2010]. From historic reports [Einarsson and Brandsdóttir, 2000]. Perhaps the most
Tómasson [1996] estimated the velocity of the flooding interesting data illustrated by a half century of moni-
discharge to be 7–10 m s−1 and the peak discharge dur- toring at Katla is Einarsson and Brandsóttir’s [2000]
ing a 1755 event to be ~300,000 m3 s−1, or similar to the notion of seasonality related to seismic dispersal. The
flow volume of the River Amazon. Tómasson [1996] authors concluded that there is an increase in earth-
also estimated that the total 3.6 km3 of sediment quake activity during the autumn (peaking in October)
deposited was moved by 8 km3 of meltwater. Fluvial due to changes in crustal pore pressure following the
modeling of the 1755 event by Duller et al. [2014] summer thaw. Soosalu et al. [2006] followed this study
determined that the coastline expanded ~4 km toward to look closer at the shallow (<2 km) movements
the ocean, but has since eroded to the original position occurring at Goðabunga. They proposed that the
mapped in 1904. Perhaps even more impressive is the Goðabunga seismic cluster might instead be the result
transport of a huge boulder (named Kötluklettur, of an acidic crytodome building in the area that accel-
64.4500, −18.6600) that is estimated to weigh 1000 tons erated toward the surface in 1999. They argued that
and was likely transported 14 km [Jónsson, 1982]. In when a cryptozone moves upward, the brittle crust
recent times, three jökulhlaups linked to seismic above it thins with time and the deformation becomes
activity underneath the ice cap occurred in 1955, 1999, localized. In 2010–2011 another seismic cluster was
and 2011. The small‐scale (M <3.7) seismicity and observed on Katla’s south flank ~4 km from the crater
geothermal activity tied to these flooding events are rim [Sgattoni et al., 2016]. Sgattoni et al. [2016]
thought to be indicators of unrest [Óladóttir et al., acknowledge the difficulty in discerning seismicity
2005; Soosalu et al., 2006; Sturkell et al., 2010]. relating to volcanic as compared to that initiated by
Sedimentary structures produced by highly turbulent glacial movement. They suggest, however, that volcanic
jöklhlaups are complex, but provide information on processes are likely the primary cause for the
the overall fluid behavior of the water [Russell et al., Goðabunga cluster as the seismicity started during a
2010]. Maizels [1989, 1991, 1992, 1993] conducted period of unrest and showed stability over time, and
detailed sedimentological investigations of the stratig- the ice cap appeared unresponsive to climate change
raphy for the 1918 event from in situ exposures, while during the episode (e.g., the glacier showed no obvious
also considering data collected on discharge, sediment signs of melting or cracking). In contrast, Jeddi et al.
supply, flow path, etc. A depositional model presented [2016] analyzed each region of Katla’s seismicity, sepa-
by Maizels [1993] indicates great variability in deposi- rating episodes into volcanotectonic, long period, or
tional style depending on location and distance from mixed, and concluded that volcanotectonic events
source (e.g., main flood path, point of flow separation, occur only at depths >4 km. Nonetheless, it is agreed
and flow around bedrock obstacles). An exposed that an eruption in the 21st century is imminent.
cliff face along the Selfjall River (63.9625, −20.4536) Likewise, Eyjafjallajökull (63.6314, −19.6083)
provides views of massive units, gravel beds, cross‐ (Photo 26), a neighbor ~25 km to the west of Katla
bedding, fining trends, etc. Additionally, she was able and within its volcanic system, has been recently active
to link lithofacies along the sandur plain with flow after nearly 200 years of dormancy. Eyjafjallajökull is
behavior thus illustrating pre‐surge (massive, clast‐ a stratovolcano (1666 m a.s.l.) that covers 100 km2 with
supported imbricated gravels), main surge (massive a 2.5 km wide elliptical summit caldera beneath 200 m
pumice granules), and post‐surge deposition (trough of glacial ice. Seismic activity at Eyjafjallajökull has
cross‐bedded pumice granules below horizontally bed- been monitored since 1973 with little activity recorded
ded pumice granules and sand) [Maizels, 1992]. Duller until 1991, which lead up to three shallow earthquake
[2007] and Duller et al. [2008] further explored the swarms in 1994, 1999, and 2009–2010 (depth to focus
main surge deposit described by Maizels [1992] as 1–13 km; most occurring at 5 km) [Sturkell et al., 2003;
massive and argued that the presence of large‐scale Gudmundsson et al., 2010]. Earthquake magnitudes in
76 VOLCANICS

Photo 26 Eyjafjallajökull, the stratovolcano in the background, is in close proximity to many farms and
farming communities. [Courtesy of Tamie Jovanelly.]

2010 were substantially larger than earlier swarms, Only three other eruptions of Eyjafjallajökull are
however, all seismicity is linked to magmatic intru- known during historic time: 1821–1823 CE, 1613 CE,
sions. Specifically, the earthquake swarms in 1994 and and 920 CE. Accounts of the 1821–1823 CE event
1999 are linked to crustal deformation and are inter- indicate small‐scale intermediate tephra emissions
preted to be the formation of a sill at 4–6 km depth from the south rim followed by a long pause of nearly
[Gudmundsson et al., 2010]. 2 years that eventually led to fluorine poisoning of
The most recent eruptive activity of Eyjafjallajökull area livestock and a jökulhlaup [Larsen, 1999; Sturkell
began on 20 March 2010. The first small eruption et al., 2003]. In less than 6 months after Eyjafjallajökull
occurred on an unglaciated eastern slope and pro- went quiet, a volcanic eruption occurred at Katla.
duced lava with little ejecta. This event was a precursor Similarly, historic accounts of Katla eruptions in
to a 3‐week effusive fissure eruption that caused little 1612–1613 CE and 920 CE followed soon after activity
disruption. On 13 April 2010 the IMO was alerted to at Eyjafjallajökull [Thoroddsen, 1925; Gudmundsson
increased seismicity beneath the summit caldera. By et al., 2005]. Additionally, a major eruption at Katla
the following morning, a VEI = 4.0 subglacial eruption occurred in 934 CE and the Eldgjá Fires from 934–940 CE.
had taken place along the southern crater rim emitting The historic accounts of these contemporaneous
an ash cloud 9 km into the atmosphere. The initial events of Eyjafjallajökull and Katla raises the question
eruption produced 1000 m3 s−1 of intermediate magma, of whether activity at Eyjafjallajökull can be a warning
and tephra deposition occurred over the course of sign of larger scale activity to follow.
3 days. Lava flows, small eruptions, seismic activity, Little is known about Eyjafjallajökull prior to his-
and flooding in the area continued until 2 May 2010 toric time, which limits the reliability of recurrence
[Gudmundsson et al., 2010]. intervals used in natural disaster planning and
Volcanics of the South and Southeastern Regions  77

assessment. Field survey by Jónsson [1998] suggested cones, spatter cones, and tuff cones [Hamilton et al.,
that some highly altered rocks intruded by dikes and 2010] (Photo 27). The vents and craters are separated
veins near Seljavellir farm (63.5667, −19.6333) may into >16 rootless cone groups within the Laki lava flow
indicate its origin to be >0.78 Ma, which would make field, which has been implied by Hamilton et al. [2010]
it among the oldest active volcanoes in Iceland [Sturkell to mean that the regional water table is high.
et al., 2003]. Typical of many volcanoes not located The Eldgjá and Laki Fires are similar in that they
within the rift zone, Eyjafjallajökull produces relatively each produced large basaltic lava fields associated with
small amounts of alkaline tephra and ash (≤0.1 km3) effusive flow; DRE 19.6 km3, and 15.1 km3, respec-
[Sturkell et al., 2003], which makes correlation by teph- tively [Thordarson, 1995]; the Eldgjá Fires being among
rochronology very challenging. At best, a thin layer of the largest flood‐basalt event recorded in historic time,
intermediate composition tephra in outcrop nearest to covering ~781 km2. Correspondingly, atmospheric
the volcano is found at the same level as Katla 1612 CE emission levels of SO2 have been estimated between
and may denote its eruption [Larsen, 1978]. Subglacial 70–122 Tg total for these two events [Thordarson
volcano eruptions, however, like those at Eyjafjallajökull, et al.,1996, 2001]. (As compared to 92 Tg emitted by
are typically associated with flooding events that can global anthropogenic sources during 2016 [EPA,
also leave deposits in the sedimentary record. Two 2017].) It should be noted here that the initial start date
flood events have been documented by Dugmore et al. of the Eldjá Fires is uncertain: recent high‐resolution
[2013] on the flanks of Eyjafjallajökull that date to the glaciochemical records from Greenland suggest 939 CE
6th and early 10th centuries. Although a silica tephra [Oppenheimer et al., 2018], and Stothers [1998] ques-
layer (Layer H) dates to 600 CE [Dugmore, 1987; tions the reliability of some documents that point to a
Kirkbride and Dugmore, 2008], there is no tephra layer 934 CE origin.
associated with the 10th century event. Undoubtedly, historic settlements were threatened
by these eruption events. The farms within the districts
9.3. ELDGJÁ AND LAKI (SKAFTÁR) FIRES of Alftaver and Síða had to be relocated during the
Eldgjá Fires, as lavas could not be prevented from
In the Katla Volcanic Zone there is a 75 km long advancing over their homesteads. Although the
active fissure system to the north of Mýdralsjökull number of direct deaths is not known for the Eldgjá
associated with eruption of flood basalts. Iceland is Fires, there is an abundance of recorded data explain-
one of the few places in the world where a recurrence ing the impact of the Laki Fires on people. As many as
interval for flood basalts has been established; an event 10,000 people may have been killed by the Laki Fires
will occur every 300–1000 years that will emit between in Iceland (referred to in Iceland as the “Haze Famine”)
5 × 106 and 9 × 106 tons of SO2 for each cubic kilo- and 60% of all grazing livestock died due to fluorino-
meter of magma erupted [Thordarson and Self, 2003]. sis. This is similar to the mortality Schmidt et al. [2011]
Although the eruption column height is usually only at found on mainland Europe, which they suggested was
tropospheric to lower stratospheric levels (i.e., polar jet as high as 17% between 1783 and 1784.
stream), the detriment to the environment is associated Additionally, historic documentation for the Eldjá
more with the duration of these effusive events, which and Laki Fires offer insight into the implications for
can last one to several years. global climate. Stothers [1998] explains how an increase
Two flood‐basalt eruption events have occurred in SO2 emissions can trigger cold weather conditions:
along this fissure, namely: Eldgjá (63.9553, −18.6200; SO2 is most harmful when it reacts with water vapor to
934–940 CE) and Laki (64.0744, −18.6200; 1783–1784 CE) form H2SO4 aerosols in the stratosphere where it can
Fires. The two systems are parallel to one another shelter incoming heat, thus inducing global cooling
although they differ in style. The older Eldgjá Fires with residence time of 1–2 years. For example,
reactivated a tectonic graben floor within the Katla Thordarson and Self [2003] suggest that in the after-
Volcanic Zone producing pseudocraters (rootless cones) math of the Laki Fires the continents of Europe and
at Landbrotshólar. The rootless cones in this area are North America experienced a mean surface cooling of
up to 70 km away from the active fissure, toward the 1.3°C that lasted 2–3 years. In parts of Europe espe-
Landbrot and Álftaver regions, as the pahoehoe‐type cially, both events have been linked to the presence of
lava was transported in lava tubes beneath the surface. sulfur fog (occurring multiple times during the Laki
The Laki Fires to the east occupied a fissure 27 km Fires), followed by unusually harsh winters that lead to
long that resulted in 140 vents and craters (also known the destruction of crops and, ultimately, famine
as the Lakagígar crater row), which include scoria [Stothers, 1998; Thordarson and Self, 2003; Schmidt
78 VOLCANICS

Photo 27 Cross‐sectional view of a crater row. [Courtesy of Tamie Jovanelly.]

et al., 2011; Sonneck et al., 2017; Oppenheimer et al., consequences to air travel and the long‐term effects on
2018]. Thordarson and Self [2003] completed detailed natural ecosystems in Europe were the greatest threat
mapping of this ~175 Mt aerosol plume of H2SO4 that if such an event occured today.
represented a mass loading from a one‐time event.
Conversely, the Eldjá Fires are not linked to global 9.4. DYRHÓLAEY LAVA CAVE
cooling despite Thordarson and Höskuldsson [2009]
classifying them as the greatest source of volcanic At present, Dryhólaey (63.3996, −19.1269) is a
pollution in recent history. Likely, this is because there 1.3 km2 tombolo connected to the mainland by two
were 15 explosive phases during a 6‐year span, as large offshore bars formed from sediment deposited
opposed to an 8‐month duration for the Laki Fires. offshore and the surrounding coastline (Photo 28).
Schmidt et al. [2011] considered how a Laki‐style However, Dryhólaey originated from a violent hydro-
eruption would affect populations at modern day. In magmatic eruption during the Surtseyan Episode
modeling the event using current atmospheric condi- 1963–1967 (see section 9.7) producing a hyaloclastic
tions against estimated eruption properties (i.e., mag- tuff cone separate from the mainland [White and
nitude, altitude, and timing of volcanic SO2 release) Houghton, 2000]. In the later stages of the eruption,
they determined a similar event could create a world- the explosions transformed into an effusive pahoehoe
wide health hazard with excess mortality to increase event, thus capping the tuff below it while also inter-
~8.2%. The authors explain that this number is less secting the shoreline to create a lava delta [Schmidt and
than those estimated for the Laki Fires because there is Schmincke, 2000; Thordarson and Höskuldsson, 2007].
less threat from starvation, increased ability to move It is in the lava delta area that a sea cave later formed
away from danger, and populations are overall by erosion mainly of hyaloclastic deposits. Gadányi
healthier. Schmidt et al. [2011] considered that the [2008] mapped the sea cave to establish that its entrance
Volcanics of the South and Southeastern Regions  79

Photo 28 At present, Dryhólaey is a tombolo connected to the mainland by two large offshore bars formed
from sediment deposited offshore and along the coastline. [Courtesy of Tamie Jovanelly.]

is 16.4 m wide, 2 m high, and 21.3 m in length. during post‐Pleistocene times; Torfajökull (63.8920,
Reynisfjara Beach (63.4053, −19.0764) nearby is −19.1220) is 25 km west and Tindfjallajökull (63.7831,
known worldwide for its black basalt beaches and its −19.7160), Eyjafjallajökull (63.6304, −19.6067), and
Plio‐Pleistocene columnar basalts formed during inter- Mýrdalsjökull (63.6467, −19.1303) are 25–50 km
glacial stages when sea level rose. southeast. In the Middle Ages, Hekla became known
as the “Gateway to Hell” by Europeans who believed
9.5. HEKLA—THE QUEEN that Hekla was one of two known entrances to Hell
(the second being the Stromboli volcano in Italy). For
Hekla (63.9923, −19.6658) is an elongated stratovol- this reason, Hekla went largely unexplored until 1750
cano that is situated on the southwestern boundary of when the first documented hikes to the summit took
the EVZ and is one of the most active volcanoes in place [Thórarinsson, 1970].
Iceland (Figure 9.3), and arguably Europe (Photo 20). The first eruption of Hekla dates to 5120 BCE
At the base of Hekla there are palagonite tuff deposits, [Óladóttir et al., 2011; Thorarinson, 1971] and marks
while the bulk of the mountain comprises many layers an important beginning to the explosive history of the
of lava and tephra. Although known for its explosively, volcano called “The Queen.” This event, referred to as
Hekla does not have a caldera, which may be the result H5 in tephrochronology, is an important marker bed
of a narrow and deep seated magma chamber [Larsen used for correlation of soil profiles as its silica‐rich
and Thórarinsson, 1977; Sverrisdottir, 2007]. Hekla is deposit can be found on ~90% of the island
part of a fissure system 5.5 km long called Heklugjá, (>80,000 km2), along with the two other equally
with the central summit standing at 1490 m a.s.l. intense events (VEI ≥ 4.0) of H4 at 2250 BCE and H3
Nearby are other major systems that have been active at 1050 BCE [Larsen and Thórarinsson, 1977]. The
80 VOLCANICS

finding of two additional tephra markers in the marine phase is typically basaltic andesite deposits and then
record further constrain the age of the H5 and H4 mafic fissure eruptions usually follow this. Since the
tephras. Specifically, Gudmundsdóttir et al. [2011] first eruption after human settlement in 874 CE, Hekla
established stratigraphical correlation of two tephra has erupted 18 times, with the last event occurring in
layers within the TFZ (Figure 2.1) on the north 2000. Up to the 1970s the recurrence interval of
Icelandic marine shelf. These marine records are Hekla eruptions was around two per century, but since
referred to as Hekla DH (4632 BCE) and Hekla Ö 1970 the volcano has erupted about every 10 years:
(4042 BCE). 1970, 1980, 1991, and again in 2000 (Table 9.1)
Here it should be noted that although all 18 recorded [Thórarinsson and Sigvaldason, 1972; Grönvold et al.,
historic eruptions (Table 9.1, Figure 9.3), as well as the 1983; Gudmundsson et al., 1992; Höskuldsson and
H3–H5 tephras, are found preserved in the tephra Olafsdottir, 2002].
record, sometimes complications can arise in distin- The duration and magnitude of the explosive phase
guishing between Hekla and nearby Katla deposits is directly correlated with the amount of time spent
[Jóhannsdóttir, 2007]. Additionally, it has been noted in dormancy prior to the eruption (Figure 9.7).
by Jagan [2010] that there are difficulties in discerning Additionally, the increasing frequency of Hekla eruptions
eruptions from the same volcano if the analysis is has been linked to the magmatic chemical composi-
based solely on major element composition. For tion of its eruption products [Sigvaldason, 1974;
example, Jagan [2010] and Larsen et al. [1999] described Höskuldsson et al., 2007]. After dormancy lasting
similarities in the mineralogical composition of the 100 years or more, high‐silica magma (rhyodacite) was
tephra deposits H1510 and H1947 (8 and 14 in ejected from the volcano (i.e. H3, H4, H5, and
Table 9.1), and only an understanding of their position 1104 CE). After the 1104 CE event there is a shift from
in the stratigraphic column can resolve any confusion. the ryhodacite composition to the less silica‐rich
A comprehensive database designated to the study of basaltic andesite, which also corresponds to an increase
tephrochronology can be found on the internet (www. in frequency of eruption (<30 year interval)
tephrabase.org) to aid in the mapping and further [Thórarinsson, 1970]. Eruptions taking place along the
understanding of tephra in Iceland and Europe. margin of Hekla, however, are always found to be
Hekla’s unique increase in eruption pattern throughout more basaltic [Thórarinsson, 1967; Jakobsson, 1979b].
its history delineates it from other Icelandic volcanoes. This again raises the question concerning the origin of
The initial phase of each Hekla eruption is always a silica‐rich magma in the EVZ. As reviewed by
highly explosive subplinan to Plinian type eruption Sverisdottir [2007] research on the chemistry and
[Janebo, 2016; Gudnason et al., 2017] that has produced petrology specific to Hekla has been extensive. Many
silica rich tephra in volumes ranging from 0.02 km3 to mechanisms have been proposed including: fractional
2.2 km3 (namely the H3 and 1104 CE events; Figure 9.3, crystallization in shallow magma chambers [Einarsson,
Table 9.1) [Larsen et al., 1999]. The second eruption 1950], assimiliation of silica rocks from unknown

Figure 9.7 Silica content as compared to repose time and Hekla eruption events. [Modified from
Thórarinsson [1970]; design credit Nathan Mennen.]
Volcanics of the South and Southeastern Regions  81

sources [Tryggvason, 1965], and mixing of two differ- these data, Pedersen et al. [2017] compared the vol-
ent magmas produced by crustal melt [Sigvaldason, umes produced with the observed changes in v­ olcano
1974a]. Although still a matter of debate, recent inves- inflation rates (coeruptive tilt change), thereby making
tigations in petrology help to support the idea of a correlations to overall eruption size.
hybrid magma that originates first from partial melting In contrast to other volcanoes in the EVZ, detailed
of tholeiitic crust to produce silica magma, which is eruption accounts exist that report on the human
held in a deep crustal reservoir until release, thereby implications and the response to Hekla [Thórarinsson,
providing the potential for mixing with other material 1970]. Particularly devastating eruptions took place in
[Sverrisdottir, 2007]. This idea of hybridization of 1300 CE, 1341 CE, 1510 CE, 1693 CE, and 1766 CE.
magma corresponds to the data collected for the For example, the event in 1300 CE produced some
majority of the 18 documented historic eruptions (e.g., 30,000 km2 of tephra, some of which was carried
1300, 1693, 1766, 1845, 1947, 1970, 1991, 1980, and northward to produce darkness for days. Some 500
2000) [Gudnasdon et al., 2017]. people died in Skagafjördur and Fljót. In 1693 it was
As described by Sigmundsson [2006], seismicity is recorded that about 60,000 m3 s−1 of tephra fell in the
typically a precursor to volcanic eruptions, as the initial hour of the eruption, devastating eight farms
movement of the intruding magma causes stress from [Thórarinsson, 1970]. Beyond the devastation that
expansion within the rock to accommodate for viscous falling ash and tephra can have on a civilization, the
flow. Different to the other volcanoes in the EVZ, how- presence of colorless gases or the distribution of harm-
ever, the modern‐day high‐intensity opening phase ful elements has proven to be just as troubling. After
eruptions monitored at Hekla occur within a 30 min the eruption in 1947, problems arose with continued
window of initial seismicity [Gudmundsson et al., 1992; CO2 emissions, which caused water contamination and
Soosalu et al., 2003; Thordarson and Larsen, 2007]. This the death of livestock in the Loddavötn area. Within
raises questions about the location and depth of the one year, the CO2 and lime precipitate in the water dis-
magma chamber. Prior to the 1991 eruption Tryggvason sipated [Thórarinsson and Sigvaldason, 1972]. The next
[1994] measured slow rates of ground inflation 4–6 km eruption on 5 May 1970 resulted in acute fluorosis of
away from the central summit, with uplift occurring at livestock, which came about through increases in the
a rate of about 3.5 cm year−1 over the course of an chemical element fluorine that had adhered to the
11 year repose. Based on these measurements, the depth falling tephra particles. The initial samples on 5 May
of source chamber should be ~6 km below the surface. contained 2000 ppm of fluorine (fatality in sheep is
Weber and Castro [2017] later confirmed this data set caused by the ingestion of 25 ppm or greater), which
by using experimental petrology constraints on magma accounted for the death of 3% of adults and 9% of the
storage for the H3 event. They concluded that a large newborn populations [Georgsson and Pétursson, 1972].
silica eruption at Hekla could be fed by a shallow, Other historic eruptions have reported the emittance
upper crustal reservoir, which may translate into less of fluorine, including those of 1341, 1693, and 1766 CE
ground displacement, which would be make it harder [Thórarinsson, 1970].
to ultimately predict. Undoubtedly, changes in magma
composition will influence the geologic structures 9.6. ÞJÓRSÁRDALUR VALLEY
found at the Earth’s surface. It has been seen at Hekla
that mixed eruptions can produce both tephra and lava Þjórsárdalur Valley is ~20 km to the north of Hekla
deposition [Thórarinsson, 1970]. After the first tephra‐ and has been labeled as the “Pompeii of Iceland” since
producing Plinian eruption (which can last between first written about in 1897 by Bruun. The deposition
one and several hours), andesitic lava forms aa lava of more than 10 m of pumice from the Hekla 1104 CE
flows that are typically followed by pahoehoe struc- eruptions enveloped 15 medieval farm sites containing
tures of alkali basalt [Thórarinsson, 1970]. one‐third of all Víking settlements in Iceland [Vésteinsson,
The past five eruptions of Hekla in the 20th century 2004; Dugmore et al., 2007]. The sites were first exca-
have been well documented by scientific communities vated in 1939 by Nordic archeaologists who collabo-
and provided both insight and further questions into rated with geologist Sigurður Þorarinsson in the very
the hazardous nature of living in its shadow. Precise preliminary stages of establishing tephrochronology.
bulk volume measurements (≤0.241 km3) completed by Specifically, a farmstead called Stöng (64.1200, −19.8261)
Pedersen et al. [2017] indicate that there has been great was found and later determined to be the home of
variability in lava production rates since the 1947 event a prominent farmer and warrior called Gaukur
(ranging from 7 × 106 to 40 × 106 m3 year−1). Using Trandilsson. The initial site assessments and subsequent
82 VOLCANICS

studies provided unclear evidence as to whether the well documented phreatomagmatic eruption has
sites were abandoned. Dugmore et al. [2007] concluded become formally classified by volcanologists as the
from historic studies that volcanic eruptions are rarely “Surtseyan‐style” eruption and has been applied to a
fatal to settlements, and if so, only for a short period number of other locations where this kind of activity
of time. Meaning, that civilization likely returned to has also occurred (e.g., Taal, Philippines; Kavachi
the area soon after. Overall, the geomorphic and teph- Volcano, Solomon Islands; Ambae, Vanuatu).
rachronologic evidence of the valley indicates Fishermen off the south coast of Iceland first
landscape destabilization (e.g., lack of soil formation) noticed smoke on the horizon during sunrise on 14
that delayed ecological recovery until post‐1300 CE November 1963. Concurrently, the residents of Vík
[Dugmore et al., 1995]. However, this maybe the result noticed the smell of hydrogen sulfide gas, and a marine
of delayed environmental rebound from the ensuing science vessel measured above‐normal sea‐surface
Hekla eruptions of 1158 CE, 1159 CE, 1206 CE, and temperatures in the North Atlantic [Thórarinsson
1300 CE. et al., 1964]. It was soon after these events that Icelandic
In the upper Þjórsárdalur Valley, near the major gla- geologist’s flying over the archipelago witnessed the
cial river Þjórsá, the Hreppar Formation is exposed growth of a new volcanic island. The growth of Surtsey
comprising a succession of lava basalts. These lava has been chronicled in detail by 85 site visits made by
units show great variability in thickness (up to 500 m) Sigurdur Thórarinsson [1964, 1966, 1968] and by innu-
as they infilled valleys and river channels, and buried merable scientists who contributed to the nine volumes
hills [Kristjánsson et al., 1998]. These volcanic deposits of the Surtsey Research Progress Reports.
were derived from two extinct volcanoes, Stóra‐Laxá Over the course of 3.5 years the island developed
and Þjórsárdalur, and are Early Pleistocene in age through three main eruption phases [Thórarinsson,
(2.6–0.78 Ma). The Hreppar Formation may be a 1966, 1968] (Figure 9.9), with the total amount of
crustal remnant that was left by the eastward movement eruptive material estimated at 1.1 km3; 70% tephra and
of the active spreading zone 2–3 Ma [Kristjánsson 30% lava [Thórarinsson, 1969]. Phase I was subaerial
et al., 1998]. On top of the Hreppar Formation near emergence of the island from 130 m below sea level.
Hrepphólar (64.0694, −20.3275) is an interglacial lava Phase II allowed for phreatic eruptions above the
flow with columnar jointing. The Hrepphólar columns ocean surface that produced unconsolidated basalt
measure an average of 0.45 m in diameter and many from 1150 to 1160°C or hotter magma dispersed in
exceed 20 m in length, although the basal contact is finely bedded layers. The newly deposited tephra was
not exposed [Kristinsdóttir, 2010]. made up of 82–88% unaltered basaltic glass (sider-
omelane; density 2.70 g cm−3), and the remaining
9.7. SURTSEY VOLCANIC ISLAND percentage comprised phenocrysts of olivine, plagio-
clase, and fragments of autogenic hyalobasalt
An extension of the EVZ is a young (<20,000 ka), [Jakobsson and Moore, 1986]. By 18 November 1963 a
mostly alkali archipelago called the Vestmannaeyjar, 550 m linear ridge, 45 m a.s.l., developed into a bifur-
or commonly the Westman Islands (63.4277, cated fissure system that gradually culminated in an
−20.2674). The largest and only inhabited island, eruption from a single vent. By 24 November the shape
Heimaey, developed over 12,000 years from a series of of the island became nearly circular (900 m long,
mixed eruptions from a central volcano, which 650 m wide) with the crater rim at 100 m a.s.l. By the
occurred under both hydromagmatic and nonhydro- end of the first five months of explosive hydromag-
magmatic influences. The other smaller islands were matic eruptions (4 April 1964) two unconsolidated tuff
developed from submergent eruptions, the most cones called Surtur I (420 m) and Surtur II (500 m)
notable being Surtsey, which is 33 km off the southern had developed [Thórarinsson et al., 1964]. During this
coast of the mainland. The island volcano was named time, eruption columns of ash rose to 13 km and
Surtsey after the fire giant “Surtr” from Norse updrafts were reported to cause severe lightning epi-
mythology, and has become one of the most docu- sodes accompanied by water spouts and hail [Decker
mented cases of volcanic development and evolu- and Decker, 2005]. Once the tuff cones grew large
tionary biology in the world. This pristine landscape enough to block the ocean water from entering the
has been protected since its origin as a United Nations vents, the eruptions transitioned to Phase III, an effu-
Educational, Scientific, and Cultural Organization sive subaerial eruption lasting 13.5 months at Surtur II
(UNESCO) World Heritage Site and continues to be a (Figure 9.8). This was a defining moment for the lon-
site of scientific discovery (Figure 9.8). Moreover, this gevity of the island as low‐viscosity lava began to
Volcanics of the South and Southeastern Regions  83

Figure 9.8 Development and formation of Surtsey volcanic island. [Modified from Decker and Decker
[2005]; design credit Nate Mennen.]

solidify on the surface to create a hard cap‐rock over i­mportant for the formation of the 460 × 106 m3 of
the loose tephra deposits that formally made up the tephra that was created during the first 4.5 months as
island [Decker and Decker, 2005]; its area would the crater rim built above sea level (Figure 9.9). The
expand to 2.5 km2. In comparison, during the “Surtla balance in creating a depositional island is having
Episodes,” occurring between December 1965 and July enough water as to not decrease temperatures, while
1966, three small volcanic cones (Syrtkingur, Jolnir, producing enough magma as to not reduce the amount
and Surtla) were formed near Surtsey along submarine of steam produced (Figure 9.10). In order for Surtsey
vents on a fissure 250–300 m long [Moore, 1985]. As to maintain a hydromagmatic steam explosion, Moore
the pseudo‐islands were small in size (<64 m), they [1985] determined that infilling of water into the
quickly eroded. Finally, the reactivation of Surtsey I Surtsey craters would have to be around 20 m3 s−1. The
(June 1967) marked the last activity when lava tunnels unusual mechanics of this continuous uprush explosion
made their way to the ocean and increased the surface cycle reflects the precise combination of variables
area to 2.65 km2. required for accretionary lapilli‐rich material to be
Amongst other revelations, the Surtsey eruption deposited [Moore, 1985]. Additionally, the presence of
provided insight into the relationship between water– diatreme pipes confirms this cycle and fits the Lorenz
magma mixtures and the explosivity of the resulting [1975] description of volcanic conduits filled with
event. Moore [1985] considered this interaction pyroclastic debris and interior wall blocks whereby the
84 VOLCANICS

improved. The initial rates of subsidence measured


in 1967–1968 were 15–20 cm year–1 and likely repre-
sents the compaction of volcanic material and the
sea‐floor underneath the island, along with litho-
spheric downwarping due to the new subaerial land-
form as 2 × 1015 g of newly added crust was deposited
on the surface [Tryggvasson, 1972; Moore et al.,
1992]. Overall, the rate of subsidence measured was
1.1 ± 0.3 m during the first 24 years [Moore et al.,
1992] and 1–2 cm year−1 thereafter, with the greatest
subsidence occurring on the eastern side of the island
near Surtsey II. Jakobbson and Moore [1986] reported
the first signs of compaction in August 1966 on the
outermost 10–15 cm of exposed tephra, although
initial microscopic examinations of these samples
did not reveal any palagonization starting. Compaction
is important to the discussion of subsidence at
Surtsey because the initial total porosity was reported
as high as 40–50% [Oddsson, 1982]. As an example, it
was estimated that by 2007, 95% of the volume of
the tephra cones was altered to compact and non-
fractured palagonite tuff [Jakobsson et al., 2009].
Thus, in addition to the height reduction of Surtsey,
erosive forces of temperature and coastal geomor-
phologic processes have caused both subaerial and
submarine destruction to reduce the area of the
island to half its original size in 1967 (~1.3 km2).
Figure 9.9 Surtsey island cross‐sectional shape and changes
over 50 years. [Adapted from Jakobsson [1972b]; design
Romagnoli and Jakobsson [2015] make comparisons
credit Nate Mennen.] of erosional and depositional changes that have
altered Surtsey’s shape over the past 50 years, while
noting that extreme rates occurred during the first
explosion carries the vesiculated material to the sur- few decades and then slowed in recent time. Perhaps
face. At Surtsey, diatremes are commonly subjected to the most drastic reduction resulting from the high‐
extensive hydrothermal alteration [Moore, 1985], energy coastal processes has been that of the uncon-
which is vapor dominated above sea level and seawater solidated tephra coastline, which eroded hundreds
dominated below sea level [Jakobsson and Moore, of meters to produce cliff faces and a submarine
1986]. Similarly, a further investigation of palago- shelf on the southern portion of the island. Here, a
nitization indicated that sideromelane and olivine boulder beach has formed in the swash zone, which
combined to form an array of secondary minerals that in turn, has protected the remaining cliff face from
crystallized both above and below sea level between wave abrasion. In contrast, on the northern coast
24° and 149°C [Jakobsson and Moore, 1986]. It was there is a migrating cuspate spit that is influenced
also determined from drill cores that virtually all of by strong currents and is dependent on the sedi-
the glass is replaced by palagonite at 120°C, which ment balance provided by erosion of other parts of
indicates that the alteration process is largely tempera- the island. The submarine morphology of the island
ture dependent [Jakobsson and Moore, 1986]. has also been monitored through bathymetric sur-
Monitoring the subsidence, consolidation, and veys from 1967 to 2007, which paid particularly
erosion of Surtsey has been an on‐going process for close attention to the three previously mentioned
more than 50 years. The initial ground surveys com- pseudo‐islands that formed below sea level.
pleted on Surtsey included the installation of 42 Interestingly, these submarine shoals are areas of
bench markers along a 1600 m east–west transect intense erosion that has led to the majority of Jólnir
[Tryggvason, 1968; Moore et al., 1992], which was and Syrtlingur being destroyed [Romagnoli and
measured frequently with markers being added to or Jakobsson, 2015].
Volcanics of the South and Southeastern Regions  85

Figure 9.10 Volumetric proportions of materials that influence the evolution of hydromagmatic eruption
craters, including temperature (°C), water (%), rubble (%), and lava (%). Diagonal pattern indicates presence
of steam. [Modified from Moore [1985]; design credit Nate Mennen.]

The evolution of Surtsey island has also contributed lichen (1970) [Fridriksson et al., 1992]. Additionally, sea
to the vast knowledge of ecological succession of plants birds (mainly gulls) started to nest on the island in
and animals. As soon as the island stabilized above sea‐ 1970, which became an intergral part of seed dispersal
level in 1965, the first evidence of vascular plants was in the island interior. Today, there are more than 69
found, which were followed by byrophytes (1967) and species of vascular plants [Magnússon et al., 2009].
10
Volcanics of the North and Northeastern Regions

The EVZ transitions to the NVZ under Vatnajökull of magma from a central volcano [Sigurðsson and
and in conjunction with MIB forms a triple junction in Sparks, 1978; Macdonald et al., 1990], whereas others
central Iceland (Figure 2.1). From central Iceland, the support ­fissure eruptions being fed by vertical magma
NVZ extends 200 km northward to the coast, where it flow not associated with a main reservoir [Hartley
connects with the TFZ. Compared with the EVZ and and Thordarson, 2012, 2013] (Figure 10.1). According
WVZ that form two subparallel rift zones in the south, to the lateral flow hypothesis, when there is a shallow
the NVZ is considered simpler as there is only one pri- magma chamber underneath a central volcano the
mary segment (50 km wide) that has been active over reservoir can become pressurized through magma
the past 6–7 Myr with overall propagation to the north replenishment that leads to lateral injections of shal-
for the past 2 Myr [Trønnes, 2002]. low dikes into a fissure swarm. Sigurðsson and Sparks
In the center of the NVZ there are five overlapping [1978] identified this process as a driving mecha-
fissure swarms arranged en echelon, each with a nism of rifting on a volcanic system. Conversely,
central volcano: Þeistareykir, Krafla, Fremrinámar, Gudmundsson [1995] suggested a magma reservoir
Askja, and Kverkfjöll (Figure 5.2). In the northern ­hypothesis where the pressurization of deep magma
part the long linear extensions of the NVZ fissure chambers (>20 km depth) can induce volcano‐
swarms are often crosscut by the TFZ (e.g., the noth- tectonic episodes that trigger subvertical dikes into
ern end of the Krafla and Fremrinámar fissure the crust. The still debatable lateral flow hypothesis
swarms). All these f­issure swarms, and particularly was first introduced as a phenomenon applied to
those areas nearest to the central volcanoes, are high‐ Icelandic rifting by Björnsson et al. [1977] to describe
temperature geothermal fields that have produced the Krafla Fires occurring in 1975–1984.
silica‐rich explosive eruptions during historic time The three‐phase cyclic deformation process (inter‐,
[Sæmundsson, 1974], however, the NVZ makes up only co‐, and postrifting) of the NVZ tectonic rifting his-
about 2% of the overall volcanism of the island and tory has influenced its volcanics [Pedersen et al., 2009].
16.5% of the total DRE produced [Thordarson and In particular, the mature central volcanoes of Askja
Höskuldsson, 2008]. Additionally, the presence of and Krafla have undergone caldera collapse, likely
numerous Holocene lava shields, hyaloclastite ridges, influenced by spreading along the plate boundary.
and fissure lavas are characteristic of the NVZ Consequently, these two volcanoes are the only ones in
[Sigvaldason et al., 1992]. the NVZ to have produced eruptions within the past
Importantly the NVZ contributes to the discussion 1000 years. The two most recent interrifting episodes
of how silica‐rich magma can be generated at MOR in the NVZ occurred between 1975 and 1984 along the
(Mid Ocean Ridge) locations (e.g., the catastrophic Krafla fissure and almost 100 years earlier, again at
eruption of Askja 1875). Historically, researchers Askja [Pedersen et al., 2009]. Sæmundsson [1991] con-
have suggested that this event was fed by lateral flow firmed that the Krafla fissure swarm has been among

Iceland: Tectonics,Volcanics, and Glacial Features, Geophysical Monograph 247, First Edition. Tamie J. Jovanelly.
© 2020 American Geophysical Union. Published 2020 by John Wiley & Sons, Inc.

87
88 VOLCANICS

Figure 10.1 Comparison between the lateral flow (top and middle panels) and magma reservoir (bottom
panel) hypotheses with regard to the generation of silica‐rich magmas at MOR locations. [Modified from
Hartley and Thordarson [2013]; design credit Nathan Mennen.]

the most active of the NVZ, with at least six major to the greatness of Hekla. “In the desert interior of the
events during the past 3000 years. Pedersen et al. [2009] island—the very fact of its being a volcano unknown
link area subsidence to weaknesses in plate boundary even to the Icelanders themselves prior to 1875—
structure and to increases in magmatic activity of this stands a volcanic mountain whose vast proportions
system (mainly Krafla and Askja) in recent time. dwarf Hekla into utter insignificance” [Lock, 1881,
Similarly, Drouin et al. [2017a] confirmed the location p. 1]. In fact, due to the rugged terrain of the interior,
of the spreading axis through Krafla and Askja central and the remoteness of the volcano, it is thought that
volcanoes and suggested that recent deformation of the Askja caldera had only been visited once prior to
the area may be linked to glacio‐isostatic adjustments the 1875 eruption [Sigurðsson and Sparks, 1978].
due to the retreat of the Vatnajökull ice cap, which has As confirmed by radiocarbon dating, the areas
been ongoing since 1890 [Björnsson et al., 2013]. eruption history begins in 6892 ± 200 BCE with a
VEI = 5 eruption that initially created the Kollur
10.1. THE DYNÁUFJÖLL MASSIF COMPLEX volcanic edifice [Sigvaldason, 1979; Sæmundsson,
AND ASKJA 1991; Smithsonian Institute, 2019] (Figure 10.2). The
Plinian‐style eruption (1–2 km3 DRE) was triggered
William George Lock wrote a book about Askja in at the time of glacial retreat and is associated with
1881. In the first pages he very abruptly describes the the glacio‐isostatic rebound of a 400–500 m crustal
greatness of this volcano, which in his opinion has depression as the 1500–2000 m thick ice sheet in
been undervalued and underestimated in comparison central Iceland melted; this coincided with volatile
Volcanics of the North and Northeastern Regions  89

Figure 10.2 Askja crater. Locations B, C, and D are described in section 10.1. [Modified from Lupi et al.
[2011]; design credit Nathan Mennen.]

supersaturation as glacial melting reduced litho- Greenland ice cores, on top of glacial deposits in
static pressure above the rhyolitic magma northeast Iceland, and in the central highlands of
chamber [Sigmundsson, 1991; Sigvaldason, 2002]. Iceland [Grönvold et al., 1995]. One of the thickest
Hyaloclastites and pillow lava ridges that dominate units of Dyngjufjöll tephra (15 cm) described on
the area were likely formed subglacially by ring‐frac- land was found 150 km to the northeast on a raised‐
ture eruptions [Sigvaldason, 1968]. Gravity survey beach in Bakkaflói (66.1250, 14.8570) [Sæmundsson,
determined that the main vent was located in the 1991]. Tephra mapping by Sigvaldason [2002] within
central part of the Dynáufjöll volcano (areal extent a 5 km radius of the central vent provided some
of 400 km2) [Brown et al., 1991]. Although the vent clues about the eruptive phases, including the likely
was destroyed by the 1875 eruption, remnants can be presence of a meltwater lake that was created by the
seen today as an embayment structure in the southern eruptions themselves. According to Sigvaldason
part of the Askja caldera (45 km2) that is nested [2002], however, erosive processes over the Holocene
within the larger system. This extensive eruption is have erased most of the tephra from the land‐based
associated with a silica‐rich tephra layer, referred to record on Iceland, limiting the possibility of map-
as the Dyngjufjöll tephra that has been found in the ping the areal extent more fully.
90 VOLCANICS

The crown of the Askja central volcano (also referred event per 50–150 years. Basaltic fissure eruptions
to as the Dyngjufjöll volcanic center) is its caldera, with volumes between 0.05 and 0.2 km3 are typical at
which was created during the eruption of 10 ka when Askja, however, large‐scale explosive silica eruptions
lava flowed over the ice‐free surfaces. The caldera (tephra volumes <2.0 km3) have been documented
formation is unusual in that it is not considered to be [IMO, 2019].
either a structural caldera (i.e., subsidence along faults) The central volcano known as Askja (65.0111,
or a topographic caldera (i.e., mass wasting along a −16.7485) (Photo 29) is just one part of a larger massif
structural depression). Instead, Sigvaldason [2002] called the Dyngjufjöll complex in central northeast
describes the caldera being formed via uplift from Iceland. This Pleistocene möberg mountain range
rapid crustal rebound and doming in the volcanic covers an area of 250 km2 and has an average relief of
center from increased magmatic pressure (forming the 600–760 m over the younger (<2500 BCE) Odáðahraun
Ö skjuop Pass, 65.0333, −16.7500), which was then fol- lava fields, which are considered the largest continuous
lowed by collapse. Since 10 ka, a minimum of 39 km3 lava desert in Iceland at 4000 km2 [Thórarinsson, 1963;
of basaltic magma has erupted from 175 effusive and Sigurðsson and Sparks, 1978]. Likely, the tholeiitic basalt
explosive eruptions that have been documented lava fields resulted from effusive pahoehoe‐style
through stratigraphic sequencing and tephrochronol- eruptions from numerous shield volcanoes in the area;
ogy [Hartley et al., 2016]. Approximately 30% of the namely, Trölladyngja (1460 m; 64.8944, −17.2497),
eruptions occurred during historic times, half of which Kollóttadyngja (1177 m; 65.2194, −16.5475), and
are associated with the Askja caldera. Ketildyngja (986 m; 65.4306, −16.6444) [Gudmundsson,
Hartley et al. [2016] suggests that magma discharge 1996]. Gudmundsson [1996] suggested that this large‐
rates at Askja were highest during postglacial times scale influx of lava may have influenced the initial
between 5182 BCE and 2282 BCE, although they were caldera collapse and associated fissure swarm (140 km
not able to associate deglaciation with peak volcanic long), some of which filled in the caldera floor from
activity here due to the low resolution of tephrochro- 1470 BCE to 782 BC. Over time, this rugged terrain
nology for the area. Of the total eruptions at Askja, 14 was highly weathered by glacially directed wind pat-
of the tephra layers found within the vicinity are terns creating the perfect pseudolunar surfaces that
considered time marker beds [Hartley et al., 2016]
­ would be used as practice grounds for the first NASA
(Table 10.1), from which eruption frequency has been lunar rovers in the late 1960s.
determined as 2–3 eruptions per 100 years on average The sheer size and complexity of Askja’s nested cal-
[IMO, 2019]. Unlike Iceland’s other major volcanoes, dera structure (50 km2) is what makes this central vol-
the repose period for Askja is highly variable; repose cano the largest in Iceland, with its three overlapping
periods can be a few years to a decade to only one crater depressions (Figure 10.2) [Sigurðsson and Sparks,
1978]. The three craters (Askja, Ö skjuvatn, and Kollur)
Table 10.1 The eruptive history of Askja with associated reside between two identical fissure swarms to the
volcanic explosivity index (VEI) west and east that delineate part of the Askja volcanic
system. Ö skjuvatn is the youngest crater (11 km2) that
Date VEI
formed during a subsidence event during the 1875
8910 BCE 5 eruption. Remote sensing of the area indicated that
2050 BCE 0
Kollur and Askja calderas are parallel to the regional
fissure swarms, whereas Ö skjuvatn caldera is perpen-
1250 BCE 0
1300 CE 1
1797 CE 0 dicular [Trippanera et al., 2018]. These differences in
1 January1875 5 structure may be important to the overall under-
1919 2 standing of magma source, caldera collapse, location
March 1921 0 of landslides, etc. [Trippanera et al., 2018]. The Askja
November 1922 0 volcanic system is 10–15 km wide and 100 km long and
15 January 1923 0 contains the volcanic flank that rises 100–500 m above
1924 0 the present caldera floor [Brown et al., 1991] and con-
15 July 1926 2 tinuously goes through cycles of inflation and defla-
19 December 1938 2 tion, with associated volcanic activity [Tryggvason,
26 October 1961 2 1989b; Camitz et al., 1995; Dalfsen et al., 2005].
Note. Adapted from the Smithsonian Institution [2019]; design Geothermal properties of the area are currently uti-
credit Nathan Mennen. lized along the eastern and southern edges of the
Volcanics of the North and Northeastern Regions  91

Photo 29 Askja crater. [Arctic images/Alamy Stock Photo.]

Ö skjuvatn caldera, specifically related to fissure vents at Askja, the Sveinagja graben (45–60 km north of
(labeled B, C, and D on Figure 10.2). Other geothermal Askja) became active from February to October 1875.
activities are utilized ~2 km north of the main caldera, Over these 8 months there were at least six eruptive epi-
and the nearby Viti Crater (60 m in depth) continues to sodes, which produced the Nýjahraun lava field [Harley
be a popular geothermal bathing spot for tourists. and Thordarson, 2013]. The Nýjahraun lava field and its
associated basalts are important to the understanding
10.1.1. The Askja 1874–1876 Event of the overall mechanics of the magma system beneath
Askja. Hartley and Thordarson [2013] found that the
After 400 years of little activity, the volcano‐tectonic petrology of the 1875 event (composed of olivine, pla-
event from 1874 to 1876 would mark the second VEI = 5 gioclase, and clinopyroxene phenocrysts) is dissimilar to
explosion recorded at Askja [Gudmundsson, 1996; the 20th century aphyric events. From this they deduced
Smithsonian Institute, 2019] (Table 10.1), leading to the that the 1875 event is not tied to the same magma
abandonment of 17 farms and the documented emigra- source as recent events; specifically, that the Nýjahraun
tion of area residents to America [Sigurðsson and Sparks, lava likely did not originate from the lateral migration
1978]. Precursory seismic activity was reported on (northward) of magma from Askja thus supporting the
17–18 April 1872 when an estimated 6–7 M earthquake magma reservoir hypothesis (Figure 10.1). The Sveinagja
occurred 100 km northwest along the TFZ (Figure 2.1) graben itself (34 km long, ~2.5 km wide) was initially
[Tryggvason, 1973; Newhall and Dzurisin, 1988]. Two active during a rifting episode that started in 1872 when
years later the Askja central volcano and associated fis- the northern part of the graben subsided between 3 and
sure swarm became active, with seismicity and eruptions 6 m [Gudmundsson and Bäckström, 1991].
that continued until early 1876 [Sigurðsson and Sparks, Plinian and phreatoplinian eruption began on 28 and
1978; Brandsdóttir, 1992]. Following the initial ­seismicity 29 March 1875 and were well documented by ­eyewitness
92 VOLCANICS

accounts that occurred both during and immediately There are two main competing hypotheses that remain
after the event [Lock, 1881]. The majority of the ash inconclusive: lateral dike injection [Sigurðsson and
produced during this 17 h event was dispersed to the Sparks, 1978] versus tectonically induced subsidence
east (Figure 10.3) and was found up to 1 mm thick as far [Hartley and Thordarson, 2012]. Viti, a large steam‐
away as 1100 km in Scandinavia [Thórarinsson, 1981a; explosion crater, also formed during this event and is
Carey et al., 2010]. The maximum column height for found on the northern margin of Ö skjuvatn. Some
the Plinian‐type explosions at Askja was estimated at active hydrothermal fumaroles are found within the
26 km, with a combined DRE total of 0.321 km3 and nested crater system today.
total tephra volume at 1.84 km3 after 8 h (Table 10.2). Well‐preserved ash deposits enable accurate docu-
The 11 km2 caldera that formed is called Ö skjuvatn, mentation of relevant parameters, although there is
and is one of the youngest collapse calderas on Earth no one locality where every layer is found [Carey
[Hartley and Thordarson, 2012] and now holds a 230 m et al., 2010]; most are documented records are from
deep lake in the center [Sigurðsson and Sparks, 1981]. the northern and eastern rims of Ö skjuvatn [Sparks
Caldera formation followed the explosive eruption and et al., 1981]. Uniquely, three drastically different
the total volume of subsidence (2.88 km3) far exceeded eruption styles occurred during the 1875 event (sub-
the volume of erupted products (0.764 km3), but it took plinian, pheatoplinian, and pyroclastic surge and
another 40 years before the caldera bottom was at Plinian) that correspond to wet versus dry magma, as
the current position of ~1100 m a.s.l. [Hartley and well as variance in eruption regime (ash fall vs. pyro-
Thordarson, 2012]. Formation of the caldera is impor- clastic surge) [Carey et al., 2010]. Changes in the
tant to understanding the Askja plumbing system. eruptive style can be caused by variations in volatile
concentrations in one or more interacting magmas
prior to eruption, or to the interaction of magma
with external water sources, as suggested by Carey
et al. [2009]. These phases were first determined by
tephrochronology work completed on pyroclastic
flow deposits by Self and Sparks [1978] where they
separated the flow into six layers (A–F) (Figure 10.4):
layers A, E, and F represent weak phreatic or hydro-
thermal activity prior to or following the main phases
represented by layers B, C1, C2, and D) [Carey et al.,
2010]. Complete descriptions for each unit are
provided in Carey et al. [2010], including details on
direction of pyroclastic flow, bedding, grain size,
etc. The phreatoplinian and Plinian ash fall units
(layers C and D) are the most widely dispersed. While
Sigurðsson and Sparks [1981] noted the varying lithol-
ogies of the pyroclastic deposits of layer D, they
determined that the overall composition of the ejecta
Figure 10.3 Askja 1875 ash fall dispersion map with is 94% white, rhyolitic pumice and ash that contain
associated thickness of deposition. [Modified from Carey less than 0.5% crystals by volume. Additionally, they
et al. [2010] and Smithsonian Institute, 2019; design credit observed that the pumice is “unusually vesicular”
Nathan Mennen.] with an average density of 0.44 g cm−3.

Table 10.2 Askja 1875 column height, dense rock equivalency (DRE), and tephra volume as compared to eruptive hours
Event Deposit Column height (km) DRE (km3) Tephra volume (km3) Eruptive time (h)
Askja 1875 Unit B Subplinian 8 0.004 0.014 1
Askja 1875 Unit C Phreatoplinian 22.8 0.104 0.45 1
Askja 1875 Unit D Plinian 26 0.213 1.37 6
Note. Modified from Carey et al. [2010]; design credit Nathan Mennen.
DRE, dense rock equivalent.
Volcanics of the North and Northeastern Regions  93

The eruption‐style and product of the Askja 1875


eruption is often compared to the rhyolitic events that
are also occasionally seen at Hekla. A major difference
in these volcanic systems, however, is that Askja is
within the divergent plate boundary of the NVZ and
Hekla is “off‐rift” (Figure 2.1). Schattel et al. [2014]
considered the Askja rift rhyolites to have originated
through extensive assimilation of high‐temperature
hydrothermally altered crust at shallow depths
(≥1.8 km). Additionally, the authors explain that these
rhyolites are hot (935–1008°C), relatively dry
(H2O <2.7 wt%), and oxidized, which is unusual
because explosive volcanism commonly seen in island‐
arc systems is associated with water‐rich pre‐eruptive
magmas, where the presence of water drives the pro-
cess [Schattel et al., 2014].

10.1.2. The Askja 1961 Event and Today

Prior to the 1961 event at Askja, several small‐scale


basaltic eruptions occurred between 1921 and 1929
near Ö skjuvatn, with an erupted volume of about
0.03 km3; including a fissure eruption 6.0 km long
opening up south of the Dyngjufjöll mountains
[Sigvaldason et al., 1992] and the creation of a small
volcanic island in Lake Ö skjuvatn in 1926.
The last major activity at Askja, however, began on
6 October 1961 with seismicity followed by the
formation of three new solfataras on the eastern cal-
dera wall discharging boiling mud at a rate of 30 L s–1
[Thórarinsson and Sigvaldason, 1962]. Narrow fissures
(5–15 cm) orientated north–south were also observed
in this area. According to Thorarinsson and
Sigvaldason, who were the two geologists at the scene,
the eruption began when an east–west fissure (0.7 km
long) opened up on 26th October. Within the first
10 h, they reported that this effusive, Hawaiian‐style
eruption had spewed ~500 m tall lava fountains and
the lava flows reached a distance of 7.5 km over an
area of approximately 6 km2. The total area of the lava
flow by the end of the 6‐week episode was mapped at
11 km2. Additionally, the viscosity of the magma
changed during the event, whereby an ore‐rich aa lava
was produced first. By 5th November a second phase
of lava containing more phenocrysts lead to the
­production of pahoehoe features [Thorarinsson and
Sigvaldason, 1962]. This event produced about 0.1 km3
Figure 10.4 Schematic lithological log of volcanic prod- basaltic lava and 0.004 km3 tephra [IMO, 2019] and
ucts from the 1874–1875 eruption of Askja. [Modified from caused subsidence in the main caldera interior along
Sparks et al. [1981]; design credit Nathan Mennen.] its ring fractures. Earthquake swarms reported on
30 January 1962 and 12 June 1962 indicated magmatic
intrusions.
94 VOLCANICS

Prior to this eruption, processes of inflation and create a tsunami‐like wave. The wave is modeled to
deflation at Askja were not monitored, nor were have that traveled about 3 km in 1–2 minutes across the
changes in lake‐level at Ö skjuvatn. This is particularly lake and inundated the shore with a vertical runup
important because the dormancy of this volcano can measuring up to 80 m [Gylfadottír et al., 2017]. Eye‐
be monitored by changes in its shape. As described by witness accounts suggest a plume of light‐colored ash
Newhall and Dzurisin [1988], a volcano at rest is com- erupted into the atmosphere, but its origin was not
monly linked to periods of inflation and deflation, determined. Small‐scale earthquakes are common
which help to determine its state of repose. Moreover, along the northeast and northwest borders of the
it has been determined that the rate of ground defor- Askja central v­ olcano, as well as in its southeast corner
mation at Askja is higher than anywhere else in [Hjartadöttir et al., 2009].
Iceland—with exception of those systems that are cur-
rently active [Tryggvason, 1989a]. The processes of 10.2. KRAFLA
deflation at Askja are also of interest because amongst
the recorded texture of all historic lava flows some The surface area of the Krafla caldera (65.715,
have been aphyric, which suggests that magma has not −16.728) extends over 64 km2 (Photo 30). This shield
pooled for any lengthy period of time before eruption volcano (818 m a.s.l.), which is capped by a caldera
[Oskarsson et al., 1982; Dalfsen et al., 2005]. This is 10 km wide, is situated within the 100-km-long Krafla
unusual because as the magma goes through the pro- fissure swarm (Figures 5.2 and 7.2). The volcano began
cess of degassing, it both cools and contracts, forming to form ~100,000 years ago with a rhyoltic eruption,
a crystallized magma [Sigmundsson et al., 1997]. which likely initiated the collapse of a 45 km2 caldera
Tryggvason initiated several forms of geodetic mon- that has since been infilled with hyaloclastite rocks
itoring at Askja in 1966 [Tryggvason, 1989b], with and lavas making it a subdued topographic feature.
subsequent studies by Camitz et al. [1995], Sturkell and Subsequently, a dome at Hágöng (65.7322, −16.6689)
Sigmundsson [2000] and Dalfsen et al. [2005]. Initial developed, and a 2.5 km3 welded air‐fall tuff layer was
deflation measurements were reported at Askja bet- deposited [Tuffen and Castro, 2009]. The detailed stra-
ween 1966–1967, 1968–1970, and 1983–1989 that were tigraphy from this layer of mixed intermediate compo-
linked to the withdrawal of magma below the caldera, sition 50 m thick (estimated DRE volume 2.4 km3)
and opposite happening during inflation. From 1983 indicates changes in eruption activity from Plinian to
to 1989 Tryggvason [1989b] measured a subsidence Stromolian [Calderone et al., 1990]. Over half this tuff
of 8 cm year−1 at the caldera center, or a volume of layer has been described as rhyolitic, which has been
about 3 × 106 m3 year−1. Camitz et al. [1995] reported hypothesized to have been generated through near‐sol-
continued subsidence of 11 ± 2.5 cm year−1 during idus fractionation of melted hydrothermally altered
1990–1993, concluding that this deflation was in basalts, rather than by extreme fractionation of pri-
agreement with decreases in pressure at 2.8 km depth. mary basaltic magma [Jónasson, 1994]. Moreover, pet-
This is in agreement with the deflation measurements rological analysis of the erupted rock suggests mixing
from 1991 to 1998 (0.7 cm year−1) by Sturkell and of melts over a range of crustal depths, including the
Sigmundsson [2000] and the microgravity measure- uppermost mantle [Maclennan, 2008; Sigmarsson et al.,
ments by Dalfsen et al. [2005], which suggested subsur- 2008]. More specifically, Winpenny and Maclennan
face mass decreases of 1.6 × 1011 kg between 1988 and [2014] described olivine‐tholeiitic rocks as being
2003. Dalfsen et al. [2005] also concluded that mag- derived from mantle‐source melts, while it is thought
matic eruption at Askja has not occurred since 1961 that rhyolites of the area originate from a crustal
because extension along the rift zone at ~1.97 cm year−1 magma chamber [Grönvold, 1976]. The second eruptive
[DeMets et al., 2010] is providing space in the ductile phase of Krafla (~100–70 ka) produced three subgla-
crust that can accommodate any magma drainage cial rhyolite ridges 300 m tall near the caldera rim
from the shallow magma chamber. This is confirmed (Gaesafjallarani, Jörundur, and Hliðarfjall). The last
by a dike intrusion occurring southeast of Askja near phase that occurred ~24 ka resulted in another rhyo-
Upptyppingar (65.0375, −16.2447) from April 2007 litic eruption (<0.05 km3) to form an obsidian
to March 2008 with an approximate volume of ridge, called Hrafntinnuhryggur (65.7036, −16.7131).
47 × 106 m3 [Hooper et al., 2011]. Hrafntinnuhryggur is particularly interesting because
According to Gylfadottír et al. [2017], the most Tuffen and Castro [2009] describe it as the first docu-
recent activity at Askja was a rockslide that fell from mented example of a rhyolitic fissure eruption beneath
its talus slopes into Lake Öskjuvatn on 21 July 2014 to firn (>55 m thick); this study started a discussion
Volcanics of the North and Northeastern Regions  95

Photo 30 Landscape view of Krafla. [Arctic images/Alamy Stock Photo.]

about the contrasting effects that variable ice thick- Table 10.3 Eight rifting episodes within the Krafla fissure
nesses can have on rhyolitic eruptions. swarm
During the Holocene, Krafla has experienced 35
Lava flows Number Date
eruptive events, mostly basaltic, that have occurred
either in the caldera or on the mountain Namafjall Krafla rifiting episode 1 1975–1984 CE
(65.6392, −16.8197), which have been confirmed Mývatn rifting episode 2 1724–1729 CE
Younger Laxá lava flow 3 182 BCE
by tephrochronology and/or historic observation
Grænavatnsbruni lava flow 4 182 BCE
[Björnsson et al., 1977, 1979; Sæmundsson, 1991; Hólseldar lava flow 5 332 BCE
Smithsonian Institute, 2019]. The total eruptive produc- Hverfjallseldar lava flow 6 782 BCE
tion during and after caldera formation is estimated at Stóraviti lava flow 7 1724 BCE
0.42 km3 kyr−1 [Ármansson et al., 1987]. Additionally, Older Laxá lava flow 8 1782 BCE
there have been eight rifting episodes (Table 10.3) along
the Krafla fissure swarm (Figure 5.2) that are separated Note. Design credit Nathan Mennen.
by long periods of dormancy. In a comparative study
of the six tholeiitic rifting episodes, Hjartadóttir et al. most notable rifting events that have occurred over the
[2012] found that most occur within 10–30 km of the past 1140 years include the Mývatn Fires (1724–1729)
central volcano and that the fissures tend to be unevenly and the instrumentally recorded volcanic‐tectonic epi-
distributed between the northern and southern por- sode, Krafla Fires (1975–1984) [Sigurðsson and Sparks,
tions of the Krafla fissure swarm. They suggest that 1978], which are discussed separately below.
this may be the result of an additional deep magma Perhaps what makes Krafla most iconic is the 2011
source south of the central volcano, or that the TFZ discovery of rhyoltic magma in two shallow (2.1 km
(Figure 2.1) may aid in the reduction of those fissures deep) exploratory geothermal wells (K‐39 and the
found in the north [Hjartadóttir et al., 2012]. The two IDDP‐1) near the central caldera [Elders et al., 2011].
96 VOLCANICS

This created an exclusive opportunity to investigate apart from the crustal movement close to the caldera
an in situ magma flow of silica‐rich composition and lateral flow of magma away from the caldera
(76.5% SiO2) in an otherwise basaltic environment. [Einarsson and Brandsdóttir, 1980]. In 1980, two large‐
It was reported that the recovered cuttings were scale eruptions took place, the first (10–18 July; 1 in
quenched sparsely phyric, vesicle‐poor glass with the Figure 10.5) at Gjástykki (65.8439, −16.7517), a large
in situ temperature of the melt between 850 and 920°C downfaulted graben with walls up to 30 m tall. A
[Zierenberg et al., 2013]. It had been previously known swarm of fissures spanning 4 km long and 500 m wide
that Iceland is an extreme example of having depleted emerged at Gjástykki over the next 8 days to produce a
levels of δ18O in its overall volcanic products, espe- 6 km2 lava field known as Snagahraun. The second event
cially when compared with other ocean island basalts, (18–23 October 1980; 4 in Figure 10.5), also north of
although it was unclear why [Harmon and Hoefs, Leirhnjúkur, produced a lava field double the size of
1995]. Here, a rhyolitic magma with a low level of δ18O Snagahraun from a fissure 7 km long. The eruptions
could be studied under specific pressure and tempera- of 30 January to 4 February 1981 and 18–23 November
ture and without the degassing of volatile contents 1981 (2 and 5 in Figure 10.5) partly covered older Krafla
that occurs during a volcanic eruption. The investiga- lava fields and totalled 18 km2 (Photo 31). The final
tion by Elders et al. [2011] concluded that assimilation eruption on 4 September 1984 (3 in Figure 10.5) inter-
of the partially melted, hydrothermally altered, rupted a dormancy of 3 years. Overall, the total area
basalts was the dominant process in forming this rhy- of the nine Krafla Fire events (Table 5.1) is 36 km2.
olite. This was later confirmed by a detailed petrology Additionally, geodetic measurements by Tryggvason
analysis by Zierenberg et al. [2013], who added that [1994] indicate that a segment of the Krafla fissure
the melt began crystallizing plagioclase, augite, swarm 80 km long widened up to 9 m at the northern
pigeonite, and titanomagnetite at temperatures near margin of the Krafla caldera.
1000°C during rise of the melt to its emplacement Continuous monitoring of the Krafla Fires provided
depth of 2.1 km. These data correspond with findings an insight into the processes of the Krafla fissure swarm
by S‐wave shadowing during the 1975–1984 eruptions and NVZ as a whole. The 3 km deep magma chamber
[Einarsson, 1978] that located the main magma [Tryggvason, 1986; Brandsd6ttir et al., 1997; Sigmundsson
chamber 3–7 km beneath the central caldera. et al., 1997] beneath the central volcano was found to
Additionally, because the exploratory wells provided be the main source of activity during the rifting episode,
a known magma position at depth, they have aided although a second deeper reservoir (>5 km) likely
in the calibration of various geophysical methods, accounted for the surplus of magma that made its way
including seismic tomography [Kim et al., 2017]. to the surface [Árnadóttir et al., 1998]. Twenty dike
injections and nine high‐temperature (1050–1250°C)
10.2.1. The Krafla Fires of 1975–1984 surface eruptions accounted for 0.25 km3 volume of
magma, which occurred as part of an inflation–deflation
After the dramatic display of the Mývatn Fires, the process of chamber infill (Figure 10.6) [Björnsson,
Krafla fissure swarm experienced more than 250 years 1985]. During these 20 cyclic events, the shallow magma
of dormancy until the Krafla Fires started in 1975. This chamber received continuous inflow of magma at
event was instrumentally monitored (e.g., leveling, tilt, ~5 m s−1, causing slow inflation that was followed by
gravity measurements) and provided the opportunity intense deflation during rifting episodes that often
for comparative studies between crustal deformation resulted in caldera subsidence of up to 2 m [Björnsson
episodes. Some would argue that the Krafla Fires are et al., 1979; Einarsson and Brandsdóttir, 1980; Buck
the best‐studied example of dike‐intrusions on a divergent et al., 2006; Heimisson et al., 2015]. The timing of the
plate boundary to date [Buck et al., 2006]. The Krafla deflation periods is inconsistent, lasting several hours
Fires began on 20 December 1975 when increasing up to 3 months. It has been determined through
seismicity was recorded near the caldera. A fissure 3 km petrology that there is not enough time in periods of
long had reactivated near the hyaloclastite ridge, deflation lasting <2 h for significant crystallization
Leirhnjúkur (593 m a.s.l., 65.7000, −16.8000). The first and fractionation to occur during the ascent [Grönvold,
eruption lasted only 12 h, but marked the beginning of 1984]. Deflation episodes were accompanied by
nearly a decade of activity along the Krafla fissure swarm. ­earthquake activity monitored up to 10 km in depth
The Krafla Fires had begun (Table 5.1, Figure 10.5). [Tryggvasson, 1984], as was the lateral movement of
Two small‐scale eruptions occurred north of dike injections along the fissure surfaces. Generally, this
Leirhnjúkur in 1977, followed by 3 years of quiescence made earthquake monitoring difficult during the Krafla
Volcanics of the North and Northeastern Regions  97

Figure 10.5 Area of Krafla Fires in proximity to Lake Mývatn. [Design credit Nathan Mennen.]

Fires [Heimisson et al., 2015], however, the largest diver- the caldera [Sturkell et al., 2008]. Postrift spreading on
gent plate boundary earthquakes (M = 5) were recorded the Krafla fissure swarm was in excess of three times the
during caldera subsidence [Einarsson, 1986]. average before returning to normal, which was not fore-
Following the final eruption phase in 1984, the Krafla seen [Foulger et al., 1992; Heki et al., 1993]. Addi­tionally,
magma chamber again inflated, returning to its previous after the Krafla Fires a significant decrease was observed
elevation. From 1989 to 1992 the land surface nearest the in earthquake activity along the TFZ (Figure 2.1) until
caldera experienced subsidence at a rate of 5 cm year−1, 2000 [Maccaferri et al., 2013]. The TFZ had historically
and then by 2.4 cm year−1 from 1992 to 1995 [Sturkell experienced several earthquakes of M = 6.0–7.0, but as
et al., 2008; Ali et al., 2014]. Subsidence is typical of a result of the Krafla Fires this transform fault was
postrifting episodes, and is likely the result of continued thought to have formed a rift‐induced stress shadow,
plate spreading and crustal cooling, as well as in which temporarily modified the way seismic energy in
response to geothermal energy production located near the area was released [Maccaferri et al., 2013].
98 VOLCANICS

Photo 31 The 1984 Krafla lava field on top of older, grass‐covered, lava flows. [Courtesy of Tamie Jovanelly.]

Figure 10.6 Inflation and deflation model of the geothermal area at Leirbotnar during the Krafla Fires
1975–1984. [Adapted from Ármannsson et al. [1987] and Björnsson [1985]; design credit Nathan Mennen.]

10.2.2. Krafla and Geothermal Energy of supporting geothermal energy for the northeast of
Iceland. Early exploration consisted of ­geological map-
Krafla is among the most studied central volcanoes in ping, aeromagnetic survey, resistivity measurements,
the NVZ due to its usefulness in providing geothermal and chemical analysis of fumarole fluids [Stefánsson,
heat from the shallow (2–3 km deep) magma chamber. 1980, 1981; Ármansson et al., 1987].
The surface presence of fumaroles and venting steam There are two main geothermal fields in the Krafla
first indicated that the 15 km2 caldera may be capable volcanic zone: one inside the Krafla caldera at
Volcanics of the North and Northeastern Regions  99

Leirbotnar (65.7031, −16.7750), and the other, to 2200 m is vapor‐dominated with temperatures above
Námafjall (65.6392, −16.8197), in a high temperature 300°C. This zoning aligns with changes in subsurface
area ~9 km south; Suðurhídar and Hvíhólar are other geology. The shallow zone is characterized by hyaloclas-
notable geothermal fields near Leirbotnar. The geo- tite formations separated by 100–200 m thick basaltic
thermal area of Leirbotnar is situated on the Daleldar lava deposits. Below, granophyric intrusions make up
crater row, which was formed ~1000 ka and is among 80–100% of the rocks and are connected to deep aquifer
the youngest eruptive products in the area. Drilling for systems where the pH 2–4 is too acidic for the fluid to be
geothermal energy began in the early 1950s with the suitable for use [Nielsen et al., 2000]. This zoning may
exploration of the Bjarnarflag field (Námafjall region). best illustrate how the area can have extraordinarily dif-
From 1963 to 1970 other exploratory wells for sulfur ferent thermodynamic and chemical properties, which
and diatomite were drilled in nearby areas but were remain in flux due to changes in magmatic movement.
mostly unsuccessful [Ólafsson et al., 2013]. As the need Overall, the Krafla geothermal fields have experi-
for energy increased in the northeast of the island dur- enced stable production during the past 30 years of
ing the 1970s, so did the protests against the expansion relative volcanic dormancy. The microseisimicity of
of hydroelectric power plants, particularly those on the Krafla’s high‐temperature geothermal fields was moni-
River Skjálfandafljót, which feeds the spectacular tored closely from 2009 to 2012 by Schuler et al. [2016],
Goðafoss waterfall. In 1974, however, the Icelandic who found, as expected, low‐level movement, where
Parliament passed legislation that allowed the building faults and fissures enable the flow of hot geothermal
of geothermal power plants up to the capacity of fluids to the surface. Although seismicity in the area
55 MWe [Nielsen et al., 2000]. Soon after, the National has not been an issue in geothermal energy production
Power Company (Landsvirkjun) began drilling explor- since 1984, Schuler et al. [2016] found that changes
atory wells and purchased two turbine generators in seismicity near the geothermal plants coincided
[Stefánsson, 1981]. The limits on the permit capacity with the intensity of energy production. In addition to
increased in subsequent years, as the need for electricity seismic monitoring, readjustment of the ground sur-
continued to grow with tourism. face in the area has also been the focus of research.
The Krafla geothermal station near Leirbotnar Modeling by Drouin et al. [2017a] confirmed significant
was the first large‐scale commercial power plant in surface deformation due to geothermal processes in
Iceland, with 33 boreholes and a capability to generate Leirbotnar and Námafjall areas. They noted that if the
500 GWe of electricity annually. Construction for the volume of geothermal fluid extracted exceeds the
power plant began in 1974, but was interrupted by rechargeability of the reservoir then the reservoir might
major volcanic eruptions occurring 2 km north when become depleted. Drouin et al. [2017a] observed that
the Krafla Fires began. This disrupted the building since 2004 about 5–8 mm year−1 of subsidence has
schedule, but nonetheless the power plant officially occurred in Leirbotnar for a net extraction rate of
opened in 1978 with 12 completed production wells. 6–9 Mtons year−1 of geothermal fluid. They deter-
As the wells at Leirbotnar became unusable due to mined that subsidence at Leirbotnar is higher than at
increased acidity in the water, other fields, namely Hellisheid̄i (see Box 3.1), which subsides about 1 mm
Suðurídar and Hvíhólar, provided the necessary per year. Subsidence of the area is commonly linked to
stream capacity [Ármansson et al., 1987]. Additionally, geothermal energy production as large amounts of
several of the initial wells in the Leirbotnar field began water are extracted during the process [Sturkell et al.,
to fail. The power plant did not operate at full capacity 2008]. The extraction rate at the Krafla geothermal
until 1984 when volcanic activity subsided, as volcanic power plant doubled after 1995 with the installation
gases severely affected the quality of the geothermal of the second turbine, which uses an average of
fluids necessary for production [Nielsen et al., 2000]. 9 Mtons year−1 [Drouin et al., 2017a]. Another geo-
Undoubtedly, geothermal energy production located thermal power plant is currently under construction in
in a highly active volcanic zone presents hazards and Þeistareykir, 20 km northwest of the Krafla caldera
complex challenges for energy extraction not seen at that has a design capacity between 90 and 200 MWe.
other geothermal fields throughout the world [Palmason
et al., 1983]. As an example, the Leirbotnar reservoir 10.3. VOLCANIC FEATURES AT MÝVATN
consists of two separate geothermal zones [Nielsen
et al., 2000; Langella et al., 2017]. The shallow zone Some of the initial activity in the Krafla volcanic
(200–1200 m depth) is a water‐dominated area that zone began shortly after the Last Ice Age during the
hosts temperatures between 190°–220°C; that extending Icelandic Lúdent Period (9000–5500 BCE). A major
100 VOLCANICS

event of this period included a phreatomagmatic Volcanic activity of ~3800 ka began to alter the
eruption that created the Lúdent tuff ring (~6000 BCE). landscape of Mývatn significantly. At this time the
This was followed by other volcanic eruptions at shield volcano Ketildyngja (65.4311, −16.6444), 25 km
Námafjall (65.6394, −16.8206) and Kröfluháls (65.8003, southeast of Mývatn on the Fremrinámur fissure belt
−16.6039). Likely, the Rauðhólar craters (65.9503, (Figure 5.2), produced a lava flow referred to as the
−16.5336) also formed at the end of the Last Ice Age. Older Laxá Lava (OLL; Thórarinsson, 1953]. The OLL
The Mývatn vent system is approximately 6 km long flowed down gradient through the Laxárdalur valley to
and is adjacent to a narrow canyon along the River the Arctic Ocean (50 km) emerging on to the Adaldalur
Jökulsá á Fjöllum and forms part of the Fremrinámur coastal plain. Thórarinsson [1953] determined the age
fissure belt (Figure 5.2). The erosive processes of gla- and size of this flow through detailed tephrochronol-
cial outburst flooding from this river system have ogy, specifically finding the light colored rhyolitic tephra
provided spectacular examples of the more resistant layer H3, a characteristic deposit from Hekla, below the
plugs (1–3 m thick) and dikes (5–8 m thick) exposed in OLL. At the time, he suggested that this was the longest
the central and southern parts of the Rauðhólar row lava flow ever created by a postglacial shield volcano,
of cone craters [Friese et al., 2013]. The most common with an approximate minimum value of 330 km2, or
morphology associated with this cone system is the 4 km3. This lava flow created a dam that initiated the
clear distinction between colonnade and entablature first Lake Mývatn basin, which had a size similar to that
cooling patterns established by the presence of water today. Additionally, an eruption at Jarðbaðshólar pro-
(Figure 7.7, Photo 32) [Friese et al., 2013]. The exposed duced a lava field between Reykjahlið and Vogar.
shallow plumbing system described by Friese et al. A second major volcanic cycle, sometimes referred
[2013] shows that magma pathways through the to as the Hverfjall Fires, began ~2500 ka along the
volcanic edifice are very complex, with incremental, Kafla fissure swarm. During this time, a hydromagmatic
repeated intrusions. eruption lasting no more than 2–3 days produced the

Photo 32 Cross‐sectional view of the Raud ̄hólar plug and dike lavas. [Courtesy of Tamie Jovanelly.]
Volcanics of the North and Northeastern Regions  101

Photo 33 Hverfjall crater. [Courtesy of Tamie Jovanelly.]

Hverfjall (65.6086, −16.8717) (Photo 33), tuff ring, closer to the magma supply (i.e., Krafla caldera), the
which is 1 km in diameter and 140 m deep. Mapping of eruption of Hverjall had already caused a reduction in
the tephra produced from Hverfjall indicated that the the overall supply (Figure 10.7).
distal extent of the airfall deposits was limited (< 20 km Around 2300 ka an eruption south of the Kafla fis-
from the vent) due to the constant presence of water sure swarm created the Þrengslaborgir–Lúdentsborgir
[Sæmundsson, 1991; Mattsson and Höskuldsson, 2011]. row of cone craters (65.5458, −16.8422) and produced
The Ludent tuff‐ring is considered to have formed the Younger Laxá Lava (YLL). The YLL is considered
under similar conditions. Lui et al. [2017], however, to represent the largest postglacial lava flow within
have determined an absence of lake sediments beneath the NVZ [Thórarinsson, 1979]: producing a volume of
these hydromagmatic deposits, which raises the question 3–4 km3 over a 200‐km2 area [Guðmundsson, 1996].
of whether proto‐Lake Mývatn extended as far as the During this phase, the YLL entered Lake Mývatn
Hverfjall tuff ring. They suggest that the more likely thereby extending it northwards to form the northern
water source for the hydromagmatic eruption was an basin [Liu et al., 2017] and to further develop the basin
influx of groundwater—Lake Mývatn today is supplied shape first created during the ~3800 ka event. The
entirely by the underground system with current dis- sequence of these events is important in consideration
charge rates of ~30 m3 s−1. Simultaneously to the of whether the proto‐Lake Mývatn extended as far as
Hverfjall event, activitivity at another cone complex, the Hverfjall tuff‐ring, thus supplying the water source
Jarðbaðshólar, began. Mattsson and Höskuldsson [2011] for the phreatomagmatic eruption [Liu et al., 2017].
linked these events from evaluation of sedimentary The YLL, however, is best known for creating the
structures present in measured stratigraphic profiles abundant pseudocraters (>1000) that are a prominent
and determined significant change in activity at the part of the Mývatn landscape, which range in size
Hverfjall vent when Jarðbaðshólar became active. Liu from 2 to 30 m in diameter, and from 1 to 15 m tall
et al. [2017] explained that although Jarðbaðshólar was [Thórarinsson, 1953]. In general, they are composed of
102 VOLCANICS

Figure 10.7 Location of Lake Mývatn in relationship to Krafla. Numbers indicate highways. [Design credit
Nathan Mennen.]

lapilli and scoria; xenoliths of glacial substratum, et al. [2016] describe these features as 1–1.5 km in
basalt bombs, and jagged fragmentation are occasion- diameter, elliptical in shape, ~30 m in height, and each
ally found in this area. Specifically, the Skutustaðagigar with a 500‐m‐wide summit depression capable of stor-
(65.5719, −16.9647) Craters, an Icelandic National ing as much as 3 × 106 m3 of lava. They attribute the
Monument, were formed on the perimeter of the lake upright and tilted pillars seen at Dimmuborgir as the
when the YLL first encountered the new wetland envi- result of collapsed sidewalls of the summit depression
ronment (Photo 34). Additionally, the Þrengslaborgir on top of the shield during drainage of the lava pond,
crater rows (65.5458, −16.8442) and Hamarshólar as seen at Kı ̄lauea in 2008. Similar features are found
(65.5336, −17.2506) formed during this time. The at Höfði (65.5789°N, 16.9558°W), but these pillars
Dimmuborgir pillars (65.5956, −16.9200), are consid- remain standing in the lake water.
ered to have formed during the YLL eruption, but this The last event prior to the Mývatn Fires was the
has recently been questioned [Skelton et al., 2016]. Dalseldar Fires (~900 CE), when the southern section
Previously, it was thought that the YLL pooled sub- of the Kafla fissure swarm became active to produce
aqueously and the pillars formed as water rose through 0.2 km3 of lava.
the molten lava as it cooled. When the pool eventually
drained the pillar formations remained. Recent LiDAR 10.3.1. Mývatn Fires of 1724–1729
and ground‐penetrating radar analysis completed by
Skelton et al. [2016] suggests that the pillars instead The Mývatn Fires are part of an eruptive fissure
represent at least two rootless shield‐like volcanic sequence that occurred between 1724 and 1729. The
structures that were fed by lava tubes along the 3 km most detailed contemporary accounts of the episode
distance of the Lúdentarborgir crater row. Skelton were recorded by Pastor Jón Sæmundsson, a local
Volcanics of the North and Northeastern Regions  103

Photo 34 Skutustad ̄agigar Craters, an Icelandic National Monument, Mývatn. [Arctic images/Alamy
Stock Photo.]

minister in Reykjahlíð (65.6428°, 16.9101°), and the been carried southwards from the crater; the deposit
description below relies heavily upon a summary consisted of country rock, ash, and scoria that ranged
provided by Grönvold [1984]. in composition from basaltic to acidic [Grönvold, 1984].
The events began on 17May 1724 when an explosive Notably, Lake Mýtan experienced drastic water‐level
eruption producing an ash column occurred on the fluctuations for the next 9 months as a result of seismic
western flanks of Krafla; the eruption has been esti- activity and vertical ground movements. A section of
mated as a VEI = 2 [Smithsonian Institute, 2019]. This the lake, known as the Ytriflói Basin, completely dried
2‐day event formed the 350 m diameter Storá‐Viti up during the initial event [Ólafsson, 1979]. In January
crater 1.5 km east of Leirhnjukur (Figure 10.7; not to of 1725, eruptions along the fissure swarms south of
be confused with the popular geothermal spot, Lake Krafla triggered a large number of intense earthquakes
Viti, near Askja), along with 14 smaller explosion cra- (it is not known how many or how large they were)
ters and activation of a fissure swarm south of Krafla. that damaged houses and notably increased geo-
Despite accounts of a heavy tephra air‐fall that morn- thermal activity in the area. Additionally, the acidity
ing, only a thin tephra layer has been identified in soil of the lake increased to the detriment of the local trout
profiles near Mývatn [Thórarinsson, 1979], and none populations [Ólafsson, 1979]. Although no activity is
have been identified from sediment cores collected recorded for 1726 or the beginning of 1727, it has been
from Lake Mýtan [Ólafsson, 1979]. The products from assumed that rifting continued in the northern part of
these explosions, however, were considered to have the fissure swarm, which was located away from local
104 VOLCANICS

communities. The next major phase of the Mývatn northern shores of Lake Mývatn (i.e., Ytriflói Basin),
Fires began on 21 August 1727 when four lava erup- again forcing farmers to evacuate. The shoreline grew
tions, of varying durations, started. The first event cre- 2.4 km because of this lava flow [Ólafsson, 1979]. This
ated a discontinuous row of craters near Leirhnjukur eruption continued until the end of September 1729.
(65.7219, −16.7878), Hrossadalur (65.2611°, 20.7928°), Over the span of 5 years, about 0.45 km3 of basaltic
and Bjarnarflag (65.6333, −16.8675); there is little lava that erupted from a discontinuous fissure 11 km
indication of how long the eruption lasted. The second long covered a 30 km2 area [Grönvold, 1984]. On 10
event followed a strong earthquake swarm occurring July 1746 a powerful but short duration eruption
on 18 April 1728, when the fissure at Leirhnjukur was recorded at Leirhnjtikur. Since then, the area of
(Figure 10.7) activated. Four hours after the initial Mývatn has been dormant.
activity began, two eruptions occurred simultaneously: In the absence of scientifically based knowledge on
one at Hrossadalur and the other at Bjarnarflag the succession of the Mývatn Fires, geologists have
(Figure 10.7). On 20 April the valley of Hlidardalur relied on evidence gathered from the more recent,
was filled with lava. Activity lulled until 18 December instrumentally monitored, Krafla Fire event (1975–
1728 when a lava channel extended towards Reykjahlið 1984) in order to make comparisons. They are able to
(65.6428, −16.9101), a small town on the northern part make these assessments with some confidence as the
of Lake Mývatn (Figure 10.7). This lava channel, now lavas that erupted from the 11 km long fissure during
referred to as the Eldhraun lava field, covered previous the Krafla Fires almost exactly coincides with the
flows, but may be best known for sparing a church on Mývatn Fires; the Krafla Fires extended 7 km further
a hill as the lava flow went around the premises (Photo to the north [Grönvold, 1984]. A petrological study by
35). The final eruptive phase again reactivated fissures Grönvold [1984] analyzed the mineral composition of
near the Leirhnjukur craters on 30 June 1729. Lava rocks produced by both episodes. He concluded that
from this event traveled 15 km south to meet the the range of compositions found in the Krafla Fires

Photo 35 The church at Reykjahlid ̄ where the lava flows from the Mývatn Fires spared the structure by
­flowing around the building. [Courtesy of Timothy Wooley.]
Volcanics of the North and Northeastern Regions  105

exceeded those from the previous eruption. Moreover, 10.4. TINNÁ CENTRAL VOLCANO AND OTHER
he found geographical differences in the fissure com- EXTINCT VOLCANOES IN THE REGION
position of the Krafla Fire event between the northern
and southern sections. Specifically, the samples ana- Well before activity at Askja or Krafla, northern
lyzed from the southern section more resembled those Iceland was a location of significant volcanism during
produced by the Mývatn Fires (i.e., basaltic); those the Tertiary as evidenced by the extinct structures found
from the northern section had a less evolved chemical in the landscape today (Figure 7.2) [Jóhannesson and
composition as compared to the Mývatn Fires and Sæmundsson, 1998]. The Neogene (5–6 Ma) Tinná
were similar to postglacial lavas erupted from single (210 km3; 65.2500, −18.8333) central volcano is located
crater events (i.e., phreatomagmatic eruptions including in the Skagafjörður Valley, (Figure 10.8). The Tinná
rhyolitic pumice and basaltic scoria) [Sæmundsson, central volcano was likely active for between 0.5
1991; Jónasson, 1994]. Perhaps most importantly, and 1 Myr according to the principles outlined by
Grönvold [1984] estimated that the difference in Gudmundsson [2000]. From various Ar/Ar dating out-
magma temperature between the most and least lined by Hjarartason [2003], the volcano was active for
evolved was only ~30°C at the time of the eruption, at least three normal polarity subchrons throughout the
indicating the close proximity of sources for all late Miocene and beginning of the Pliocene. A ­caldera
magmas involved. collapse occurred during the final stages of volcanic

Figure 10.8 Simplified geology of the Tinná central volcano. The domain of the volcano includes all the acid
and intermediate lavas related to it. The approximate limits of the Skati Dome and the proposed alignment
of the caldera are based on Hjartason [2003]. [Design credit Nathan Mennen.]
106 VOLCANICS

activity before it was buried under younger piles of lava considered not to have changed much since the time of
as it drifted west away from the active NVZ. the eruption [Hjartarson, 2003].
The activity at Tinná produced four rhyolitic lava The Skati tephra is just one of 11 units in the
domes and is thought to have originated on‐rift in the succession of the Tinná Group that together describe
NVZ. The largest of the four domes (80 km2) is Skati changes in volcanism of the area over time. The units
(500 m tall) and it has been referred to as the most range in thickness from 18 to 300 m and compositions
­voluminous monogenetic rhyolite formation in Iceland include lignite, tholeiite, andesite, and rhyolite. Within
defined in terms of both lava (8 km3) and tephra depo- the succession dike intensity becomes greater with
sition (18 km3) that were likely deposited simulta- increasing depth; occupying 10–20% of rock volume at
neously [Hjartarson, 2003]. Although generally rare, depth. Individual dikes that penetrate lava piles range
Hjartarson [2003] describes the Skati dome to be an in thickness between 1 and 5 m, although they have
order of magnitude larger than any other found in been mapped up to 40 m in thickness. Additionally,
Iceland and argues that the size may have been deter- dike swarms have not been found in the Skagafjörður
mined by a shallow magma chamber thus allowing acid Valley [Hjartarson, 2003].
lavas to easily access the surface; although the absence Other notable Neogene features include several
of a caldera and high‐temperature alteration may dis- extinct volcanoes located in the area between
pute this claim. The initial Plinian eruption was fol- Borgarfjörður Eystri (65.5475, −13.7366) and
lowed by intermediate volcanism and later accumulation Loðmundarfjöður (65.3883, −13.9536), including
of tholeiitic lava piles, which primarily flowed to the Dyrfjöll, Breiðavík, Kækjuskörð, Herfell, and Álftavík‐
east for over 8 km into the central Tinnárdalur Valley Seydisfjörður. At Dyrfjöll (65.5247, −13.9575) Berg
and to the west 7 km toward Mount Hofsfjall. et al. [2014] determined at least two eruptive centers
Perhaps most significant is the contribution that the and that the volcano had produced at least five erup-
Skati tephra has made as a marker bed in deep‐sea tive phases, including two that were predominately
cores, and hence to the Ocean Drilling Project (ODP). silica‐rich explosive eruptions. These large eruptions
The ~5.2 Ma Skati tephra has been recovered in 100% lead to caldera collapse and the production of large
of the cores downwind from Iceland along the north volumes of ignimbrites. The Neogene volcanoes
and northeast coasts. This discrete tephra layer contains became dormant over the course of the next 20 Myr,
75% SiO2 and has reversed magnetic polarity [Lacasse as they moved away from the NVZ axial rift system.
and Garbe‐Schönberg, 2001]. Moreover, from further Subsequently, these extinct central cones became
investigation of the dispersal patterns, thickness, and buried beneath plateau basalts from fissure eruptions
grain‐size analysis they concur that the Skati tephra was in the area, which can now be seen as basalt flats across
derived from one of the largest explosive eruptions that the northern landscape. The burial metamorphism
ever occurred in the Neogene rift zones of Iceland. The that resulted from lava piles in the NVZ produced zeo-
Skati tephra was found in ODP core 907, located 550 km lite zones that were first traced throughout the deposits
from the Tinná volcano in the Arctic Ocean, a distance by Walker [1960].
11
Volcanics of the Western Region

Historically, the volcanism on the Snæfellsnes to rift, as evidenced by locations such as Þingvellir (see
Peninsula in western Iceland has always been of sections 3.4 and 8.5). Sigurðsson [1970] suggests that
interest, featuring anomalous patterns unseen on other the stored energy of the north was then displaced,
parts of the island. First, the Snæfellsnes syncline rep- which resulted in the transcurrent faulting (sometimes
resents an extinct volcanic zone that was active at least referred to as the Þingvellir–Langjökull zone) of the
from 16 to 6 Ma (Figure 6.1) [Sæmundsson, 1974; Snæfellsnes Peninsula westwardly to the Icelandic
Jóhannesson, 1980]. The majority of the Holocene Shelf (Figure 11.2) where a dense concentration (>8)
volcanic rocks now exposed at the surface rest uncon- of extinct offshore central volcanoes (>6 Ma) have
formably on top of Matuyama–Brunhes transition been mapped by magnetic and gravity reflection data
Pleistocene basalts (~840 ka produced by this once [Kristjansson, 1979]; essentially, the SVB was a tempo-
active rift zone [Sæmundsson, 1979]. All that otherwise rary continuation of the MIB (Figures 2.1 and 11.2).
remains from this period of paleovolcanism are the To date, what remains unexplained by this hypothesis
centers of three volcanoes, Setberg I (64.0630, is the dissipation of activity at the eastern end of the
−21.9272), Ellidi (64.8780, −22.9980), and Hrappsey SVB, along with the question of dormancy versus
(65.1191, −22.5997) further to the north (Figure 11.1). extinction of the SVB.
Volcanic activity on this rift ceased after 6 Ma, With the onset of the Plio‐Pleistocene, three central
although the significant Froda‐Breiðavík (64.8652, volcanoes were formed along the east–west orientated
−23.6541; Figure 11.1) intruded between 3.8 and 4 Ma. SVB: Snæfellsnes, Hellgrindur, and Ljósufjöll (Figures
The next activity on the Snæfellsnes Peninsula was 7.2 and 11.1). Little is known about the deep structure
evolution of the modern SVB (Figure 2.1), which is of the Snæfellsnes Peninsula, although computer mod-
~120 km in length and extends from Nordurárdalur eling suggests that the crustal thickness in the area is
(64.9777, −23.8969) to Borgarfjórdur (64.3004, the thinnest on Iceland (15–20 km) [Allen et al., 2002a;
−22.0032). Although still debatable, it is likely that the Du et al., 2002] and that the estimated depth to the
SVB was initiated from differential spreading rates Mohorovičić discontinuity beneath the peninsula is
between the northern and southern portions of the 20–24 km [Darbyshire et al., 2000].
island during the Early Pleistocene (~2.6 Ma) Overall contribution of SVB volcanism is minimal
[Sigurðsson, 1970]. More specifically, during this time during historic times (0.5%) compared with other
the EVZ was connected to the WVZ via the SISZ and volcanic zones in Iceland [Thordarson and Larsen,
the MIB transform boundaries in the south, thereby 2007]. The estimated total volume of erupted magma
creating the Hepprar microplate. Simultaneously, the for the SVB is 6.7 km3: composed of 0.2 km3 lava and
NVZ was moving away from the KR to create the TFZ 6.5 km3 tephra. Each of the three volcanoes on the
(Figure 2.1). By the late Pleistocene, movement along peninsula, however, is known to have been active
the NVZ ceased, while the southern portion continued during the Holocene, depositing both alkaline and
­

Iceland: Tectonics,Volcanics, and Glacial Features, Geophysical Monograph 247, First Edition. Tamie J. Jovanelly.
© 2020 American Geophysical Union. Published 2020 by John Wiley & Sons, Inc.

107
Figure 11.1 Paleo central volcanoes (Setberg I, Ellidi, and Hrappsey) of the Snæfellsnes Peninsula, the
Snæfellsnes syncline axis, and the Snæfellsnes, Helgríndur (Lýsuskard), and Ljosufjöll volcanic systems. The
Froda‐Breiðavík intrusion is found between Snæfellsjökull and Helgríndur. [Modified from Hardarson and
Fitton [1991]; design credit Nathan Mennen.]

Figure 11.2 The Langjökull and Þingvellir transform faults are important to the features and activity on the
Snæfellsnes Peninsula. [Modified from Sigurdsson [1970]; design credit Nate Mennen.]
Volcanics of the Western Region 109

tholeiitic basalts on the surface (Figure 7.1) that con- 11.1. THE SNÆFELLSJÖKULL VOLCANIC SYSTEM:
tribute significantly to the area’s tephrochronology A JOURNEY TO THE CENTER OF EARTH
record [Jóhannesson et al., 1981]. Specifically, the dom-
inant rhyolitic volcanism found on the peninsula can The Snæfellsjökull volcano (64.9777, −23.8969) has
be best attributed to at least 10 Holocene eruptions at played a central role in the Pleistocene volcanic history
both the Snæfellsjökull and Ljósufjöll volcanoes of the SVB, which is 30 km in length and covers the
[Smithsonian Institute, 2019]. Although rifting epi- majority of the western peninsula (Figures 6.1 and 11.1).
sodes are not found on the Snæfellsnes Peninsula, Pleistocene deposits (hyaloclastites) and interglacial
three distinct, parallel bands of ~60 craters have been lava flows (≤ 800 m) that tend to show striations, thus
mapped (Figure 11.3). The highest concentration of distinguishing them from postglacial deposits
craters is found in the extreme west dating to the last [Harðarson, 1993], form the easternmost part of the
glacial period, indicating reactivation of the early volcanic system. The ability to delineate strata based
Pleistocene volcanotectonic line (see transect C–C’, on glacial scouring is particularly important to the
Figure 11.3) [Sigurðsson, 1970]. geology of this area, as there is evidence for at least 20

Figure 11.3 Geology of the Snæfellsnes Peninsula highlighting three main areas of volcanic zonation (A‐A’,
Ljosufjöll; B‐B’, Helgrindur; C‐C’, Ljosufjöll) that mimic the parallel relationship of paleocraters. [Modified
from Sigurdsson [1970]; design credit Nathan Mennen.]
110 VOLCANICS

glacial–interglacial episodes in the Icelandic record age and includes hyaloclastites, however, its structure
over the past 3.1 Myr and on this peninsula evidence also incorporates pillow lavas and a lava cap.
for 10 glaciations have been observed [Jóhannesson Snæfellsjökull (1446 m) is located on the western-
et al., 1981]. most tip of the peninsula and may be known best for
Common to the SVB are the numerous subglacially its fictional description by Jules Verne in the 1864
formed tindars and tuyas (Figure 7.5) that are found novel Journey to the Centre of the Earth in which the
around Snæfellsjökull dating to ~60,000 BCE, or the central vent is thought to lead scientists into the
last glaciation. Notable examples include: Sandkulur underworld. According to Thordarson and Larsen
(844 m; 64.8300, −23.7047), Geldingafell (824 m; [2007], Snæfellsjökull (Photo 36) is one of only five
64.8461, −23.7488), and Ólafsvíkurenni (519 m; volcanoes in Iceland that fit the classification of stra-
64.8580, −23.6722). Temporally equivalent hyaloclas- tovolcano (470 km2) as it has a summit crater (200 m
tite deposits containing palagonite and columnar deep) and is lined with a lava dome along the rim.
basalts indicating submarine or subglacial eruptions The horseshoe shape of the crater may indicate an
occur on the southern coast, from Londrangar to asymmetrical explosive eruption leading off the
Breiðavík (Figures 7.6 and 11.1) [Harðarson, 1993]. In northern flank with characteristics similar to the
contrast to these mafic effusive structures, Pleistocene 1980 Mount St Helens eruption [Jóhannesson et al.,
rhyolitic (ranging from acidic to intermediate) deposits 1981]. Seismic tomography by Obermann et al. [2016]
also can be found in the SVB. As described by found three anomalies that reveal the internal struc-
Jóhnnesson et al. [1981], Maelifell (64.7444, −23.7241) ture of the volcano: a solidified magma chamber at
is a rhyolitic dome of mixed composition that likely between 3.3 and 5.5 km depth, a gabbro intrusion
formed during the last glaciation. Skridan (64.8169, at 2.5 km depth, and a shallow magmatic reservoir at
−23.9058) is similar to Maelifell as it is of equivalent 1.5 km depth.

Photo 36 The stratovolcano Snæfellsjökull. [Courtesy of Tamie Jovanelly.]


Volcanics of the Western Region 111

This 700 ka volcano [Hoernle et al., 2009] is also Harðarson [1993] established that the majority of
sometimes referred to as a stratovolcano–tuya hybrid Snæfellsjökull rock samples had a phenocryst content
as it likely formed under deep deposits of glacial ice of <20% by volume, which were analyzed in order to
(Figures 7.3 and 7.5), which would explain its com- illustrate complex morphologies and zonal patterns
plex volcanic history [Hards et al., 2000; Evans et al., used to provide evidence of disequilibrium conditions
2016] (Figure 11.4). For example, Snæfellsjökull has resulting from magma mixing. Interestingly, Martin
had ~20 explosive and effusive eruptions that have and Sigmarsson [2007] found that both basic and acid
­produced sizable pyroclastic craters on its flanks (e.g., magmas from Snæfellsjökull have identical Sr and Nd
Holatindar, 64.7947, −23.8425), as well as mildly alka- isotope compositions, which could indicate that they
lic Holocene lava flows that extend over the entire were derived from a common source (i.e., fractional
western side of the volcano and peninsula [Smithsonian crystallization of magma). Three of the last 10
Institute, 2019). It has been estimated that the evolu- Holocene eruptions are classified as Plinian style (VEI
tion of complex magma compositions took ~190 kyr estimated to be >4.0), producing trachytic tephra
and more than 300 kyr before trachybasalts (i.e., layers and peralkalic rhyolites similar to those in the
hawaiite) were erupted from the volcano [Harðarson, EVZ at Torfajökull [Hards et al., 2000; Jóhannesson
1993]. The hawaiite lava flow (referred to as SNB31 in et al., 1982] (Table 7.1). Comparison of the unique
tephrochronology), which indicates the earliest evolu- chemical composition of the deposits found in the
tion of magma in the SVB, can be seen just east of SVB with those in the EVZ has enabled an assessment
Olafsvík in Yti Bugur (64.8738, −23.6427). of silica magma genesis as a function of geographic
location. This comparison further illustrates that
fractional crystallization is linked to off‐rift locations
with lower geothermal gradients (40–60°C km−1) and
to crustal melt processes at on‐rift locations [Martin
and Sigmarsson, 2007].
Three important postglacial rhyolitic tephra deposits
from Snæsfellsjökull have radiocarbon ages of 200 BP,
2010 BP, and 6050 BP, and are often referred to as the
Snæfellsnes Peninsula marker beds, SN‐1, SN‐2, and
SN‐3, respectively [Steinthorsson, 1967; Jóhannesson
et al., 1981]. From tephra isopach maps Jóhannesson
et al. [1981] interpret that the primary axis of deposition
for these three events is orientated to the northeast of
the volcano. Additionally, they found that the thickness
of the deposits reduces with distance from >5 m thick at
the volcano base to 1 cm thick 45 km away. Based on
radiocarbon dating and geochemical analysis, the SN‐2
tephra layer has been identified in central Sweden and is
thought to correlate with a deposit that locally is referred
to as LBA‐2 [Wastegard et al., 2009]. The thickest of
these tephra deposits along the northeast margins have
been mined during the 20th century for building
construction additives [Evans et al., 2016].
The SN‐1 layer is the most investigated unit of the
three as it documents the last eruption occurring at
Snæfellsjökull, which deposited 0.11 km3 of volcanic
material. Investigations of jökulhlaup events associ-
ated with the 200 BP event revealed them to be
relatively small in comparison with the historic floods
of Katla (Figure 9.6) and Grímsvötn [Smith and
Figure 11.4 Geology and chronostratigraphy of Snæfell­ Sigmundsson, 2010]. Importantly, from a hazard miti-
sjökull. [Modified from Evans et al. [2016]; design credit gation point of view, they found that the more exten-
Nathan Mennen.] sive floods flowed to the north and east out of the
112 VOLCANICS

main outlet glaciers that drain from the caldera. Also volcanic zone (i.e., greater values in the west) to be the
important for hazard mitigation and to evacuation result of greater mantle partial melting closer to the
planning was their recognition of fluvially transported SVB in the east.
pumice deposits on the southern and western slopes, While smaller in both height (1063 m) and area
which occurred post‐eruption when rain‐triggered (720 km2) than Snæfellsjökull, Ljósufjöll has provided
lahars remobilized the unconsolidated airfall pumice. the most recent volcanic activity on the peninsula,
The last volcanic eruption on the peninsula occurred dating to 960 CE [Fuchs et al., 2013]. Flude et al. [2008]
1750 years ago, which is relevant to questions suggest that this volcano represents the largest outcrop
concerning the repose of the volcano and the SVB. (by volume) of Quaternary silica material on the penin-
Currently, the peninsula is not seismically monitored, sula. Through field observation and analysis of geo-
nor do historic records suggest seismicity beyond a chemical signatures, 13 silica units representing separate
low‐level (M <3.0) intraplate earthquake event that eruptive events have been identified in the succession
occurred at the easternmost part of the SVB in 1974 near Ljósufjöll. Flude et al. [2010] describe the geo-
[Einarsson et al., 1977]. An extensive study by Fuchs chemistry of these eruptions further by identifying
et al. [2013], however, determined that the southern them as six peralkaline rhyolites (displaying strong
and northeast flanks have measureable microseismic depletions in Ba, Sr, and Eu), six trachytes, and one
activity (Ml ≤1.0, where ML is is local magnitude) at alkalic rhyolite. The geochemistry of these off‐rift
depths ranging 8–15 km. They suggest that the seis- ­volcanics has important relevance in the comparison
micity is associated with fluid‐induced fracturing in with similar deposits found for the Torfajökull central
the shallow, ductile crust, related to one or more volcano in the SISZ (Figure 2.1) [Martin and Sigmarsson,
­magmatic reservoirs. It has been postulated that the 2007]. In their study on the petrogenesis of the rhyolites
seismicity of the peninsula will escalate with increasing at Ljósufjöll, Martin and Sigmarsson [2007] determined
global temperatures that induce glacial melt. Bakker that ~95% of the samples formed via fractional
et al. [2014] used Snæfellsjökull to demonstrate this crystallization from a parent basaltic magma. Also of
hypothesis as the glacier has been measured by
­ interest, the lava samples from Ljósufjöll contained lit-
Jóhannesson et al. [2011] to be melting at tle water despite their subglacial origin. These findings
1.25 m.w.e. year−1, where w.e. is water equivalent. Their can be compared with the geologic maps of the area
study indicates that it is easier for shallow magmatic by Jóhannesson and Sæmundsson [1995] that show that
fluids to upwell during times of accelerated glacial the volcano is surrounded by Pleistocene subglacial to
retreat as the mantle surface is devoid of stresses and is interglacial lava fields (ranging from basaltic to
undergoing isostatic release of pressure. Moreover, intermediate) that were rapidly deposited over the
Harðarson and Fitton [1991] argue that melting of the course of 800–100 ka. Additionally, some of the largest
glacial ice on Snæfellsjökull causes a reduction of of these deposits can be found in two valleys, Hitardalur
pressure in the mantle sufficient to affect magma com- (64.7786, −22.0650) and Hnappadalur (64.8927,
position. This phenomeon may help to answer why −22.2488), and they indicate an eruption period 3000–
there are regional changes in chemistry along the pen- 2000 ka [Einarsson, 1970]. Bersekjahraun (64.9694,
insula whereby rocks become more undersaturated −22.9320), Horgsholtshraun (64.9041, −22.6611), and
from west to east along the SVB [Harðarson, 1993]. Gullborgarhraun (64.8844, −22.2558) are significant
craters associated with the Ljósufjöll volcanic zone.
11.2. LJÓSUFJÖLL
11.3. HELGRINDUR
Another active volcano on the Snaefellsnes Peninsula
is Ljósufjöll (64.9313, −22.5658; Figure 11.1) of the The Lýsuskard volcanic system is the smallest of
Ljósufjöll volcanic zone, which is approximately 90 km the three systems depicted in Figure 11.1 and
in length and 10–20 km wide. In contrast to the SVB, spans 30 km from Tröllatindar (64.8925, −22.9858)
there is a fault that runs from Ljósufjöll to a crater to Búlandshöfdi (64.9516, −23.4405). The most
called Grábrók (64.8233, −21.5516). Movement along significant feature in the Lýsukard system is the vol-
the fault seems to coincide with volcanic eruptions, cano Helgrindur (647 m tall, area 220 km2), which
although the eruptions are less silica rich in composi- sometimes is referred to as Lýsukard or Lysuhóll.
tion as compared with Snæfellsjökull [Jóhannesson, The flanks of this volcano provide one of the only
1982; Sigurddson, 1970]. Harðarson [1993] considers places on the Snaefellsnes Peninsula where at least 10
the variability of alkalinity along the Ljósufjöll glacial cycles can be identified [Jóhannesson, 1982].
Volcanics of the Western Region 113

Figure 11.5 Cross‐section and plan view of Vatnshellir Cave system. [Modified from Stefánsson [2010];
design credit Nathan Mennen.]

Craters associated with the Lýsuskard volcanic Characteristically, the floors of the cave are lined with
system have produced prominent Holocene lava pahoehoe flow structures and the lowermost parts of
fields, including Bláfeldarhraun (64.8211, −23.1702) the walls are lined with benches marking the height of
and Hraunsmulahraun (64.8225, −23.0772). the last lava flow.
Additionally, the Grábók Lava Field (64.8022, Members of the conservation committee of the
−21.5608) is one of the only examples of aa lava in Icelandic Speleological Society have extensively sur-
the Snaefellsnes Peninsula due to the low‐viscosity veyed the cave for the purpose of providing both a safe
magma generated by the event. access for the visiting public and tourists and a means
of conserving the cave from further destruction
11.4. VATNSHELLIR CAVE [Stefánsson, 2010]. The middle floor is 12–23 m below
the surface and about 100 m long, and there is a 12 m
The Vatnshellir Cave (64.7483, −23.8177) is a 205 m vertical lava fall that leads to the lowermost level,
long lava system in the SVB that has linear tubes found about 35 m below the surface. The cave, one of the
on three levels (Figure 11.5). These linear tubes were jewels of the Snæfelljsökull National Park, is run by a
formed by the slow movement of low‐viscosity magma local tourist firm (Summitguides) and is now a popular
that began to form an exterior crust, thus insulating public and tourist destination on the Snæfellsnes
the warm magma inside, preventing it from solidifying. Peninsula.
Part III: Glacial Features
12
Overview of Glacial Features in Iceland

12.1. GLACIAL SETTING synchronous over the whole island. At 2.9 Ma the
mean air temperature is estimated to have been at 0°C
As evidenced by the deep fjords, jagged mountain [Símonarson, 1980] and the first regional glacier expan-
peaks, tabletop mountains, and U‐shaped valleys sion occurred ~2.6 Ma [Geirsdóttir and Eiríksson,
found in Iceland, the alpine landscape has largely been 1994]. It took at least 200 kyr thereafter for ice to
carved by expansion and retreat of glaciers over extend across Iceland from the north towards the east
geologic time. Although much is known about the and west and, lastly, to the south [Geirsdóttir and
island’s past glaciations, the Miocene through Eiríksson, 1994]. In line with this idea of continued ice‐
Pleistocene terrestrial record in Iceland is considered sheet growth is the appearance of glacial deposits
to be the most continuous and detailed in the Northern found everywhere on the island that date to 2.4 Ma.
Hemisphere [Geirsdóttir and Eiríksson, 1996]. The first From these records it has been established that there
tillite can be found in Lower Pliocene (5–3.5 Ma) have been at least 15–23 glaciations in Iceland during
deposits [Símonarson, 1980]. This compares with the past 3 Myr [Einarsson and Albertsson, 1988]. Since
paleomagnetic stratigraphic records of bracketing lava between 2.2 and 2.0 Ma, effusive eruptions are notable
flows, and the K–AR dating of sediments, that deter- during interglacial phases, and phases of glacial expan-
mined that glaciation began before the normal polarity sion are represented across the island as tabletop
Gauss Chron (~5 Ma) when a small icecap formed mountains and pillow lavas.
near the center of Iceland [Geirsdóttir and Eiríksson, The offshore record confirms onland glacial expan-
1994] (Figure 2.2). This deterioration in climate in the sion, with a dramatic increase in the abundance of ice‐
Late Pliocene prompted a switch in flora from decid- rafted debris (IRD) deposits being identified in North
uous to boreal forest regimes [Einarsson, 1994]. Atlantic deep‐sea cores [Shackleton et al, 1984; Raymo
Evidence of this can be found from fossil leaf imprints et al., 1989]. The record of initial local onset is fol-
of Dryas octopetala, Betula nana and Empetrum lowed by evidence for onset of global ice expansion at
nigrum [Friedrich, 1966; Rungred and Ingólfsson, 1999] about 2.6 Ma and establishment of glacial–interglacial
(Photo 37). Similarly, as indicated by the Sleggjulækur cycles (Figure 2.2; Box 12.1) [Eiríksson and Geirsdóttir,
flora in western Iceland and the Pliocene Tjörnes flora 1991]. Closer to the shore the eroded glacial products
(Box 2.1) in northern Iceland, a later Late Pliocene are preserved as cyclothems with sequences indicating
transition occurred where Betula and Salix shrubs and transgressive or regressive sea‐level changes [Einarsson
grasses became dominant as the forests declined and Albertsson, 1988].
[Einarsson, 1994; Rungred and Ingólfsson, 1999]. By the Last Glacial Maximum (LGM; ~26 ka) the
Although there is evidence for local mountain ice Weichselian ice sheet grew to cover the entire island
caps of Miocene age in the southeast [Geirsdóttir and (Figure 2.2; Figure 12.1). As evidenced by end
Eiríksson, 1994], it is considered that the onset of this (terminal) moraines deposited offshore, during the
extensive glacial period was gradual rather than LGM the ice sheet expanded towards the shelf slope

Iceland: Tectonics,Volcanics, and Glacial Features, Geophysical Monograph 247, First Edition. Tamie J. Jovanelly.
© 2020 American Geophysical Union. Published 2020 by John Wiley & Sons, Inc.

117
118 GLACIAL FEATURES

Photo 37 Small birch trees in Iceland at present are similar to the boreal forest vegetation seen during the
Early Pleistocene climatic transition. [Tim Graham/Alamy Stock Photo.]

break, ~130 km off the coastline of the western fjords reducing the ice sheet to about 25% of the LGM size by
at Breiðafjörður (65.2500, −23.2500) [Olafsdóttir, 1975; 13.8 ka [Pétursson et al., 2015]. The influence of glacio-
Ingólfsson and Norðdahl, 1994; Andrews et al., 2000]. isostatic adjustment (GIA) is also seen through the
From trimlines on tabletop mountains formed during transgressive shorelines found at 70 m a.s.l. in the
the Saalian (penultimate) glaciation (Figure 12.1), southwest and 90 m a.s.l. in the northwest [Pétursson
Einarsson and Albertsson [1988] extrapolated the et al., 2015]. Oxygen isotope ratios ­measured in fossil
maximum altitude of the ice‐sheet surface during this marine sediments found along the North Atlantic coast-
period to be between 2000 and 2500 m; a value Hubbard line from the Bölling–Alleröd Interstadial suggest that
et al. [2006] later confirmed by modeling. Additionally, sea‐surface temperatures were similar to present
Hubbard et al. [2006] estimated the ice‐sheet surface (Box 12.2). Ruddiman and MacIntyre [1981] proposed
area covering Iceland during the LGM to be that final deglaciation was initiated by a drastic
330,000 km2 or 30 times larger than today. change that occurred when the marine polar front
Starting at about 15 ka, the large and continuous ice entered the North Atlantic. Specifically, low winter
sheet covering Iceland began a step‐wise retreat to its insolation and the formation of low‐salinity meltwater
current size, as indicated from radiocarbon dating caused the density‐driven ocean current system of the
of both glaciomarine deposits and lake sediments North Atlantic to slow, which resulted in a moisture
[Ingólfsson and Norðdahl, 1994; Jennings et al., 2000] starvation of the ice sheet and subsequent disintegra-
(Figure 12.2). The climate in Iceland during the inter- tion [Ruddiman and MacIntyre, 1981]. This late
stadials following the LGM was probably similar to Pleistocene deglaciation is linked to a 30‐fold increase in
present, or 5–10°C warmer than the stadials [Símonarson, volcanism for about 2 kyr [Maclennan et al., 2002].
1979]. The influence of the Bølling–Allerød Interstadial Interrupting the overall melting trend, however, are
(~14,700 ka to ~12,700 ka) in Iceland was significant; some well‐documented ice‐sheet readvances that have
Box 12.1 The Use of Oxygen Isotopes as Climate Change Indicators
Emily Larrimore
Elucidating details about past climatic changes gram of sediment (IRD gsed−1) is used, utilizing siev-
often requires the use of climate proxies. One such ing to remove coarse material from the sample in
proxy is ice rafted debris (IRD), or sediment trans- order to reduce noise in the data [Fronval and
ported within icebergs that become entrained in Jansen, 1996]. Using this method enables small
ocean circulation patterns [Phleger, 1949]. The ice- pulses of IRD to be determined and the results to
bergs eventually degrade, depositing the sediment be statistically significant. A high concentration of
on the ocean floor. This sediment often has a IRD is an indicator that icebergs, sea ice, or both
specific mineralogical signature and grain size that were present at a time interval indicated by the
allows its source material to be identified. Although depositional layer of the sediment [Phleger, 1949].
the IRD usually has a finer grain size distribution During the Quaternary, it is estimated that 70% of
comparable to glacial till, coarser material is an the total IRD deposited globally was deposited in the
indicator that the source was likely a glacial subpolar North Atlantic, south of Iceland [Ruddiman,
moraine rather than a sandur plain [Grobe, 1987]. 1977]. In terms of volume of IRD, it is estimated that
Off Iceland, IRD has been used to determine the in the past 3 Ma, approximately 200,000 km3 of
distribution of icebergs into the North Atlantic, unconsolidated sediment has been moved from
the volume of continental erosion caused by ice ­terrestrial locations to the deep Atlantic [Ruddiman,
movement, and the rate of glacial melting 1977]. The major ocean currents responsible for
[Ruddiman, 1977]. The information derived allows transporting icebergs, and subsequently IRD, to the
researchers to draw specific conclusions about the mid‐latitude Atlantic are the Labrador Current and
regional and global climate over both recent and the East Iceland Current [Andrews et al., 2000].
geologic times [Grobe, 1987]. Information from IRD Many ice cores are collected from the North
is especially important regarding Heinrich Events, or Icelandic shelf, an area with high oceanographic
events occurring every 8000–10,000 years, in which sensitivity to climate changes. The location of the
large fluxes of iceberg armadas enter the ocean North Icelandic shelf is between the North Atlantic
circulation, influencing the thermohaline circulation and Polar oceanographic and atmospheric fronts,
and, ultimately, global climate [Heinrich, 1988]. which are very strong fronts that dictate tempera-
During a Heinrich Event, it is possible for enough ture, salinity, and Arctic conditions. Even small
meltwater volume (~2.3 million km3) to be intro- shifts in the input or output of water masses has a
duced to the density‐driven ocean current system to pronounced effect on the system, and this is
cause mixing of warm waters to reduce as evidenced reflected in the sediment found in marine sediment
in the North Atlantic during the Last Glacial cores [Knudsen et al., 2009]. Collecting marine
Maximum [Roche et al., 2004; Obrochta et al., sediment cores from the North Icelandic shelf pro-
2014]. For this reason, Heinrich Events initiate the vides favorable temporal resolution for past climate
onset of extreme cooling. and oceanographic changes because the sedimen-
To analyze IRD, marine sediment cores are col- tation rates of the sedimentary basins located here
lected in the North Atlantic, Antarctic, or Arctic are very high [Knudsen et al., 2009]. Also, the
Ocean, where glaciers have formed. Material North Icelandic shelf has the distinction of being a
eroded by the glacier is transported at the base of sample area within a volcanically active region.
the ice where it remains when icebergs are formed Tephra layers deposited in the sediment allow for
at calving ice margins. The icebergs may become more reliable age determination and avoid issues
entrained in sea ice before being released into the with spatial and temporal increase in the marine
open ocean, but in either state when basal melting reservoir dating caused by incursions of Arctic
occurs debris in the ice descends to the ocean (or water masses [Knudsen et al., 2009]. The presence
lake or fjord) floor to be recognized as IRD in sed- of volcanic and non‐volcanic source material aids
iment cores [Phleger, 1949]. Grain analysis is con- researchers in ascertaining the origin of the IRD
ducted to evaluate the content of IRD influxes. The whether it be Iceland, Greenland, or the Arctic
IRD material is typically rounded and too dense to Ocean [Knudsen et al., 2009]. Furthermore,
have been transported by ocean currents, thus it is because the bedrock of Iceland does not contain
concluded the material must have been deposited sufficient amounts of quartz grains, IRD containing
from melting icebergs [Heinrich, 1988]. Typically, rounded quartz grains are assigned to a different
a measurement of the total minerogenic grains per source region [Eiríksson et al., 2000].
Figure 12.1 Maximum Weichselian ice sheet extent in Iceland as indicated by end moraines mapped
­offshore, and the Álftanes (A) and Búd ̄i (B) readvance stages. The altitudes of Saalian tabletop mountains
have been used to extrapolate glacial ice thickness. [Modified from Einarsson and Albertsson [1988]; design
credit Nathan Mennen.]

Figure 12.2 Distribution and areal extent of Iceland’s present ice caps. [Design credit Nathan Mennen.]
OVERVIEW OF GLACIAL FEATURES IN ICELAND 121

Box 12.2 Oxygen Isotope Fractionation


Emily Larrimore
Observational data regarding temperature and Isotopic analysis can be used to determine past
­climate are extremely limited prior to 1860 CE changes in climate by investigating terrestrial and
[Schöne et al., 2005]. To extrapolate high‐resolu- ocean records. Ice cores are considered a valuable
tion and long‐term data about climate, proxies are way to describe the land surface because as
used to create more accurate and spatiotemporally incremental layers of firn accumulate, atmospheric
complete climate models [Schöne et al., 2005]. To gases are also trapped in the ice, thereby preserving
quantitatively assess past climates, isotope ratios can a sample of the surface air during the time of
be used as a proxy for air temperature, sea‐surface formation [Johnsen et al., 2001]. In addition to using
temperature, and climate. As an example, oxygen the water molecules themselves as proxies for cli-
isotopes are often used for these purposes. The mate change, data for atmospheric carbon dioxide
ratio of O18 to O16 is expressed as δ18O or levels can also be extracted from the ice. Ocean
Delta‐O‐18, and is measured in units of parts per sediment cores containing microfossils with calcium
thousands, denoted as ‰ [Dansgaard, 1964]. carbonate (CaCO3) and siliceous (SiO2) exteriors are
An isotope is two or more elements all with the used to describe changes in the ocean because these
same number of protons, but a different number of organisms are particularly sensitive to sea‐surface
neutrons. While O16 is the most abundant oxygen temperatures and salinity. The value of δ18O precipi-
isotope in nature, two others also exist (O17 and tated by the carbonate microfossils is a function of
O18) [Meyer, 2005]. Individual water molecules can the δ18O in the water when the carbonate shell was
contain different isotopes of oxygen. When a water formed, and the temperature of the water at the time
molecule contains O18 it is described as being of shell formation [Leng and Lewis, 2016]. Isotope
“heavy” because it has a greater molecular weight ratio mass spectrometry, or the process of separating
than O17, for example [Dansgaard, 1964]. These different isotopes based on their mass‐to‐charge
isotopes act as proxies for climate change because ratios, is used to obtain an accurate isotopic
­
they reflect the amount of heat energy available in ­composition of the sample [Leng and Lewis, 2016].
a system. This means that in warmer climates there In Iceland, oxygen isotope ratios were measured
is enough energy to evaporate higher concentra- from ice cores extracted from the Hofsjökull glacier,
tions of the heavier O18 water molecules from the Iceland’s third largest glacier covering 880 km2 in the
ocean. Once evaporated into the atmosphere, the southwest region of the country [Thorsteinsson et al.,
O18 water molecules can then be precipitated as 2004]. In the central area of the summit caldera of
snow onto land thereby creating a “heavy” O18 sig- Hofsjökull the ice is 750 m thick with an estimated
nature in ice cores. In turn, the “lighter” O16 is left age of 1000 years old [Thorsteinsson et al., 2004].
in the water column and can then be incorportated An average δ18O of –12.9‰ was found, signaling a
into the exteriors of the microfossils living within warming climate that corresponds with the settling of
that system [Dansgaard, 1964]. The opposite effect the Víkings [Thorsteinsson et al., 2004]. Furthermore,
occurs to the oxygen isotope ratios when the cli- it is believed that future ice‐core drilling at Hofsjökull
mate is cool. Additionally, an inverse relationship is could yield higher resolution isotopic records to
established when comparing oxygen isotope levels further explain LIA conditions in the North Atlantic
in ice cores to sediment cores. region [Thorsteinsson et al., 2004].

been studied in Iceland, including the Álftanes Stage 1988]. Moraine sequences found in southwest Iceland
(~12 ka correlating with the Younger Dryas) and, (64.1006, −22.0290 and 64.8402, −23.3407, respec-
­following the Bölling–Alleröd Interstadial, the Búði tively) mark readvance limits that define both the
Stage (~9.8–9.6 ka) (Figure 12.1). These readvances Álftanes and Búði Stages [Einarsson, 1968; Ingólfsson,
represent just two of nine (minimum) that occurred 1988; Norðahl et al., 2008]. Hagen [1995] concluded
between the LGM and early Preboreal time [Ingólfsson, from isotopic analysis and planktonic foraminiferal
122 GLACIAL FEATURES

Figure 12.3 The extent of Iceland’s ice sheet during the Younger Dryas Stadial (~12 ka). [Modified from
Pétursson et al. [2015]; design credit Nathan Mennen.]

assemblages that the usually warm Irminger Current 2.5 ka [Einarsson, 1985], leading to the expansion of
that sits off the southwest coast of Iceland likely mountain glaciers and ice caps that persist today
cooled by 6°C during the Búði Stage. [Einarsson and Albertsson, 1988]. Paleoceanographic
The Iceland ice sheet grew significantly during investigations of the North Atlantic using marine
the Younger Dryas Stadial (Figure 12.3). The appearance microfauna assemblages [Andrews and Giraudeau,
of artic molluscs (Portlandia arctica, Buccinum groen- 2003], drift ice [Massé et al., 2008], and ocean currents
landicum) in the Fossvogur sedimentary sequence [Eiríksson et al., 2004] indicate a period of fluctuating
(64.1166, −21.9333) near the end of the Bölling–Alleröd sea‐surface and bottom‐water temperatures with an
Interstadial (~12 ka) illustrates the shifting climate amplitude of ±3°C for the past 2 kyr [Jiang et al., 2005].
[Pétursson et al., 2015]. Ice‐contact deltas and truncated The Víking Ingólfur Arnarson was the first to settle
shorelines attributed to the Younger Dryas Stadial are in Iceland, in 874 CE, but it was likely the onset of the
dated to 12 ka [Ingólfsson et al., 1997]. Ruddiman Medieval Warm Period (~950–1250 CE) that most
and MacIntyre [1981] explains that the Younger Dryas influenced settlement of the island. Approximately
Stadial is likely the consequence of an influx of tabular 3000 Víkings traveled to Iceland in search of new land
icebergs derived from a disintegrating Arctic ice shelf that and resources, and they found a landscape resembling
provided cold‐water pulses to the North Atlantic Ocean. their Norse origins [Bryson and Murray, 1977].
By the onset of the early Holocene, the Younger The Little Ice Age (LIA; 1850–1930 CE) is the most
Dryas ice sheet had retreated to half its size. The periods documented of the cool periods to occur in Iceland.
of intense glacial melt led to rapid crustal rebound bet- The LIA likely arrived in Iceland earlier (~1750–
ween 9 and 8 ka, as testified by the entire coastline of 1800 CE or earlier) [Bradwell et al., 2006], however,
Iceland being higher than today due to the glacioiso- as radiocarbon dates obtained from landscape fea-
static adjustment (GIA) of a low viscosity mantle tures reflect glacial maximums in the 18th and 19th
beneath Iceland [Einarsson and Albertsson, 1988; centuries. In comparison with Europe, however, the
Pétursson et al., 2015]. The highest shorelines mapped data reflect the onset of the LIA in Iceland occurring
along the southern coast are 100 m above present sea about 500 years later, which is perplexing due to
level; 30–60 m on other coastlines [Einarsson and its northern latitude [Gordon and Sharp, 1983].
Albertsson, 1988]. Birch trees also began to colonize at Establishing the LIA glacial maximum on Iceland has
9 ka and persisted until a sudden cold event occurred at not been straightforward. With only two exceptions,
OVERVIEW OF GLACIAL FEATURES IN ICELAND 123

radiocarbon dates from Vatnajökull outlet glaciers 12.2. MODERN CLIMATE OVERVIEW
record a mid‐ to late 19th century date, but inland and
steeper mountain glaciers record a date almost Today, approximately 10% of the Earth’s surface is
100 years earlier [Björnsson, 1979; Bradwell et al., in a solid liquid state forming a region called the cryo-
2006]. Using lichenometry, however, Kirkbride and sphere. The cryosphere includes all forms of frozen
Dugmore [2001] found Eyjafjallajökull ice‐cap water including sea ice, lake ice, snow cover, glaciers,
moraines in southern Iceland dating from the mid‐ ice caps, ice sheets, and permafrost. The land cover
18th century. In addition, even earlier dates (1727– represented by the cryosphere is small and only ~1.7%
1755 CE) have been reported using tephrochronolgy of freshwater frozen within the system represents gla-
along Eyjafjallajökull’s outlet glaciers: Gigjökull and cial ice, but this percentage translates into a water
Steinholtsjökull. Overall, a late 19th century date for volume of 24,064,000 km3, or 68.7% of the global dis-
the LIA glacial maximum in Iceland has been the most tribution of freshwater [Gleick, 1996]. Currently, the
widely accepted [Björnsson, 1979; Gudmundsson, 1997; planet is experiencing an interglacial climate phase
Evans et al., 1999; Bradwell et al., 2006]. when there are small volumes of solid water. Slight
Undoubtedly, the glaciers and small ice caps in changes in average global temperature (±2°C), how-
temperate environments, such as Iceland, are extremely ever, or changes in mid‐latitude insolation, can have a
sensitive indicators of climate change that can seem to large impact on the glacial–interglacial cycle, as docu-
lack the synchrony often seen on larger ice sheets, like mented over both historic and geologic time.
Greenland. As found by Kirkbride and Dugmore [2001] Iceland is uniquely positioned at polar latitudes at a
the lack of temporal agreement found in the LIA records point where both atmospheric and oceanic currents
could reflect unreliable geomorphological preservation interact to support Tertiary and Quaternary glacia-
rather than a reliable chronostratigraphy, and earlier tions, as reflected in the ages of various deposits
cooling may have occurred as the written record suggests. (Figure 2.2). Today, along the southwest coast the air
Additionally, the situation is complicated by the presence temperature on average is about 2°C, but can rise to
of subglacial volcanic centers under Öræfajökull, above 11°C during summer months; overall, colder air
Eyjafjallajökull, and Mýrdalsjökull that could result in temperatures are found in the north. These tempera-
magma generated ice‐sheet advance or retreat, rather tures correspond with subartic and tundra climatic
than a climatic signature of growth or melt. regimes (as found in the central highland area) and are
Regardless of the LIA glacial maximum in Iceland, best illustrated by the low diversity of sparse vegeta-
historic accounts indicate that the cooling period tion found in the landscape (e.g., Festuca spp., Salix
severely affected the inhabitants. A poem published spp., Agrostis spp.). Historically, this has resulted in
pastor Olafur Einarsson (1573–1659) may best describe little opportunity for agricultural development due to
the strife felt by the people during this cool period the infertility of soils beyond the cultivated outwash
[Bryson and Murray, 1977]. plains and paleo ground moraines.
Iceland’s mild maritime climate is due to the warming
Formerly the earth produced all sorts
of the North Atlantic by the Irminger Oceanic Current
of fruit, plants and roots.
(IOC) (5°–12°C), which travels west along the southern
But now almost nothing grows….
Then the floods, the lakes and the blue waves coastline to eventually collide with the East Greenland
Brought abundant fish. Current (EGC) (<0°C) coming off the Arctic Ocean.
But now hardly one can be seen. This collision results in the formation of low‐pressure
The misery increases more. systems that can bring heavy snowfalls (annual precip-
The same applies to other goods…. itation ~500 cm year−1) in mountainous regions, with
Frost and cold torment people the heaviest snowfall occurring in the months of
The good years are rare. February and September. At present, the precipitation
If everything should be put in a verse on the southern Vatnajökull above the snowline at
Only a few take care of the miserable…. 1000 m is 400 cm, but on Mýrdalsjökull the precipita-
Additionally, thorough investigations by Lamb tion is 600 cm. The colder EGC promotes the formation
[2002] indicate population declines from 77,500 in of sea ice that can be found along the northern coast-
1095 to 72,000 in 1311, to 50,000 in 1703, and to 38,000 line, but rarely in the south. In some instances, the
in 1780. He correlates the falling population to lack of EGC can block the movement of the ICO at the
grain harvested and grown on the island due to the northwest peninsula, causing the north and east coasts
onset of colder climates [Lamb, 2002]. to be dominated by Arctic seawater [Stefánsson, 1981;
124 GLACIAL FEATURES

Einarsson and Albertsson, 1988]. As this example illus- with episodes of glacier readvance [Hannesdóttir et al.,
trates, it is the dynamic setting of oceanic and atmo- 2014]. The most notable readvance, in the early to mid‐
spheric polar fronts that makes Iceland an ideal setting 1990s, is thought to have corresponded with a positive
to study fluctuations of the Quaternary glacial–inter- North Atlantic Oscillation index [Bradwell et al., 2006;
glacial cycles. Evans et al., 2017]. As Evans et al. [2017] explain,
Currently, 10% (10,500 km2) of the country is covered Icelandic glaciers are particularly sensitive to interan-
in glacial ice that is predominately in the form of ice nual variations in summer average air temperatures,
caps (Figure 12.2). Björnsson [2017] suggested that the which can quickly induce either accumulation or aba-
amount of glacial ice found on Iceland today repre- lation processes. The pioneering Icelandic glaciologist
sents the amount of total precipitation over the past Helgi Björnsson, who has conducted research for the
20 years, and if the remaining glacial ice on the island University of Iceland since the 1960s, is quoted as
were to melt, then global sea‐level would increase by saying that the modern rate of ice lost in Iceland is
1 cm. The glacial systems on Iceland have been known among the highest on Earth’s surface [Katz, 2013].
to respond rapidly to shifts in climate change [Geirsdóttir Moreover, Björnsson predicts that Iceland could lose
and Eiríksson, 1993], as well as to the influence of geo- 30% of its glacial mass by 2050 if climate change
thermal heat and subglacial eruptions that are initiated effects continue to intensify [Katz, 2013].
by the hot spot currently under Iceland’s largest glacier
(Vatnajökull), which has been in that position for the
12.3. GLACIAL MORPHOLOGY
past 15 Myr. Approximately 60% of the glacial ice on
Iceland covers active volcanoes [Björnsson, 2017]. The 12.3.1. Formation of Ice
melting of the glaciers from such activity hasbeen
known to create surges of glacial flooding, called jökul- Ice is unique in that it can be differentiated from
hlaups, as seen during historic times at Grímsvötn and water by its variances in physical properties and
Katla (Figure 9.6) [Björnsson, 2017]. isotopic composition. As the phase diagram for water
Recent decades, from 1960 to 2010, have been char- illustrates (Figure 12.4), if solid ice is kept at a steady
acterized by slow, but continual, melting intermixed temperature as pressure increases, then it will undergo

Figure 12.4 Water phase diagram. The diagram shows that all three phases of water can exist under condi-
tions at the Earth’s surface. [Modified from Bierman and Montgomery [2014]; design credit Nathan
Mennen.]
OVERVIEW OF GLACIAL FEATURES IN ICELAND 125

a phenomenon called pressure melting (this often cre- in a transitional zone between snow and ice (typically
ates basal sliding at the contact between the base of 20–30 m below surface); having a higher porosity and
the glacier and the bedrock.). Additionally, ice at 0°C lesser density. This is important because as compac-
and 1 atm has a volume of 1.0906 cm3 g−1 and a specific tion (i.e., metamorphism) continues to reduce the
gravity of 0.91689 g cm−3, where both the volume porosity, the ability for air to move freely between the
(1 cm3 g−1) and specific gravity (1 g cm−3) of water is grains begins to decrease, reducing the ability for
slightly different. Moreover, there are eight varieties of exchange with atmospheric air. Thus, during the
crystalline ice that fit the definition of a mineral, transition from firn to ice, atmospheric gases impor-
whereby the forms that exist on our planet occur bet- tant to climate change processes (e.g., carbon dioxide)
ween 0°C and −70°C and at atmospheric pressure up can be trapped between the pore spaces. As compac-
to 2000 atm [Martini et al., 2001]. The isotopic compo- tion occurs over time, the age of the occluded air
sition of water and ice will vary depending on the bubble will be less (ranging from a few hundred to sev-
amounts of hydrogen (H2 or H1) and oxygen (O18, O17, eral thousand years) than the age of the surrounding
O16) present during the evaporation or solidification ice [Center for Ice and Climate, 2019]. Typically,
process, which is often used as a proxy to determine empirical models based on Icelandic densification
paleoclimate (Box 12.1). rates from snow to ice are used to determine the age of
Water, once solidified, has a specific chemical com- the trapped gases.
position that is arranged in an orderly and repetitative
pattern, and is considered a mineral that can go 12.3.2. Glaciers and Associated Landforms
through processes to change its form into a rock. Ice
rock can be shaped by two main processes: the Glaciers found on Iceland are grouped into three
concentration of water via freezing or the compaction main categories: ice sheets, ice caps, and valley glaciers.
of snow from pressure resulting in metamorphism to An ice sheet is an area of glacial ice greater than
produce firn first, then ice (Figure 12.5). Firn is found 50,000 km wide and the last time that there was an ice
sheet covering the island was ~26 ka during the LGM
(Figures 2.2 and 12.3). When an ice sheet recedes, ice
caps (<50,000 km2) are often left behind in moun-
tainous regions. The highest point of an ice cap is
sometimes referred to as the massif or dome, from
which ice flows down gradient (Figure 12.6). The
highest massif in Iceland is Hvannadalshnúkur
­
(64.0158, −16.6747), which forms part of Öræfajökull’s
summit crater. This massif also represents a feature
called a glacial horn, formed when three or more
­glaciers converge on a central point to create a steep,
pyramid‐shaped peak.
Currently, there are six ice‐caps on Iceland
(Figure 12.3) that contain an estimated 3400 km3 of
water: Vatnajökull (64.4220, −16.7902; 8300 km2),
Langjökull (64.6562, −20.1531; 950 km2), Hofsjökull
(64.8167, −18.8167; 923 km2), Mýrdalsjökull (63.6467,
−19.1303; 600 km2), Drangajökull (66.1500, −22.2500;
160 km2), and Eyjafjallajökull (63.6304, −19.6067;
80 km2). Each of these central ice‐caps have several
valley or outlet glaciers, that begin at a high‐elevation
cirque, before it descends down a confined valley
(Photo 38, Figure 12.6). As an example, Vatnajökull
has 11 outlet glaciers that either end at a lake or an
Figure 12.5 Transition of unconsolidated snow pack as it outwash (or sandur) plain (Figure 12.6). Much is
turns into glacial ice. The percentage of air volume is indi- understood about Iceland’s valley glaciers as they have
cated on the right and depth on the left. [Design credit been systematically monitored since 1930, providing
Nathan Mennen.] some ideas of ice‐front fluctuations for 70 years
126 GLACIAL FEATURES

Figure 12.6 Depositional and erosional geomorphologic features as glacial ice retreats towards the massif
or dome. [Design credit Nathan Mennen.]

[Sigurðsson, 1998]. Valley glaciers affect life at a scouring of bedrock from the weight of a glacier can
regional scale, thus influencing the distribution of also have an impact on the bedrock itself by leaving
hydroelectric power, irrigation, road construction, striations in the direction of movement (Photo 41).
and tourism. For example, the reduction in volume of Additionally, some of the bedrock scoured (and other
glaciers in Iceland during the first half of the 20th landscape debris) will become entrained and carried in
century amounted to a reduction in discharge by the ice as long as the glacier is advancing.
20 L s−1 km2, and such a reduction could have pro- Once the glacier begins to melt, the material carried
found effects on long‐term energy generation in the will be deposited either along the sides of the valley
country [Björnsson, 1980]. (lateral moraines) or at the toe (terminal moraine)
As glaciers move down gradient, they create both (Figure 12.6). The elevation of the lateral moraines
erosional and depositional features that remain as presents a proxy to determine an approximate level for
indicators of past climate over geologic time. Common the zone of ablation. Likewise, terminal moraines
to the southern side of Vatnajökull are U‐shaped (Photo 42) are particularly important in determining
valleys (Photo 39) that result from the scouring the position of past glacial advance, but often they are
movement of a confined valley glacier. If the initial destroyed by outwash streams that discharge meltwater
topography is steep, like that of northern coastline of from a glacier. The terminal moraines found at the six
Iceland, then the final result from valley glacier glaciers remaining in Iceland date back to the maximum
movement will be a fjord that will be flooded by ocean advance of the LIA; those dated from 1890 still have ice
water once the glacier has disappeared (Photo 40). The remaining at their centers [Björnsson, 1980]. The size of
OVERVIEW OF GLACIAL FEATURES IN ICELAND 127

Photo 38 Skeid ̄arárjökull valley glacier ending on an outwash plain. [Courtesy of Tamie Jovanelly.]

the terminal moraine will depend on how long the gla- terminate in this area and the heavily laden meltwater
cial ice‐front had been at that position and, ultimately, develops into braided fluvial streams that continu-
how much debris the ice was carrying. The moraine ously change their position across the outwash plains
sediment is called glacial till and is characterized as from one side to the other. Skeiðarársandur (63.7508,
being unconsolidated and poorly sorted with grain −17.5319) is the largest outwash plain (1000 km2) in
sizes ranging from silt to boulders (Photo 43). Iceland, which extends up to 30 km in front of
In addition to the volume of debris calculated from Skeiðarájökull (Photo 44). The sediment load brought
moraine deposits, the erosive capacity of a glacier can to this outwash plain is largely the result of jökul-
be determined using several other means. Specifically, hlaups from the subglacial volcano Grímsvötn (Figure
the rate of erosion of the basaltic regions southeast of 12.2). For example, during a 1934 eruption, the total
Vatnajökull has been quantified using sediment load sediment transport down to Skeiðarársandur was on
of glacial meltwater; the land denudation rate esti- the order of 150 × 106 tons [Björnsson, 1980]. To date,
mated from sediment load transport for Vatnajökull is however, Skeiðarársandur is being eroded at the rate
3.2 mm year−1, but the rate differs in other areas of 8.5 m year−1, which threatens power lines and the
(Table 12.1) [Björnsson, 1980]. Along the southern major road (Golden Circle Road, Highway 1) [Landl
coast the transported sediment load is deposited to et al., 2003]. This is likely due to the formation of pro-
form the broad, low‐lying landscape (average slope glacial lakes that are trapping sediment within their
5 m km−1) of the outwash (sandur) delta plains. As basins, such as those in front of Skeiðarájökull and
Björnsson [1980] explained, wide glacier margins often Breidamerkurjökull.
Photo 39 A U‐shaped valley on the southern side Vatnajökull. [Courtesy of Tamie Jovanelly.]

Photo 40 A glacial fjord on the northern coast of Iceland. [Courtesy of Tamie Jovanelly.]
Photo 41 An example of glacial striations (linear structures) that result from gouges cut into bedrock by
glacial abrasion over time. [Courtesy of Tamie Jovanelly.]

Photo 42 A terminal moraine in front of a glacial snout along the southern coast of Iceland. [Courtesy of
Tamie Jovanelly.]
130 GLACIAL FEATURES

Photo 43 A glacial till deposit next to a glacial lake on the southern coast of Iceland. [Courtesy of Tamie
Jovanelly.]

Table 12.1 Denudation rates estimated for Iceland glaciers the portion of the glacier that is moving plastically.
Typically, the weight of the ice limits the depths of cre-
Glacier Rates of denudation (mm year−1) vasses to <30 m. Crevasses can result from three main
Mýrdalsjökull 4.5 processes: (a) release of stress when two glacial lobes
Vatnajökull 3.2 meet that are moving at different rates; (b) release of
Langjökull 0.4 stress when a glacier is moving over uneven basement
Hofsjökull 0.9 topography; (c) release of stress when a glacier is
Other Glaciers 0.3
expanding to accommodate the width of a valley (i.e.,
Unglaciated areas 0.1
longitudinal crevasses) or the valley is becoming steeper
Note. Modified from Björnsson [1979]. (i.e., transverse crevasses) (Figure 12.7). Additionally,
extensional faulting and heat from magma upwelling
similarly can cause large gaps within glacial walls.
Although understanding of erosional and deposi- Studies of ice fracturing and the formation of cre-
tional processes occurring under the ice provides insight vasses have increased over recent decades with the
into glacier movement, there is also much that can be awareness that these processes can promote ice calving
learned from the exploration of the ice surface itself. at the terminus of glaciers [Mottram and Benn, 2009].
The upper layer undergoes brittle deformation (e.g., Additionally, they may become meltwater paths that
crevasses) as it has not yet become incorporated within can temporarily accelerate flow by providing a conduit
OVERVIEW OF GLACIAL FEATURES IN ICELAND 131

Photo 44 Skeid ̄arásandur, an example of the southern sandur plain coastline. [Courtesy of Tamie Jovanelly.]

Figure 12.7 Types of crevasses that form on glaciers. [Design credit Nathan Mennen.]
132 GLACIAL FEATURES

for basal lubrication [Benn et al., 2009]. Landl et al. snowfall, and ice‐flow rate [Björnsson and Pálsson, 2008;
[2003] explained the impact of calving glaciers by quan- Bradwell et al., 2013]. As described by Einarrson [1994]
tifying the volume lost annually (260 × 106 m3 year−1) at the rate of flow of large valley glaciers in Iceland is
the outlet glacier Breiðamerkurjökull, where it enters about 1 m day−1, although it is largely dependent on
into the Jökulsarlon lagoon (64.0784, −16.2306). glacier thickness, slope, and accumulation area.
The mass balance of glaciers and ice sheets is found
12.3.3. Periglacial Environments by identifying the difference between ice gain
(accumulation) and loss (ablation) over time. Net
Periglacial environments are found at the edge of accumulation occurs above an area referred to as the
glaciers where saturated ground and freezing tempera- snow line, or equilibrium line, during the winter season
tures can support the development of permafrost. (Figure 12.8). Below the snowline, no matter how
Permafrost distribution is primarily controlled by a much snow has been deposited over a season, the gla-
constant mean annual air temperature of around 0°C. cier’s budget in that location will not be net positive.
Data gathered from regional maps by Etzelműller et al. The position of the snow line is dependent on air tem-
[2007] indicate that mountain permafrost found at ele- perature and precipitation, whereby it establishes the
vations 800–900 m a.s.l. covers ~8000 km2 of Iceland’s division between the firn area of the glacier and that of
surface, or about 8% of the country. This does not ablation. Specifically, melting mainly occurs below the
include the permafrost locations described by Priesnitz snowline and becomes more rapid with decreasing ele-
and Schunke [1978], who focused on that found in the vation [Einarsson, 1994]. On average, melting of firn is
central highlands, where they estimated that the eleva- slow, however, if dirt or volcanic ash are intermixed
tion of the permafrost was at a lower elevation, 420– the process can be increased to induce a loss of 1–2 cm
720 m a.s.l., and covered an area of 180 km2. In the annually [Einarsson, 1994]. If the snow line does not
central highlands, geomorphic features related to per- change over time, the glacier is in steady state and its
mafrost development, like palsas, patterned‐ground, mass balance is stable. If the snow line rises, however,
and ice wedges were documented. the glacier will lose mass and retreat or vice versa.
The elevation of the snow line is variable throughout
12.3.4. Glacial Mass Balance Iceland. As noted by Einarrson [1994], on the southern
side of Vatnajökull the snow line is between 1000 and
A glacier is composed of densely packed ice that 1100 m, while the average height found along the
forms when the accumulation of snow exceeds ablation northern side is between 1300 and 1400 m. The highest
processes, namely sublimation and melting over time elevation of the snow line on Vatnajökull is in the north
(Figure 12.8). Empirical studies have shown that at Hyannadalshnúkur (64.0158, −16.6747; 2109 m).
Iceland’s glaciers are particularly sensitive to ablation The higher elevation of the snowline along the northern
processes at the annual to decadal scale [Bradwell et al., edge of Vatnajökull is likely the result of being further
2013]. More specifically, fluctuations in the mass balance away from coastal precipitation systems that can induce
and areal extent are a direct result of the interactions melting during summer months thereby decreasing
between radiative flux, ice accumulation through overall glacier growth. It has been seen that an increase

Figure 12.8 Mass balance of an outlet glacier as it moves down gradient. [Design credit Nathan Mennen.]
OVERVIEW OF GLACIAL FEATURES IN ICELAND 133

in mean annual temperature of 1°C can influence the The rate of uplift is small in comparison to the esti-
snow line in Iceland to retreat 200 m, with the opposite mates of subsidence that occurred throughout the
effect if cooler [Einarrson, 1994]. Pleistocene; a glacier 2000 m thick can produce 600 m
To date, the overall mass balance of Iceland’s glaciers isostatic depression [Biessy et al., 2008].
is negative. From 1995 to 2010, the island’s ice caps lost Undoubtedly, understanding the mass balance of
on average 9.5 ± 1.5 Gt year−1 despite large fluxes in the glacial systems is fundamental to the interpretation of
interannual variability of precipitation [Björnsson climate change. For example, without looking at the
et al., 2013], which is similar to that estimated by system as a whole, a glacial advance or surge in
Gardner et al. [2013] for the period 2003–2009. The response to shear stress induced by gravity could be
largest ice cap, Vatnajökull, is estimated to have lost mistaken as glacial growth or a positive accumulation
435 km3 of ice from 1890–2004 [Pagli et al., 2007], with of ice due to climate. Modeling shows that firn begins
a deglaciation rate of 25.0–82.17 cm year−1; the data for to behave plastically at a depth of 30 km enabling the
Iceland’s other glaciers can be found in Table 12.2. system to creep forward, generally in response to
Additionally, between 1890 and 2010 the elevation of changes in topography at the base of the ice. Basal
the valley glaciers of Vatnajökull have lowered 150– sliding, however, is amplified by warming tempera-
270 m near their termini; negligible downwasting was tures as meltwater enters at the surface via crevasses to
observed above ~1500–1700 m [Hannesdóttir et al., create a positive feedback cycle. This translates to the
2014]. Other valley glaciers (e.g., Heinabergsjökull, glacier base (and hence ice‐front) moving more quickly
Hoffellsjökull, and Lambatungnajökull) have experi- downgradient, thereby resulting in an overall reduction
enced significant downwasting during this same time of ice thickness.
frame, retreating close to 3 km since the LIA [Einarsson, Surge‐type glaciers are characterized by gently slop-
1994; Hannesdóttir et al., 2014] (Figure 12.9). As a ing surfaces (1.6–4°) that move too slowly to remain in
consequence of loss of ice mass, Iceland is undergoing balance with associated accumulation rates [Björnsson
GIA, with uplift rates between 25 and 29 mm year−1 et al., 2003]. (Surges that occur in glaciers steeper than
during the 20th century [Árnadóttir et al., 2009; Auriac 3° typically affect only the low‐slope ablation areas
et al., 2013] and with 0.23 km3 year−1 of new magma [Björnsson et al., 2003].) All major ice caps in Iceland
being produced beneath Iceland [Schmidt et al., 2013]. contain surge‐type outlet glaciers. Since 1625, about 47
surge advances have been recorded at Vatnajökull, rang-
Table 12.2 A comparison of deglaciation of major Iceland ing from 0.2 to 10 km (Figure 12.10) [Thórarinsson,
glaciers 1930 to present day 1969; Björnsson et al., 2003]. The majority of
Vatnajökull’s outlet glaciers, however, experienced
Glacier Area (km2) Volume loss (km2) surges during a glacial retreat period that occurred bet-
Vatnajökull 8428 458 ween 1890 and 1964, which affected ~40% of the ice cap
Hofsjökull 976 76.1 [Thórarinsson, 1969; Björnsson et al., 2003] (Table 12.2,
Langjökull 1024 79.9 Figure 12.10). Thórarinsson [1969] and Björnsson et al.
Eyjafjallajökull 96 7.5 [2003] agree that these surges are likely associated with
Myrdalsjökull 688 53.7
increased basal water lubrication due to warming,
Smaller glaciers 230 17.9
Total 11 442 693.1
rather than being triggered by seismicity. There have
been 80 surges in total recorded in Iceland since 2003
Note. Design credit Nathan Mennen. [Björnsson, 1980]. Through a comparison of 74 years of

Figure 12.9 Cross‐section and longitudinal profiles of Hoffellsjökull indicating glacier thinning related to
retreat. [Modified from Einarsson [1994] and Ahlmann and Thórarinsson [1937]; design credit Nathan
Mennen.]
134 GLACIAL FEATURES

Figure 12.10 Geographical distribution of known surge‐type glaciers within Vatnajökull and their associ-
ated dates (data from Björnsson et al. [2003]). [Design credit Nathan Mennen.]

data, Thórarinsson [1969] concluded that the surging has surged at 30‐year intervals since 1930 (Figure 12.10).
glaciers in Iceland typically have flat ablation zones and Additionally, Dyngjujökull (64.7641, −16.7775) has
shallow concave basins that broaden toward the glacier advanced at 20–30‐year intervals, and Brúarjökull
terminus. Additionally, he noted that stress is not likely (64.7207, −16.1769) every 80–100 years (Figure 12.10).
accumulated along all parts of the glacier, but instead More often, however, it is the case that glacier surges
will build up until it reaches a critical limit before sliding are sporadic. This may be best illustrated by
occurs, which can result in catastrophe. As an example, Breiðamerkurjökull (64.0922, −16.2566), which has
over the course of 1 year (1963–1964) the ice front of surged 11 times between 1794 and 1969, at intervals of
Brúarjökull (64.6817, −16.1567) advanced 8 km with 6–38 years [Björnsson et al., 2003]. Between valley gla-
average velocities measured at 5 m h−1; peak flow rate cier surge events the crevasses that have formed will
was 25 m in 24 h [Einarsson, 1994]. This broad‐lobed close as the glacier retreats until the next phase of for-
outlet glacier also had an episode in 1890 when the toe ward movement [Einarsson, 1994]. Better monitoring
advanced 10 km in a surge and created a push moraine has revealed that velocity builds over the course of
20 m high [Björnsson, 1980]. 2–3 years from initial movement in the upper
In general, it is thought that surge frequency cannot accumulation zone before its energy is perpetuated
be truly considered cyclical and is not related to size or downslope [Sigurðsson, 1998]. It has been determined
mass balance [Björnsson et al., 2003], although historic that the abrupt movement of ice from the accumulation
observations for some glaciers suggest that surging has zone downward ultimately changes the overall geom-
occurred regularly. Siðujökull, (64.1127, −17.8050), a etry of the glacier, creating new challenges for deter-
western lobe on Vatnajökull, is such an example as it mining mass balance.
OVERVIEW OF GLACIAL FEATURES IN ICELAND 135

12.3.5. Glaciation and Volcanics sition throughout Iceland’s neovolcanic zones,


including the NVZ, WVZ, and EVZ [Eason et al.,
Glaciovolcanic deposits (e.g., tuyas, tindars; Figure 2015] (Figure 2.1). According to Maclennan et al.
7.5) of the Late Pleistocene cover about 11,200 km2 of [2002], the average eruption rates during the deglacia-
the surface of Iceland, indicating an abundance of tion period (~12 ka) were 100 times higher than those
intraglacial volcanism on the island [Jakobsson and during both glacial and recent times. In addition, these
Gudmundsson, 2008]. Additionally, Jakobsson and high eruption rates persisted for <1.5 kyr after the
Johnson [2012] estimated the number of Pleistocene deglaciation of each area.
monogenetic volcanoes that started under or within an Increases in volcanic activity can amplify crustal
ice sheet to be in the thousands. Jakobbson and Johnson melting and are sometimes associated with heightened
[2012] described three main eruptive phases associated mantle decompression resulting from glacioisostatic
with intraglacial volcanic events: the formation of unloading [Maclennan et al., 2002; Sinton et al., 2005;
basal pillow lavas in deep‐water; explosive shallow‐ Eason et al., 2015]. Maclennan et al. [2002] considered
water hyaloclastites; and effusive subaerial capping that postglacial bursts in eruption rates could be used
lava flows. They further hypothesized that monoge- to estimate the rate of melt transported in the mantle,
netic systems likely were deposited in lakes that were and accordingly determined melt extraction velocities
melted within the ice sheets. Overall, the chemical of more than 50 m year−1 for the last deglaciation. This
composition of the Late Pleistocene intraglacial increase in the rate of both magma melt and transport
volcanic rocks is tholeiitic (Figure 7.1). results in temporal variations in geochemistry (namely
Deglaciation following the LGM induced unusually lower MgO content), which has been observed at
high volcanic productivity and changes in lava compo- Kraftla and on the Reykjanes Peninsula.
13
Glacial Features of the Reykjanes Peninsula
and Southwestern Region

During the Late Pleistocene from 26 ka through They identified specific geomorphic characteristics
~15 ka, the island of Iceland was entirely covered in that they used to separate the tuyas and tindars of the
glacial ice; this includes the Reykjanes Peninsula and WVZ into those that formed prior to and after the
the WVZ (Figures 2.1 and 2.2). As the ice sheet Younger Dryas readvance.
retreated the Langjökull and Hofsjökull formed rem- Biessy et al. [2008] studied the postglacial rebound
nant ice caps (Figure 12.3) and are still present today. of southwest Iceland since the Younger Dryas in detail
Through Ar–Ar dating of lava flows in the area it has and determined that uplift occured in two major
been determined that volcanic activity on the WVZ phases during the Holocene. First, the vertical uplift
during the Pleistocene was relatively steady, but it was of the Reykjanes Peninsula between 10 and 8.5 ka was
less active than that on the Reykjanes Peninsula determined to be 67.5–157.5 m from global positioning
[Jakobsson and Johnson, 2012]. Uniquely, although system (GPS) profiles, morphological observations of
Langjökull sits on top of the volcanically active WVZ terminal moraines, paleobeaches, glacial striations,
there have not been any postglacial eruptions. The sub- and marine cliff faces. This implies rapid uplift rates of
glacial topography beneath Langjökull reflects its 4.5–10 cm year−1 as a response to the fast‐paced glacial
volcanic heritage (Figure 13.1). melting that occurred as climate warmed. This rise in
A common Pleistocene feature of the WVZ are the ground surface stimulated volcanic activity in the area
>250 basaltic monogenetic volcanoes that cover an as two lava domes formed beneath the Langjökull ice
area of 1350 km2. Although the height of these ­features cap, Leggjabrjótur (1010 m; 64.6922, −19.8213) and
that dominate the landscape are typically between 200 Hagafell (920 m; 64.4986, −20.3469). The second
and 500 m at modern day, it has been estimated they phase starting at 8.5 ka initiated 10–20 m of uplift, but
have lost 40% of their volume due to glacial and post- was more likely the result of tectonic activity occurring
glacial erosive processes. The erosional products have along the WVZ.
been deposited either in the ocean or have contributed
to shoreline expansion along the peninsula [Jakobsson 13.1. LANGJÖKULL
and Johnson, 2012]. Eason et al. [2015], however, points
out that ice‐sheet mechanics can also influence erosion, Today, the Langjökull ice cap (64.6562, −20.1531) is
as those features created nearest the stable interior of the second largest in Iceland with an estimated volume
the ice sheet will erode less than those created towards of 195 km3, and a mean ice thickness of 210 m and
the more mobile margins. Similarly, from detailed maximum of 650 m (Photo 45) [Björnsson and Pálsson,
investigations of tuyas and tinders in the WVZ, 2008]. The elevation of the ice cap ranges from ~1400 m
Jakobsson and Johnson [2012] determined that the at its dome to ~460 m at the base, with the equilibrium
shape of the intraglacial features was dictated by the line currently between 1100 and1200 m. Although the
thickness of the ice sheet at the time of the eruptions. ice cap has continued to shrink in size since the LIA,

Iceland: Tectonics,Volcanics, and Glacial Features, Geophysical Monograph 247, First Edition. Tamie J. Jovanelly.
© 2020 American Geophysical Union. Published 2020 by John Wiley & Sons, Inc.

137
138 GLACIAL FEATURES

Figure 13.1 Cross-section of the subglacial topography of Langjökull. [Modified from Bjornsson [2017];
design credit Nathan Mennen.]

Langjökull practically disappeared entirely on two wide. Outlet glaciers on the northwest side of Þórisjökull
occasions in the Holocene (namely between 10.2 and make an important contribution to the discharge of the
8.7 ka and 7.35–5.5 ka), indicating that temperatures River Geitá, and to the River Þórisdalur from the east.
in south‐central Iceland were warmer than today Langjökull has 19 outlet glaciers, but there are 6
[Black, 2008]. primary lobes: two on the southern margin—
­
The oblong ice cap is oriented SW–NE and is 55 km Hagafellsjökull Vestari (137 km2; 64.5066, −20.2619),
long from its snout in Hagavatn (64.4805, −20.2847) to Hagafellsjökull Eystri (111 km2; 64.5100, −20.4275);
the northeast terminus in Hundadalur (64.9066, two on the eastern margin—Norðurjökull (62 km2;
−19.7836) and 30 km at its widest (Figure 13.2). This 64.5986, −19.9436) and Suðurjökull (54 km2; 64.7547,
shape is largely dictated by the variability found in the −19.8241); and two on the western margin—
amounts of precipitation on Langjökull. For example, Svartárjökull (39 km2; 64.6716, −20.5127) and
less precipitation in the south results in steeper and Þrístapajökull (104 km2; 64.7400, −20.2002). Of these
faster moving outlet glaciers and north‐flowing glaciers six outlet glaciers, three are classified as surging and
terminate in broad fronts at higher altitudes [Pope et al., have advanced during the past 50 years: Hagafellsjökull
2016]. Langjökull is positioned in the Central Highlands Vestari (1971, 1980), Hagafellsjökull Eystri (1974,
and is at considerably higher elevation than most other 1980), and Suðurjökull (1828–1930, 1994, 1999)
ice caps as the tallest peak is 1450 m. The tuya (Figure 13.2, Table 13.1). Although each of these three
Geitlandsjökull (63.6000, −20.6000; 5 km length) outlet glaciers may have had other episodes not
beneath the ice cap is among the tallest peaks at 1395 m. reported here, past surging events have only been con-
There are two smaller ice caps nearby that were once firmed at Suðurjökull through varved lake sediments
likely contiguous with the Langjökull ice cap: Eírksjökull from nearby Lake Hvitárvatn (Photo 45) [Larsen et al.,
(21 km2; 64.7658, −20.4077) and Þórisjökull (25 km2; 2013]. Eight surge episodes were described by Larsen
64.5455, −20.7147). Unusual for Iceland ice caps, et al. [2013] that occurred between 1828 and 1930,
Eírksjökull has a smooth and convex appearance as it is when the glacial front advanced up to 1.6 km per event
the result of a Quaternary volcanic eruption that pro- before terminating into the lake. The rapid advance of
duced a palagonite tuya and was then capped by a dol- these specific glaciers is attributed to subglacial
erite lava dome; the mountain’s base covers 40 km2 braided‐river‐type channels intermittently distributing
[Björnsson, 2017]. Þórisjökull is also the result of a tuya meltwater, as observed by Palmer et al. [2009] and
formation and is 10 km long from east to west and 6 km Björnsson et al. [2003].
Glacial Features of the Reykjanes Peninsula and Southwestern Region 139

Photo 45 A view of Langjökull from Lake Hvitárvatn. [Courtesy of Simon Vaughan/Alamy Stock Photo.]

From historic accounts, Norðurjökull and for obtaining detailed stratigraphy of glacial activity
Suðurjökull have retreated approximately 5000 m since and corresponding environmental change over the
1900 as they once covered the proglacial lake basin of past 3 kyr [Larsen et al., 2011]. Varve thicknesses at
Hvitárvatn (64.6069°N, 19.8408°W) [Sigurðsson, Lake Hvitárvatn represent variations in sedimentation
1998]. Unlike the other outlet glaciers of Langjökull, rates determined by local glacial erosion of a large
these two eastern lobes are thought to have maintained watershed (~820 km2) that remains partially (~33%)
their maximum Holocene extents over the past covered by the Langjökull ice cap [Larsen et al., 2011].
400 years instead of experiencing multiple advance Sedimentation rates at Lake Hvitárvatn have intensi-
and retreat episodes [Flowers et al., 2007]. The fied over time, with the last 300 years (estimated
maximum ice cover of the lake likely occurred between volume of 1.5 × 1011 kg) [Flowers et al., 2007] repre-
1800 and 1900 which can either reflect the impact of senting 75% of the core [Larsen et al., 2011]. At pre-
the LIA or a glacial surge event [Flowers et al., 2007]. sent, the eroded material produced by the ice cap first
From sonar bathymetry of the lake bottom and passes through Hvítá, an outlet stream that removes
moraines still visible on land, Geirsdóttir et al. [2016] the majority of sediment before reaching the lake, thus
determined that Norðurjökull retreated at a slower explaining its clear appearance [Black et al., 2004].
rate than Suðurjökull, such that the toe of Norðurjökull About 70% of the lake’s present water budget derives
remained in the lake until the summer of 2009. The from Langjökull meltwater, with two‐thirds trans-
recessed eastern margin of the glacier has exposed the ported as groundwater [Flowers et al., 2007]. The high
14 km long Jarlhettur (64.5308, −20.1158) palagonite contribution of meltwater to groundwater flow is sup-
mountain range, which is the product of a 15 km long ported by the widespread presence of moulins found
fissure vent underneath the glacier during the on the surface of outlet glaciers that provide a conduit
Pleistocene. Lake Hvitárvatn (422 m a.s.l.) is 84 m deep to the base of the glacier, and thereby to meltwater
and 36 km2 in area, and is considered an ideal archive channels [Eyre et al., 2005]. Flowers et al. [2007],
140 GLACIAL FEATURES

Figure 13.2 Location of Langjökull with Lake Hvítárvatn and drainage divides. [Modified from Bjornsson
[2017] and Bjornsson et al. [2003]; design credit Nathan Mennen.]

­ owever, estimated subglacial discharge to ground-


h disappear within 100–150 years (Björnsson, 2017]. This
water to be ~7% greater during the LIA. Following the was projected by determining that with each 100 m rise
gradient of the modern landscape, Lake Hvitárvatn of the firn line, the mean mass balance of the glacier
drains southeast and has an average monthly discharge shrunks by 0.75 m year−1. Additionally, a map of the
of 45 m3 s−1 with an average suspended sediment load basement topography of Langjökull reveals that only
of 35 mg L−1 [Larsen et al., 2011]. Moreover, due to the 10% of the mountain tops reach higher than the snow-
area being underlain by porous lava, it is suspected line in the current climate [Björnsson and Palson, 2008]
that subsurface water flow generated from the ice cap (Figure 13.1). If the Langjökull ice cap were to disap-
as a whole travels much further south to enter Lake pear then it is projected that only minor glaciers would
Þingvallavatn (section 3.4) [Björnsson, 2017]. settle on the highest mountains above a height of
The area of Langjökull ice cap has been surveyed 1100–1200 m. The melting of the Langjökull ice cap
since 1890. Björnsson [2017] reported that from 1900 to would translate to 0.5 mm of eustatic sea‐level rise
2004, Langjökull lost 40 km3 or 21.3% of its current [Pálsson et al., 2012].
mass volume, with more intervals of losses than gains
since. Over the past 100 years, the ice cap’s mass 13.2. HOFSJÖKULL
balance has been negative, reducing by more than half
over this interval, from 2000 to 800 mm year−1 [Pálsson The Hofsjökull ice cap (1800 m; 64.8313°N,
et al., 2012]. This is a reflection of below‐average 18.8230°W) is considered the third largest in Iceland
winter precipitation (1000–2000 mm) and higher with an area of 925 km2, a volume of ~200 km3, and an
summer melting (2000–4000 mm) [Björnsson, 2017]. average thickness of 225 m [Björnsson, 2017]. About
Assuming a rise in temperature of 2.0°C per decade as two‐thirds of the ice cap is above 1000 m, but at pre-
projected by the International Panel on Climate Change sent only 10% is above the snow line, which has been
[2016], the glaciers of the central highlands will probably between 1200 and1300 m since 1988 [Björnsson, 1988;
Glacial Features of the Reykjanes Peninsula and Southwestern Region 141

Table 13.1 Langjökull (total area 8100 km2) reported glacial surge advances
Area Advance Area affected
Glacier (km2) Slope (°) Surge years (km) (km2) Years between surges
Síðujökull 380 1.3 1893, 1934, 1963, 1994 0.5–1.2 500 41, 29, 31
Tungnaárjökull 308 1.4 1920, 1945, 1994 1.2–2.0 550 25, 49
Sylgjujökull 175 2.2 1945, 1996 0.5 150 51
Köldukvjslarjökull 313 2.6 1975, 1992 0.2 100 17
Dyngjujökull 1040 1.6 1900, 1934, 1951, 1977, 1.5 400–800 34, 17, 26, 22
1999
Brúarjökull 1500 1.7 1625, 1730, 1775, 1810, 8–10 1500 105, 45, 35, 80, 73
1890, 1963
Brúarjökull East 1938
Eylabakkajökull 100 2.7 1890, 1894, 1931, 1938, 0.6–2.8 110 4, 37, 37, 34
1972
Skeiðararjökull 1 2.0 1787, 1812, 1857, 1873, 1.0 1200 25, 45, 16, 56, 56, 6
1921, 1985, 1991
Breiðamerkurjökull 540 2.5 1794, 1815, 1821, 1861, 0.4–1.0 600 21, 8, 38, 8, 6, 17,
1869, 1875, 1892, 1912 20, 8, 35, 15
Eastern stream 1919, 1954, 1969
Middle stream 210 3.2 1978
Note. As described by Björnsson et al. [2003]; design credit Nathan Mennen.

Thorsteinsson and Bjornsson, 2012]. The temperate ice and helps to explain why the rates of erosion at
cap at Hofsjökull began to form at 4 ka and has likely Hofsjökull are higher (0.9 mm year−1), than at nearby
formed at a faster rate than the Langjökull ice cap due Langjökull (0.4 mm year−1) [Björnsson, 1980].
to its higher elevation. The gently sloping flanks of the shield‐type volcano
It is unique in that its circular shape covering 925 km2 (~3.4°) beneath Hofsjökull (Figure 13.3) are conducive
is dictated by the subglacial caldera 650 m deep to forming outlet glaciers. Due to this low slope, how-
(~40 km2) of Iceland’s largest volcano in terms of ever, glacial surges occur only in the flat ablation areas
volume [Björnsson, 2017; Thorsteinsson and Bjornsson, and have not been observed to propagate upglacier,
2012] (Figure 12.9). The schistose rhyolitic Hásteinar which is considered atypical [Björnsson et al., 2003].
mountain (1630 m), which rises up through the glacial The five main outlet glaciers are: Blágnípujökull
ice on the eastern caldera rim, was described by (52 km2; 64.8166, −19.1166) and Blautukvíslarjökull
Björnsson [2017], who considered that ice was not likely (73 km2; 64.7936, −19.1230) on the southeast side;
to have covered the volcano when it formed during the Nauthagajökull (9 km2; 64.7580, −19.3069), which
LGM. Subsequent eruptions have not produced bifurcates to form Mulajökull (91 km2; 64.6627,
enough material to infill the caldera thus leaving a deep −18.7144), in the south; and Sátujökull (72 km2;
cavity now occupied entirely by the ice cap. The topog- 64.9522, −18.6269) in the north (Figure 13.4). Of these,
raphy of the caldera on the western edge shows a slight Mulajökull is the largest lobe and descends from the
depression that could occupy a lake should the glacier highest portion of the dome (1800 m a.s.l.) through a
melt. Björnsson [2017] describes this scenario as being narrow passage 2 km wide between rhyolite peaks, from
similar to the formation of Lake Öskjuvatn in the Askja where it fans out into the low‐lying Þjórsárver plains
caldera (see section 10.1). Basaltic lava fields found on (610 m a.s.l.) as a piedmont glacier. The margin of this
the north, east, and south sides indicate that the vol- outlet is 12 km‐long, and the last ~1 km has a gradient
cano has been active during postglacial times, however, between 10 and 12° that flattens to a slope of <3°. At
it is not likely that it has been active since Víking present, the maximum thickness of the ice lobe is bet-
settlement of the island. Other local peaks include the ween 250 and 100 m, thinning at its snout. Due to its
two nunataks Gægir and Tanni. A tomography survey steep descent, Mulajökull has experienced at least seven
conducted in 1988 revealed basement features of the glacial surges between 1924 and 1992 that tend to occur
volcano including a ravine 250 m deep that is currently about every 10 years to affect an area ~10 km2. These
under an outlet glacier [Blautukvíslarjökull). This fea- glacial surges have been described in detail from the
ture likely formed by catastrophic jökulhlaup flooding sedimentary succession found in the forefield, which is
142 GLACIAL FEATURES

Figure 13.3 Cross‐section of the subglacial topography of Hofsjökull. [Modified from Bjornsson [2017];
design credit Nathan Mennen.]

Figure 13.4 Hofsjökull outlet glaciers and ice divides. [Modified from Bjornsson [2017] and Bjornsson
et al. [2003]; design credit Nathan Mennen.]
Glacial Features of the Reykjanes Peninsula and Southwestern Region 143

dominated by till layers [Benediktsson et al., 2009]. The known Víking homestead inhabited by Stöng (64.1200,
four to five till layers present within drumlins in the −19.8202). This valley today remains the largest con-
forefield are considered by Benediktsson et al. [2009] to tinuous “oasis of vegetation” in the Icelandic central
have been deposited from surge events. Here, it should highlands and provides a good example of permafrost
be noted, that the Múlajökull drumlin field is unusual features [Björnsson, 2017] (Photo 46). Changes in
because it is a feature rarely seen at other contemporary Þjórsá River runoff induced by glacial melt are a
glaciers. These drumlins have provided a modern‐day ­concern due to its use as a hydroelectric power source.
analog for glaciologists to make comparisons with Ye et al. [2003] predict that the meltwater discharge of
Pleistocene examples [Benediktsson et al., 2009]. This Hofsjökull as a whole will increase by 50% near the
forefield is also important because it contains evidence end of the this century, and then decline as the glacier
of 10–12 end moraines with 5–20 m spacing that illus- continues to shrink. The firn layer is also expected to
trate the retreat of the margin in the past 10–12 years decrease in depth as the temperate glacier is exposed to
[Benediktsson et al., 2009]; the outermost end moraine, higher temperatures. Woul et al. [2006] have empha-
Arnarfellsmúlar, is 5–10 m tall, 100 m wide, and con- sized the role of the firn layer in delaying water flow
tains sequences of loess, peat, and tephra [Benediktsson through glaciers and its influence on seasonality.
et al., 2009]. The near circular shape of Hofsjökull, and in turn,
Additionally, the Mulajökull lobe is particularly its radial flow pattern has made this glacier an ideal
important because it contains the headwaters for candidate for the study of mass‐balance modeling
Iceland’s longest river, Þjórsá (230 km), which starts at [Jóhannesson, 1991; Jóhannesson et al., 2006]. Survey
the Bergvatnskvísl tributary. The River Þjórsá passes records of the major outlet glaciers extend back to
through the Þjórsádalur Valley and alongside a well‐ 1937, however, Jóhannesson et al. [2006] did not begin

Photo 46 Stöng homestead alongside the River Þjórsá fed by the runoff of the Hofsjökull glacier, which
passes through the Þjórsádalur Valley. [Courtesy of Imagebroker/Alamy Stock Photo.]
144 GLACIAL FEATURES

annual measurements until in 1988, which have been by flow through the glacial hydrological system. In
Jóhannesson et al. [2006] and has continued since. comparison, applying a climate‐model‐based tempera-
Jóhannesson et al. [2006] reported that the surface area ture and precipitation scenario for Iceland until 2050
of Hofsjökull reduced by 3–4 m.w.e. from 1988 to 2005; results in higher runoff throughout the year, increasing
the mean annual mass decreases by 0.6 m.w.e. C−1 of total runoff by roughly one‐third. The results empha-
increased global warming. According to Björnsson size the role of the firn layer in delaying water flow
[2017] the Hofsjökull glacier has shrunk by 16% (from through glaciers, and the influence on discharge sea-
996 km2 to 830 km2) over the past 70 years, as evidenced sonality when firn areas shrink in response to climate‐
by the most recent terminal moraines being 2–3 km change‐induced glacier wastage.
from the oldest most distal moraines. This has been
accompanied by an increasing rate of run off, with 25% 13.3. KERLINGARFJÖLL
higher annual discharge, or by 0.6 m year−1 [Jóhannesson
et al., 2006]. With continued warming, Björnsson [2017] Explosive activity of the WVZ formed the rhyolitic
predicts that by 2100 the surface area of Hofsjökull will Kerlingarfjöll mountains (1500 m; 64.6500, −19.2644)
decrease by 50% and Múlajökull will completely disap- during the LGM (Figure 7.2). At this time, two calderas
pear, and the ravine at Blautukvíslarjökull will be com- formed: the eastern (17 km2) and western (14 km2).
pletely exposed. Currently, the eastern caldera hosts 10 glaciers (total
Reducing the firn layer under otherwise likewise surface area 3 km2) at an elevation of 1200–1300 m
conditions results in almost unchanged total runoff [Björnsson, 2017]. The largest glaciers are:
volumes, but yields a redistribution of discharge within Loðmundarjökull Eystri (64.6452, −19.1916), Jökulkinn
the year. Early summer discharge (June to mid‐August) (64.6336, −19.1913), and Mænisjökull (64.6525,
is amplified by roughly 5–10%, whereas late‐summer −19.3238). The Kerlingarfjöll mountains may best be
to autumn discharge (mid‐August to November) is known, however, as the divide between the area’s two
reduced by 15–20% as a result of accelerated water major rivers: Hvítá and Þjórsá.
14
Glacial Features of the South and Southeastern Regions

With the onset of colder temperatures ~5 Ma came As they discussed, however, with only two exceptions,
the initial expanse of the Vatnajökull ice cap (Figure 2.2). the dates of terminal moraines around the Vatnajökull
The Vatnajökull ice cap (7800 km2) sits on top of a ice cap are about 100 years younger (1850–1900 CE).
mantle plume that has been in the approximate same The use of tephrochronology from Vatnajökull’s
location for the past 13 Myr and today marks the tri- southernmost volcano (Öræfajökull) constrained the
ple‐poimt junction between the NVZ, EVZ, and MIB LIA maxima to before 1873 CE [McKinzey et al.,
(Figure 2.1). Since ~3 Ma the EVZ has been propa- 2005].
gating eastward at a rate of 2–5 cm year−1 [Einarsson, The debate regarding the timing of deglaciation at
1991, 2008], and has both initiated and influenced the Vatnajökull following the LGM is still ongoing.
transform movement of the MIB. As explained by Undeniably, identifying the causes of advances and
Helgason and Duncan [2001], beginning at 2.6 Ma both retreats of Vatnajökull’s outlet glaciers is somewhat
the intensity and frequency of glaciations increased complicated by the presence of subglacial volcanic
significantly along the southern portions of centers and the response of the ice to eruptions from
Vatnajökull. Mapping cliff profiles in the Skaftafell them; thus the possibility of a synchronous retreat
region (64.0704, −16.9752) provided estimated rates of ­following the LGM may be unlikely.
topographic uplift in the area between 100 m and It is along the EVZ and MIB that the complicated
2 km, with rapid increases occurring at 0.8 Ma story of fire and ice unfolds. Over the past 800 years
[Helgason and Duncan, 2001]. The change in elevation approximately 80 subglacial or proglacial volcanic
amplified the downgradient movement of the ice cap, eruptions have occurred along the EVZ and MIB
which resulted in deeply eroded valleys that became [Larsen et al., 1998]. Björnsson and Einarrson [1990]
permanent ice flow channels during successive glacial ascertained from radio echo soundings and earth-
intervals and can still be seen in this region (Photo 40). quake and jökulhlaup data that there are six active
As a whole, the 30 outlet glaciers that make up the volcano systems beneath Vatnajökull: Grímsvötn,
­
Vatnajökull ice cap continued to grow from the time Veiðivötn, Bárðarbunga, Kverfjöll, Esjufjöll, and
the Víkings settled in 874 CE through the 19th century Öræfajökull (Figure 7.2). These systems are described
[Björnsson, 2017]. Life along the southern coast was by Björnsson and Einarrson [1990], and in Chapters 4
challenging for the initial Víking settlements as surging and 9. Bárðarbunga is in the center of a volcanic system
glaciers destroyed newly established farms during the that includes fissure swarms of Veiðivötn; Grímsvötn
LIA with their uncontrollable advance of between 10 is the center of a volcanic system that comprises the
and 15 km. Lichenometric techniques used on Laki ­fissure swarm in the southwest and extends for
moraines from one of Vatnajökull’s outlet glaciers, unknown distances to the northeast of the volcano;
Lambatungajökull, found that the most extensive the Kverfjöll system extends 10 km southwest of the
surges were likely in the 1780s [Bradwell et al., 1999]. central volcano; and Esjufjöll and Öræfajökull behave

Iceland: Tectonics,Volcanics, and Glacial Features, Geophysical Monograph 247, First Edition. Tamie J. Jovanelly.
© 2020 American Geophysical Union. Published 2020 by John Wiley & Sons, Inc.

145
146 GLACIAL FEATURES

independently. Of these, Bárðarbunga, Grímsvötn, from the sea to 25 km inland [Skeiðarársandur,


and Kverfjöll lie directly on either the EVZ or the MIB 1000 km2; Photo 44). These outwash plains have
and pose the most threat for phreatic eruptions at pre- become characteristic of the southern coast where the
sent (see section 9.1). For example, Bárðarbunga is wide glacier margins terminate on broad lowland areas
considered the second most active volcano throughout with meltwater discharging to the North Atlantic via
historic time, and Grímsvötn has erupted four times braided river channels (Photo 47). The thickest portion
in the last 15 years (1996, 1998, 2004, 2011) after a 60‐ of the Skeiðarársandur (63.7877, −17.2780) is in the
year repose period. Nearby, there are four other middle of the outwash plain (200–250 m thick) and it
smaller ice caps found along the southern coast that thins toward the terminus of the glacier.
are adjacent to the EVZ rift zone and also conceal The meltwater runoff from Vatnajökull feeds the
active volcanoes beneath the ice (Mýrdalsjökull, largest rivers in Iceland, which encompass half the
596 km2; Eyjafjallajökull, 78 km2; Tindfjallajökull, country’s catchment areas (Figure 12.10). At some
19 km2; Torafajökull, 15 km2; Figure 7.2). The outlet glaciers (e.g., Skeiðarájökull, Breiðamerkurjökull),
Mýrdalsjökull ice cap conceals the Katla volcano, the erosive capacity of their discharge, in terms of both
which shares a magma chamber with Eyjfjallajökull. frequency and power, has eroded channels up to 300 m
The last activity at Eyjfjallajökull was in 2010 when a below sea level [Björnsson, 2017]. As an example, River
VEI = 4.0 eruption produced an ash cloud 9 km tall Jökulsá á Brú was monitored during a surging event at
(see section 9.2). Historically, when Eyjfjallajökull has Brúarjökull in 1963–1964 that likely was initiated by
erupted, Katla has soon followed. the emptying of subglacial melt from the Kverfjöll
The retreat of Vatnajökull has been instrumental in ice‐filled caldera [Björnsson and Einarsson, 1990]. The
understanding the connections between glacial retreat suspended load of sediment from this event was
and changes in magmatic activity. It has been esti- upwards of 6470 mg L−1. Additionally, Björnsson [1980]
mated that during the 20th century Vatnajökull lost reported that the total transport of sediment during the
about 10% of its ice mass (~450 km3), which through year was 25 × 106 tons and discharge in the order of
modeling has been correlated to an increase in magma 10,000 m3 s−1, which corresponded to a denudation rate
production beneath the glacier of 0.014 km3 year−1 of 3.7 mm year−1 (Table 12.1). These values may be low,
based on present day uplift of 25 mm/year [Pagli and however, when compared with the 150 × 106 tons of
Sigmundsson, 2008]. The increase in magmatic produc- sediment estimated to have been transported to
tion of 1.4 km3 100 year−1 modeled by Pagli and Skeiðarársandur during the 1934 eruption of Grímsvötn
Sigmundsson [2008] was comparable with historic [Björnsson, 1980]. The dynamically changing system of
eruption estimates quantified at Vatnajökull by the Skeiðarársandur limited transport communications
Thordarson and Larsen [2007]. If glacial retreat con- from eastern to western Iceland along the southern
tinues Pagli and Sigmundsson [2008] suggest that coast until the first main national highway was built in
increased volcanic activity may be expected at 1974. Unfortunately, however, part of the Skeiðará
Vatnajökull in the future. Swindles et al. [2018], how- Bridge (63.9558, −17.4647; Photo 48) was destroyed in
ever, determined a lag of ~600 years between Icelandic 1996 by large calving icebergs and the extreme river dis-
volcanism and climate change, which concurs with charge rate of 45,000 m3 s−1. The dramatically reshaped
Schmidt et al. [2013] who quantified a huge (100–135%) coastline resulted in US$10 million of damage to the
increase in mantle melt production due to deglaciation country’s infrastructure [Russell et al., 2000].
between 1890 and 2010 CE. Schmidt et al. [2013] sug- The amount of water from a jökulhlaup event at
gest that the recent increases in volcanic activity on Vatnajökull is much greater than the contribution of
Iceland may instead be the result of GIA in response seasonal melt runoff in a water budget. Water gener-
to LIA deglaciation. ated via seasonal melt, however, is an important factor
Undoubtedly, subglacial eruptions have led to cata- in determining the overall mass balance of a glacier
strophic jökulhlaups. The shear presence of geothermal whereby warming temperatures can induce a positive
heat at the base of these glaciers due to their proximity feedback loop by increasing the amount of meltwater
to the hot spot, however, creates meltwater that can paths (e.g., crevasses; Figure 12.7). Meltwater follows
accumulate in subglacial lakes and produce outburst drainage divides that have been established for the sur-
floods (jökulhlaups) in the absence of a volcanic face of the glacier from topographic surveys and water
eruption. Regardless of origin, flooding events over the pressure potential maps (Figure 12.10). Steffensen
past 10,000 years have produced enough sediment to et al. [2008] modeled the surface climate and energy
create the world’s largest sandur plain that extends balance of the Vatnajökull ice cap from 1980 to 2016,
Photo 47 Braided river channel of a sandur plain discharging meltwater from Vatnajökull into the North
Atlantic. [Courtesy of Imagemaker/Alamy Stock Photo.]

Photo 48 Skeiđará Bridge remains near Skaftafell National Park. [Courtesy of Tamie Jovanelly.]
148 GLACIAL FEATURES

as well as the sensitivity of runoff to the spring snow thickness has been maintained at ~1000 m, although
cover. They found that the glacial runoff is controlled the average thickness is closer to 400 m. Using adio
mostly by summer weather (mid‐June through August) echo soundings Björnsson [1982] and Eythórsson
when daylight persists for up to 22 h day−1. The amount [1952] revealed that the central ice mass of Vatnajökull
of runoff can increase between 22 and 72% if there is is resting on a highland plateau at 700–800 m altitude
spring snow cover on the ice cap that can contribute to with only a few peaks rising above 1100 m (e.g.,
the water budget. This is because newly accumulated Hvannadalshnúkur, 2110 m; 64.0147°N, 16.6749°W).
snow melts faster than that compacted into firn (Figure Due to its expanse and basement topography domi-
12.5). The 37 years of modeled runoff data was nated by volcanoes, the undulating Vatnajökull ice
­compared with field observations from flow gauges at cap has four main domes: Bárðarbunga (2000 m),
glacial outlet rivers (e.g., Brűarjökull), with analogous Kverkfjöll (1920 m), Breiðabunga (1500 m), and
results showing an overall negative mass balance for Háabunga (1730 m).
Vatnajökull. Likewise, Wittman et al. [2017] calculated Vatnajökull has at least 30 outlet glaciers, but the
the snow albedo effect in order to predict seasonal following 11 are the major lobes separated by topog-
snowmelt and run‐off rates, which were used in calcu- raphy and/or drainage divides: Bárðarbunga
lating energy budgets at Vatnajökull. They concluded (1000 km2; 64.6410, −17.5280), Dyngjujökull
that impurities in the snow (e.g., dust, ash, loess) from (500 km2; 66.1500, −22.2500), Brúarjökull (1600 km2;
specific events (e.g., Gjálp eruption of 1996) have 64.6817, −16.1567°), Eyjabakkajökull (109 km2;
reduced the albedo affect so that heat is absorbed 63.6304, −19.6067), Breiðamerkurjökull (1700 km2;
instead of being reflected off the ice sheet surface. 64.1531, −16.4000), Öræfajökull (356 km2; 63.9822,
Since 1991 the glacier has lost ~3% of its mass, or −16.6536), Skeiðarárjökull (1300 km2; 64.0558,
92.7 km3 of ice volume. Additional modeling c­ ompleted −17.2081), Siðujökull (500 km2; 64.1178, −17.8525),
by Hannesdóttir et al. [2014] using aerial photographs Tungnaárjökull (235 km2; 64.3186, −18.0222)
and digital elevation models from 1890 to 2010 charac- Sylgjujökull (300 km2; 64.4472, −17.9205),
terized two major periods of high mass loss at Köldukvíslarjökull (200 km2; 64.5350, −17.8438)
Vatnajökull: 2002–2010 and 1904–1945. Conversely, (Figure 12.10; Table 14.1). Here it should be clarified
data collected for 2014–2015 indicated the first positive that the popular tourist destination of Skaftafellsjökull
mass balance in over two decades of monitoring National Park is a subset, or arm, of the larger
[Gudmundsson et al., 2016]. This was attributed to Skeiðarárjökull (Figure 12.10). Of those listed above,
higher than normal precipitation during the winter the glaciers can be categorized as surge, nonsurge,
season, and an unusually cold June and July. and mixed types. Those outlets encompassed by
Eyjabakkajökull on the southeast margin are smaller,
14.1. VATNAJÖKULL steeper, and surge at a rate of about once every 35 years
[Sigurðsson, 1998]. Eyjabakkajökull’s advance/retreat
On 7 June 2008, Vatnajökull National Park record, however, may not be as reliable as other surg-
(64.4220, −16,7902) was established. With a total ing lobes on Vatnajökull because two different systems
area of 14,141 km2, it covers ~14% of the island and meet here to influence the flux patterns (i.e., valley
makes it the second largest national park in Europe. ­glacier meets outlet glacier [Björnsson, 1998]).
The focus of the national park is Vatnajökull, which The outlet glaciers of southeast Vatnajökull are
may best be known as Iceland’s and Europe’s largest located in the warmest and wettest area of Iceland
glacier at 7800 km2 (Figures 12.3 and 14.1). At pre- and descend down to the lowlands. Results from ice
sent the ice cap covers approximately 8% of the dynamics and hydrology models indicate that these
island. The protected landscapes around the glacier glaciers are the most sensitive to future warming of all
take up the other 6% of area, including: Jökulsárgljúfur the outlets of Vatnajökull [Hannesdóttir et al., 2014].
fjords (see ­section 5.4; 66.0285, −16.4871), Lakagígar In particular, Hannesdóttir et al. [2014] consider that
crater rows (see section 9.1; 64.0707, −18.2378), they are especially vulnerable to warming climate
Langisjor highlands (64.2154, −18.2184), Jökulsárlón conditions because their beds lie 100–300 m below the
lagoon (64.0784, −16.2306), and Skaftafell (64.0704, elevation of the current terminus. Of these, the most
−16.9752). popularly visited is Skeiðarájökull, which is easily
Being the largest of the remaining ice caps on the viewed from Highway 1. Its 12 km wide base spreads
island, Vatnajökull continues to provide the most onto the Skeiðarársandur plain and makes it the larg-
detailed and continuous ice record as maximum est lobe of Vatnajökull.
Glacial Features of the South and Southeastern Regions  149

Figure 14.1 Details of the Vatnajökull ice cap including location of the Grimsvotn, Kverfjöll, and Bárđabunga
volcanoes. [Modified from Bjornsson (1979]; design credit Nathan Mennen.]

Table 14.1 Vatnajökull lobes and their associated surge events by date
Glacial lobe Volume (km2) Surge dates
Bárdarbunga 1000 –
Dyngjujökull 500 1900, 1934, 1951, 1977, 1999
Brúarjökull 1600 1625, 1730, 1775, 1810, 1890, 1963
Eyjabakkajökull 109 1890, 1894, 1931, 1938, 1972
Breidamerkurjökull 1700 1794, 1815, 1823, 1861, 1869, 1875, 1892, 1912, 1919, 1954, 1969, 1978
Oræfajökull 356 –
Skeidarárjökull 1300 1787, 1812, 1857, 1873, 1929, 1985, 1991
Sidujökull 500 1893, 1934, 1963, 1994
Tungnaárj ökull 235 1920, 1945, 1994
Sylgiujökull 300 1945, 1996
Köldulvíslarjökull 200 1975, 1992
150 GLACIAL FEATURES

The Öræfajökull glacier covers the largest stratovol- (Skeiðará, Gígja or Sandgígjukvísl, and Súla). Of the
cano in Iceland (Figure 12.10), and is a noticeable three rivers, Skeiðará has the largest average summer
landmark viewed from sea‐level at the Skeiðarársandur. flow (200–400 m3 s−1) and runs along the snout of the
The ice‐filled caldera (400 m deep) is 4–5 km in diam- glacier before connecting with Gígja and discharging
eter and has a surface area of 14 km2 and is capped into sea. To date, Breiðamerkurjökull has retreated
with ~ 550 m of ice. Öræfajökull provides a unique sce- 4 km since 1890 and areas beneath 200 m of glacial ice
nario in that during eruptions downgradient slides of in 1904 are now ice free [Björnsson et al., 2001].
ice are often promoted [Björnsson, 2003]. Öræfajökull’s Skaftafellsjökull (64.0654, −16.8635) is another pop-
highest peak, Hvannadalshnúkur (2110 m; 64.0147, ularly visited outlet glacier of Vatnajökull because of
−16.6749), is located on the northwest edge of the the proximity of Skaftafell National Park located east
­caldera rim although this is just one of six peaks above of its snout. In total, Skaftafellsjökull has retreated
1700 m. Although listed above as part of Vatnajökull, ~2300 m since 1904, when the first terminus approxima-
Öræfajökull is sometimes considered to be an tions were derived from early topographic maps. Annual
independent ice cap that has merged with the south field measurements began in 1932, with increasing
side of Vatnajökull, on the eastern side of ­precision and consistency beginning in 1942. The data
Breiðamerkurjökull (Figure 12.10). indicate that the average rate of retreat was 72 m year−1
Breiðamerkurjökull is separated by two distinct from 1932 to 1942, reducing to 30 m year−1 from 1942 to
medial moraines, with the eastern branch entering 1970. The subsequent 50 years reflects a complicated
Breiðárlón (64.0680, −16.3713) proglacial lagoon and history of advance and retreat at Skaftafellsjökull:
the western branch terminating at Jökulsárlón (64.0816, 1971–1988 the glacier advanced 96 m; 1989–1995 the
−16.2319 tidal lagoon (Photo 49). As described by glacier retreated 136 m; 1996–1998 the glacier advanced
Sigurðsson [1998], the two medial moraines have led to 90 m; 1999 to present the glacier has retreated on average
the development of three different meltwater channels 53 m year−1 [Marren and Toomath, 2014].

Photo 49 Jökulsárlón glacial tidal lagoon in front of the main glacier. [Courtesy of Tamie Jovanelly.]
Glacial Features of the South and Southeastern Regions  151

14.2. FJALLSARLÓN, BREIÐÁRLÓN, Discharge of these influent streams caused rapid


AND JÖKULSÁRLÓN expanse of the lake from 2000–2009, such that
Jökulsárlón increased by 6 km2 [Schomacker, 2010].
Today, there are three proglacial lakes in front of With a continued trend of negative net mass balance at
Breiðamerkurjökull: Fjallsarlón (4 km2; 64.0163, Breiðamerkurjökull into the future, Canas et al. [2015]
−16.4002), Breiðárlón (9 km2; 64.0605, −16.3663), and project that Jökulsárlón will continue to expand, reach-
Jökulsárlón (25 km2; 64.0686, −16.2255) (Figure 12.10). ing 56.3 km2 by 2055.
The basins for Breiðárlón and Jökulsárlón were likely As the base of Breiðamerkurjökull is on unlithified
created as the glacier retreated during the LIA [Boulton sediment, rather than bedrock, pronounced ice‐
et al., 1982]. Breiðárlón filled during the early 1920s marginal overdeepening has developed at Jökulsárlón
with influx from the River Breiðá. During 1954 the [Boulton et al., 1982; Tweed et al., 2005]. Thus, glacial
course of the river moved west to feed Fjallsarlón retreat and influx from rivers has caused deep under-
(Photo 50). According to Björnsson [2017], the first cutting into the lake bottom, resulting in a basin depth
appearance of Jökulsárlón was in 1934 when the glacier of ~300 m below sea level. Björnsson [1992] confirmed
snout was about 1 km from the Atlantic Ocean; today this in 1991 using radio echo sounding to map a deep
the distance is 8 km. Of the three lakes, Jökulsárlón is trench 2–5 km wide by 20 km in length at the toe of
the largest and has consistently grown in area since Breiðamerkurjökull. Jökulsárlón is considered the
1960. In 1992, a nearby small (2 km2) glacial lake called deepest lake in Iceland and provides a sink for sedi-
Stemmulón converged with Jökulsárlón thereby ment entering via the river systems, which recently has
combining discharge through all of Breiðamerkurjökull’s become a problem because the coastline is not being
central and eastern drainage systems including: Vestri‐ renourished. According to Björnsson [1998] the
Stemma, Eystri‐Stemma, Brennhólakvísl, and Veðurá. ­coastline is eroding at a rate of 8.5 m year−1, which is a

Photo 50 Fjallsarlón lagoon discharging into the ocean via a braided river system. [Courtesy of Tamie
Jovanelly.]
152 GLACIAL FEATURES

concern to Icelanders as it threatens the location of remains within a few degrees of freezing until the lake
Highway 1 and electricity infrastructure. bottom is approached. Within 10 m of the caldera
During the 1990s the rate at which icebergs were floor, the temperature of the water can increase
calving notably increased and was quantified as a ­between 1° and 6°C [Jóhannesson et al., 2007]. The
loss of 600 m year−1 from the glacier terminus, which movement and distribution of the accumulated melt-
caused the lake area to expand by 0.5 km2 year−1 water in a caldera is not uniform, but has been
[Björnsson et al., 2001; Björnsson, 2017]. For added described to move in convection patterns [Jóhannesson
context, the height of the glacial snout at the et al., 2007]. Data on mixing are important because
Breiðamerkurjökull terminus is about 30 m above temperature of the meltwater controls basal melting
sea level, with an additional 60 m below the surface. and further widening of subglacial conduits, and hence
In addition to being larger than the other two lakes, can induce flooding. Increases in the temperature of
Jökulsárlón is the only lake connected to the ocean the meltwater can cause an ice cauldron to form, which
and is thereby influenced by brackish conditions and is a surface depression causing circular surface cracks
tidal fluctuations, which produces great variability in in the ice above the subglacial lake prior to, or just
salinity distribution (vertically and spatially) at dif- after, a subglacial lake is drained (Figure 14.2). If the
ferent times of the year [Landl et al., 2003]. As an formation of a cauldron is spotted early by noticing
example, during spring the maximum value of the described surficial changes in the ice, it can be used
salinity at the bottom of the lake is 12%, but during as a warning indicator that a flooding and/or a volcanic
the summer months the meltwater flushes the salt- event may soon occur. As an example, rapid
water out of the lake, thereby decreasing the salinity (50 cm day−1) observations of glacial surface subsi-
to 5%. As previously mentioned, because the coast- dence above the Bárdarbunga cauldron just before the
line is continually eroding the distance between caldera collapsed on 16 August 2014. At Vatnajökull
Jökulsárlón and the ocean is decreasing, resulting in there are three known active subglacial lakes:
increasing amounts of seawater being able to flow Grímvotn, Kverjökull, and Barðarbunga (Figure 14.1).
into the lake at high tide [Landl et al., 2003]. This is Of the three, Grímsvötn has the largest subglacial
a concern as the denser seawater sinks below the less lake with a caldera 10 km in diameter and depth of
dense lake water and density driven systems can 300 m. The caldera is confined by Mount Grímsfjall
cause stratification in the water column, which can on the south and west, but the lake can expand to the
affect flow conditions. Ultimately, the connection north and northeast as the water level rises. The
with the sea, or lack thereof, affects the appearance volcanic structure of Grímsvötn creates a localized
of the lakes. For example, Jökulsárlón is clear topographic high for the ice cap, thus causing a dome
because suspended sediment is continually moved effect that will influence drainage patterns. The
out of the basin by the tide, whereas Breiðárlón and majority of the subglacial flow drains south onto the
Fjallsarlón are turbid. Skeiðarársandur plain (Figure 14.1). Three other
smaller ice cauldrons (i.e., Skaftár cauldrons) in the
14.3. LAKES UNDER THE ICE: GRÍMSVÖTN, Grímsvötn volcanic zone, however, drain elsewhere;
KVERJÖKULL, AND BARÐARBUNGA two on the eastern side drain to the River Skaftá and
the westernmost ice cauldron drains to proglacial Lake
Radio echo soundings first revealed several surface Hamarslón, which feeds River Kaldakvísl [Björnsson,
depressions on Vatnajökull that are created by subgla- 1982, 1988]. Today, the ice above the Grímsotn lake is
cial geothermal activity and are underlain by water‐ ~260 m thick and has grown continuously in size and
filled vaults that accumulate until drained by a volume since the 1960s owing to an overall decrease in
jökulhlaup [Björnsson, 1974, 1975, 1977, 1980]. As area geothermal activity [Björnsson and Einarsson,
described by Björnsson [2003] the release of meltwater 1990; Björnsson and Gudmundsson, 1993]. In turn, the
from subglacial lakes can occur by three different total volume of water discharging from the lake has
mechanisms: (a) rapid meltwater accumulation via also decreased over the past 60 years.
proximal volcanic activity; (b) the lake level can rise In historic times, the Grímsvötn volcanic zone has
until the ice dam is penetrated; (c) drainage can begin had the highest frequency of eruptions in Iceland, with
on the lake bottom when the water begins to melt con- at least 40 eruptions occurring since 905 CE
duits that slowly expand with friction and contact with [Thóraarinsson, 1974; Björnsson and Einarsson, 1990].
water. Although a stratified temperature column is Of these events, half have been accompanied by a
typically observed in a subglacial lake, the temperature jökulhlaup [Björnsson and Einarsson, 1990]. As described
Glacial Features of the South and Southeastern Regions  153

Figure 14.2 Formation of ice cauldron: a depression at the surface of a glacier can be formed when ice
melts at the glacier base to form a subglacial lake. [Modified from Bjornsson [2003]; design credit Nathan
Mennen.]

Figure 14.3 Discharge relating to jökulhlaups as distributed onto the Skeiđararsandur. [Modified from
Bjornsson [2017] design credit Nathan Mennen.]

by Björnsson [2003], in general, the longer the subgla- [Björnsson, 2003]. Most surprising was a tsunami‐type
cial eruption the more meltwater will be produced. wave that inundated Skeiðarársandur after 10.5 h of
Caldera eruptions do not cause jökulhlaups because initial meltwater release and caused a washout of the
the water level does not rise in the lake when the floating main road (Photo 47). The peak d ­ ischarge reached
ice cover melts. At Grimsvotn, however, nearby fissure 4 × 104 m3 s−1 in 16 h and fell to zero 27 h later [Björnsson,
eruptions under >500 m of ice have caused meltwater 2003]. Also unusual, this rapid discharge of water initi-
to quickly fill up the caldera lake. This is what hap- ated surficial changes to the ice sheet whereby the roof
pened at Grímsvötn in 1996 to produce the most rapid of a subglacial tunnel collapsed resulting in a canyon‐
jökulhlaup ever recorded when 3.2 km3 of water drained like feature 100 m deep, 800 m wide, and 6 km long
from the lake within a period of 40 h (Figure 14.3) [Aðalgeirsdóttir et al., 2000].
154 GLACIAL FEATURES

Volcanic‐induced flooding, however, is not the pri- case, the major rivers (e.g., Jökulsá á Fjöllum) flow
mary mechanism at Grímvötn. Instead, geothermal north to provide a major pulse of freshwater and sedi-
heating continuously generates meltwater within the ment onto the North Icelandic Shelf and Arctic Ocean.
subglacial lake. At Grímsvötn, once the meltwater Historically, Bárdarbunga is the second most active
reaches a critical level, typically when it rises 80–110 m, volcano in Iceland with about five eruptions every 100
it will breach the basin, thereby producing a flood years since 7.6 ka, as identified from the 87 tephra
event (Figure 14.4) [Björnsson, 2002]. The initial emp- layers around Vatnajökull [Thordarson and Larsen,
tying of the lake is accompanied by icequakes, subsi- 2007; Óladóttir et al., 2011]. At least 13 jökulhlaups
dence of the ice cap above the lake, and a sulfurous relating to Bárdarbunga have been identified in the
odor at the glacial margin where the water is discharged depositional record since 900 CE [Björnsson and
[Björnsson, 2002]. The majority of the Grímsvötn Einarsson, 1990]. Its radial drainage pattern enables
jökulhlaups will flow ~50 km beneath the glacier to meltwater to flow in all directions off the volcanic
reach the outwash plain at Skeiðarársandur. After the peak (2118 m), and thence to enter the Jökulsá á
flooding event, systematically, the remaining ice will Fjöllum, Skjálfandafljót, Köldukvísl, Tungnaá, and
dam the flow and the cycle will start again. At Skaftá rivers, or even find subglacial routes towards
Grímsvötn, the subglacial lake tends to regenerate Grímsvötn and out to the Skeiðarársandur, as was
itself every 5 years, leading to another flood event observed in 1996 [Björnsson, 2017].
[Björnsson, 1992, 1998; Gudmundsson et al., 1995]. The Kverkfjöll is further away from the hot spot and
flooding events can last between 2 days and 4 weeks therefore has been relatively docile volcanically during
with peak discharge ranging from 600 to 4 × 104 m3 s−1 historic time; zero to three eruptions every100 years. it
at the terminus. The shorter, outburst floods discharge has two relatively small calderas (8 km long and 5 km
through a subglacial tunnel at the glacier’s base. If the wide), however, where vigorous geothermal activity is
discharge exceeds 3000 m3 s−1 the extra meltwater will suppressed under the Gengissig (100 m deep) and
be accommodated by other tunnels at the central part Galtarlón (500 m deep) ice‐dammed lakes, called. As
of the terminus and join the River Gígja. both Bárdabunga and Kverkfjöll are on the northern
According to Björnsson [2002], the largest and most side of Vatnajökull where the drainage divide directs
catastrophic jökulhlaups in Iceland may be caused by water toward the Central Highlands, it is likely that
eruptions in the voluminous ice‐filled calderas of subglacial eruptions of either could produce associ-
Bárdarbunga (64.6411, −17.5280) and Kverkfjöll ated flooding events via the River Jökulsá á Fjöllum.
(64.6725, −16.6902) (Figures 2.1 and 14.1). This is Carrivick and Twigg [2005] used geomorphological
­particularly important to discussions of past climate and sedimentological evidence at the Kverkfjöll ice
change as it is thought that jökulhlaups that have a margin to identify possible jökulhlaup flowpaths. This
capacity of >1.0 × 106 m3 s−1 could have the potential included erosional features (gorges, cataracts, spill-
to trigger temporary, but abrupt, changes in climate as ways) and depositional features (wash limits, boulder
seen by events disrupting the North Atlantic thermo- bars, terraces) found in the proglacial area of
haline circulation [Manabe and Stouffer, 1995; Kverkfjallarani (64.7755, −16.4866). Carrivick et al.
Rahmstorf, 1997; Carrivick and Twigg, 2005]. In this [2004] have compared these geomorphologic features,

Figure 14.4 Changes in water levels in Lake Grímsvötn. [Modified from Bjornsson [2017]; design credit
Nathan Mennen.]
Glacial Features of the South and Southeastern Regions  155

which often relate to intense flooding events, with 14.4. MÝRDALSJÖKULL


those related to well‐documented jökulhlaups in
Iceland. They found that the Kverkfjöll flooding events Mýrdalsjökull is the fourth largest ice cap in Iceland
tend to behave as more turbulent and supercritical (590 km2) rising to an elevation of 1480 m a.s.l. Today,
flows, with higher amounts of sediment concentra- only about 20% of the ice cap lies between the firn line
tions and spatially variable hydraulic properties. In of 1000 m and 1200 m, which largely influences the
addition, from stratigraphical relationships between amount of meltwater discharge at its steep outlet gla-
jökulhlaup surfaces and other landforms at Kverkfjöll ciers (Figure 12.8) [Björnsson, 2017] .This ice cap situ-
they concluded that there have been at least three ated at the southern end of the EVZ may best be
events during the Holocene [Carrivick et al., 2004]. As known for the Katla volcano that resides beneath it
described by Carrivick et al. [2002] the individual flood (section 9.2). The 700 m deep caldera is centered under
deposits are up to 50 cm thick and moderately well Mýrdalsjökull and spans 10 km in diameter. Owing to
sorted. Similarly, Helgason [1987] suggested as many the shallowness of Katla’s magma chamber (<3 km),
as four paleojökulhlaups with discharges upwards of hydrothermal activity is common as a result of the
4 × 106 m3 s−1. From historic records, Thórarinsson interaction of ice with the upwelling magma.
[1959] suggested many smaller magnitude events have Few would refute that Katla, although hidden from
occurred and more recently in 2002 a jökulhlaup was view beneath 400 m of ice, is the most dangerous vol-
observed on Jökulsá á Fjöllum originating from cano in Iceland. The glaciated Katla volcano has
Kverkfjöll [Sigurðsson et al., 1992]. exploded 20 times since the settlement of the Víkings
The question remains as to whether or not meltwater in 874 CE [Larsen, 2000]. With a recurrence interval of
events at Bárðarbunga or Kverkfjöll could have had two eruptions per century, the Katla volcano delivers
the capacity to erode the Jökulsárgljúfur canyon at some of the largest ever recorded releases in terms of
~9 ka (section 5.4, Figure 5.3, Photo 14). According to volumes of ash (Figures 9.5 and 9.6) and discharge
Carrivick [2007], the Holocene flooding at Kverkfjöll produced (i.e., 1.0 × 105 to 3.0 × 105 m3 s−1 for the 1918
was comparable to the Missoula, Bonneville, and Altai jökulhlaup event) [Björnsson, 2003; Tómasson, 2002;
“megafloods” (Table 14.2). Although the comparative Russell et al., 2010]. Katla, however, is different to
study shows that the flow depth, flow width, and dis- those calderas that host glacial lakes like Grimsvotn,
charges were orders of magnitudes less than the floods Bardabunga, and Kverkfjöll (section 14.3). According
mentioned, the geomorphology of the area seemingly to Björnsson [2003], at Katla faults within the caldera
made up for it. Specifically, if the source water was facilitate rapid vertical transport of meltwater before it
generated at the high‐elevation (1600 m) cauldron(s) is able to collect in the depression; although there are
and then descended down the steep flanking sides 12 small (20–50 m deep, 500–1000 m wide) ice caul-
(900 m over 6 km) of the Kverkfjöll stratovolcano in a drons on Mýrdalsjökull where water can accumulate
confined canyon (like Hraundalur) then it may be pos- over time (Figure 14.5). The cauldrons are unequally
sible. Equally, the Bárdarbunga caldera is known to distributed over the three major catchment basins that
produce a greater volume of meltwater due to its sheer supply water to the outwash plains Mýrdalssandur,
size (caldera width 10 km, 700 m deep) and minimum Sólheimasandí, and Markarfljót (Figure 14.5).
850 m thick ice cover. Although the meltwater accumulates in the cauldrons,

Table 14.2 Comparison of the hydraulic parameters of the largest Kverkfjöll jökulhlaup with recent estimates of the
largest known terrestrial megafloods
Missoula flood Bonneville flood Altai flood Kverkfjöll Jökulhlaup
Width (km) 6 NR 2.5 1.6
Depth (m) 150 156 400–500 5.9
Velocity (m s−1) 25 41 20–45 15
Discharge (m3 s−1) 17 × 106 1 × 106 18 × 106 2 × 105
Bed Shear Stress (N m−2) 1 × 104 2.9 × 103 2 × 105 1 × 104
Stream power (W m−2) 2.5 × 105 1.2 × 105 1 × 106 1 × 105
Channel Slope 0.01 0.0095 0.01 0.02–0.36
Roughness (Manning’s n) 0.04–0.1 0.03–0.006 0.03–0.07 0.03–0.07
Note. Modified from Carrivick [2007]; design credit Nathan Mennen.
156 GLACIAL FEATURES

Figure 14.5 Details of features on Mýrdalsjökull. [Modified from Bjornsson [2003]; design credit Nathan
Mennen.]

eventually draining as small jökulhlaup events glacier spanning 12 km across. Uniquely, those surges
[Björnsson et al., 2000; Russell et al., 2000], a persistent observed on the northern side of Mýrdalsjökull ice cap
hydrogen sulfide odor at Sólheimasandí also indicates appear to be generated as localized movements starting
that meltwater is continuously released there. The mas- near the glacier snout. In turn, deformation of the ice was
sive meltwater releases that Katla is known for then are not seen throughout the entire outlet glacier or near the
initiated by powerful eruptions that can break through central ice divides. This may be attributable to the
the 400 m of ice cover within 1–2 h [Björnsson, 2003] basement substrate beneath Sléttjökull and Öldufellsjökull
to produce water‐transported volcanic debris up to being porous lavas [Björnsson et al., 2003].
2 × 109 tons per event [Tómasson, 1996; Larsen, 2000]. Sólheimajökull and Kötlujökull are steep valley gla-
Additionally, computer models predict that the volume ciers that flow off the caldera and emerge onto the
of water produced by glacial outbursts could generate southern coastline. These two outlet glaciers hace been
tsunami‐like waves entering into the North Atlantic studied extensively because over the past six centuries
and traveling towards the Vestmannaeyjar Islands with all major jökulhlaups generated from Katla eruptions
wave heights of 2–3 m [Elíasson, 2008]. have been released through them [Sigurðsson, 1998;
Mýrdalsjökull has four notable outlet glaciers: Russell et al., 2000] (Figure 9.6). As an example, the last
Sólheimajökull (44 km2; 63.5569, −19.3028), Kötlujökull eruption in 1918 transported enough sediment to
(133 km2, 63.5661, −18.8672), Sléttjökull (72 km2; extend the coastline of Mýrdalssandur by 3.0 km sea-
63.7202, −19.0188) and Öldufellsjökull (40 km2; 63.5969, ward [Tómasson, 1996]. A similar process of meltwater
−18.8811) (Figure 14.5). Of these, the gently sloping generation occurs at these lobe systems as described at
northern outlets of Sléttjökull and Öldufellsjökull are Breiðamerkurjökull in section 14.2, which also termi-
considered surge‐type glaciers, with both advancing in nates on the southern coastline. At these outlet glaciers,
1992; surges on Öldufellsjökull have occurred approxi- 25% of the meltwater is generated at higher elevations
mately every 10 years since 1974. The more frequent within the accumulation zone (Figure 12.8) thus pro-
advance of Öldufellsjökull is likely due to its confinement ducing large amounts of water to be transported great
in a valley, whereas Sléttjökull is a broad‐lobed piedmont distances downgradient to the ablation areas.
Glacial Features of the South and Southeastern Regions  157

Like outlet glaciers at Vatnajökull, Sólheimajökull ocean. The Eyjafjöll sandur plain separates the cliff
has had an extreme history of glacial advance and face from the ocean. Additionally, there are 10 small
retreat since 1960. In particular, between 1960 and outlet glaciers that descend from 1200 to 900 m, of
1996 the glacier advanced 400 m and increased in area which the two most studied are Gigjökull (7.5 km2;
by 0.61 km2, but has since retreated 450 m [Staines 63.6256, −19.6790) and Steinholtsjökull (3 km2;
et al., 2015]. The influence of glacier advance and 63.5941, −19.5333). Gigjökull originates from the sum-
retreat on proglacial landscapes in this area has been mit where a breach in the caldera wall creates an 800 m
documented to reveal the significant interplay between gap for ice to flow downgradient; the average slope
glacial meltwater runoff regimes, glacier terminus changes from 14° between 1000 and 800 m, to 26° from
position, sediment supply, and episodic fluvial flood- 800 to 200 m. Gígjökulslón proglacial lake is situated
ing events [Staines et al., 2015]. The proglacial systems in front of the glacier snout. Similar to Mýrdalsjökull,
of Sólheimajökull were relatively stable from 1960– the firn line is ~1100 m, but has a greater impact on this
1999 before the area was disturbed by a jökulhlaup glacier due to its lower overall elevation. The east–west
event. The flooding that took place in 1999 along the trending Fimmvörðuháls ridge (1100 m, 63.6316,
River Jökulsá smoothed the longitudinal profile of −19.4420) connects the Eyjafjallajökull and
meltwater channels and changed both the channel Mýrdalsjökull ice caps. Although Fimmvörðuháls
slope and reach‐based volume. Staines et al. [2015] maintained a glaciated surface throughout the 19th
also highlight increased channel braiding and down- century, current warming climates and ascending firn
stream incision from 2001 to 2010 as a result of line make this glacier unstable and consequently is an
increased stream power due to glacial ablation and a area closely monitored by glaciologists.
reduction of sediment supply with the development of The most recent and well‐documented eruptive activity
an ice‐marginal lake. Likewise, as a result of the sedi- of Eyjafjallajökull began on 20 March 2010 with a lat-
ment starvation the number of channel bars decreased eral ejection from the main caldera and, concurrently, a
by 50% [Staines et al., 2015]. Russell [2018] considered 500 m fissure opened up in an ice‐free area on
rates of incision and aggradation of the River Jökulsá Fimmvörðuháls 2–3 km away. As described by Oddson
during the LIA and although many scenarios could et al. [2016], an early indicator was the formation of ice
have evolved, they suggest that during periods of gla- cauldrons around the active vents in the western part of
cial advance damming effects of proglacial moraine the ice‐filled caldera where the ice was 200 m thick.
and ridges can place constraints on channelized water Subsequently, from 18 April to 4 May, a subglacial lava
and sediment flows thereby also reducing aggradation. flow 3.2 km long (0.55 km2) occurred on the northern
This is important because glaciofluvial deposits in the flanks of Gígjökull. As described by Oddson et al. [2016],
area prior to the LIA may not have been buried or during the first 6 days the subglacial lava flow piled
eroded from the paleorecord. 80–100 m thick below the ice and advanced at
<100 m day–1. By day 7, the surface ice above the lava
14.5. EYJAFJALLAJÖKULL flow completely melted and an ice canyon formed on
Gígjökull. Uniquely, the ice–magma contact produced a
Eyjafjallajökull (63.6200, −19.6133), a smaller very fine trachyandesitic ash that became a hazard for air
(80 km2) ice‐capped stratovolcano to the west of travel as it became entrained in the eastward jet stream
Mýrdalsjökull, is also located in the EVZ (Figure 2.1, towards Europe (Figure 9.4) [Gislason et al., 2011].
section 9.2). The caldera basin is 2.5 km in diameter The distribution of meltwater during this event was
and is covered by a maximum glacial thickness of unusual because five new flowpaths were formed in
400 m in the eastern part, but about 200 m in the west- areas that had pre‐existing crevasses above bedrock
ern part (where the 2010 eruption took place). The cal- undulations [Oddson et al., 2016]. As suggested by
dera has two main peaks: Hámundur (1650 m; 63.6205, Magnússon et al. [2012], this likely occurred because
−19.6108) and Guðnasteinn (1500 m, 63.6366, the event began during the Icelandic winter season
−19.6113). Steep cliff faces present on the southside of when the main meltwater channels under Gígjökull
Eyjafjallajökull provide evidence of wave‐scoured were still blocked with ice. Episodically, as the melt-
benches denoting sea levels during the Holocene water began to infill the caldera basin, it was released
[Björnsson, 2017]. Three popular waterfalls, Skógafoss in hyperconcentrated flows of pyroclastic mixtures
(63.5319, −19.5113), Seljalandsfoss (63.6188, beneath Gígjökull to inundate Gígjökulslón proglacial
−19.9883), and Gljúfuráfoss (63.6841, −19.9750), flow lake before being directed to the Markarfljót sandur
down these cliff faces, which are now ~5 km from the plain. The largest of the jökulhlaup events occurred
158 GLACIAL FEATURES

Photo 51 Farmstead at the base of Eyjafjallajökull during the 2010 eruption. [Courtesy of Arctic Images/
Alamy Stock Photo.]

during the first 3 days of subglacial lava movement Three other Holocene flooding events in
and waned in discharge and sediment transport there- the Markarfljót Valley have been described and are
after. The total volume of ice melt is estimated to be thought to have had a dominant role in creating the
10–13 × 107 m3 [Odsson et al., 2016] and the peak dis- terraces and surface sediments in the area [Smith and
charge monitored during the event was ~2000 m3 s−1 Dugmore, 2006]. Additionally, the jökulhlaup events
[Björnsson, 2017]. About 800 inhabitants of the local that occurred after human settlement likely had a role
farmsteads (Photo 51) were evacuated during this in deciding the lack of settlement on the northern side
event and no lives were lost. of Eyjafjallajökull.
15
Glacial Features of the Northern and Western Regions

The complicated story of Iceland during the LGM 3. The Langhóll Stadial ~8000 BCE, which represents a
continues with the exploration of Iceland’s peninsulas readvance of the ice sheet in the valleys adjacent to the
in the north (Vestfirðir, Skagi, and Tröllaskagi) and Eyjafjördur (Figure 15.1). Brynjólfsson et al. [2015] dated
west (Snæfellsnes) (Figure 15.1). The pattern of degla­ an outer moraine at the Leirufjördur (Figure 15.1) using
ciation from the LGM ice sheet maximum, especially
36
Cl to reveal a similar age. They suggested that ice‐sheet
in the north, is not well constrained spatially as glacial growth was a response to a cooler climate forced by a
geological studies that provided indirect dating from reduction of the Atlantic Meridional Overturning
sediment cores lack chronological control due to the Circulation (AMOC). The AMOC supplies the higher
absence of volcanic ash in the record to provide marker latitudes of the Atlantic with warmer surface ocean water
beds [Geirsdóttir et al., 2007, 2009,]. Hence, challeng­ which makes the North Atlantic climate relatively mild,
considering its high latitude. This final, although local­
ing questions remain about the configuration, thick­
ized, stadial event marked the termination of the LGM in
ness, and ice dynamics of the area. The situation has
northern Iceland.
been improved upon with the application of cosmo­
There is an ongoing debate as to whether the LGM
genic exposure dating, which estimates how long a
ice sheet extended to cover the whole of Vestfirðir
rock surface has been exposed to cosmic radiation on
Peninsula or if the peninsula maintained an
the Earth’s surface. Here it should be noted that gener­
independent ice cap, similar to Öræfajökull at present
ally 36Cl is the only nuclide applied on Icelandic basalt
(section 14.4, Figure 14.5). Furthermore, glacial con­
except if it contains substantial olivine phenocrysts
ditions in northern Iceland in the LGM are also
that enable 3He to be applied.
debated, which is important to the discussion as to
Norðdahl [1981] divides the Weichselian glaciation in
whether the ice sheet was cold based and nonerosive or
northern Iceland into three main stages.
warm based with movement induced by basal sliding;
1. The maximum stage when the ice sheet covered the
at present all Icelandic glaciers are considered warm
island of Grímsey (66.5423, −17.9961) with 100 m
(minimum) of ice. This remote island is 40 km north of based and all have surge‐type outlet glaciers [Björnsson
Iceland marks the maximum ice‐sheet extent, as evidenced and Pálsson, 2008; Brynjólfsson, 2015]. Quantification
by glacial striae and till described by Hoppe [1968]. of geomorphological erosional patterns and density of
2. The ice‐lake stage when a series of ice‐dammed ice‐scour lakes found on Vestfirðir Peninsula by
lakes formed in Fnjóskadalur (65.7563, −17.8994) dur­ Principato and Johnson [2009] support previous views
ing a period of glacial retreat. This acts as an indicator [Hoppe, 1982; Norðahl, 1990] that there were two
as the formation of lakes dammed by ice is usually a independent ice sheets present on Iceland during the
result of the thickening or advance of the damming ice LGM: one forming on the mainland and one forming
margin, which acts as a physical barrier whereby trap­ on Vestfirðir Peninsula. Specifically, glacially scoured
ping meltwater. and nonscoured landscapes represent the presence or

Iceland: Tectonics,Volcanics, and Glacial Features, Geophysical Monograph 247, First Edition. Tamie J. Jovanelly.
© 2020 American Geophysical Union. Published 2020 by John Wiley & Sons, Inc.

159
160 GLACIAL FEATURES

Figure 15.1 Location of Vesfird ī r, Snæfellsnes, Skagi, and Trollaskagi Peninsulas with associated fjord
l­ocations. [Design credit Nathan Mennen.]

absence, respectively, of moving ice, and the density of that the LGM ice‐sheet extended into the Arctic Ocean,
lake basins are commonly used as a proxy for the ~90 km from the Eyjafjördur (Figure 15.1). Optimal
occurrence of glacial scouring. Norðahl [1990] initially LGM modeling completed by Hubbard et al. [2006]
proposed localized and alpine glaciation centers in suggests that the overextension of the ice sheet across
northern Iceland (i.e., Tröllaskagi and Flateyjarskagi the northern shelf (Figure 2.2) occurred due to extreme
mountains) on the basis of: (a) trimlines that distin­ aridity across the region and that further advance may
guish ice‐free zones from nunataks; and (b) areas with have been limited by the topography of the continental
dendritic glacial drainage patterns could not have been shelf‐break.
occupied by glaciers from an inland ice sheet. In these The latitudinal position of the Polar Front continu­
northern mountain plateaus ~24 ka ice thinning and ally fluctuated throughout the Holocene, and conse­
deglaciation likely began and continued gradually, quently both the cooler waters of the sea‐ice‐bearing
albeit asynchronously, until 12 ka; the majority of the Arctic Ocean and the warm Atlantic currents have
outlets at Iceland’s northernmost glacier, Drangajökull, influenced Iceland’s climate (Figure 15.2). Although at
were occupied until 7 ka [Brynjólfsson, 2015]. present the Irminger Current dominates surface waters
If, and the extent to which, parts of Iceland were ice‐ along the northern and western coastlines, the winds
free ~21 Ma has implications for estimates of the affecting these peninsulas mostly come out of the
overall thickness of the ice sheet covering Iceland. northeast where cooler, polar waters dominate, thus
Hubbard et al. [2006] estimated that the mean thick­ making nearby sea‐surface temperatures ~2°C cooler
ness of the ice sheet was 940 m, with a plateau eleva­ than those of the south at any given time of the year.
tion of ~2000 m that was likely breached by nunataks, Historically, this has translated to an overall thinner
as described by Norðahl [1990]. Offshore evidence layer of glacial ice in the northern and western
acquired by seismic stratigraphy and accelerator mass portions of Iceland due to less precipitation
spectrometry radiocarbon dating by Andrews et al. (~3000 mm year−1) from the direction of the Arctic
[2000] and Norðdahl and Pétursson [2005] c­onfirms Circle than that from the warmer waters of the North
Glacial Features of the Northern and Western Regions  161

Figure 15.2 Present ocean current systems around Iceland. [Design credit Nathan Mennen.]

Atlantic. Additionally, despite cold periods persisting natural sites of net sediment accumulation. Sedimentary
through to the 20th century, such as the LIA, precipi­ infill of fjords, however, is relatively absent in the north
tation in the northern and western regions was not and west of Iceland. Unlike on the southern coast
sufficient to maintain glacier expansion as it did in the (section 14.3, Figures 9.6 and 14.2), the fjords of the
south at Vatnajökull [Björnsson, 2017]. Due to the northern peninsulas described here are not located at
colder air temperatures in the western fjords, however, or near active volcanic zones, with Vestfirðir penin­
glacier limits extend down to 600–700 m altitude, sula being the furthest away at more than 150 km.
which are the lowest in the country [Björnsson and Thus there is an absence of sediment‐laden jökulhlaup
Pálsson, 2008]. events to transport sediment down glacial valleys to
Major differences between the landscapes of the the coast, hence the low sediment supply to these
northern and western coasts versus those of the fjords. Often an underwater ridge at the mouth of
southern coast are the presence of glacial fjords (gen­ the fjord forms a subbasin in front of the glacial toe,
erally <50 m deep) and the absence of proglacial lakes which acts as a shallow barrier to the open ocean
and sandur plains (Photo 40). A fjord is formed when (Figure 15.3). This barrier forms when the glacier
a glaciated U‐shaped valley that is carved below sea ­terminates at the ocean, reducing its ability to erode.
level becomes flooded during, and after, ice retreat Often, deposition of glacial sediment on top of the
[Martini et al., 2001]. As a geomorphological feature, bedrock ridge can cause it to heighten over time,
fjords are considered immature, nonsteady‐state sys­ which may provide challenges to large ship traffic.
tems, evolving and changing over relatively short time­ Additionally, icebergs can be trapped behind the ridge.
scales; like those formed here since the LGM [Syvitski As with all estuaries, fjords are transitional regions
et al., 1987]. Additionally, as Syvitski et al. [1987] between land and open ocean, such that stratification
describes, whether presently associated with glaciers or of the water column can result as freshwater from
not, fjords are also immature estuaries, and hence glacial melt cannot easily mix.
162 GLACIAL FEATURES

Figure 15.3 Longitudinal profile of a glacial fjord. [After Martini et al. [2001]; design credit Nathan
Mennen.]

Also cirques, arȇtes, and nunataks are more com­ saline Irminger Current follows the west coast, whereas
monly found on the northern and western peninsulas the cooler, lower‐salinity East Greenland and East
(Photo 52), where they are accentuated by the area’s Iceland currents control the waters to the north and
climatic regime. Snow can accumulate in valleys through east (Figure 15.2). Vestfirðir Peninsula is known for its
avalanches or snow‐blow drifts and begin to pile up in 30 fjords formed during past glaciations; the two fjords
valley ravines. Increases in summer temperatures do not on the east and west of the isthmus are Gilsfjörður and
cause ablation so readily in these deep valleys, especially Bitrufjörður. The largest fjord is called Ísafjarðardjúp
if they are facing northward where they are shaded from (66.1200, −23.0763), which dissects the center of the
direct sunlight. Perennial snowfields typically occur on Vestfirðir Peninsula and is thought to have contained
the leeward sides of mountains or ridges sheltered from an ice stream during the LGM [Bourgeois et al., 2000;
the northeasterly winds across each peninsula. These Geirsdóttir et al., 2002] (Figure 15.1). This fjord sep­
features are found on the northern tip of Vestfirðir arates two upland plateaus with mean elevations
Peninsula at Hornstranðir (66.3719, −22.5583), and between ~400 and ~800 m a.s.l., with peaks at
about 150 valley and cirque glaciers are located on the 1200 m a.s.l. As described by Principato and Johnson
Tröllaskagi Peninsula in central north Iceland. One of [2009], the geomorphology of the uplands, valleys, and
the most photographed locations in Iceland is Kirkjufell fjords on Vestfirðir Peninsula is typical of Icelandic
Mountain (463 m, 64.9422, −23.3061), a nunatak on the glacial terrain, which is an area of ice accumulation on
Snæfellsnes Peninsula near Grundarfjodur (Figure 15.1). a high plateau that drains down steep icefalls into
Here, during the LGM, two glaciers converged to erode valleys that dissect the plateau. There are 10 cirque gla­
either side of the steep mountain that now survives. ciers on the peninsula rising to a height of 600–700 m.
The four major areas important to understanding Included in this number is the Drangajökull ice cap
past glacial activity in the north and west are (66.1494, −22.2497) on the eastern plateau. There are
the,Vestfirðir (22,271 km2), Tröllaskagi, (1,622 km2), three nunataks rising through the Drangajökull ice
Skagi (1,300 km2), and Snæfellsnes (2,540 km2; cap: Hrolleifsborg (851 m), Hljóðabunga (825 m), and
Figure 15.1) peninsulas. Reyðarbunga (778 m). The highest portion of the
western plateau at Glama (65.8402, −22.9961) does
15.1. VESTFIRÐIR PENINSULA not currently have permanent ice. Vestfirðir Peninsula,
AND DRANGAJÖKULL also referred to as the Western Fjords of Iceland, is
among the most northern latitudes of the island and is
Vestfirðir Peninsula is connected to the mainland by separated from Greenland by the Denmark Strait.
an isthmus that is less than 10 km wide but it compli­ Historic data of glacial activity has been chronicled
cates the passage of ocean currents. The warm and in church books and local annuals at Drangajökull for
Glacial Features of the Northern and Western Regions  163

Photo 52 Glacial nunatuk Kirkjufell on the Snæfellsnes Peninsula near Grundarfjordur. [Courtesy of Tamie
Jovanelly.]

the past three centuries. This archive has proven to be and Schoof [2016]. The subsequent Holocene history
important in discussions of climate change near the of the Drangajökull ice cap is poorly known. It is
Arctic Circle, as it has documented unique surging thought that the melting of ice shelf areas off of
behavior different to all other ice caps in Iceland. Vestfirðir Peninsula was likely synchronous with that
Additionally, due to the remoteness of Drangajökull, of the Iceland ice sheet as a whole, in response to
few field studies had been performed until more recent warming ocean currents and rapidly rising sea levels
years, so there is heavy reliance on human documenta­ [Norðdahl and Ingólfsson, 2015]. Records of IRD in
tion in this area of the country. sediment cores collected in the fjords indicate contin­
Iceland’s northernmost ice cap, Drangajökull (925 uous deposition until ~10,000 BCE when the ice sheet
m; 66.1500, −22.2500), is located on the eastern high­ retreated onto land [Geirsdóttir et al., 2002]. This cor­
land plateau of the Vestfirðir Peninsula and has had a responds to data suggesting ice streams flowing out of
dynamic history since the Late Weichselian glacial major fjords, like Isafjardardjúp, and then perhaps
stage to present (Figure 15.4). According to reaching as far as the submarine trench, Djúpáll
Brynjólfsson [2015], during the LGM the mountains [Stokes and Clark, 2001]. Dating of the Saksunarvatn
beneath Drangajökull were likely covered by cold‐ tephra in Lake Skoravatn (177 m a.s.l.; 66.2563,
based nonerosive sectors of the ice sheet, while fjords −22.3244) at ~8000 BCE indicates that at this time the
and valleys were occupied by warm‐based ice prone to northern limit of Drangajökull had receded to the size
erosive basal sliding. Hence, the ice sheet is considered seen at present, or smaller [Schomacker et al., 2016].
to have been polythermal in that it contained both cold This is significant because it suggests that unlike
ice below its pressure melting temperature, and Langjökull (section 13.1) and Vatnajökull (section 14.1)
temperate ice, which is a two‐phase mixture of ice and in the south, which were considerably smaller by
water at the melting temperature, as defined by Hewitt ~7000 BCE, Drangajökull maintained its size into the
164 GLACIAL FEATURES

Figure 15.4 Details and location of Drangajökull ice cap. [Modified from Bjornsson and Palsson [2008];
design credit Nathan Mennen.]

Holocene Thermal Maximum. Harning et al. [2016] Brynjólfsson, 2015]. Snow accumulation in the west,
produced a multiproxy record dating back to 3 ka. however, is the result of the accumulation of drifting
They dated terminal moraines, lacustrine deposits snow driven by northeasterly winds. Also unique to
from proglacial lakes (e.g., Skeifuvatn, Djúpipollur, Drangajökull ice cap is the fact that it is the only one
Ljótárvatn), and vegetation that had newly emerged in Icelandic to lie completely below an altitude of
during Drangajökull’s present recession in order to 1000 m, terminating just below 200 m; the average
determine recent growth and decline of the ice sheet. equilibrium line altitude (ELA) is between 550–
Their research revealed that the ice cap went through 650 m a.s.l. (Figure 12.8); ELA limits at other Icelandic
numerous stages of advance (~320 BCE, 180 CE, ice caps usually lie above 1100 m a.s.l. To date, 70% of
560 CE, 950 CE, 1400 CE to LIA) and retreat the ice cap’s surface area is above 600 m a.s.l. It is the
(~450 CE, 1250 CE, 1850 CE). Through geomorpho­ overall colder temperatures on Vestfirðir Peninsula
logical mapping and the use of aerial photography it (annual average temperature 2.5–4°C; summer 6–8°C)
has been estimated that the size of the ice sheet in the due to its proximity to Greenland and the cold polar
LIA maximum was approximately 25–50% larger than East Greenland Current (Figure 15.2), its low insola­
today [Sigurðsson et al., 2013; Harning et al., 2016]. tion during summer months, and high precipitation in
At present, the dome‐shaped Drangajökull ice cap is the area, which have sustained the ice cap at lower alti­
Iceland’s fifth largest by area at 160 km2 and has a tudes. Together with the Breiðamerkurjökull outlet of
maximum ice thickness of 284 m (at Kaldalónsjökull) Vatnajökull (Figure 12.10), Drangajökull extends
and an average ice thickness of 107 m [Magnússon closest to the sea of any of the country’s glaciers.
et al., 2016]. Interestingly, the majority of the ice (71%) Drangajökull has three major surging outlet glaciers
is stored on the western side of the ice cap. This is due that have been monitored by locals since 1886.
to unevenly distributed rainfall patterns and the domi­ Eythórsson (1935) began scientific monitoring in 1931
nance of northeasterly winds. The northeast coast of with the installation of permanent markers at the ter­
Vestfirðir Peninsula receives about 1100 mm annual mini of Leirufjarðarjökull (27 km2; 66.1588, −22.1569),
average precipitation, whereas the west coast of the ice Kaldalónsjökull (37 km2; 66.1936, −22.1903), and
cap receives about 580 mm, reflecting an orographic Reykjarðarjökull (22 km2; 66.1938, −22.1972;
effect on the peninsula [Crochet et al., 2007; Figure 15.4). In contrast to most other surge‐type
Glacial Features of the Northern and Western Regions  165

­ utlets from Iceland ice caps, those at Drangajökull


o Leirufjarðarjökull occurred with about 10‐year intervals
are confined within valleys, which affect the forefield during the period of ~1840 to 1898 CE, but140 years
geomorphology and can contain glaciofluvial land­ passed with no surge recorded from 1700 to 1840 CE,
forms, moraines (10–15 m high), flutes (up to 100 m which could reflect a lack of preservation rather than
long), and boulder fields (up to 1 m diameter) inactivity.
[Brynjólfsson et al., 2014]. The absence of other glacial Notwithstanding longer surge durations, the flow
structures typical of forefields (e.g., crevasse‐fill ridges, velocities of Svalbard glaciers are considerably (flow
eskers) is likely due to the impermeable Neogene pla­ rates at Svalbard glaciers have been observed to be bet­
teau basalt strata beneath the Drangajökull ice cap ween 0.1 m day−1 and 7.0 m day−1), i.e., at least 10
[Ingólfsson, 2013]. All three outlet glaciers surged asyn­ times, and as much as 100 times, faster than normal
chronously between 1700 and 1846 CE to meet their glaciers. As an example, the highest annual average
LIA maximum extents, which are well constrained by advance rate of the Leirufjarðarjökull margin was
end moraines and historic data around the northern measured in 1995 at 2.0 m day−1; the rate at
perimeter of the ice cap. At Kaldalónsjökull, rock Reykjarfjarðarjökull peaked at 0.2 m day−1 in 2003
exposure dating of the most distal end moraines [Sigurðsson, 1998]. This means that mass is transferred
yielded 36Cl ages of 9700 BCE, correlating them to the downglacier more quickly, and over a considerably
Younger Dryas Stadial [Principato et al., 2006]. longer period. Brynjólfsson [2015] observed a distinct
Measurements made from the maximum LIA terminal ice discharge pattern that occurs during surges at
moraines indicate an overall retreat between 3000 and Drangajökull, where the zones of accumulation
4000 m. undergo surface thinning between 10 and 30 m and the
According to Brynjólfsson [2015], the amount of ablation areas can thicken between 10 and 120 m
times the outlet glaciers have surged, the duration of (Figure 12.8). Here a comparison can be drawn to the
surging events, and the intervals between surge events 1963–1964 surge of Brúarjökull, an outlet glacier of
at the Drangajökull ice cap tend to resemble character­ Vatnajökull, that flowed up to 125 m day−1 (Chapter 14).
istics of Svalbard surge‐type glaciers (where 25% gla­ Although moving at a much faster pace, the event was
ciers are surging at speeds rarely seen in other parts of limited and induced by a jökulhlaup that would prove
the world) rather than the surges of the larger ice caps to be one of the largest recorded in Iceland.
in central and southern Iceland. The reasons for differ­ It is the long quiescent phases of these Svalbard
ences in surge propagation are still unknown, but surge‐type glaciers at Drangajökull that makes corre­
Brynjólfsson [2015] attributes the high alpine relief lating climate to this glacial landscape difficult. For
setting of northern Drangajökull compared to the flat this reason, Brynjólfsson [2015] does not indicate clear
highland plateau of Vatnajökull as a main contributor. relationships between surge initiation or periodicity
Categorically, all These three of outlet glaciers at and climate. Conversely, all outlets did increase surge
Drangajökull have been recognized historically as frequency during the 19th and early 20th centuries
surging between two and four times, however, compared with the relatively cool 18th century, and the
Brynjólfsson [2015] reconstructed five surges at warmer late 20th century. Again, it is worth noting
Reykjarðarjökull, six surges at Kaldalónsjökull, and that this may reflect incomplete data, rather than a
seven surges at Leirufjarðarjökull. The most recent lengthy pause in surge movement. At present the
events that have been monitored in detail occurred ­glaciers are thought to be in an inactive phase, with
from 1995 to 2000 at Leirufjarðarjökull and thickening of accumulation areas 0.5–0.7 m year−1
Kaldalónsjökull, and 2002–2006 at Reykjarðarjökull [Brynjólfsson, 2015]. According to Brynjólfsson [2015],
(Figure 15.5) [Sigurðsson, 1998, 2003]. Details of new contributions to reservoir areas may be steepening
advances and net losses for Leirufjarðarjökull, the overall profile of the glacier with the potential that
Kaldalónsjökull, and Reykjarðarjökull, respectively, this could initiate future surges, in 45–65 years time; an
are as follows: 1150 m, −0.205 ± 0.074 km3; 150 m, not unstable glacier could lead to hazardous snow ava­
available; 227 m, −0.076 ± 0.053 km3 [Brynjólfsson lanches or debris flows (Figure 15.6, see Box 15.1).
et al., 2016]. Also, rather than surge events lasting bet­ The area of Drangajökull, except the three surging
ween 1 and 2 years, these outlet glaciers experience outlets, appears to have been relatively stable during the
surge durations of 4–7 years with periods of inactivity past 50–80 years, which appears to be in stark contrast to
lasting much longer (50–140 years) than is considered the negative mass balance of Iceland’s other glaciers
typical behavior [Dowdeswell et al., 1991; Björnsson [Brynjólfsson et al., 2014]. Nonetheless, the total ice‐cap
and Pálsson, 2008]. As an example, surges at volume from 1946 to 2011 (about half of which was lost
166 GLACIAL FEATURES

Figure 15.5 Fluctuations of the three Drangajökull surge‐type outlet glaciers over recent times. [Modified
from Brynjólfsson et al. [2016]; design credit Nathan Mennen.]

Figure 15.6 Snow avalanche (left) and debris flow (right) maps of Vestfird ̄ir Peninsula. [Modified from
Decaulne [2007]; design credit Nathan Mennen.]

between 1994 and 2011) was reduced from 18 km3 to not match the trend of increasing air and sea‐surface
15 km3 [Magnússon et al., 2016], with a mean mass temperatures recorded in the Arctic. Starting with the
balance rate of ice‐cap depletion over those 65 years at estimated size of Drangajökull during the LIA it has
−0.25 ± 0.04 m.w.e. year−1. Magnússon et al. [2016] also been determined that the ice cap is thinning and its sur­
observed high decadal variability, however, with positive face area has decreased by ~55 km2 over the previous
mass balance rates in 1975–1985 (0.07 ± 0.08 m.w.e. year−1) century [Jóhannesson et al., 2013; Brynjólfsson et al.,
and 1985–1994 (0.26 ± 0.11 m.w.e. year−1), which does 2014, 2015; Magnússon et al., 2016].
Glacial Features of the Northern and Western Regions  167

Box 15.1 Snow avalanches and debris flows

Throughout the past 1000 years snow avalanches which mostly included the clearing and rebuilding
have caused more fatalities than any other single of roads, and the replacement of infrastructure for
type of natural hazard in Iceland [Bjornsson, 1980]. heat and electricity [IMO, 2019b].
Jónsson and Rist [1972] combined the first data Predicting downslope movement of glacial ice in
sets in 1957 from annuals that have recorded ava- the fjords is important to the local communities
lanche scenarios since the 13th century, making it because of the known snow‐avalanche and debris‐
among the longest natural disaster chronology in flow hazards in the area. As described by Bjornsson
the country. [1980], 80–90% of avalanches in Iceland fall
Vestfirðir, also known as the Westernfjords, is within what is referred to as an “avalanche cycle,”
considered one of the five primary locations prone i.e., within a period of a few days a large number
to snow avalanches and debris flows in Iceland of avalanches occur in avalanche‐hazard areas.
(Figure 15.6) [Gundmundsson, 1997]. Here, about The cycle is directly linked to wind direction,
25% of all reported avalanche deaths in Iceland snowfall, and slope instability. Thus, once a cycle
have been recorded and it is estimated that 65% of begins, an alert can be issued by the IMO. The IMO
the inhabitants on Vestfirðir live in proximity to recognizes, however, that avalanche protective
these hazardous conditions. Catastrophic ava- measures or land‐use changes are necessary for a
lanches occurred in two villages in 1995, Súðavík permanent solution to this problem. As an example,
and Flateyri, resulting in the deaths of 34 people supporting structures for preventing snow ava-
[Decaulne, 2007]. An earlier avalanche in 1994 at lanches have been installed at Siglufjörður
Seljalandsdalur (66.0702, −23.2419) in Isafjordur (66.2052, −18.9061) and Ólasvík (64.8975,
(Figure 15.1) caused one death and destroyed 40 −23.7091). The support structures are built up
homes. Overall, during the 20th century slope haz- slope with the intention of trapping the falling
ards have led to the loss of 193 lives; 166 of which material behind walls, thereby protecting the low‐
are due to snow avalanches [Sæmundsson et al., lying communities. In trial runs held in Tvisteinahlið
2003]. The financial cost to Iceland between 1974 the support structures held back 5000–10,000 m3
and 2000 amounted to 3.3 billion IKR or US$27000, of snow.

15.2. SKAGI AND TRÖLLASKAGI PENINSULAS majority of the southern region is not glaciated today,
AND NORÐURLANDSJÖKLAR the movement of past glaciers is evidenced by stria­
tions (Photo 41) that can be seen along valley slopes
Two peninsulas that are important discussion of cli­ from 1000 m down to sea level where the valley ends in
mate change are presented in this section (Skagi and the low relief of the northern region, nearest the coast­
Tröllaskagi), along with the ice caps known collectively line, which hosts numerous isolated lake basins.
as Norðurlandsjöklar (Figure 15.7). Skagi (66.2327,
−21.4566) is a northern peninsula connected to the
mainland of Iceland and is located across Húnaflói
Bay, ~50 km east of Vestfirðir Peninsula. The
Tröllaskagi Peninsula (66.2111, −18.9330) is east of
the Skagi Peninsula, between two fjords Eyjafjörður
and Skagafjörður (Figure 15.1). These two fjords were
likely carved by valley head glaciers that originated
from Hofsjökull (section 13.2) in central Iceland dur­
ing the Weischselian glaciation. Tröllaskagi Peninsula
is divided into two regions (south and north) that have
clearly different topographies. The uplifted and steep
topography of the southern region supports ice caps Figure 15.7 Location of Nord ̄urlandsjöklar. [Modified from
and fosters dendritic watershed patterns that become a Bjornsson and Palsson [2008]; design credit Nathan
dominant feature in the landscape. Whereas the Mennen.]
168 GLACIAL FEATURES

The mountain range within the Norðurlandsjöklar the lower latitudes, however, typically melts any
region spanning the fjords of the Tröllaskagi Peninsula accumulation; the ELA limit of the glaciers in the
form the largest highlands section in the country, from northern region is usually between 1000 and 800 m
85 km north to south and 35 km east to west. Within (Figure 12.8), where permafrost features are com­
this mountain range there are numerous alpine cirques monly found. In some years, heavy precipitation
containing between 100 and 250 small glaciers (40– ­coming off the Arctic Ocean can deposit snow along
100 km2) that are predominately categorized as debris‐ the entire mountain chain, thereby expanding the
covered or rock glaciers, and for which the timing of accumulation zone as seen for 3 years between 1965
origin (pre‐LIA or during) is under debate [Hamilton and 1968. Conversely, the warm temperatures of 1984
and Whalley, 1995; Kellerer‐Pirklbauer et al., 2008]. caused entire ice caps in the Norðurlandsjöklar region
This is the highest concentration of these types of gla­ to completely melt.
ciers in the country, and practically the only area of As summarized by Andrés et al. [2016], the northern
Iceland where they exist [Andrés et al., 2016]. The region of Iceland was likely deglaciated for a longer
presence of rock glaciers makes it more challenging to period of time during the Pleistocene and Holocene
reconstruct paleoclimate in the north as these glaciers than any other part of Iceland, leading to an alpine
are not often associated with terminal moraine deposi­ landscape predisposed to eolian and periglacial
tion. In most cases, the glaciers are found on the north‐ processes, in addition to glaciation. Extensive research
facing slope where they are sheltered by steep mountain in the 20th century on pollen [Einarsson, 1963], vegeta­
slopes that reduce the incoming solar radiation, which tion biodiversity [Steindorsson, 1963], and geomor­
can be intense on the black Tertiary plateau basalts. phology [Sigbjarnsson, 1983] indicated ice‐free or
Typically, the lower valley heads are the main locations ice‐limited situations for the peninsulas’ summit peaks
of snow accumulation that result from avalanches during the LGM. Norðdahl [1990, 1991] emphasized
(Figure 15.6, Box 15.1) and snow drifts. A group of ice that the largest valleys on the peninsula are north–
caps in the area that have been continuously monitored south trending outlets that could easily facilitate flow
include: Gljúfurárjökull (3 km2; 65.7175, −18.6611), away from the highland areas. By the Bølling‐Allerød
Hálsjökull (0.5 km2; 65.8563, −17.7625), Tungnahry­ Interstadial (12,700–10,700 BCE), however, it was
ggsjökull (4 km2; 65.7186, −18.8522), Barkárdalsjökull mostly agreed that the northern region was already
(3 km2; 65.6708, −18.8327), Bægisárjökull (3 km2; deglaciated, with estimates of mean annual tempera­
65.6122, −18.3627), and Grímslandsjökull (2 km2; tures reaching as high as 10°C [Andrés et al., 2016].
65.9047, −17.7625) (Figure 15.7). Of these peaks, The immediately succeeding Younger Dryas Stadial
Gljúfurárjökull (1300 m a.s.l.) has the longest valley initiated new ice accumulation and the readvance of
glacier as it descends down to 580 m with a maximum any of the remaining ice caps from the uplands, which
ice thickness of 120 m [Björnsson, 2017]. Similar to the reached the limits of the present coastline and fjords;
glaciers in central Iceland, Caseldine [1985] established in many cases obstructing tributary valleys and form­
that LIA maximum extents at Gljúfurárjökull and ing lakes [Geirsdóttir et al., 2002; Norðdahl and
Tungnahryggsjökull occurred in the later part of Pétursson, 2005]. After glacial retreat at the end of
the 19th century, although they were asynchronous Younger Dryas, glaciers readvanced toward the coast
(1898–1903 and 1868, respectively). During this time one more time in the Early Preboreal Stage, around
the estimated area of these two glaciers combined was 9200 BCE [Ingólfsson et al., 1997]. This correlates with
18 km2, or more than double of that today [Andrés et al., isolated lake‐basin stratigraphy on the Skagi Peninsula
2016]. Different to the outlet glaciers of Drangajökull recorded by Rundgren et al. [1997], who determined a
ice cap, there have only been four reported surges from relative sea‐level fall of 45 m, which is still evident
three glaciers in the Norðurlandsjöklar area [Björnsson today by the presence of relict marine terraces and
et al., 2012]. raised beaches seen along the coastline (Photo 53).
As well documented, all of these relatively small ice This transgression was somewhat counteracted by a
caps respond quickly, and dramatically, to fluctuations mean absolute uplift rate of 6.9 cm year−1 that
in the climate making it nearly impossible to quantity Ingólfsson et al. [1995] ascribes to a combination of
their change using a mass balance approach as applied rapid Preboreal deglaciation and low asthenosphere
to Iceland’s major glaciers (Vatnajökull, Drangajökull, viscosities below Iceland. According to historic docu­
etc.). At present the northern fjords receive consider­ mentation, the area’s ice sheet during the LIA was suf­
ably more precipitation (2000–3000 mm year−1) than ficiently stable that hiking trails were preferred as the
the nearby Vestfirðir Peninsula. The warm summers at main transit between the fjords rather than travel
Glacial Features of the Northern and Western Regions  169

Photo 53 Marine terraces along the eastern coast of Iceland. [Courtesy of Tamie Jovanelly.]

by boat (Björnsson, 2017]. Modeling by Fernández‐ the glacier remains at 1000 m [Gudmundsson et al.,
Fernández et al [2018] indicates an overall reduction in 2017]. Although small in size, it is one of the most
area (25%) and volume (33%) of the Norðurlandsjöklar visited glaciers in Iceland due to its proximity to
glaciers as a result of the warming that began at the Reykjavík, and on a clear day the volcano can be seen
end of the LIA. across the bay. The volcano is located at the western
end of the Snæfellsnes Peninsula (Figure 2.1, section 11.1)
15.3. SNÆSFELLNES PENINSULA and hosts five major outlet glaciers: Hyrningsjökull
AND SNÆFELLSJÖKULL (2 km2; 64.7994, −23.7491), Jökulháls (2.5 km2;64.7944,
−23.7552), Norðurkinn (1.5 km2; 64.8255, −23.7625),
The Snæfellsnes Peninsula is 160 km north of Blágilsjökull (3 km2; 64.7980, −23.8000), and
Reykjavík and Breðafjordur, which joins the Arctic Hólatindajökull (1 km2; 64.8075, −23.8100) (Figure 15.8).
Ocean, and Faxaflói, which joins the North Atlantic Of these outlet glaciers, perhaps the most reliable for
Ocean (Figure 15.1). Similar to Mýrdalsjökull (sec­ indicating climate change is Hyrningsjökull because it is
tion 14.4), the Snæsfellsjökull ice cap covers the a nonsurging type that has been monitored since the
summit (11 km2) and fills the caldera of an active 1930s [Sigurðsson et al., 2007]. Variations in glacial ter­
­stratovolcano that reaches a maximum elevation of mini at Hyrningsjökull from 1930 to 2005 have been
1446 m. The volcanic materials that create the slopes compared with four other nonsurging glaciers on
(~22°) of Snæsfellsjökull are prone to rapid postdepo­ Vatnajökull and Oræfajökull (section 14.1), and
sitional weathering, especially at altitudes where Mýrdalsjökull (section 14.4) (Figure 15.9). As shown in
freeze–thaw is prevalent [Evans et al., 2016] which is Figure 15.9 the major shifts in advances and retreats are
just below the ELA at 1200–1000 m. At present, only more or less synchronous for all the glaciers [Sigurðsson
10% of the glacier remains above 1200 m, while 40% of et al., 2007]. Additionally, Sigurðsson et al. [2007] compared
170 GLACIAL FEATURES

the mass balance of Hyrningsjökull with Sólheimajökull,


an outlet glacier at Mýrdalsjökull, and found both were
highly correlated in time. They related this link to the
predominant influence of summer temperatures, which
affect rates of accumulation or ablation over short
periods of time that lag by only a few years. Interestingly,
a comparison of temperature and precipitation data
shows that there had been more annual and decadal
variations in precipitation than in temperature over the
past 75 years on the western peninsula [Sigurðsson et al.,
2007]. Weather from the North Atlantic Ocean is the
moisture source for Snæsfellsjökull and annually has
produced 800 mm on average since 1930 when moni­
toring began [Jóhannesson et al., 2011]. It is thought that
these weather patterns developed during the LIA when
the ice cap is estimated to have been double its current
size. Prominent ice‐cored moraines found at all of the
outlet glaciers represent the glacial maximum of the
LIA, however, the exact age of these is not known.
Meltwater channels likely extended past the LIA limits
onto existing and newly developed sandur plains [Evans
et al., 2016].
Figure 15.8 Outlet glaciers on Snæsfellsjökull. [Design credit
Nathan Mennen.]

Figure 15.9 Glacier variations at Snæfellsjökull from 1930 to 1995. [Modified from Sigurdsson [1998];
design credit Nathan Mennen.]
Glacial Features of the Northern and Western Regions  171

Figure 15.10 Cumulative variations of the termini of five nonsurge‐type glaciers during the period 1930–
2005. [Modified from Sigurdsson et al. [2007]; design credit Nathan Mennen.]

Although the country is concerned about the ice thickness of this glacier is 50 m or less, which
influence of climate change on all of its glaciers, the emphasizes the sensitivity of its mass balance. As
impact that a continued warming trend could have on expected, the thickest ice is in the caldera (70 m) and
Snæsfellsjökull may mean its demise within the next around the crater rim (100 m). Thicker ice can also be
few decades. According to maps of Iceland published found on the main outlets to the west and east of the
in 1910, the area of the ice cap was estimated to be summit (e.g., Blágilsjökull, Jökulháls, and Hyrningsjökull)
22 km2 in 1910. If this value is considered correct, then where average ice thickness is between 40 and 50 m.
Snæsfellsjökull has lost 50% of its ice volume since the The concern, however, is that thinner portions of the
beginning of the 20th century [Jóhannesson et al., ice sheet will melt first, causing downwasting and ice‐
2011]. Additionally, all glacial outlets at Snæsfellsjökull free zones as had occurred between 2002 and 2008 on
have retreated between 600 and 800 m over the past the northern flank [Jóhannesson et al., 2011].
100 years [Björnsson, 2017]. This is not to say that the Climate change can also have an impact on the
extent of this ice cap has not been variable: a rapid landscape and volcanism at Snæfellsjökull. As
retreat was documented during a warm period from described by Bakker et al. [2014a,b], glacial unloading
1930 to 1965, and was followed by an advance during may have two major effects. First, it may lead to
a cooler period from 1970 to 1995, that created push significant stress release at the base of the volcanic edi­
moraines still present in the landscape (Figure 15.10). fice that could destabilize the flanks and generate land­
Since 1995, however, the ice cap has continued to slides (Box 15.1). Next, they considered that pressure
shrink rapidly at 1.25 m.w.e. year−1. A detailed map­ changes around the magma chamber could lead to
ping project using airborne LiDAR imagery estimated volcanic unrest. In particular, melting of the ice cap
that the ice cap lowered 14.0 m from 1999 to 2008 could trigger movement in the shallow magma chamber
[Jóhannesson et al., 2011]. Additionally, radio echo as it responds to isostatic rebound. The last eruptions
soundings completed in 2003 revealed that the average were about 4000 years ago.
GLOSSARY

aa: Hawaiian name for mafic lava flows, such as basalt, cold‐based glacier: glaciers in polar regions where the
that characteristically form rough and fragmented temperature of the snow fall is at or below 0°C and
surfaces the ice of the glacier remains at below zero
anastomosing river: a type of river typically consisting throughout the year; stays frozen to the bedrock all
of multithreaded, clay‐ and slit‐based channels that year with little ice movement and little erosion
occur in a low‐gradient environment columnar basalt: a type of basalt structure formed by
aphanitic: a type of texture used to describe igneous the rapid cooling of a thick lava flow with an exten-
rocks that cooled quickly (extrusively) and thus sive fracture network that causes columns to form
composed of fine grains composite volcano: also known as a stratovolcano, a
asthenosphere: upper layer of Earth’s mantle that lies type of volcano characterized by steep sides and a
below the lithosphere and contains convection cells large height that is formed by the build up of several
which allow it to be ductile and have fluid motion layers of hardened lava, tephra, and pumice over
basalt: a mafic, aphanitic igneous rock that comprises time; also characterized by violent/explosive erup-
oceanic crust and is characteristic of nonviolent tions and felsic lava
volcanic eruptions, such as those found in Hawaii country rock: any rock native to an area
bookshelf faulting: occurs when stress is released in crater: a circular‐shaped depression in the ground
transform fault zones during strike‐slip earthquakes, formed by volcanic activity or meteor impact
causing the transform fault to be transverse to the crust: the outermost layer of Earth
fault zone and blocks between the fault zones to sub- diatreme: a fracture in Earth’s crust formed in
sequently rotate conjunction with a maar when magma makes
Bowen’s reaction series: a Y‐shaped diagram that contact with a body of ground water, increasing
explains the temperature of crystallization (cooling pressure and causing rapid expansion of gases and a
magma) of common silicate minerals out of magma series of explosions
such as olivine, biotite, plagioclase, and quartz; dike: a type of igneous intrusion that cuts vertically
describes the temperatures at which mafic, into older rock layers
intermediate, and felsic magma are found within drumlin: an oval‐shaped (streamlined) hill formed by
Earth the movement of glaciers over rock debris, charac-
braided river: a river channel typically comprising a terized by a steep, upglacier (stoss) side and a more
network of small channels separated by smaller gentle, downglacier side, thus indicating the direction
islands called bars that are characterized by a high of ice movement
stream and channel gradient and a high sediment edifice: a volcano’s cone‐shaped structure built over-
load time by the accumulation of erupted lava, pyro-
cinder‐cone volcano: also known as scoria cone, a type clastic flows, tephra, and other volcanic deposits
of volcano that is formed from large mounds of around a central vent system
pyroclastic material, such as cinders and volcano esker: a sinuous ridge, typically composed of gravel
ash, that has been built up around a volcanic vent by and finer sediment that is left after the retreat of a
violent eruptions glacier or ice sheet
cirque: amphitheater‐shaped basin with steep walls fault: a fracture or discontinuity in a rock/sediment
found at the head of a glacial valley formed by caused by extensional, compressional, and rota-
erosion at the base of a glacier tional movement, e.g., related to plate tectonics,
climate change: any long‐term shift in global or magmatic or diapiric intrusions, sediment flow, and
regional weather patterns gravity loading

Iceland: Tectonics,Volcanics, and Glacial Features, Geophysical Monograph 247, First Edition. Tamie J. Jovanelly.
© 2020 American Geophysical Union. Published 2020 by John Wiley & Sons, Inc.
173
174 GLOSSARY

felsic magma: high‐silica magma that is formed at shal- hyaloclastite: hydrated tuff‐like breccia rich in black
low depths in the Earth leading to violent/explosive volcanic glass, formed during volcanic eruptions
eruptions resulting in pyroclastic flows, pumice, and under water, under ice, or where subaerial flows
volcanic ash; also creates light‐colored igneous rocks reach bodies of water; appearance of angular flat
such as pumice and rhyolite fragments
firn: granular snow found on the upper part of a gla- ice‐cored moraine: a transitional landform of glacial
cier where it has yet to be compressed into ice sediment cored by glacial ice
fissure swarm: an area composing several fissures, or ice‐rafted debris (IRD): various clasts embedded in
fractures in the bedrock, due to tensional forces ice that are transported until the ice melts when
within volcanic regions; usually sites of geothermal they are then deposited; a primary characteristic
energy systems of subaqueous deposition during glacial
gabbro: a mafic, phaneritic igneous rock usually dark episodes
green to black in color and is the most abundant igneous rock: one of the three main rock types; formed
rock in the deep oceanic crust by the cooling and solidification of magma
geologic time: the period of time since the formation of inner core: the innermost part of Earth, believed to be
Earth; divided into eons, eras, periods, and epochs solid and composed of an iron–nickel alloy
geologic time eras: the second largest geologic subdivi- intraplate volcanic belt: a chain of hot spots within
sion of time that typically spans a few hundred tectonic plate interiors rather than on the plate
­
­million years in length; marked by extinctions or boundaries
other major geologic events isotopic dating: technique used to date materials such
geologic time periods: a subdivision of geologic time as rocks or carbon through the decay of radioactive
often associated with formation of a particular rock elements
system; typically spans tens of millions of years jökulhlaup: Icelandic word for “glacier run;” a type of
geysir: a hot spring that discharges water and steam glacial outburst flood triggered by geothermal
intermittedly heating or a volcanic subglacial eruption
glacial flow: downglacier movement of ice due to gravity Kaiser effect: describes the absence of acoustic
glacial surging: cyclic glacial flow characterized by emission or seismic events until the initial seismic
switches between fast (active) and slow (quiescent) response load is exceeded in a body of rock
phases kettle lake: a shallow, body of water formed by retreat-
glacial till: unsorted sediment (diamicton) deposited ing glaciers, especially by melting of buried ice
by ice and derived from the erosion of material by a knickpoints: a sharp change in the channel slope of a
glacier river, such as a lake or waterfall, possibly caused by
glacier: persistent body of dense ice formed when the erosion from glaciation or variance in lithology
accumulation of snow exceeds the melt of the snow lateral moraine: product of rock fall onto the margin
(ablation) that is continuously moving under its own of a glacier caused by rock slope failure in response
weight to frost weathering of the rock wall and oversteepen-
glacio‐isostatic rebound: readjustment of Earth’s crust ing of cliffs by glacial erosion; melting of the glacier
following depression by loading imposed by ice leaves a ridge/bench of blocky debris on the flank of
sheets during glaciations the valley
global warming: gradual increase in average tempera- lava: magma that reaches the surface of Earth
ture of Earth’s atmosphere and its oceans that per- lignite: a sedimentary rock formed of alternating layers
manently changes the climate of decomposed organic matter, such as compressed
ground moraine: till‐covered areas with irregular topog- peat—the first stage in coalification
raphy and no ridges that often form gently rolling lithosphere: rigid outer part of Earth divided into tec-
hills or plains; accumulated a the base of ice or tonic plates that contains the crust and the upper
deposited as a glacier retreats mantle; lies above the asthenosphere
horst and graben: regions that lie between normal maar volcano: a broad, low‐relief volcanic crater
faults; the horst is the uplifted area and the graben is caused by an explosion that results in groundwater
the downfaulted area coming into contact with magma (phreatomagmatic
hot spot: portion of the surface of the earth that expe- eruption) to form a relatively shallow crater lake
riences volcano activity due to a rising mantle plume mafic magma: magma that lies deep below the surface
in conjunction with plate tectonics of Earth and results in gentle, nonviolent volcanic
GLOSSARY 175

eruptions; also creates dark‐colored igneous rocks palsa: a periglacial feature of earth that has been
such as basalt and olivine pushed up into a mound in which perennial ice is
magma: mixture of molten rock, volatiles, and solids formed
that is found beneath Earth’s surface permafrost: a layer of soil that remains frozen for two
mantle: the thickest layer of Earth found between the or more years
crust and outer core; it is both solid and plastic and piedmont glacier: formed by merging valley glaciers as
divided into the upper and lower mantle they emerge from mountains
marker bed: a bed of rock readily distinguishable by its pillow basalt: lavas that contain characteristic thick
physical characteristics and traceable over large sequences of pillow‐shaped structures caused by the
horizontal distances; also known as a key bed extrusion of lava under water
meandering river: a river that flows over gently sloping plate tectonics: the theory explaining the structure of
gound and begins to curve back and forth across a Earth’s crust and associated phenomena resulting
landscape; erodes sediment from the outer curve of from the interaction of rigid lithospheric plates as
each meander bend and deposits it on an inner curve they move slowly over the underlying mantle
further down stream positive feedback cycle: a system that contains a
mid‐ocean ridge: underwater mountain range formed function that is fed back into the system in way that
by plate tectonics; center of seafloor spreading increases the magnitude of the output
möberg ridge: long ridges produced by fissure erup- pumice: extremely light weight porous, felsic igneous
tions beneath ice rock formed from explosive magma that cools
moulin: a vertical cavity within a glacier that serves as almost instantaneously; typical of pyroclastic
a tunnel for surface meltwater flows
negative feedback cycle: a system that contains a pyroclastic flow: a destructive fast‐flowing mass com-
function that is fed back into the system in way that posed of hot ash, lava fragments, and toxic gases
reduces the magnitude of the output from an explosive volcanic eruption
normal fault: a fault where rock drops down on one side quartz: the most abundant mineral found at Earth’s
relative to the other side; characteristic of extensional surface; the simplest silicate composed of one part
movement such as at divergent plate boundaries silicon and two parts oxygen (SiO2)
nunatak: a mountain peak or ridge that is nor covered regressive boundary: sedimentary rock sequence dis-
by a glacier or ice sheet playing a decrease in sea level over time
obsidian: volcanic glass formed as an extrusive igneous reverse fault: a fault where the rock on one side is
rock; produced when lava from a volcano cools rapidly uplifted; characteristic of convergent movement
olivine: a rock‐forming mineral that is typically found such as at plate boundaries
in mafic igneous rocks such as basalt, gabbro, dunite, rhyolite: a felsic igneous rock with a glassy to aphanitic
diabase, and peridotite where plagioclase and texture; typically faint pink/salmon in color
pyroxene are also present; usually green in color ridge jump: caused by a propagating rift; in a ridge–
outer core: a fluid layer of Earth composed of mostly transform–ridge system propagation can occur from
iron and nickel that lies above the solid inner core a dying ridge to an active ridge, perhaps jumping
and below the mantle from one plate to the other
outwash plain: the flat area found in the forefield of a rift valley: a linear valley formed by faulting on a diver-
glacier where large quantities of water have flowed gent plate boundary
from the melting glacier ice depositing layers of sand rock cycle: the process in which igneous, sedimentary,
and other fine sediments and metamorphic rocks are formed and deformed;
pahoe pahoe: Hawaiian name for mafic lava flows, such igneous rocks are eroded into sediment to be lithi-
as basalt, that characteristically form smooth ropy fied into sedimentary rocks or transformed by heat
masses and pressure into metamorphic rocks; metamorphic
Pangea: a supercontinent that was formed of all rocks melt into magma which then cools to form
Earth’s continents more than 300 Ma and began to igneous rocks
break apart about 160 Myr later rock glacier: a mass of rock, ice, snow, mud, and/or
phaneritic: a type of texture used to describe coarse‐ water that moves slowly downslope due to gravity
grained igneous rocks that cooled slowly (intrusively) rootless cone: volcanic landform that resemble a
palagonite: alteration product resulting from inter- volcanic crater but does not contain an actual
actions of water with basalt melt or basaltic glass volcanic vent and therefore lacks a magma conduit
176 GLOSSARY

connecting beneath the surface; formed by steam stratigraphy: the sequential ordering of rock strata; particularly
explosions as flowing lava crosses a wet surface and strata classification and sequence relationships
creates a crater sulfur: a bright yellow mineral, light in weight, com-
sag ponds: depressions form along a strike‐slip fault posed of elemental sulfur (atomic number 16) with a
where freshwater collects in the lowest portion to distinct “rotten‐egg” odor
create a permanent pond tephra: any airborne rock fragment or particle, such as
scoria: highly vesicular, mafic igneous rock formed ash/lapilli/volcanic bombs, that is ejected by an
when magma containing abundant dissolved gas erupting volcano
flows from a volcano and cools; typically dark brown thrust fault: a low‐angle (between the fault and horizontal
to red in color plane) reverse fault that results from thrusting
sea‐floor spreading: process that occurs at mid‐ocean tillite: a lithified glacial till
ridges where new oceanic crust is formed through tombolo: one or more sandbars occurring in shallow
volcanic activity and gradually moves away from the waters that connect an island to the mainland
ridge causing new crust to be found at the center of transgressive boundary: sedimentary rock sequence dis-
the ridge and old crust to be found at the edges near playing an increase in sea level over time
continents; this explains continental drift in the triple point junction: a point where three tectonic plate
theory of plate tectonics margins (of variable type) intersect
sedimentary rock: lithified accumulations of nonor- tumulus: a mound commonly found on pahoehoe
ganic minerals and rock fragments, organic material, surfaces, formed by the buckling of crust as slow‐
and chemical precipitates moving magma upwells and increases pressure
shield volcano: a large, gently sloping volcano that usu- beneath the crust
ally erupts very fluid, mafic, basaltic lava u‐shaped valley: a valley with a cross‐section resembling
sill: an igneous intrusion that forms parallel to existing the letter “U”; formed by glacial erosion resulting in
bedrock (see also dike) steep sides and a flat bottom
sinter: siliceous or calcareous deposits created by viscosity: a liquid’s resistance to flow; informally
springs, accumulating around the vent of a geyser referred to as “thickness”
solfataras: a volcanic steam vent characterized by warm‐based glacier: glaciers in which water is present
hot vapors and high concentrations of sulfur throughout the ice mass for a period of the year or
gases all year and allows for much greater rates of
steam explosion: eruption caused by violent boiling of movement and thus more erosion
water into steam from contact with flowing lava waterfall: vertical descent of water; often formed where
strath terraces: the former valley flat of a rock‐floored rocks are exposed with different resistances to erosion
river valley that remains following downcutting into xenolith: a fragment of pre‐existing rock contained
the bedrock by the river within an igneous body
REFERENCES

Aðalgeirsdóttir, G., Gudmundsson, G. H., and Björnsson, near a hotspot: Earthquake mechanisms and regional
H. (2000), The response of a glacier to a surface distur­ stress in the south Iceland Seismic Zone, Tectonophysics,
bance: a case study on Vatnajökull ice cap, Iceland, Ann. 447(1–4), 95–116.
Glaciol., 31, 104–110. Árnadóttir, T., Geirsson, H., and Einarsson, P. (2004),
Ahlmann, H. W. and Thórarinsson, S. (1937), Previous inves­ Coseismic stress changes and crustal deformation on the
tigations of Vatnajökull, marginal oscillations of its outlet‐ Reykjanes Peninsula due to triggered earthquakes on 17
glaciers, and general description of its morphology. June 2000, J. Geophys. Res. Solid Earth, 109(B9).
Vatnajökull, scientific results of the Swedish–Icelandic Árnadóttir, T., Jiang, W., Feigl, K. L., Geirsson, H., and
investigastions 1936–1937, Geogr. Ann., 19(3–4), 176–211. Sturkell, E. (2006), Kinematic models of plate boundary
Ali, S. T., Feigl, K. L., Carr, B. B., Masterlark, T., and deformation in southwest Iceland derived from GPS
Sigmundsson, F. (2014), Geodetic measurements and observations, J. Geophys. Res. Solid Earth, 111(B7).
numerical models of rifting in northern Iceland for 1993– Árnadóttir, T., Sigmundsson, F., and Delaney, P. T. (1998),
2008, Geophys. J. Int., 196(3), 1267–1280. Sources of crustal deformation associated with the Krafla,
Allen, R. M., Nolet, G., Morgan, W. J., Vogfjörd, K., Iceland, eruption of September 1984, Geophys. Res. Lett.,
Bergsson, B. H., Erlendsson, P., et al. (2002a), Imaging the 25(7), 1043–1046.
mantle beneath Iceland using integrated seismological Ármannsson, H. and Kristmannsdóttir, H. (1992), Geothermal
techniques, J. Geophys. Res. Solid Earth, 107(B12). environmental impact, Geothermics, 21, 869–880.
Allen, R. M., Nolet, G., Morgan, W. J., Vogfjörd, K., Nettles, Ármannsson, H. and Thórhallsson, S. (1996), Krýsuvík, an
M., Ekström, G., et al. (2002b), Plume‐driven plumbing overview of previous exploration and exploitation and utili­
and crustal formation in Iceland, J. Geophys. Res. Solid zation possibilities, along with proposals for further explo­
Earth, 107(B8), 105–111. ration, Report OS‐96012/JHD‐06B, Orkustofnun,
Andrés, N., Tanarro, L. M., Fernández, J. M., and Palacios, Reykjavík, 25 pp. [In Icelandic.]
D. (2016), The origin of glacial alpine landscape in Ármannsson, H., Gudmundsson, A., and Steingrímsson, B.
Tröllaskagi Peninsula (north Iceland), Cuad. Invest. S. (1987), Exploration and development of the Krafla
Geogr., 42(2), 341–368. geothermal area, Jökull, 37, 13–30.
Andrew, R. E. and Gudmundsson, A. (2007), Distribution, Bakker, R., Lupi, M., and Frehner, M. (2014a), Effects of
structure, and formation of Holocene lava shields in rapid icecap melting on a shallow magma chamber: A multi‐
Iceland, J. Volcanol. Geotherm. Res., 168(1–4), 137–154. discplinary case study of Snæfellsjökull volcano, Western
Andrews, J. T. S. and Giraudeau, J. (2003), Multi‐proxy Iceland, 12th Swiss Geoscience Meeting, Fribourg.
records showing significant Holocene environmental vari­ Bakker, R., Lupi, M., Frehner, M., Berger, J., and Fuchs, F.
ability: the inner N. Iceland shelf (Húnaflói), Quat. Sci., (2014b), Volcanic unrest primed by ice cap melting: A case
22(2–4), 175–193. study of Snæfellsjökull volcano, Western Iceland, in EGU
Andrews, J. T., Hardardóttir, J., Helgadóttir, G., Jennings, A. General Assembly Conference Abstracts, Vol. 16.
E., Geirsdóttir, Á., Sveinbjörnsdóttir, Á. E., et al. (2000), Baynes, E. R., Attal, M., Dugmore, A. J., Kirstein, L. A.,
The N and W Iceland Shelf: insights into Last Glacial and Whaler, K. A. (2015), Catastrophic impact of extreme
Maximum ice extent and deglaciation based on acoustic flood events on the morphology and evolution of the
stratigraphy and basal radiocarbon AMS dates, Quat. Sci. lower Jökulsá á Fjöllum (northeast Iceland) during the
Rev., 19(7), 619–631. Holocene, Geomorphology, 250, 422–436.
Angelier, J., Bergerat, F., Dauteuil, O., and Villemin, T. Benediktsson, Í. Ö., Ingólfsson, Ó., Schomacker, A., and
(1997), Effective tension–shear relationships in exten­ Kjær, K. H. (2009), Formation of submarginal and pro­
sional fissure swarms, axial rift zone of northeastern glacial end moraines: Implications of ice‐flow mechanism
Iceland, J. Struct. Geol., 19(5), 673–685. during the 1963–64 surge of Brúarjökull,Iceland, Boreas,
Angelier, J., Bergerat, F., and Homberg, C. (2000), Variable 38, 440–457.
coupling across weak oceanic transform fault: Benn, D.I., Gulley, J.D., Luckman, A., Adamek, A., and
Flateyjarskagi, Iceland, Terra Nova, 12(3), 97–101. Glowacki, P. (2009), Englacial drainage systems formed
Angelier, J., Bergerat, F., Stefansson, M., and Bellou, M. by hydrologically driven crevasse propagation, J. Glaciol.,
(2008), Seismotectonics of a newly formed transform zone 55(191), 513–523.

Iceland: Tectonics,Volcanics, and Glacial Features, Geophysical Monograph 247, First Edition. Tamie J. Jovanelly.
© 2020 American Geophysical Union. Published 2020 by John Wiley & Sons, Inc.
177
178 REFERENCES

Berg, S.E., Troll, V.R., Burchardt, S., Riishuss, M.S., Björnsson, H. and Guðmundsson, M. T. (1993), Variations
Krumbholz, M., and Gústafsson, L.E. (2014), Iceland’s in the thermal output of the subglacial Grímsvötn c­ aldera,
best kept secret, Geol. Today, 30(2), 54–60. Iceland, Geophys. Res. Lett., 20(19), 2127–2130.
Bergerat, F. and Angelier, J. (2000), The South Iceland Björnsson, H. and Pálsson, F. (2008), Icelandic glaciers,
Seismic Zone: tectonic and seismotectonic analyses Jökull, 58, 365–386.
revealing the evolution from rifting to transform motion, Björnsson, H., Pálsson, F., and Guðmundsson, M. T. (2000),
J. Geodyn., 29(3–5), 211–231. Surface and bedrock topography of the Mýrdalsjökull ice
Bergerat, F. and Angelier, J. (2008), Immature and mature cap, Jökull, 49, 29–46.
transform zones near a hot spot: the South Iceland Seismic Bjornsson, H., Rott, H., Gudmundsson, S., Fischer, A.,
Zone and the Tjörnes Fracture Zone (Iceland), Tectonophysics, Siegel, A., and Gudmundsson, M.T. (2001), Glacier–
447(1–4), 142–154. volcano interactions deduced by SAR interferometry.
Bierman, P. R. and Montgomery, D. R. (2014), Key concepts J. Glaciol., 47(156), 58–70.
in geomorphology, W.H. Freeman and Company, 494 pp. Björnsson, H., Pálsson, F., Sigurðsson, O., and Flowers, G.
Birks, H. H., Gulliksen, S., Haflidason, H., Mangerud, J., E. (2003), Surges of glaciers in Iceland, Ann. Glaciol., 36,
and Possnert, G. (1996), New radiocarbon dates for the 82–90.
Vedde Ash and the Saksunarvatn Ash from western Björnsson, H., Pálsson, F., Gudmundsson, S., Magnússon,
Norway, Quat. Res., 45(2), 119–127. E., Adalgeirsdóttir, G., Jóhannesson, T., et al. (2013),
Bjarnason, J. Ö. (2000), A note on chemical composition of Contribution of Icelandic ice caps to sea level rise: Trends
geothermal steam from well KR‐9 in Krýsuvík, and variability since the Little Ice Age, Geophys. Res.
southwestern Iceland, Orkustofnun, 1 pp. Lett., 40(8), 1546–1550.
Björnsson, A. (1985), Dynamics of crustal rifting in NE Iceland, Björnsson, Ó. G., Ámason, A., Gudmunosson, S., Jensson, Ó.,
J. Geophys. Res. Solid Earth, 90(B12), 10151–10162. Ólafsson, S., and Valdimarsson, H. (1978), Macroglobulinaemia
Björnsson, A., Sæmundsson, K., Einarsson, P., Tryggvason, in an Icelandic family, J. Internal Med., 203(1–6), 283–288.
E., and Grönvold, K. (1977), Current rifting episode in Black, J. (2008), Holocene climate change in south‐central
north Iceland, Nature, 266(5600), 318. Iceland: A multi‐proxy lacustrine record from glacial lake
Björnsson, A., Johnsen, G., Sigurðsson, S., Thorbergsson, Hvítárvatn, University of Colorado at Boulder, ProQuest
G., and Tryggvason, E. (1979), Rifting of the plate Dissertations Publishing.
boundary in North Iceland 1975–1978, J. Geophys. Res. Black, J., Miller, G. H., Geirsdóttir, Á., Manley, W., and
Solid Earth, 84(B6), 3029–3038. Björnsson, H. (2004), Sediment thickness and Holocene
Bjornsson, H. (1974), Explanation of jökulhlaups from erosion rates derived from a seismic survey of Hvitarvatn,
Grımsvotn, Vatnajokull, Iceland, Jökull, 24, 1–26. central Iceland, Jökull, 54, 37–56.
Bjornsson, H. (1975), Subglacial water reservoirs, jökul­ Blockley, S. P. E., Lane, C. S., Lotter, A. F., and Pollard,
hlaups and volcanic eruptions, Jökull, 25, 1–14. A. M. (2007), Evidence for the presence of the Vedde
Bjornsson, H. (1977), The cause of jökulhlaups in the Skafta Ash in Central Europe, Quat. Sci. Rev., 26(25–28),
river, Vatnajokull. Jökull, 27, 71–78. 3030–3036.
Björnsson, H. (1979), Glaciers in Iceland, Jökull, 29, 74–80. Böðvarsson, G. and Walker, G. P. L. (1964), Crustal drift in
Björnsson, H. (1980), Avalanche activity in Iceland, climatic Iceland, Geophys. J. Int., 8(3), 285–300.
conditions, and terrain features, J. Glaciol., 26(94), 13–23. Bonafede, M., Ferrari, C., Maccaferri, F., and Stefánsson,
Björnsson, H. (1982), Drainage basins on Vatnajökull R. (2007), On the preparatory processes of the M 6.6
mapped by radio echo soundings, Hydrol. Res., 13(4), earthquake of June 17th, 2000, in Iceland, Geophys. Res.
213–232. Lett., 34(24).
Björnsson, H. (1988), Hydrology of ice caps in volcanic Boulton, G. S., Harris, P. W. V., and Jarvis, J. (1982), Stratigraphy
regions, Soc. Sci. Island., 45. and structure of a coastal sediment wedge of glacial origin
Björnsson, H. (1992), Jökulhlaups in Iceland: prediction, inferred from sparker measurements in glacial Lake
characteristics and simulation, Ann. Glaciol., 16, 95–106. Jökulsarlon in southeastern Iceland, Jökull, (32), 37–47.
Björnsson, H. (1998), Hydrological characteristics of the Bourgeois, O., Dauteuil, O., and Vliet‐Lanoë, B. V. (2000),
drainage system beneath a surging glacier, Nature, 395, Geothermal control on flow patterns in the Last Glacial
771–774. Maximum ice sheet of Iceland, Earth Surf. Proc. Land.,
Björnsson, H. (2003), Subglacial lakes and jökulhlaups in 25(1), 59–76.
Iceland, Global Planet. Change, 35(3–4), 255–271. Bradwell, T., Dugmore, A. J., and Sugden, D. E. (2006), The
Björnsson, H. (2017), The glaciers of Iceland: A historical, Little Ice Age glacier maximum in Iceland and the North
cultural, and scientific overview, 1st edn, Atlantic Press, Atlantic Oscillation: evidence from Lambatungnajökull,
613 pp. southeast Iceland, Boreas, 35(1), 61–80.
Björnsson, H. and Einarsson, P. (1990), Volcanoes beneath Bradwell, T., Siðurdsson, Ó., and Everest, J. (2013), Recent,
Vatnajökull, Iceland: Evidence from radio echo‐sounding, very rapid retreat of a temperate glacier in SE Iceland,
earthquakes and jökulhlaups, Jökull, 40, 147–168. Boreas, 43(4), 953–979.
REFERENCES 179

Brandsdóttir, B. (1992), Historical accounts of earthquakes Canas, D., Chan, W. M., Chiu, A., Jung‐Ritchie, L., Leung,
associated with eruptive activity in the Askja volcanic M., Pillay, L., and Waltham, B. (2015), Potential environ­
system, Jökull, 42, 1–12. mental effects of expanding Lake Jökulsárlón in response
Brandsdottir, B., and Einarsson, P. (1979), Seismic activity to melting of Breiðamerkurjökull, Iceland, Cartograph.
associated with the September 1977 deflation of the Int. J. Geogr. Inform. Geovis., 50(3), 204–213.
Krafla central volcano in northeastern Iceland, J. Carey, R. J., Houghton, B. F., and Thordarson, T. (2009),
Volcanol. Geotherm. Res., 6(3–4), 197–212. Abrupt shifts between wet and dry phases of the 1875
Brandsdóttir, B., Menke, W., Einarsson, P., White, R. S., and eruption of Askja Volcano: microscopic evidence for mac­
Staples, R. K. (1997), Färoe–Iceland ridge experiment 2. roscopic dynamics. J. Volcanol. Geotherm. Res., 184(3–4),
Crustal structure of the Krafla central volcano, J. Geophys. 256–270.
Res. Solid Earth, 102(B4), 7867–7886. Carey, R. J., Houghton, B. F., and Thordarson, T. (2010),
Brooke, R.D., Rajagopalan, S., Popelli, C.A., Brook, J.R., Tephra dispersal and eruption dynamics of wet and dry
Bhatnagar, A., Diez‐Roux, A.V., et al. (2010), Particulate phases of the 1875 eruption of Askja Volcano, Iceland,
matter air pollution and cardiovascular disease, Am. Bull. Volcanol., 72(3), 259–278.
Health Assoc. Sci. Statement, 12, 2331–2378. Carrivick, J. L. (2007), Hydrodynamics and geomorphic work
Brown, G. C., Everett, S. P., Rymer, H., McGarvie, D. W., of jökulhlaups (glacial outburst floods) from Kverkfjöll vol­
and Foster, I. (1991), New light on caldera evolution— cano, Iceland, Hydrol. Process., 21(6), 725–740.
Askja, Iceland, Geology, 19(4), 352–355. Carrivick, J. L. and Twigg, D. R. (2005), Jökulhlaup‐influenced
Brunn. (1897), Ancient remains and modern homes in Iceland, topography and geomorphology at Kverkfjöll, Iceland,
Nordiske Forlag, Kobenhavn. J. Maps, 1(1), 7–17.
Brynjólfsson, S. (2015), Dynamics and glacial history of the Carrivick, J. L., Russell, A. J., and Tweed, F. S. (2004),
Drangajökull ice cap, northwest Iceland, Doctoral disserta­ Geomorphological evidence for jökulhlaups from
tion, [Retrieved from Open visindi.] Kverkfjöll volcano, Iceland, Geomorphology, 63(1–2),
Brynjólfsson, S., Ingólfsson, Ó., and Schomacker, A. (2012), 81–102.
Surge fingerprinting of cirque glaciers at the Tröllaskagi Carrivick, J. L., Russell, A. J., Tweed, F. S., and Knudsen, O.
peninsula, north Iceland, Jökull, 62, 151–166. (2002), Determining the routeways and flow characteris­
Brynjólfsson, S., Schomacker, A., and Ingólfsson, Ó. (2014), tics of jökulhlaups from Kverkfjöll, Iceland, in V.
Geomorphology and the Little Ice Age extent of the Thondycraft, G. Benito, M. Barriendos, and M. C. Llasat
Drangajökull ice cap, NW Iceland, with focus on its three (eds), Abstracts of palaeofloods, historical data and climatic
surge‐type outlets, Geomorphology, 213, 292–304. variability: applications in flood risk assessment, 16–19
Brynjólfsson, S., Schomacker, A., Ingólfsson, Ó., and October, Barcelona.
Keiding, J. K. (2015), Cosmogenic 36Cl exposure ages Casadevall, T. J. (1994), The 1989–1990 eruption of Redoubt
reveal a 9.3 ka BP glacier advance and the Late Volcano, Alaska: impacts on aircraft operations,
Weichselian–Early Holocene glacial history of the J. Volcanol. Geotherm. Res., 62(1–4), 301–316.
Drangajökull region, northwest Iceland, Quat. Sci. Rev., Caseldine, C. J. (1985), The extent of some glaciers in
126, 140–157. northern Iceland during the Little Ice Age and the nature
Brynjólfsson, S., Schomacker, A., Korsgaard, N. J., and of recent deglaciation, Geogr. J., 151, 215–227.
Ingólfsson, Ó. (2016), Surges of outlet glaciers from the Center for Ice and Climate (2019), Niels Bohr Institute,
Drangajökull ice cap, northwest Iceland, Earth Planet. http://www.iceandclimate.nbi.ku.dk/ [Accessed on 17
Sci. Lett., 450, 140–151. September 2019.]
Bryson, R. A. and Murray, T. J. (1977), Climates of Hunger: Clifton, A. E., Sigmundsson, F., Feigl, K. L., Guðmundsson,
Mankind and the World’s Changing Weather, The G., and Árnadóttir, T. (2002), Surface effects of faulting
University of Wisconsin Press, 171 pp. and deformation resulting from magma accumulation
Buck, W. R., Einarsson, P., and Brandsdóttir, B. (2006), at the Hengill triple junction, SW Iceland, 1994–1998,
Tectonic stress and magma chamber size as controls on J. Volcanol. Geotherm. Res., 115(1), 233–255.
dike propagation: Constraints from the 1975–1984 Krafla Clifton, A. E., Pagli, C., Jónsdóttir, J. F., Eythorsdóttir, K.,
rifting episode, J. Geophys. Res. Solid Earth, 111(B12). and Vogfjörð, K. (2003), Surface effects of triggered fault
Calderone, G. M., Grönvold, K., and Oskarsson, N. (1990), slip on Reykjanes Peninsula, SW Iceland, Tectonophysics,
The welded air‐fall tuff layer at Krafla, northern Iceland: 369(3–4), 145–154.
a composite eruption triggered by injection of basaltic Crochet, P., Jóhannesson, T., Jónsson, T., Sigurðsson, O.,
magma, J. Volcanol. Geotherm. Res., 44(3–4), 303–314. Björnsson, H., Pálsson, F., and Barstad, I. (2007),
Camitz, J., Sigmundsson, F., Foulger, G., Jahn, C. H., Estimating the spatial distribution of precipitation in
Völksen, C., and Einarsson, P. (1995), Plate boundary Iceland using a linear model of orographic precipitation,
deformation and continuing deflation of the Askja vol­ J. Hydrometeorol., 8(6), 1285–1306.
cano, north Iceland, determined with GPS, 1987–1993, Dalfsen, E., Pedersen, R., Sigmundsson, F., and Pagli. C.
Bull. Volcanol., 57(2), 136–145. (2004), Deep accumulation of magma near crust–mantle
180 REFERENCES

boundary at the Krafla volcanic system, Iceland: Evidence Dugmore, A., Shore, J.S., Cook, G.T., Newton, A.J.,
from satellite radar interferometry 1993–1999, Geophys. Edwards, K.J., and Larson, G. (1995), The radiocarbon
Res. Lett., 31, L13611. dating of tephra layers in Britain and Iceland, Radiocarbon,
Dalfsen, E., Rymer, H., Sigmundsson, F., and Sturkell, E. 37, 286–295.
(2005), Net gravity decrease at Askja volcano, Iceland: Dugmore, A., Church, M. J., Mairs, K., McGovern, T. H.,
constraints on processes responsible for continuous cal­ Perdikaris, S., and Vésteinsson, O. (2007), Abandoned
dera deflation, 1988–2003, J. Volcanol. Geotherm. Res., farms, volcanic impacts, and woodland management:
139(3–4), 227–239. Revisiting Þjórsárdalur, the “Pompeii of Iceland,” Artic
Dansgaard, W. (1964), Stable isotopes in precipitation, Anthropol., 44, 1–11
Tellus, XVI(4), 1–10. Dugmore, A. J., Newton, A. J., Smith, K. T., and Mairs, K.
Darbyshire, F. A., White, R. S., and Priestley, K. F. (2000), A. (2013), Tephrochronology and the late Holocene
Structure of the crust and uppermost mantle of Iceland volcanic and flood history of Eyjafjallajökull, Iceland,
from a combined seismic and gravity study, Earth Planet. J. Quat. Sci., 28(3), 237–247.
Sci. Lett., 181(3), 409–428. Duller, R. (2007), Depositional processes associated with vol­
Davies, S.M. (2015), Cryptotephras: the revolution in corre­ caniclastic jokulhlaups Myrdalssandur Iceland, Doctoral
lation and precision dating, J. Quat. Sci., 30(2), 114–130. dissertation, Keele University.
Decaulne, A. (2007), Snow‐avalanche and debris‐flow haz­ Duller, R. A., Mountney, N. P., Russell, A. J., and Cassidy,
ards in the fjords of north‐western Iceland, mitigation N. C. (2008), Architectural analysis of a volcaniclastic
and prevention, Natural hazards, 41(1), 81–98. jökulhlaup deposit, southern Iceland: sedimentary
Decker, R. and Decker, B. (2005), Volcanoes, W.H. Freeman. evidence for supercritical flow, Sedimentology, 55(4),
­
Fourth Edition. 939–964.
Decriem, J., Árnadóttir, T., Hooper, A., Geirsson, H., Duller, R. A., Warner, N. H., McGonigle, C., De Angelis, S.,
Sigmundsson, F., Keiding, M., et al. (2010), The 2008 Russell, A. J., and Mountney, N. P. (2014), Landscape
May 29 earthquake doublet in SW Iceland, Geophys. reaction, response, and recovery following the catastrophic
J. Int., 181(2), 1128–1146. 1918 Katla jökulhlaup, southern Iceland, Geophys. Res.
DeMets, C., Gordon, R. G., Argus, D. F., and Stein, S. Lett., 41(12), 4214–4221.
(1994), Effect of recent revisions to the geomagnetic Eason, D.E., Sinton, J.M., Grönvold, K., Kurz, M.D. (2015),
reversal time scale on estimates of current plate motions, Effects of deglaciation on the petrology and eruptive his­
Geophys. Res. Lett., 21(20), 2191–2194. tory of the Western Volcanic Zone, Iceland, Bull.
DeMets, C., Gordon, R. G., and Argus, D. F. (2010), Geologically Volcanol., 77, 47–57.
current plate motions, Geophys. J. Int., 181(1), 1–80. Einarsson, P. (1968), Geology, history of rocks and land.
Denk T., Grimsson F., Zetter R. and Simonarson L. (2011), Language and Culture, Reykjavik.
Late Cainozoic floras of Iceland. 15 million years of veg­ Einarsson, P. (1978), S‐wave shadows in the Krafla caldera
etation and climate history in the northern North Atlantic, in NE‐Iceland, evidence for a magma chamber in the
Topics Geol., 35. crust. Bull. Volcanol., 41(3), 187–195.
Dowdeswell, J. A., Hamilton, G. S., and Hagen, J. O. (1991), Einarsson, P. (1986), Seismicity along the eastern margin
The duration of the active phase on surge‐type glaciers: of the North American plate, in P.R. Vogt and B.E.
contrasts between Svalbard and other regions, J. Glaciol., Tu‐cholke (eds), The geology of North America, Vol. M,
37(127), 388–400. The western North Atlantic region, pp. 99–116, Geological
Drouin, V., Sigmundsson, F., Ófeigsson, B. G., Hreinsdóttir, Society of America, Boulder, CO.
S., Sturkell, E., and Einarsson, P. (2017a), Deformation in Einarsson, P. (1991), Earthquakes and present‐day tecto­
the Northern Volcanic Zone of Iceland 2008–2014: An nism in Iceland, Tectonophysics, 189(1–4), 261–279.
interplay of tectonic, magmatic, and glacial isostatic defor­ Einarsson, P. (2001), Structure and evolution of the Iceland
mation, J. Geophys. Res. Solid Earth, 122(4), 3158–3178. hotspot. Deutsch. Geophys. Gesell. Mitt., 1, 11–14.
Drouin, V., Sigmundsson, F., Verhagen, S., Ófeigsson, B. G., Einarsson, P. (2008), Plate boundaries, rifts and transforms
Spaans, K., and Hreinsdóttir, S. (2017b), Deformation at in Iceland, Jökull, 58(12), 35–58.
Krafla and Bjarnarflag geothermal areas, Northern Einarsson, P. and Brandsdóttir, B. (1980), Seismological evi­
Volcanic Zone of Iceland, 1993–2015, J. Volcanol. dence for lateral magma intrusion during the July 1978
Geotherm. Res., 344, 92–105. deflation of the Krafla volcano in NE‐Iceland, Report
Du, Z., Foulger, G. R., Julian, B. R., Allen, R. M., Nolet, G., UI‐79‐9‐7, University of Iceland, Reykjavík.
Morgan, W. J., et al. (2002), Crustal structure beneath Einarsson, P. and Brandsdóttir, B. (2000), Earthquakes in the
western and eastern Iceland from surface waves and Myrdalsjökull area, Iceland, 1978–1985; seasonal correla­
receiver functions, Geophys. J. Int., 149(2), 349–363. tion and connection with volcanoes, Jökull, 49, 59–73.
Dugmore, A. J. (1987), Holocene glacier fluctuations around Einarsson, P., Klein, F. W., and Björnsson, S. (1977), The
Eyjafjallajökull, south Iceland: a tephrochronological Borgarfjördur earthquakes of 1974 in West Iceland, Bull.
study, Doctoral dissertation, University of Aberdeen. Seismol. Soc. Am., 67, 197–208.
REFERENCES 181

Einarsson, P., Björnsson, S., Foulger, G., Stefánsson, R., Evans, D. J. A., Archer, S., and Wilson, D. J. H. (1999), A
and Skaftadóttir, T. (1981), Seismicity pattern in the South comparison of the lichenometric and Schmidt hammer
Iceland seismic zone, Earthq. Predict., 4, 141–151. dating techniques based on data from the proglacial areas
Einarsson, Þ. (1985), Geology: landscape and culture, of some Icelandic glaciers, Quat. Sci. Rev., 18, 13–41.
Reykjavik. 233 pp. Evans, D. J., Ewertowski, M., Orton, C., Harris, C., and
Einarson, Þ. (1994), Geology of Iceland: Rocks and land­ Guðmundsson, S. (2016), Snæfellsjökull volcano‐centred
scapes, Mál og menning, Reykjavík, 309 pp. ice cap landsystem, West Iceland, J. Maps, 12(5),
Einarsson, T. and Albertsson, K. J. (1988), The glacial his­ 1128–1137.
tory of Iceland during the past three million years, Philos. Evans, D. J. A., Ewertowski, M., and Orton, C. (2017).
Trans. R. Soc. B, 318(1191), 659–701. Skaftafellsjökull, Iceland: glacial geomorphology
Einarsson, T. (1950), Chemical analyses and differentiation recording glacier recession since the Little Ice Age,
of Hekla’s magma: The basic mechanism of volcanic erup­ J. Maps, 13(2), 1–12.
tions and the ultimate causes of volcanism, Vísindafélag Eyre, N. S. (2005), The use of salt injection and conductivity
islendinga. monitoring to infer nearmargin hydrological conditions on
Einarsson, T. (1960), The plateau basalt areas in Iceland. In Vestari‐Hagafellsjokull, Iceland, Ann. Glaciol., 40, 83–88.
On the geology and geophysics of Iceland (Guide to Eythorsson, J. (1935), On the variations of glaciers in
Excursion A2), pp. 5–20, XXI International Geological Iceland: some studies made in 1931, Geogr. Ann., 17(1–2),
Congress, Copenhagen. 121–137.
Einarsson, T. (1963), Pollen‐analytical studies on vegetation Eythórsson, J. (1952), The land below Vatnajökull, Jökull, 2,
and climate history of Iceland in late and post times, in, A. 1–4.
Löve and D. Löve, (eds), North Atlantic biota and their Fernandez‐Fernandez, J.M., Andres, N., Saemundsson, P.,
history, pp. 355–365, Pergamon. Brynjolfsson, S., and Palacios, D. (2018), High sensitivity
Eiríksson, J. (2008), Glaciation events in the Pliocene– of North Iceland (Tröllaskagi) debris‐free glaciers to
Pleistocene volcanic succession of Iceland, Jökull, 58, climatic change from the ‘Little Ice Age’ to the present,
315–329. The Holocene, 27(8), 1187–1200.
Eiriksson, J. and Geirsdóttir, A. (1991), A record of Pliocene Field, D. J., Boessenecker, R., Racicot, R. A., Ásbjörnsdóttir,
and Pleistocene glaciations and climate changes in North L., Jónasson, K., Hsiang, A. Y., et al. (2017), The oldest
Atlantic based on variations in volcanic and sedimentary marine vertebrate fossil from the volcanic island of
facies in Iceland, Mar. Geol., 101, 147–159. Iceland: a partial right whale skull from the high latitude
Eirikisson, J., Larsen, G., Knudsen, K. L., Heinemejer, J., Pliocene Tjörnes Formation, Palaeontology, 60(2),
and Simonarson, L. A. (2004), Marine reservoir age vari­ 141–148.
ability and water mass distribution in the Iceland Sea, Fitton, J. G., Saunders, A. D., Norry, M. J., Hardarson, B.
Quat. Sci. Rev., 23(20–22), 2247–2268. S., and Taylor, R. N. (1997), Thermal and chemical struc­
Elders, W. A., Friðleifsson, G. Ó., Zierenberg, R. A., Pope, ture of the Iceland plume, Earth Planet. Sci. Lett.,
E. C., Mortensen, A. K., Guðmundsson, Á., et al. (2011), 153(3–4), 197–208.
Origin of a rhyolite that intruded a geothermal well while Flóvenz, Ó. G., and Gunnarsson, K. (1991), Seismic crustal
drilling at the Krafla volcano, Iceland, Geology, 39(3), structure in Iceland and surrounding area, Tectonophysics,
231–234. 189(1), 1–17.
Elíasson, J. (2008), A glacial burst tsunami near Flowers, G. E., Björnsson, H., Geirsdóttir, Á., Miller, G. H.,
Vestmannaeyjar, Iceland, J. Coast. Res., 24(1), 13–20. and Clark, G. K. C. (2007), Glacier fluctuation and
Eliasson, S. (1974), Eldsumbrot i Jökulsárgljúfrum, Náttúrufræð­ inferred climatology of Langjokull ice cap through the
ingurinn, 44, 52–70. little ice age, Quat. Sci. Rev., 26, 2337–2353.
Eliasson, S. (1977), Molar un jokulsarhlaup og asbyrgi, Flude, S., Burgess, R., and McGarvie, D. W. (2008), Silicic
Naturalist, 47, 67–79. volcanism at Ljósufjöll, Iceland: Insights into evolution
Ellis, D. and Stoker, M. S. (2014), The Faroe–Shetland and eruptive history from Ar–Ar dating, J. Volcanol.
Basin: A regional perspective from the Paleocene to the Geotherm. Res., 169(3–4), 154–175.
present day and its relationship to the opening of the Flude, S., McGarvie, D. W., Burgess, R., and Tindle, A. G.
North Atlantic Ocean, Geol. Soc. Lond. Spec. Publ., (2010), Rhyolites at Kerlingarfjöll, Iceland: the evolution
397(1), 11–31. and lifespan of silicic central volcanoes, Bull. Volcanol.,
EPA (2017), Sulfur dioxide trends, Environmental Protection 72(5), 523–538.
Agency, Retrieved from https://www.epa.gov/air‐trends/ Forslund, T. and Gudmundsson, A. (1991), Crustal
sulfur‐dioxide‐trends. spreading due to dikes and faults in southwest Iceland,
Etzelmüller, B., Farbrot, H., Gudmundsson, A., Humlum, J. Struct. Geol., 13(4), 443–457.
O., Tveito, O.E., and Bjornsson, H. (2007), The regional Forsyth, D. A., Morel‐a‐l’Huissier, P., Asudeh, I., and
distribution of mountain permafrost in Iceland, Green, A. G. (1986), Alpha Ridge and Iceland—products
Permafrost Periglacial Proc., 18(2), 8–14. of the same plume? J. Geodyn., 6(1–4), 197–214.
182 REFERENCES

Foulger, G. R. (2006), Older crust underlies Iceland, Geirsdóttir, Á., Andrews, J. T., Ólafsdóttir, S., Helgadóttir,
Geophys. J. Int., 165(2), 672–676. G., and Hardardóttir, J. (2002), A 36 ky record of iceberg
Foulger, G. R. and Natland, J. H. (2003), Is “hotspot” volca­ rafting and sedimentation from north‐west Iceland, Polar
nism a consequence of plate tectonics? Science, 300(5621), Res., 21(2), 291–298.
921–922. Geirsdóttir, Á., Miller, G. H., and Andrews, J. T. (2007),
Foulger, G. R., Jahn, C. H., Seeber, G., Einarsson, P., Julian, Glaciation, erosion, and landscape evolution of Iceland,
B. R., and Heki, K. (1992), Post‐rifting stress relaxation at J. Geodyn., 43(1), 170–186.
the divergent plate boundary in northeast Iceland, Nature, Geirsdóttir, Á., Miller, G. H., Axford, Y., and Ólafsdóttir, S.
358(6386), 488. (2009), Holocene and latest Pleistocene climate and glacier fluc­
Franzson, H. (2000), Hydrothermal evolution of the Nesjavellir tuations in Iceland, Quat. Sci. Rev., 28(21–22), 2107–2118.
high‐temperature system, Iceland, in Proceedings of the Geirsdóttir, Á., Miller, G. H., and Larsen, D. J. (2016),
World Geothermal Congress, pp. 2075–2080. Landforms in Hvítárvatn, central Iceland, produced by
Franzson, H., Thordarson, S., Björnsson, G., Gudlaugsson, recent advances of surging and non‐surging glaciers, Geol.
S. T., Richter, B., Fridleifsson, G. O., and Thorhallsson, S. Soc. Lon. Mem., 46, 143–146.
(2002), Reykjanes high‐temperature field, SW‐Iceland. Geirsson, H., Árnadóttir, T., Völksen, C., Jiang, W., Sturkell,
Geology and hydrothermal alteration of well RN‐10, in E., Villemin, T., et al. (2006), Current plate movements
Workshop on Geothermal Reservoir Engineering, Vol. 27, across the Mid‐Atlantic Ridge determined from 5 years of
pp. 233–240. continuous GPS measurements in Iceland, J. Geophys.
Fridriksson, S. (1992), Vascular plants on Surtsey 1981– Res. Solid Earth, 111(B9).
1990, Surtsey Res., 10, pp. 17–30. Georgsson, G. and Pétursson, G. (1972), Fluorosis of sheep
Friedrich, W. L. (1966). On the geology of Brjanslaekur caused by the Hekla eruption in 1970, Fluoride, 5, 58–66.
(northwest Iceland) with special regard to fossil flora, Gíslason, G. (1973), Study of high‐temperature hydrothermal
Special Publication, Institut für Geologie und Mineralogie, alteration in Krísuvík and Námafjall, Unpublished BSc
Universität zu Köln. thesis, University of Iceland. [In Icelandic.]
Friese, N. (2008) Brittle tectonics of the Þingvellir and Gíslason, S. R., Hassenkam, T., Nedel, S., Bovet, N.,
Hengill volcanic systems, southwest Iceland: field studies Eiriksdottir, E. S., Alfredsson, H. A., et al. (2011),
and numerical modeling, Geodinam. Acta, 21(4), Characterization of Eyjafjallajökull volcanic ash particles
169–185. and a protocol for rapid risk assessment, Proc. Nat. Acad.
Friese, N., Bense, F. A., Tanner, D. C., Gústafsson, L. E., Sci., 108(18), 7307–7312.
and Siegesmund, S. (2013), From feeder dykes to scoria Gleick, P. H. (1996), Basic water requirements for human
cones: the tectonically controlled plumbing system of the activities: Meeting basic needs, Water Int., 21, 83–92.
Rauðhólar volcanic chain, Northern Volcanic Zone, Gordon, J. E. and Sharp, M. (1983), Lichenometry in dating
Iceland, Bull. Volcanol., 75(6), 717. recent glacial landforms and deposits, southeast Iceland,
Fronval, T. and Jansen, E. (1996), Late Neogene paleocli­ Boreas, 12(3), 192–200.
mates and paleo‐ceanography in the Iceland–Norwegian Grímsson, F., and Símonarson, L. A. (2008), Upper Tertiary
Sea: Evidence from the Iceland and Vøring Plateaus, Proc. non‐marine environments and climatic changes in Iceland.
Ocean Drill. Progr. Sci. Results, 151, 455–468. Jökull, 58, 303–314.
Fuchs, F., Lupi, M., Jakobsdóttir, S. S., Thordarson, T., and Grobe, H. (1987), A simple method for the determination of
Miller, S. A. (2013), Seismicity observed under the icerafted debris in sediment cores, Polarforschung, 57,
Snæfellsjökull volcano, Jökull, 63, 105–112. 123–126.
Garcia, S., Arnaud, N. O., Angelier, J., Bergerat, F., and Grönvold, K. (1972), Structural and petrochemical studies in
Homberg, C. (2003), Rift jump process in northern the Kerlingarfjöll region, southwest Iceland.
Iceland since 10 Ma from 40Ar/39Ar geochronology, Earth Grönvold, K. (1976), Variation and origin of magma types
Planet. Sci. Lett., 214(3–4), 529–544. in the Namafjall area, North Iceland. Bulletin de la Société
Gardner A S., Moholdt, S., Cogley, G., Wouters, J., Arendt, Géologique de France, 7(4), 869–870.
B., Wahr, J., et al. (2013), A reconciled estimate of glacier Grönvold, K. (1984), Myvatn fires 1724–1729:Chemical
contributions to sea level rise: 2003 to 2009, Science, 340, composition of the lava, Professional Paper 8401, Nordic
852–857. Volcanologicai Institute, Reykjavík, 30 pp.
Gardner, J. S. (1885), The Tertiary basaltic formation in Grönvold, K., Larsen, G., Einarsson, P., Thórarinsson, S.,
Iceland, Q. J. Geol. Soc., 41(1–4), 93–101. and Sæmundsson, K. (1983), The Hekla eruption 1980–
Geirsdóttir, A. and Eiríksson, J. (1994), Growth of an inter­ 1981. Bull. Volcanol., 46(4), 349–363.
mittent ice sheet in Iceland during late Pliocene and early Grönvold, K., Óskarsson, N., Johnsen, S. J., Clausen, H. B.,
Pleistocene, Quat. Res., 42, 115–130. Hammer, C. U., Bond, G., and Bard, E. (1995), Ash layers
Geirsdóttir, A. and Eiríksson, J. (1996), A review of studies from Iceland in the Greenland GRIP ice core correlated
of the earliest glaciation of Iceland, Terra Nova, 8(5), with oceanic and land sediments, Earth Planet. Sci. Lett.,
400–414. 135(1–4), 149–155.
REFERENCES 183

Gudmundsdóttir, E. R., Eiríksson, J., and Larsen, G. (2011), Gudmundsson, M. T. and Högnadóttir, T. (2007), Volcanic
Identification and definition of primary and reworked systems and calderas in the Vatnajökull region, central
tephra in Late Glacial and Holocene marine shelf sedi­ Iceland: Constraints on crustal structure from gravity
ments off North Iceland, J. Quat. Sci., 26(6), 589–602. data, J. Geodyn., 43(1), 153–169.
Gudmundsóttir, M., Brynjólfsdóttir, A., and Albertsson, A. Gudmundsson, M. T. and Milsom, J. (1997), Gravity and
(2010), The history of the blue lagoon in Svartsengi, in magnetic studies of the subglacial Grímsvötn volcano,
Proceedings of the World Geothermal Congress. Iceland: implications for crustal and thermal structure,
Gudmundsson, A. (1987a) Lateral magma flow, caldera col­ J. Geophys. Res. Solid Earth, 102(B4), 7691–7704.
lapse, and a new mechanism of large eruptions in Iceland, Gudmundsson, M. T., Björnsson, H., and Pálsson, F. (1995),
J. Volcanol. Geotherm. Res., 34, 65–78. Changes in jökulhlaup sizes in Grímsvötn, Vatnajökull,
Gudmundsson, A. (1987b), Tectonics of the Thingvellir fis­ Iceland, 1934–91, deduced from in‐situ measurements of
sure swarm, SW Iceland, J. Struct. Geol., 9(1), 61–69. subglacial lake volume, J. Glaciol., 41(138), 263–272.
Gudmundsson, A. (1992), The crustal structure of the sub­ Gudmundsson, M. T., Elíasson, J., Larsen, G., Gylfason, Á.
glacial Grimsvotn Volcano, VatnaJökull, Iceland, from mul­ G., Einarsson, P., Jóhannesson, T., Hákonardóttir, K. M.,
tiparameter geophysical surveys, PhD thesis, University of and Torfason, H. (2005), Overview on the dangers of
London, UK. volcanic eruptions and floods from the western part of
Gudmundsson, A. (1995), Ocean‐ridge discontinuities in Mýrdalsjökull and Eyjafjallajökull, in M. T. Gudmundsson
Iceland, J. Geol. Soc., 152, 1011–1015. and Á. G. Gylfason (eds), Hættumat vegan eldgosa og
Gudmundsson, A. T. (1997), A lively neighbour—occasion­ hlaupa frá vestanverdum Mýrdalsjökli og Eyjafjallajökli,
ally bad tempered. Atlantica, 2, 16–22. Ríkislögreglustjórninn, Háskólaútgáfan, pp. 11–44. [In
Gudmundsson, A. (1996), Volcanoes in Iceland: 10,000 years Icelandic.]
of volcanic history, Vaka‐Helgafell, Reykjavík. Gudmundsson, M. T., Pedersen, R., Vogfjörd, K.,
Gudmundsson, A. (2000), Dynamics of volcanic systems in Thorbjarnardóttir, B., Jakobsdóttir, S., and Roberts, M. J.
Iceland: example of tectonism and volcanism at juxta­ (2010), Eruptions of Eyjafjallajökull Volcano, Iceland,
posed hot spot and mid‐ocean ridge systems, Ann. Rev. Eos, Trans. Am. Geophys. Union, 91(21), 190–191.
Earth Planet. Sci., 28(1), 107–140. Gudmundsson, M. T., Jónsdóttir, K., Hooper, A., Holohan,
Gudmundsson, A. and Bäckström, K. (1991), Structure and E. P., Halldórsson, S. A., Ófeigsson, B. G., et al. (2016),
development of the Sveinagja graben, northeast Iceland, Gradual caldera collapse at Bárdarbunga volcano,
Tectonophysics, 200(1–3), 111–125. Iceland, regulated by lateral magma outflow, Science,
Gudmundsson, A., Brynjolfsson, S., and Jonsson, M. T. 353(6296).
(1993), Structural analysis of a transform fault‐rift zone Gudmundsson, O., Brandsdottir, B., Menke, W., and
junction in North Iceland, Tectonophysics, 220(1–4), Sigvaldason, G. E. (1994), The crustal magma chamber of
205–221. the Katla volcano in south Iceland revealed by 2‐D seismic
Gudmundsson, A., Friese, N., Galindo, I., and Philipp, S. L. undershooting, Geophys. J. Int., 119(1), 277–296.
(2008), Dike‐induced reverse faulting in a graben, Geology, Gudnason, J., Thordarson, T., Houghton, B. F., and Larsen,
36(2), 123–126. G. (2017), The opening subplinian phase of the Hekla
Gudmundsson, A., Oskarsson, N., Gronvold, K., 1991 eruption: properties of the tephra fall deposit, Bull.
Sæmundsson, K., Sigurðsson, O., Stefansson, R., et al. Volcanol., 79(5), 34.
(1992), The 1991 eruption of Hekla, Iceland, Bull. Gunnarsson, B., Marsh, B. D., and Taylor, H. P. (1998),
Volcanol., 54(3), 238–246. Generation of Icelandic rhyolites: silicic lavas from the
Gudmundsson, B. T. and Arnórsson, S. (2005), Secondary Torfajökull central volcano, J. Volcanol. Geotherm. Res.,
mineral–fluid equilibria in the Krafla and Námafjall 83(1), 1–45.
geothermal systems, Iceland, Appl. Geochem., 20(9),
­ Gunnlaugsson, E. and Gíslason, G. (2005), Preparation for a
1607–1625. new power plant in the Hengill geothermal area, Iceland, in
Guðmundsson, H. J. (1997), A review of the Holocene envi­ Proceedings World Geothermal Congress.
ronmental history of Iceland, Quat. Sci. Rev., 16, 81–92. Gylfadóttir, S. S., Kim, J., Helgason, J. K., Brynjólfsson, S.,
Gudmundsson, H., Bjornsson, H., and Palsson, E. (2017), Höskuldsson, Á., Jóhannesson, T., et al. (2017), The 2014
Changes of Breiðamerkurjökull glacier, SE‐Iceland, from Lake Askja rockslide‐induced tsunami: Optimization of
its late nineteenth century maximum to the present, Geogr. numerical tsunami model using observed data, J. Geophys.
Ann. Ser. A Phys. Geogr., 99(4), 338–352. Res. Oceans, 122(5), 4110–4122.
Gudmundsson, M. T. (2005), Subglacial volcanic activity in Hagen, S. (1995), Watermass characteristics and climate in
Iceland, in C. Caseldine, et al. (eds), Iceland—Modern the Nordic Sea during the last 10,200 years: based on pale­
Processes and Past Environments, pp. 127–151, oceanographic proxies retrieved from sediment cores from
Developments in Quaternary Sciences, Vol. 5, Elsevier. the continental shelves off northern Norway and
Gudmundsson, M. T. and Björnsson, H. (1991), Eruptions in southwestern Iceland, Doctoral Thesis, University of
Grímsvötn, Vatnajökull, Iceland, 1934–1991, Jökull, 41, 21–45. Tromso. 114 pp.
184 REFERENCES

Hamilton, S. J. and Whalley, W. B. (1995), Preliminary Hauksson, E. (1983), Episodic rifting and volcanism at
results from the lichenometric study of the Nautardálur Krafla in north Iceland: Growth of large ground fissures
rock glacier, Tröllaskagi, northern Iceland, Geomorphology, along the plate boundary, J. Geophys. Res. Solid Earth,
12(2), 123–132. 88(B1), 625–636.
Hamilton, C. W., Thordarson, T., and Fagents, S. A. (2010), Heiken, G. and Wohletz, K. (1985), Volcanic ash, University
Explosive lava–water interactions I: architecture and Presses of California, Chicago, Harvard and MIT.
emplacement chronology of volcanic rootless cone groups Heimisson, E. R., Einarsson, P., Sigmundsson, F., and
in the 1783–1784 Laki lava flow, Iceland, Bull. Volcanol., Brandsdóttir, B. (2015), Kilometer‐scale Kaiser effect
72(4), 449–467. identified in Krafla volcano, Iceland, Geophys. Res. Lett.,
Hannesdóttir, H., Björnsson, H., Pálsson, F., Adalgeirsdottir, G., 42(19), 7958–7965.
and Gudmundsson, S. (2014), Area, volume and mass changes Heki, K., Foulger, G. R., Julian, B. R., and Jahn, C. H.
of southeast Vatnajökull ice cap, Iceland, from the Little Ice Age (1993), Plate dynamics near divergent boundaries:
maximum in the late 19th century to 2010, American Geophysical implications of postrifting crustal deforma­
Geophysical Union, Fall Meeting, abstract id. C42A‐05. tion in NE Iceland, J. Geophys. Res. Solid Earth, 98(B8),
Hansell, A. and Oppenheimer, C. (2004), Health hazards 14279–14297.
from volcanic gases: A systematic literature review. Arch. Helgason, J. and Duncan, R. A. (2001), Glacial–interglacial
Environ. Health, 59, 628–639. history of the Skaftafell region, southeast Iceland, 0–5
Harðarson, B. S. (1993), Alkalic rocks in Iceland with special Ma, Geology, 29(2), 179–182.
reference to the Snæfellsjökull volcanic system, PhD thesis, Hewitt, I. J. and Schoof, C. (2017), Models for polythermal
University of Edinburgh. ice sheets and glaciers, The Cryosphere, 11(1), 541–551.
Harðarson, B. S. (2014), Geothermal exploration of the Hjartardóttir, Á.R., Einarsson, P., and Sigurdsson, H.
Hengill high‐temperature field, Presented at Short Course (2009), The fissure swarm of the Askja volcanic system
IX on Exploration for Geothermal Resources, 001374011, along the divergent plate boundary of N Iceland, Bull.
Lake Bogoria and Lake Naivasha, Kenya, 2–23 November. Volcanol., 71, 961.
Harðarson, B. S. and Fitton, J. G. (1991), Increased mantle Hjartardóttir, Á. R., Einarsson, P., Bramham, E., and
melting beneath Snæfellsjökull volcano during late Wright, T. J. (2012), The Krafla fissure swarm, Iceland,
Pleistocene deglaciation, Nature, 353, 62–64. and its formation by rifting events. Bull. Volcanol., 74(9),
Harðarson, B. S., Fitton, J. G., Ellam, R. M., and Pringle, 2139–2153.
M. S. (1997), Rift relocation—a geochemical and geo­ Hjartardóttir, Á. R., Einarsson, P., Gudmundsson, M. T., and
chronological investigation of a palaeo‐rift in northwest Högnadóttir, T. (2016a), Fracture movements and graben
Iceland, Earth Planet. Sci. Lett., 153(3), 181–196. subsidence during the 2014 Bárðarbunga dike intrusion in
Hards, V. L., Kempton, P. D., Thompson, R. N., and Iceland, J. Volcanol. Geotherm. Res., 310, 242–252.
Greenwood, P. B. (2000), The magmatic evolution of the Hjartardóttir, Á. R., Einarsson, P., Magnúsdóttir, S.,
Snaefell volcanic centre; an example of volcanism during Björnsdóttir, Þ., and Brandsdóttir, B. (2016b), Fracture
incipient rifting in Iceland, J. Volcanol. Geotherm. Res., systems of the Northern Volcanic Rift Zone, Iceland: an
99(1), 97–121. onshore part of the Mid‐Atlantic plate boundary, Geol.
Harmon, R. S. and Hoefs, J. (1995), Oxygen isotope hetero­ Soc. Lond. Spec. Publ., 420(1), 297–314.
geneity of the mantle deduced from global 18O systematics Hjartarson, Á. (2005), The Late Miocene Tinná Central
of basalts from different geotectonic settings, Contrib. Volcano, North Iceland, Jökull, 55, 33–48.
Mineral. Petrol., 120(1), 95–114. Hjartarson, Á. (2003), Skagafjörður unconformity: North
Harning, D. J., Geirsdóttir, Á., Miller, G. H., and Anderson, Iceland and its geological history, PhD thesis, University of
L. (2016), Episodic expansion of Drangajökull, Vestfirðir, Copenhagen.
Iceland, over the last 3 ka culminating in its maximum Hjartarson, Á., Friðleifsson, G. Ó., and Hafstað, Þ. H.
dimension during the Little Ice Age, Quat. Sci. Rev., 152, (1997), Berggrunnur í Skagafjarðardölum og jarðganga­
118–131. leiðir, Orkustofnun.
Hartley, M. E. and Thordarson, T. (2012), Formation of Hjartarson, Á., Erlendsson, Ö., and Blischke, A. (2017), The
Öskjuvatn caldera at Askja, North Iceland: Mechanism Greenland–Iceland–Faroe Ridge Complex, Geol. Soc.
of caldera collapse and implications for the lateral flow Lond. Spec. Publ., 447(1), 127–148.
hypothesis, J. Volcanol. Geotherm. Res., 227, 85–101. Hjort, C., Ingólfsson, Ó., and Norðdahl, H. (1985), Late
Hartley, M. E., and Thordarson, T. (2013), The 1874–1876 Quaternary geology and glacial history of Hornstrandir,
volcano‐tectonic episode at Askja, North Iceland: Lateral northwest Iceland: a reconnaissance study, Jökull, 35(9),
flow revisited. Geochem., Geophys., Geosyst., 14(7), e29.
2286–2309. Hlodversdottir, A., Petrusdottir, G., Carlsen, H.K., and
Hartley, M.E., Thordarson, T., and de Joux, A. (2016), Gislason, T. (2016), Long‐term health effects of the
Postglacial eruptive history of the Askja region, North Eyjafjallajökull volcanic eruption: a prospective cohort
Iceland, Bull. Volcanol., 78, 28. study in 2010 and 2013, Public Health, 6(9), 1–13.
REFERENCES 185

Hoernle, K., Lundstrom, C., Hauff, F., and van den Bogaard, Ingólfsson, Ó. (2008), Outline of the physical geography and
C. (2009), Time‐scales for magmatic differentiation at the geology of Svalbard.
Snæfellsjökull central volcano, western Iceland: con­ Ingólfsson, Ó. (2013), Surging glaciers in Iceland—research
straints from U–Th–Pa–Ra disequilibria in post‐glacial status and future challenges, in EGU General Assembly
lavas, Geochim. Cosmochim. Acta, 73(4), 1120–1144. Conference Abstracts, Vol. 15.
Holbrook, W. S., Larsen, H. C., Korenaga, J., Dahl‐Jensen, Ingolfsson, Ó. and Norddahl, H. (1994), A review of the
T., Reid, I. D., Kelemen, P. B., et al. (2001), Mantle environmental history of Iceland, 13,000–9,000 yr BP,
thermal structure and active upwelling during continental J. Quat. Sci., 9, 147–150.
breakup in the North Atlantic, Earth Planet. Sci. Lett., Ingolfsson, O., Nordahl, H., and Haflidasaon, H. (1995),
190(3), 251–266. Rapid isostatic rebound in southwestern Iceland at the
Hollingsworth, J., Leprince, S., Ayoub, F., and Avouac, J. P. end of the last glaciation, Boreas, 14, 245–259.
(2013), New constraints on dike injection and fault slip Ingólfsson, Ó., Björck, S., Haflidason, H., and Rundgren, M.
during the 1975–1984 Krafla rift crisis, NE Iceland, J. (1997), Glacial and climatic events in Iceland reflecting
Geophys. Res. Solid Earth, 118(7), 3707–3727. regional North Atlantic climatic shifts during the Pleistocene‐
Hooper, A., Ófeigsson, B., Sigmundsson, F., Lund, B., Holocene transition, Quat. Sci. Rev., 16(10), 1135–1144.
Einarsson, P., Geirsson, H., and Sturkell, E. (2011), Intergovernmental Panel on Climate Change (2016), AR5
Increased capture of magma in the crust promoted by ice‐ Synthesis Report: Climate Change 2014, 169 pp.
cap retreat in Iceland, Nature Geoscience, 4(11), 783. Islam, M. T. and Sturkell, E. (2017), Rheological responses
Hoppe, G. (1968), Grímsey and the maximum extent of the to plate boundary deformation at the Eastern Volcanic
last glaciation of Iceland, Geogr. Ann. Ser. A, Phys. Geogr., Zone in Iceland, Tectonophysics, 717, 16–26.
50(1), 16–24. Islam, M. T., Sturkell, E., LaFemina, P., Geirsson, H.,
Hoppe, G. (1982), The extent of the last inland ice sheet of Sigmundsson, F., and Ólafsson, H. (2016), Continuous
Iceland, Jökull, (32), 3–11. subsidence in the Thingvellir rift graben, Iceland: Geodetic
Horwell, C. J. and Baxter, P. J. 2006. The respiratory health observations since 1967 compared to rheological models
hazards of volcanic ash: A review for volcanic risk mitiga­ of plate spreading, J. Geophys. Res. Solid Earth, 121(1),
tion, Bull. Volcanol., 69, 1–24. 321–338.
Höskuldsson, Ó. R., and Ólafsdóttir, R. (2002), Pyroclastic Jagan, A. (2010), Tephra stratigraphy and geochemistry from
flows formed in the eruption of Hekla 2000, in Proceedings, three Icelandic lake cores: a new method for determining
25th Nordic Geological Winter Meeting, Reykjavík, source volcano of tephra layers, MSc dissertation, University
Abstract. of Edinburgh.
Höskuldsson, Á., Óskarsson, N., Pedersen, R., Grönvold, Jakobsson, S. P. (1972a), Chemistry and distribution pattern
K., Vogfjörð, K., and Ólafsdóttir, R. (2007), The millen­ of recent basaltic rocks in Iceland, Lithos, 5(4), 365–386.
nium eruption of Hekla in February 2000, Bull. Volcanol., Jakobsson, S. P. (1972b), On the consolidation and palago­
70(2), 169–182. nitization of the tephra of the Surtsey volcanic island,
Houghton, B. F., Swanson, D. A., Rausch, J., Carey, R. J., Iceland, Surtsey Res., 6, 1–8.
Fagents, S. A., and Orr, T. R. (2013), Pushing the Volcanic Jakobsson, S. P. (1979a), Outline of the petrology of Iceland,
Explosivity Index to its limit and beyond: Constraints Department of Geology and Geography, Museum of
from exceptionally weak explosive eruptions at Kilauea in Natural History, Reykjavík.
2008, Geology, 41(6), 627–630. Jakobsson, S. P. (1979b), Petrology of recent basalts of the
Hubbard, A., Sugden, D., Dugmore, A., Norddahl, H., and Eastern Volcanic Zone, Iceland, Acta Natural. Islan., 26,
Pétursson, H. G. (2006), A modelling insight into the 1–103
Icelandic Last Glacial Maximum ice sheet, Quat. Sci. Jakobsson, S. P. and Gudmundsson, M. T. (2008), Subglacial
Rev., 25(17–18), 2283–2296. and intraglacial volcanic formations in Iceland, Jökull, 58,
IMO (2018), Iceland Meteorological Office, Available at: 179–196.
http://en.vedur.is/ [Accessed 2018.] Jakobsson, S. P. and Johnson, G. L. (2012), Intraglacial
IMO (2019a), Iceland Meteorological Office, Northeast Iceland, ­volcanism in the Western Volcanic Zone, Iceland, Bull.
Askja. https://en.vedur.is/ [Accessed on 4 September 2019.] Volcanol., 74(5), 1141–1160.
IMO (2019b), Avalanches: Avalanches in Iceland. [Accessed Jakobsson, S. P. and Moore, J. G. (1986), Hydrothermal
20 September 2019.] minerals and alteration rates at Surtsey volcano, Iceland,
Iceland Tourism Bureau (2018), Available at: https://iceland. Geol. Soc. Am. Bull., 97(5), 648–659.
nordicvisitor.com [Accessed 2018.] Jakobsson, S. P., Jónsson, J., and Shido, F. (1978), Petrology
Imsland, P. (1973), The geology of Sveifluháls, BSc disserta­ of the western Reykjanes peninsula, Iceland, J. Petrol.,
tion, University of Iceland, 87 pp. 19(4), 669–705.
Ingolfsson, Ó. (1988), Glacial history of the lower Jakobsson, S. P., Jónasson, K., and Sigurðsson, I. A. (2008),
Borgarfjordur area, western Iceland, Geol. For., 110(4), The three igneous rock series of Iceland, Jökull, 58,
293–309. 117–138.
186 REFERENCES

Jakobsson, S. P., Thors, K., Vésteinsson, Á. T., and namics in a subglacial geothermal lake under the Western
Ásbjörnsdóttir, L. (2009), Some aspects of the seafloor Skaftá cauldron of the Vatnajökull ice cap, Iceland,
morphology at Surtsey volcano: The new multibeam Geophys. Res. Lett., 34(19).
bathymetric survey of 2007, Surtsey Res., 12, 99–108. Jóhannesson, T., Björnsson, H., Pálsson, F., Sigurðsson, O.,
Janebo, M. H. (2016), Historical explosive eruptions in Hekla and Þorsteinsson, Þ. (2011), LiDAR mapping of the
and Askja volcanoes; eruption dynamics and source param­ Snæfellsjökull ice cap, western Iceland, Jökull, 61, 19–32.
eters, Doctoral dissertation, University of Hawaii at Jóhannesson, T., Björnsson, H., Magnússon, E.,
Manoa, Honolulu. Guðmundsson, S., Pálsson, F., Sigurðsson, O., et al.
Jeddi, Z., Tryggvason, A., and Gundmundsson, O. (2016), (2013), Ice‐volume changes, bias estimation of mass‐
The Katla volcanic system imaged using local earthquakes balance measurements and changes in subglacial lakes
recorded with a temporary seismic network, J. Geophys. derived by lidar mapping of the surface of Icelandic
Res. Solid Earth, 121(10), 6989–7703. ­glaciers, Ann. Glaciol., 54(63), 63–74.
Jennings, A. E., Syvitski, J., Gerson, L., Gronvold, K., Jóhannsdóttir, G. E. (2007), Mid Holocene to late glacial teph­
Geirsdottir, A., Hardardóttir, J., Andrews, J. and Hagen, rochronology in West Iceland as revealed in three lacustrine
S. (2000), Chronology and paleoenvironments during the environments, M.S. thesis, University of Iceland, Reykjavík.
late Weichselian deglaciation of the southwest Iceland Johnsen, G. V., Bjornsson, A., and Sigurðsson, S. (1980),
shelf, Boreas, 29, 167–183. Gravity and elevation changes caused by magma
Jennings, A. E., Grönvold, K., Hilberman, R., Smith, M., movement beneath the Krafla caldera, northeast Iceland,
and Hald, M. (2002), High‐resolution study of Icelandic J. Geophys., 47(1–3), 132–140.
tephras in the Kangerlussuaq Trough, southeast Johnsen, S. J., Dahl‐Jensen, D., Gundestrup, N., Steffensen,
Greenland, during the last deglaciation, J. Quat. Sci., J. P., Clausen, H. B., Miller H., et al. (2001), Oxygen iso­
17(8), 747–757. tope and palaeotemperature records from six Greenland
Jiang, H., Eiriksson, J., Schultz, M., Knudsen, K.L., and ice‐core stations: Camp Century, Dye‐3, GRIP, GISP2,
Seidenkrantz M. S. (2005), Evidence for solar forcing of Renland and NorthGRIP, J. Quat. Sci., 16(4), 299–307.
sea surface temperature on the North Icelandic Shelf dur­ Johnston, S. T. and Thorkelson, D. J. (2000), Continental
ing the late Holocene, Geology, 33, 73–76. flood basalts: episodic magmatism above long‐lived
Jóhannesson, H. (1980), Stratigraphy and the development hotspots, Earth Planet. Sci. Lett., 175(3), 247–256.
of rift zones in West Iceland, Náttúrufrædingurinn, 50, Jónasson, K. (1994), Rhyolite volcanism in the Krafla
13–31. central volcano, north‐east Iceland, Bull. Volcanol.,
Johannesson, H. (1982), Summary of the geology of 56(6–7), 516–528.
Snaefellsnes, Arbok Ferdafelags Islands, 1, 151–172. Jones, B., Renaut, R. W., Torfason, H., and Owen, R. B.
Jóhannesson, H. and Sæmundsson, K. (1998), Geological (2007), The geological history of Geysir, Iceland: a teph­
map of Iceland, 1:500,000, bedrock geology, Icelandic rochronological approach to the dating of sinter, J. Geol.
Institute of Natural History and Iceland Geodetic Survey, Soc., 164(6), 1241–1252.
Reykjavík. Jónsson, J. (1982), Notes on the Katla volcanoglacial debris
Jóhannesson, H. and Sæmundsson, K. (2009), Geological flows, Jökull, (32), 61–68.
Map of Iceland. 1:600 000. Bedrock Geology, 1st edn, Jónsson, J. (1978), Geological map of the Reykjanes
Icelandic Institute of Natural History, Reykjavík. Peninsula, Orkustofnun, Report OS‐JHD 7831, 332 pp.
Jóhannesson, H., Jónsson, J., and Flores, R. M. (1981), A [In Icelandic.]
short account of the Holocene tephrochronology of the Jónsson, J. (1998), Eyjafjöll: draft geology. Lower ace.
Snæfellsjökull central volcano, western Iceland, Jökull, Jónsson, Ó. and Rist, S. (1972), Distribution and risk of
31, 23–30. snow avalanches in Iceland, Jökull, 21, 24–44.
Jóhannesson, H., Jakobsson, S. P., and Sæmundsson, K. Jónsson, S., Einarsson, P., and Sigmundsson, F. (1997),
(1982), Geological map of Iceland, sheet 6, south Iceland, Extension across a divergent plate boundary, the Eastern
1:250,000 geol map, 2nd edition. Icelandic Museum of Volcanic Rift Zone, south Iceland, 1967–1994, observed
Natural History and Iceland Geodetic Survey. with GPS and electronic distance measurements, J. Geophys.
Johannesson, T. (1991), Modelling the effect of climatic Res. Solid Earth, 102(B6), 11913–11929.
warming on the Hofsjökull Ice Cap, central Iceland, Jude‐Eton, T. C., Thordarson, T., Gudmundsson, M. T., and
Hydrol. Res., 22(2), 81–94. Oddsson, B. (2012), Dynamics, stratigraphy and proximal
Johannesson, T., Siggurdsson, O., Einarsson, B. and dispersal of supraglacial tephra during the ice‐confined
Thorsteinsson, T. (2006), Mass balance modeling of 2004 eruption at Grímsvötn Volcano, Iceland, Bull.
HofsJökull ice cap based on data from 1988–2004, Volcanol., 74(5), 1057–1082.
Orkustofnun Project Report, Hydrological Service Katz, R. W. (2013), Statistics of extremes in climate change,
Division, OS‐2006/004, 42 pp. Climatic Change, 100(1), 71–76.
Jóhannesson, T., Thorsteinsson, T., Stefánsson, A., Gaidos, Keiding, M., Arnadottir, T., Lund, B., Sturkell, E., Geirsson,
E. J., and Einarsson, B. (2007), Circulation and thermody­ H., and Slunga, R. (2006), Strain and stress along the
REFERENCES 187

Reykjanes Peninsula oblique spreading ridge, south Lacasse, C., Sigurðsson, H., Jóhannesson, H., Paterne, M.,
Iceland, in AGU Fall Meeting Abstracts. and Carey, S. (1995), Source of ash zone 1 in the North
Keiding, M., Árnadóttir, T., Sturkell, E., Geirsson, H., and Atlantic, Bull. Volcanol., 57(1), 18–32.
Lund, B. (2008), Strain accumulation along an oblique Lacasse, C., Sigurðsson, H., Carey, S. N., Jóhannesson, H.,
plate boundary: the Reykjanes Peninsula, southwest Thomas, L. E., and Rogers, N. W. (2007), Bimodal volca­
Iceland, Geophys. J. Int., 172(2), 861–872. nism at the Katla subglacial caldera, Iceland: insight into
Keiding, M., Lund, B., and Árnadóttir, T. (2009), the geochemistry and petrogenesis of rhyolitic magmas,
Earthquakes, stress, and strain along an obliquely diver­ Bull. Volcanol., 69(4), 373.
gent plate boundary: Reykjanes Peninsula, southwest LaFemina, P. C., Dixon, T. H., Malservisi, R., Árnadóttir,
Iceland. J. Geophys. Res. Solid Earth, 114(B9). T., Sturkell, E., Sigmundsson, F., and Einarsson, P. (2005),
Kellerer‐Pirklbauer, A., Wangensteen, B., Farbrot, H., and Geodetic GPS measurements in south Iceland: Strain
Etzelmüller, B. (2008), Relative surface age‐dating of rock accumulation and partitioning in a propagating ridge
glacier systems near Hólar in Hjaltadalur, Northern system, J. Geophys. Res. Solid Earth, 110(B11).
Iceland, J. Quat. Sci., 23(2), 137–151. Lamb, H. H. (2002), Climate, history and the modern world,
Kim, D., Brown, L. D., Árnason, K., Águstsson, K., and 2nd edn, Taylor and Francis Publishing, 46 pp.
Blanck, H. (2017), Magma reflection imaging in Krafla, Langella, G., Paoletti, V., DiPippo, R., Amoresano, A.,
Iceland, using microearthquake sources, J. Geophys. Res. Steinunnardottir, K., and Milano, M. (2017), Krafla geo­
Solid Earth, 122(7), 5228–5242. thermal system, northeastern Iceland: Performance assessment
Kirkbride, M. P. and Dugmore, A. J. (2001), Can the late of alternative plant configurations, Geothermics, 69, 74–92.
“Little Ice Age” glacial maximum in Iceland be dated by Landl, B., Björnsson, H., and Kuhn, M. (2003), The energy
lichenometry? Climatic Change, 48, 151–167. balance of calved ice in Lake Jökulsarlón, Iceland, Arctic
Kirkbride, M. P. and Dugmore, A. J. (2008), Two millennia Antarctic Alpine Res., 35(4), 475–481.
of glacier advances from southern Iceland dated by teph­ Larsen, D. J., Miller, G. H., Geirsdóttir, Á., and Thordarson,
rochronology, Quat. Res., 70(3), 398–411. T. (2011). A 3000‐year varved record of glacier activity
Kjartansson, G. (1960), Geological map of Iceland, 1: and climate change from the proglacial Lake Hvítárvatn,
250.000, Iceland Survey Department, Reykjavík. Quat. Sci. Rev., 30, 2715–2731.
Klein, E. M. and Langmuir, C. H. (1987), Global correla­ Larsen, D. J., Miller, G. H., Geirsdottir, A., and Thordarson,
tions of ocean ridge basalt chemistry with axial depth and T. (2013), A 3000‐year varved record of glacier activity
crustal thickness, J. Geophys. Res. Solid Earth, 92(B8), and climate change from the proglacial lake Hvítárvatn,
8089–8115. Iceland, Quat. Sci. Rev., 30(19–20), 2715–2731.
Knudsen, K. L., Eirksson, J., Jiang, H., and Jonsdottir, I. Larsen, G. (1984), Recent volcanic history of the Veidivötn
(2009). Palaeoceanography and climate changes off north fissure swarm, southern Iceland—an approach to volcanic
Iceland during the last millennium: comparison of fora­ risk assessment. J. Volcanol. Geotherm. Res., 22(1–2),
minifera, diatoms and ice‐rafted debris with instrumental 33–58.
and documentary data, J. Quat. Sci., 24, 457–468. Larsen, G. (1999), The eruption of the Eyjafjallajökull vol­
Kodaira, S., Mjelde, R., Gunnarsson, K., Shiobara, H., and cano in 1821–1823, Research Report RH‐28‐99. Science
Shimamura, H. (1998), Evolution of oceanic crust on the Institute, Reykjavík.
Kolbeinsey Ridge, north of Iceland, over the past 22 Myr, Larsen, G. (2000), Holocene eruptions within the Katla
Terra Nova, 10, 27–31. volcanic system, south Iceland: characteristics and envi­
Kristinsdóttir, B. D. (2010), Comparison of fractional stress ronmental impact, Jökull, 49, 1–28.
and single‐axis fracture strength of rock samples from Larsen, G. (2002), A brief overview of eruptions from ice‐
Helguvík and Hólahnúkar. Bachelor dissertation, Sigillum covered and ice‐capped volcanic systems in Iceland during
Universitatis Islandiae. the past 11 centuries: frequency, periodicity and implica­
Kristjánsson, L. (1979), The shelf area around Iceland, tions, Geol. Soc. Lond. Spec. Publ., 202(1), 81–90.
Jökull. 29, 3–6. Larsen, G. (2010), Katla: tephrochronology and eruption
Kristjánsson, L. and McDougall, I. (1982), Some aspects of history, in Developments in Quaternary Sciences, Vol. 13,
the late Tertiary geomagnetic field in Iceland, Geophys. J. pp. 23–49, Elsevier.
Int., 68(2), 273–294. Larsen, G. and Eiríksson, J. (2008a), Late Quaternary terres­
Kristjansson, L., Duncan, R.A., and Gudmundsson, R.A. trial tephrochronology of Iceland—frequency of explosive
(1998), Stratigraphy, palaeomagnetism and age of volca­ eruptions, type and volume of tephra deposits, J. Quat.
nics in the upper regions of ÞJórsárdalur valley, central Sci., 23(2), 109–120.
southern Iceland, Boreas, 27(1), 1–13. Larsen, G. and Eiríksson, J. (2008b), Holocene tephra
Lacasse, C. and Garbe‐Schonberg, D. (2001), Explosive archives and tephrochronology in Iceland—a brief over­
silicic volcanism in Iceland and the Jan Mayen area during view, Jökull, 58, 229–250.
the last 6 Ma: Sources and timing of major eruptions, Larsen, G. and Thórarinsson, S. (1977), H4 and other acidic
J. Volcanol. Geotherm. Res., 107(1), 113–147. Hekla tephra layers, Jökull, 27, 28–46.
188 REFERENCES

Larsen, G., Gudmundsson, M. T., and Björnsson, H. (1998), eruption of Askja volcano, Iceland: combined fractional
Eight centuries of periodic volcanism at the center of the crystallization and selective contamination in the genera­
Iceland hotspot revealed by glacier tephrostratigraphy, tion of rhyolitic magma, Mineral. Mag., 51(360),
Geology, 26(10), 943–946. 183–202.
Larsen, G., Dugmore, A., and Newton, A. (1999), Maclennan, J. (2008), Concurrent mixing and cooling of
Geochemistry of historical‐age silicic tephras in Iceland, melts under Iceland, J. Petrol., 49(11), 1931–1953.
The Holocene, 9(4), 463–471. Maclennan, J., Jull, M., McKenzie, D., Slater, L., and
Larsen, G., Newton, A. J., Dugmore, A. J., and Gronvold, K. (2002), The link between volcanism and
Vilmundardottir, E. G. (2001), Geochemistry, dispersal, deglaciation in Iceland, Geochem., Geophys., Geosyst.,
volumes and chronology of Holocene silicic tephra layers 3(11), 1–25.
from the Katla volcanic system, Iceland, J. Quat. Sci., Magnúsdóttir, S. and Brandsdóttir, B. (2011), Tectonics of
16(2), 119–132. the Þeistareykir fissure swarm, Jökull, 61, 65–79.
Larsen, H. C. (1978), Offshore continuation of East Magnúsdóttir, S., Brandsdóttir, B., Driscoll, N., and Detrick,
Greenland dyke swarm and North Atlantic Ocean R. (2015), Postglacial tectonic activity within the
formation, Nature, 274(5668), 220. Skjálfandadjúp Basin, Tjörnes Fracture Zone, offshore
Lawver, L. A. and Müller, R. D. (1994), Iceland hotspot Northern Iceland, based on high resolution seismic stra­
track, Geology, 22(4), 311–314. tigraphy, Mar. Geol., 367, 159–170.
Leadbetter, S. J. and Hort, M. C. (2011), Volcanic ash hazard Magnússon, B., Magnússon, S. H., and Fridriksson, S.
climatology for an eruption of Hekla Volcano, Iceland, (2009), Development in plant colonization and succession
J. Volcanol. Geotherm. Res., 199(3–4), 230–241. on Surtsey during 1999–2008, Surtsey Res., 12, 57–76.
Leng, M. and Lewis, J. P. (2016), Oxygen isotopes in mol­ Magnússon, E., Gudmundsson, M. T., Roberts, M. J.,
luscan shell: Applications in environmental archaeology, Sigurðsson, G., Höskuldsson, F., and Oddsson, B. (2012),
J. Human Palaeoecol., 21(3), 295–306. Ice–volcano interactions during the 2010 Eyjafjallajökull
Licciardi, J. M., Kurz, M. D., and Curtice, J. M. (2007), eruption, as revealed by airborne imaging radar, J. Geophys.
Glacial and volcanic history of Icelandic table mountains Res. Solid Earth, 117(B7).
from cosmogenic 3He exposure ages, Quat. Sci. Rev., Magnússon, E., Belart, J. M., Pálsson, F., Anderson, L. S.,
26(11–12), 1529–1546. Gunnlaugsson, Á. Þ., Berthier, E., et al. (2016), The sub­
Liu, E. J., Cashman, K. V., Rust, A. C., and Hoskuldsson, A. glacial topography of Drangajökull ice cap, NW‐Iceland,
(2017), Contrasting mechanisms of magma fragmentation deduced from dense RES‐profiling, Jökull, 66, 1–16.
during coeval magmatic and hydromagmatic activity: the Maizels, J. (1989), Sedimentology, paleoflow dynamics and
Hverfjall Fires fissure eruption, Iceland, Bull. Volcanol., flood history of jokulhlaup deposits: paleohydrology of
79, 65–68. Holocene sediment sequences in southern Iceland sandur
Lock, W.G. (1881), Askja, Iceland’s largest volcano, Translated deposits, J. Sed. Res. 59(2).
at https://bookdome.com/nature/Askja‐Volcano/index. Maizels, J. (1991), The origin and evolution of Holocene
html#.XXAYZuhKhPY. [Accessed 4 September 2019.] sandur deposits in areas of jökulhlaup drainage, Iceland,
Lorenz, V. (1975), Formation of phreatomagmatic maar– in Environmental change in Iceland: Past and present, pp.
diatreme volcanoes and its relevance to kimberlite dia­ 267–302, Springer, Dordrecht.
tremes, Phy. Chem. Earth, 9, 17–27. Maizels, J. (1992), Boulder ring structures produced during
Lund, J. W., Bjelm, L., Bloomquist, G., and Mortensen, A. jökulhlaup flows: origin and hydraulic significance, Geogr.
K. (2008), Characteristics, development and utilization of Ann. Ser. A, Phys. Geogr., 74(1), 21–33.
geothermal resources‐a Nordic perspective, Episodes, Maizels, J. (1993), Lithofacies variations within sandur
31(1), 140–147. deposits: The role of runoff regime, flow dynamics and
Lupi, M., Geiger, S., Carey, R. J., and Thordarson, T. (2011), sediment supply characteristics, Sed. Geol., 85(1–4),
A model for syn‐eruptive groundwater flow during the 299–325.
phreatoplinian phase of the 28–29 March 1875 Askja vol­ Manabe, S. and Stouffer, R. J. (1995), Simulation of abrupt
cano eruption, Iceland, J. Volcanol., 203(3–4), 146–157. climate change induced by freshwater input to the North
Maccaferri, F., Rivalta, E., Passarelli, L., and Jonsson, S. Atlantic Ocean, Nature, 378(6553), 165.
(2013), The stress shadow induced by the 1975–1984 Martin, E., Paquette, J. L., Bosse, V., Ruffet, G., Tiepolo,
Krafla rifting episode, J. Geophys. Res. Solid Earth, 118(3), M., and Sigmarsson, O. (2011), Geodynamics of rift–
1109–1121. plume interaction in Iceland as constrained by new
MacDonald, R., McGarvie, D. W., Pinkerton, H., Smith, R. 40
Ar/39Ar and in situ U–Pb zircon ages, Earth Planet. Sci.
L., and Palacz, A. (1990), Petrogenetic evolution of the Lett., 311(1–2), 28–38.
Torfajökull Volcanic Complex, Iceland I. Relationship Martin, E. and Sigmarsson, O. (2007), Crustal thermal state
between the magma types, J. Petrol., 31(2), 429–459. and origin of silicic magma in Iceland: the case of
MacDonald, R., Sparks, R. S. J., Sigurðsson, H., Mattey, D. Torfajökull, Ljósufjöll and Snæfellsjökull volcanoes,
P., McGarvie, D. W., and Smith, R. L. (1987), The 1875 Contrib. Mineral. Petrol., 153(5), 593–605.
REFERENCES 189

Martini, I. P., Brookfield, M. E., and Sadura, S. (2001), Moore, J. G., Jakobsson, S., and Holmjarn, J. (1992),
Principles of glacial geomorphology and geology, College Subsidence of Surtsey volcano 1967–1991, Bull. Volcanol.,
Division, Pearson. 55(1–2), 17–24.
Marren, P. M. and Toomath, S. C. (2014), Channel pattern Morales, J. R. V. (1992), Geology and geothermal consider­
of proglacial rivers: topographic forcing due to glacier ations of Krisuvík valley, Reykjanes peninsula, Iceland,
retreat, Earth Surf. Process. Landf., 39(7), 943–951. United Nations University.
Massé, G., Rowland, S. J., Sicre, M. A., Jacob, J., Eystein, J., Mottram, D. and Benn, D.I. (2009), Testing crevasse‐depth
Simon, J. and Bell, T. (2008), Abrupt climate changes for models: a field study at Breiðamerkurjokull, Iceland,
Iceland during the last millennium: Evidence from high J. Glaciol., 55(192), 746–752.
resolution sea ice reconstructions, Earth Planet. Sci. Lett., Muehlenbachs, K., Anderson Jr, A. T., and Sigvaldason, G.
269(3–4), 565–569. E. (1974), Low 18O basalts from Iceland, Geochim.
Mattsson, H. B., and Höskuldsson, Á. (2011), Contempo­ Cosmochim. Acta, 38(4), 577–588.
raneous phreatomagmatic and effusive activity along the Nelson, S. A. (2003), Igneous rock and plate tectonics,
Hverfjall eruptive fissure, north Iceland: Eruption chro­ University of Tulane.
nology and resulting deposits, J. Volcanol. Geotherm. Res., Németh, K. and Smith, I. E. M. (2017), Source to surface
201(1–4), 241–252. model of monogenetic volcanism: a critical review, Geol.
Mawejje, P. (2007), Geothermal exploration and geological Soc. Lond. Spec. Publ., 446(1), 1–28.
mapping at Seltun in Krysuvík geothermal field, Reykjanes Newhall, C. G. and Dzurisin, D. (1988), Historical unrest at
peninsula, SW‐Iceland, Geothermal training programme the large calderas of the world, Vol. 2, No. 1855, Department
report, pp. 257–276. of the Interior, US Geological Survey.
McDougall, I., Kristjansson, L., and Sæmundsson, K. Newhall, C. G. and Self, S. (1982), The Volcanic Explosivity
(1984), Magnetostratigraphy and geochronology of Index (VEI): An estimate of explosive magnitude for his­
northwest Iceland, J. Geophys. Res. Solid Earth, 89(B8), torical volcanism, J. Geophys. Res., 87(2), 1231–1238.
7029–7060. Newhall, C.G., Self, S., Robock, A. (2018), Anticipating
McKinzey, K. M., Orwin, J. F., and Bradwell, T. (2005), A future Volcanic Explosivity Index (VEI) 7 eruptions and
revised chronology of key Vatnajökull (Iceland) outlet their chilling impacts, Geosphere, 14(2), 572–603.
glaciers during the Little Ice Age, Ann. Glaciol., 42, Nielsen, G., Maack, R., Gudmundsson, A., and Gunnarsson,
171–179. G. I. (2000), Completion of Krafla geothermal power plant,
McPeek, T., Wang, X. Z., Brown, K., and Filipelli, G. M. in Proceedings of the World Geothermal Congress, pp.
(2007), Rates of carbon ingrowth and nutrient release 3259–3264.
from young Icelandic basalts, Jökull, 57, 37–44. Nielsen, N. (1937), Renewed activity of Great Geysir, Geogr.
Metzger, S., Jónsson, S., and Geirsson, H. (2011), Locking J., 84, 451–454.
depth and slip‐rate of the Húsavík Flatey fault, North Norðdahl, H. (1981), A prediction of minimum age for the
Iceland, derived from continuous GPS data 2006–2010, Weichselian maximum glaciation in North Iceland,
Geophys. J. Int., 187(2), 564–576. Boreas, 10(4), 471–476.
Metzger, S., Jónsson, S., Danielsen, G., Hreinsdóttir, S., Norðdahl, H. (1990), Late Weichselian and early Holocene
Jouanne, F., Giardini, D., and Villemin, T. (2012), Present deglaciation history of Iceland, Jökull, 40, 27–50.
kinematics of the Tjörnes Fracture Zone, North Iceland, Norddahl, H. (1991), A review of glaciation maximum con­
from campaign and continuous GPS measurements, cept and the deglaciation of Eyjafjodur, North Iceland, in
Geophys. J. Int., 192(2), 441–455. J. K. Maizels and C. Caseldine (eds), Environmental change
Meyer, B. S. (2005), in A. N. Krot, E. R. D. Scott, and B. in Iceland: past and present, pp. 31–47, Kluwer Academic,
Reipurth (eds), Chondrites and the protoplanetary disk, Dordrecht.
ASP Conference Series 341, San Francisco, 515 pp. Norðdahl, H. and Haflidason, H. (1992), The Skogar tephra,
Meyer, P. S., Sigurðsson, H., and Schilling, J. G. (1985), a Younger Dryas marker in north Iceland, Boreas, 21(1),
Petrological and geochemical variations along Iceland’s 23–41.
neovolcanic zones. J. Geophys. Res. Solid Earth, 90(B12), Norðdahl, H. and Ingólfsson, Ó. (2015), Collapse of the
10043–10072. Icelandic ice sheet controlled by sea‐level rise? Arktos,
Mittelstaedt, E., Ito, G., and Behn, M.D. (2008), Mid‐ocean 1(1), 13.
ridge jumps associated with hotspot magmatism, Earth Norðdahl, H. and Pétursson, G. P. (2005) Relative sea level
Planet. Sci. Lett., 266(3–4), 256–270 changes in Iceland: new aspects of the Weichselian deglaci­
Moorbath, S., Sigurðsson, H., and Goodwin, R. (1968), ation of Iceland, in C. Caseldine, C. Russel, A. Hardardottir,
K–Ar ages of the oldest exposed rocks in Iceland, Earth and J. Knudsend (eds), Iceland: Modern processes and past
Planet. Sci. Lett., 4(3), 197–205. environments, pp. 25–78, Elsevier, Amsterdam.
Moore, J. G. (1985), Structure and eruptive mechanisms at Norðdahl, H., Ingólfsson, Ó., Pétursson, H. G., and
Surtsey Volcano, Iceland, Geological Magazine, 122(6), Hallsdóttir, M. (2008), Late Weichselian and Holocene
649–661. environmental history of Iceland, Jökull, 58, 343–364.
190 REFERENCES

Obermann, A., Lupi, M., Mordret, A., Jakobsdottir, S., and tonics, chemical fractionation and isotope evolution of the
Miller, S.A. (2016), 3D‐ambient noise Rayleigh wave crust, J. Geophys. Res. Solid Earth, 90(B12), 10011–10025.
tomography of Snæfellsjökull volcano, Iceland, Bull. Óskarsson, F., Armannsson, H., Olafsson, M.,
Volcanol., 317, 42–52. Sveinbjornsdottir, A.E., and Markusson, S.H. (2013), The
Obrochta, T., Crowley, J., Channel, E.T., Hodell, D.A., Theistareykir Geothermal Field, NE Iceland: Fluid chem­
Baker, P.A., Seki, A., and Yokoyama, Y. (2014), Climate istry and production properties, Proc. Earth Planet. Sci.,
variability and ice‐sheet dynamics during the last three 7, 644–647.
glaciations, Earth Planet. Sci. Lett., 406, 198–212. Ostra, B. (1984), A search for a threshold in the relationship
Oddsson, B. (1982), Rock quality designation and drilling of air pollution to mortality—a reanalysis of data on
rate correlated with lithology and degree of alteration in London winters, Environ, Health Perspect., 58, 397–399.
volcanic rocks from the 1979 Surtsey drill hole, IX, pp. 94– Oxford Economics (2010), The economic impacts of air travel
97, Surtsey Research Society, Reykjavık. restrictions due to volcanic ash, A report prepared for
Oddsson, B. (2007), The Grímsvötn eruption in 2004: Airbus, 15 pp.
Dispersal and total mass of tephra and comparison with Pagli, C. and Sigmundsson, F. (2008), Will present day gla­
plume transport models, MSc thesis, University of Iceland. cier retreat increase volcanic activity? Stress induced by
Oddsson, B., Gudmundsson, M. T., Edwards, B. R., recent glacier retreat and its effect on magmatism at the
Thordarson, T., Magnússon, E., and Sigurðsson, G. (2016), Vatnajökull ice cap, Iceland, Geophys. Res. Lett., 35(9).
Subglacial lava propagation, ice melting and heat transfer Pagli, C., Sigmundsson, F., Lund, B., Sturkel, E., Geirsson,
during emplacement of an intermediate lava flow in the H., Einarsson, P., Arnadottir, T., and Hreinsdottir, S.
2010 Eyjafjallajökull eruption, Bull. Volcanol., 78(7), 48. (2007), Glacio‐isostatic deformation around the
Óladóttir, B. A., Larsen, G., Thordarson, T., and Sigmarsson, Vatnajökull ice cap, Iceland, induced by recent climate
O. (2005), The Katla volcano S‐Iceland: Holocene tephra warming: GPS observations and finite element modeling,
stratigraphy and eruption frequency, Jökull, 55, 53–74. J. Geophys. Res. Solid Earth, 112(B8), 1–12.
Óladóttir, B. A., Sigmarsson, O., Larsen, G., and Thordarson, Palmason, G., Arnórsson, S., Fridleifsson, I. B.,
T. (2008), Katla volcano, Iceland: magma composition, Kristmannsdóttir, H., Sæmundsson, K., Stefansson, V.
dynamics and eruption frequency as recorded by Holocene and Kristjánsson, L. (1979), The Iceland crust: evidence
tephra layers, Bull. Volcanol., 70(4), 475–493. from drillhole data on structure and processes, pp. 43–65,
Óladóttir, B. A., Sigmarsson, O., Larsen, G., and Devidal, J. American Geophysical Union.
L. (2011), Provenance of basaltic tephra from Vatnajökull Pálmason, G., Stefansson, V., Thorhallsson, S., and
subglacial volcanoes, Iceland, as determined by major‐and Thorsteinsson, T. (1983), Geothermal field developments in
trace‐element analyses, The Holocene, 21(7), 1037–1048. Iceland, in Proceedings, Ninth workshop on geothermal
Ólafsdóttir, R. and Dowling, R. (2014), Geotourism and reservoir engineering, pp. 37–52.
geoparks—a tool for geoconservation and rural Pálmason, G., Johnsen, G. V., Torfason, H., Sæmundsson,
development in vulnerable environments: a case study K., Ragnars, K., Haraldsson, G. I., and Halldórsson, G.
from Iceland, Geoheritage, 6(1), 71–87. K. (1985), Assessment of geothermal energy in Iceland,
Ólafsdóttir, T. (1975), A moraine ridge on the Iceland shelf, Report OS‐85076/JHD‐10, National Energy Authority.
west of Breidafjördur. Natturufredinggurin, 45, 31–37. Palmer, S., Shepherd, A., Björnsson, H., and Pálsson, F.
Ólafsson, J. (1979), Physical characteristics of lake Mývatn (2009), Ice velocity measurements of Langjökull, Iceland,
and river Laxá, Oikos, 38–66. from interferometric synthetic aperture radar (inSAR),
O’Nions, R. K., Evensen, N. M., Carter, S. R., and Hamilton, J. Glaciol., 5(5), 834–838.
P. J. (1979), Isotope geochemical studies of North Atlantic Palsson, F., Gudmundsson, S., Bjornsson, H., Berthier, E.,
Ocean basalts and their implications for mantle evolution Magnunsson, E., Gudmundsson, S. and Haraldsson, H.
(pp. 342–351), American Geophysical Union. (2012), Mass and volume changes of Langjökull ice cap,
Opheim, J. A. and Gudmundsson, A. (1989), Formation and Iceland, ∼1890 to 2009, deduced from old maps, satellite
geometry of fractures, and related volcanism, of the images and in situ mass balance measurements, Jökull, 62,
Krafla fissure swarm, northeast Iceland, Geol. Soc. Am. 81–93.
Bull., 101(12), 1608–1622. Pálsson, F., Gunnarsson, A., Jónsson, Þ., Steinþórsson, S.,
Oppenheimer, C., Orchard, A., Stoffel, M., Newfield, T. P., and Pálsson, H. S. (2016), Vatnajökull: Mass balance,
Guillet, S., Corona, C., et al. (2018), The Eldgjá eruption: meltwater drainage and surface velocity of the glacial year
timing, long‐range impacts and influence on the 2014–15.
Christianisation of Iceland, Climatic Change, 147, 1–13. Pálsson, S., Eyþórsson, J., Hannesson, P., and Steindórsson,
Oskarsson, N., Sigvaldason, G. E., and Steinthorsson, S. S. (1983), Svein Pálsson’s travel book: diaries and essays
(1982), A dynamic model of rift zone petrogenesis and the 1791–1797, Snæland issue.
regional petrology of Iceland, J. Petrol., 23(1), 28–74. Passarelli, L. and Brodsky, E. (2012), The correlation bet­
Oskarsson, N., Steinthorsson, S., and Sigvaldason, G. E. ween run‐up and repose times of volcanic eruptions,
(1985), Iceland geochemical anomaly: origin, volcanotec­ Geophys. J. Int., 188, 1025–1045.
REFERENCES 191

Pasvanoglu, S., Kristmannsdóttir, H., Bjönsson, S., and Principato, S. M. and Johnson, J. S. (2009), Using a GIS to
Torfason, H. (1998), Geochemical study of the Geysir geo­ quantify patterns of glacial erosion on Northwest Iceland:
thermal field in Haukadalur, S‐Iceland. United Nations implications for independent ice sheets, Arctic Antarctic
University. Alpine Res., 41(1), 128–137.
Pedersen, G., Vilmundardóttir, O. K., Falco, N., Principato, S. M., Geirsdóttir, Á., Jóhannsdóttir, G. E.,
Sigurmundsson, F. S., Rustowicz, R., Belart, J. M. C., and and Andrews, J. T. (2006), Late Quaternary glacial and
Benediktsson, J. A. (2016), Environmental mapping and deglacial history of eastern Vestfirðir, Iceland using cos­
monitoring of Iceland by remote sensing (EMMIRS), in mogenic isotope (36Cl) exposure ages and marine cores,
EGU General Assembly Conference Abstracts, Vol. 18, J. Quat. Sci., 21(3), 271–285.
p. 12432. Ragnarsson, Á. (2005), Geothermal development in Iceland
Pedersen, G. B. M., Höskuldsson, A., Dürig, T., Thordarson, 2000–2004, Fish Farm., 4(9).
T., Jonsdottir, I., Riishuus, M. S., et al. (2017), Lava field Ragnarsson, A. (2015), Geothermal development in Iceland
evolution and emplacement dynamics of the 2014–2015 2010‐2014. Proceedings World Geothermal Congress,
basaltic fissure eruption at Holuhraun, Iceland, J. Volcanol. Melbourne, pp. 1–15.
Geotherm. Res., 340, 155–169. Rahmstorf, S. (1997), Risk of sea‐change in the Atlantic,
Pedersen, R., Sigmundsson, F., Feigl, K. L., and Árnadóttir, Nature, 388(6645), 825.
T. (2001), Coseismic interferograms of two Ms= 6.6 earth­ Raymo, M. E., Ruddiman, W. F., Backman, J., Clement, B.
quakes in the south Iceland seismic zone, June 2000, M., and Martinson, D. G. (1989). Late Pliocene variation
Geophys. Res. Lett., 28(17), 3341–3344. in northern hemisphere ice sheets and North Atlantic
Pedersen, R., Sigmundsson, F., and Masterlark, T. (2009), deep water circulation, Paleoceanography, 4(4), 413–444.
Rheologic controls on inter‐rifting deformation of the Rickers, F., Fichtner, A., and Trampert, J. (2013), The
Northern Volcanic Zone, Iceland, Earth Planet. Sci. Lett., Iceland–Jan Mayen plume system and its impact on
281(1), 14–26. mantle dynamics in the North Atlantic region: evidence
Perlt, J., Heinert, M., and Niemeier, W. (2008), The from full‐waveform inversion, Earth Planet. Sci. Lett.,
continental margin in Iceland—a snapshot derived from 367, 39–51.
combined GPS networks, Tectonophysics, 447(1–4), Roaldset, E. (1983), Tertiary (Miocene‐Pliocene) interbasalt
155–166. sediments, NW‐and W‐Iceland. Jökull, (33), 39–56.
Petersen, G. N., Bjornsson, P., Arason, P., and von Lewis, S. Roberts, D. G. and Hunter, M. (1979), Bathymetry of the
(2012a), Two weather radar time series of the altitude of northeast Atlantic: continental margin around the British
the volcanic plume during the May 2011 eruption of Isles, Deep Sea Res. Part A Oceanogr. Res. Pap., 26(4),
Grimsvotn, Iceland, Earth Syst. Sci. Data, 4(1), 121–127. 417–428.
Petersen, G. N., Bjornsson, H., and Arason, P. (2012b), The Roche, D., Paillard, D., and Cortijo, E. (2004), Constraints
impact of the atmosphere on the Eyjafjallajökull 2010 on the duration and freshwater release of Heinrich event 4
eruption plume. Journal of Geophysical Research: through isotope modelling, Nature, 432, 379–382.
Atmospheres, 117(D20). Rögnvaldsson, S. T., Gudmundsson, A., and Slunga, R.
Pétursson, H. Norðdahl, O., and Ingólfsson, O. (2015), Late (1998), Seismotectonic analysis of the Tjörnes Fracture
Weichselian history of relative sea level changes in Iceland Zone, an active transform fault in north Iceland,
during a collapse and subsequent retreat of marine based J. Geophys. Res. Solid Earth, 103(B12), 30117–30129.
ice sheet, Cuad. Invest. Geogr., 41(2), 261–277. Romagnoli, C. and Jakobsson, S. P. (2015), Post‐eruptive
Phleger, F. (1949), Submarine geology and Pleistocene morphological evolution of island volcanoes: Surtsey as a
research, Bull. Geol. Soc. Am., 60, 1457–1459. modern case study, Geomorphology, 250, 384–396.
Pope, A. and Dockery, D. W. (2006), Health effects of fine Rossi, M. J. (1996), Morphology and mechanism of eruption
particulate air pollution: lines that connect, J. Air Waste of postglacial shield volcanoes in Iceland, Bull. Volcanol.,
Man. Assoc., 56(6), 709–742. 57(7), 530–540.
Pope, L., Willis, I. C., Pope, A., Miles, E. S., Arnold, N. S., Rubin, A.M. (1995), Propagation of magma‐filled cracks,
and Rees, W.G. (2016), Contrasting snow and ice albedos Ann. Rev. Earth Planet. Sci., 23, 287–336.
derived from MODIS, Landsat ETM+ and airborne data Ruddiman, W.F. (1977). North Atlantic Ice‐Rafting: A
from Langjökull, Iceland, Remoting Sens. Environ., 175, Major Change at 75,000 Years Before the Present. Science,
183–195. 196(4295): 1208–1211.
Priesnitz, K. and Schunke. E. (1978), An approach to the Ruddiman, W. F. and MacIntyre, A. (1981), The North
ecology of permafrost in central Iceland, Proceedings of Atlantic Ocean during the last deglaciation, Palaeogeogr.,
the Third International Conference on Permafrost, Palaeoclimatol., Palaeoecol., 35, 145–214.
Edmonton. Rundgren, M., Ingólfsson, Ó., Björck, S., Jiang, H., and
Principato, S. M. (2008), Geomorphic evidence for Holocene Haflidason, H. (1997), Dynamic sea‐level change during
glacial advances and sea level fluctuations on eastern the last deglaciation of northern Iceland, Boreas, 26(3),
Vestfirðir, northwest Iceland, Boreas, 37(1), 132–145. 201–215.
192 REFERENCES

Rungred, M. and Ingolfsson, O. (1999), Plant survival in deglaciation in Iceland on mantle melt production rates,
Iceland during glacial periods? J. Biogeogr., 26(2), J. Geophys. Res. Solid Earth, 118(7), 3366–3379.
387–396. Schmidt, R. and Schmincke, H. (2000), Seamounts and
Russell, A. J. (2018), Glacifluvial and glacilacustrine land­ island building, in H. Sigurðsson (ed.), Encyclopedia of
forms and deposits of Skeiðarárjökull and Skeiðarársandur, volcanoes, pp. 383–402, Academic Press, New York.
in D.G.A Evans (ed.), Glacial landsystems of southeast Schmidt, A., Ostro, B., Carslaw, K. S., Wilson, M.,
Iceland: Quaternary applications, pp. 86–101, Field Guide. Thordarson, T., Mann, G. W., and Simmons, A. J. (2011),
Quaternary Research Association, London. Excess mortality in Europe following a future Laki‐style
Russell, A. J., Tweed, F. S., and Knudsen, Ó. (2000), Flash Icelandic eruption,Proc. Nat. Acad. Sci., 108(38),
flood at Sólheimajökull heralds the reawakening of 15710–15715.
an Icelandic subglacial volcano, Geol. Today, 16(3), Schomacker, A. (2010), Expansion of ice‐marginal lakes at
102–106. the Vatnajökull ice cap, Iceland, from 1999 to 2009,
Russell, A. J., Tweed, F. S., Roberts, M. J., Harris, T. D., Geomorphology, 119(3–4), 232–236.
Gudmundsson, M. T., Knudsen, Ó., and Marren, P. M. Schomacker, A., Brynjolfsson, J. M., Anderassen, J.M.,
(2010), An unusual jökulhlaup resulting from subglacial Gudmunsdotir, E.R., Olsen, J., Odgaard, B.V., Hakansson,
volcanism, Sólheimajökull, Iceland, Quat. Sci. Rev., L., et al. (2016). The Drangajökull ice cap, northwest
29(11–12), 1363–1381. Iceland, persisted into the early‐mid Holocene, Quat. Sci.
Russell, J. K., Edwards, B. R., Porritt, L., and Ryane, C. Rev., 148, 68–84.
(2014), Tuyas: a descriptive genetic classification, Quat. Schone, B. R., Pfeiffer, M., Pohlmann, T., and Siegismund,
Sci. Rev., 87, 70–81. F. (2005), A seasonally resolved bottom‐water tempera­
Sæmundsson, K. (1967), Outline of the structure of SW‐ ture record for the period AD 1866–2002 based on shells
Iceland, Visindafelag Islendinga. of Arctica islandica (Mollusca, North Sea), Int. J.
Sæmundsson, K. (1973), Straumrákaðarklappir í kringum Climatol., 25, 947–962.
Ásbyrgi, Náttúrufræðingurinn, 43, 52–60. Schuler, J., Greenfield, T., White, R. S., Roecker, S. W.,
Sæmundsson, K. (1974), Evolution of the axial rifting zone Brandsdóttir, B., Stock, J. M., et al. (2015), Seismic
in northern Iceland and the Tjornes fracture zone, Geol. imaging of the shallow crust beneath the Krafla central
Soc. Am. Bull., 85(4), 495–504. volcano, NE Iceland, J. Geophys. Res. Solid Earth, 120(10),
Sæmundsson, K. (1978), Fissure swarms and central vol­ 7156–7173.
canoes of the neovolcanic zones of Iceland, Geol. J. Spec. Selbekk, R. S. and Trønnes, R. G. (2007), The 1362 AD
Issue, 10, 415–432. Öræfajökull eruption, Iceland: Petrology and geochem­
Sæmundsson, K. (1979), Outline of the geology of Iceland, istry of large‐volume homogeneous rhyolite, J. Volcanol.
Jökull, 29, 7–28. Geotherm. Res., 160(1), 42–58.
Sæmundsson, K. (1986), Subaerial volcanism in the western Self, S. and Sparks, R. S. J. (1978), Characteristics of wide­
North Atlantic, The Geology of North America, 1000, 69–86. spread pyroclastic deposits formed by the interaction of
Saemundsson, K. (1991). Geology of the Krafla system, in silicic magma and water, Bull. Volcanol., 41(3), 196.
A. Gardarsson and Á. Einarsson (eds), Náttúra Mývatns. Sella, G. F., Dixon, T. H., and Mao, A. (2002), REVEL: A
pp. 25–95, Hid Íslenska Náttúrufraedifélag, Reykjavík. model for recent plate velocities from space geodesy, J.
Sæmundsson, K. and Karson, J. A. (2006), Stratigraphy and Geophys. Res., 107, ETG 11‐1‐31, doi:10.1029/2000JB000033
tectonics of the Húsavík‐Western Tjörnes Area, Iceland Shackleton, N. J., Backman, J., Zimmerman, H., Kent, D.V.,
Geosurvey. Hall, M.A., Roberts, D.G., et al. (1984), Oxygen isotope
Sæmundsson, Th., Petursson, H.G., and Decaulne, A. calibration of the onset of ice‐rafting and history of glaci­
(2003), Triggering factors for rapid mass movements in ation in the North Atlantic region. Nature, 307, 620–623.
Iceland, in: D. Rickenman and C. I. Chen (eds), Debris‐ Sheath, H. C. and Canon‐Tapia, E. (2015), Are flood basalt
flow hazards and mitigation: mechanics, prediction, and eruptions monogenetic or polygenetic? Int. J. Earth Sci.,
assessment, pp. 167–178, Mill Press, Rotterdam. 104(8), 2147–2162.
Saunders, A. D., Jones, S. M., Morgan, L. A., Pierce, K., Shorttle, O. and Maclennan, J. (2011), Compositional trends
Widdowson, M., and Xu, Y. G. (2007), Regional uplift of Icelandic basalts: Implications for short–length scale
associated with continental large igneous provinces: The lithological heterogeneity in mantle plumes, Geochem.,
roles of mantle plumes and the lithosphere, Chem. Geol., Geophys., Geosyst., 12(11).
241(3–4), 282–318. Sigbjarnarson, G. (1983), The Quaternary alpine glaciation
Schattel, N., Portnyagin, M., Golowin, R., Hoernle, K., and and marine erosion in Iceland, Jökull, 33, 87–98.
Bindeman, I. (2014), Contrasting conditions of rift and Sigmarsson, O., Karlsson, H. R., and Larsen, G. (2000), The
off‐rift silicic magma origin on Iceland, Geophys. Res. 1996 and 1998 subglacial eruptions beneath the
Lett., 41(16), 5813–5820. Vatnajökull ice sheet in Iceland: Contrasting geochemical
Schmidt, P., Lund, B., Hieronymus, C., Maclennan, J., and geophysical inferences on magma migration, Bull.
Árnadóttir, T., and Pagli, C. (2013), Effects of present‐day Volcanol., 61(7), 468–476.
REFERENCES 193

Sigmarsson, O., Maclennan, J., and Carpentier, M. (2008), Sigvaldason, G. E. (1968), Structure and products of sub­
Geochemistry of igneous rocks in Iceland: a review, aquatic volcanoes in Iceland, Contr. Mineral. Petrol., 18,
Jökull,58, 139–160. 1–16.
Sigmarsson, O., Haddadi, B., Carn, S., Moune, S., Sigvaldason, G. E. (1974a), Basalts from the centre of the
Gudnason, J., Yang, K., and Clarisse, L. (2013), The assumed Icelandic mantle plume. J. Petrol., 15(3),
sulfur budget of the 2011 Grímsvötn eruption, Iceland, 497–524.
Geophys. Res. Lett., 40(23), 6095–6100. Sigvaldason, G. E. (1974b), The petrology of Hekla and
Sigmundsson, F. (1991), Post‐glacial rebound and astheno­ origin of silicic rocks in Iceland. HF Leiftur.
sphere viscosity in Iceland, Geophys. Res. Lett., 18, Sigvaldason, G. E. (1979), Rifting, magmatic activity and
1131–1134. interaction between acid and basic liquids, Report 79(03),
Sigmundsson, F. (2006), Iceland geodynamics: crustal defor­ Nordic Volcanological Institute.
mation and divergent plate tectonics, Springer Science and Sigvaldason, G. E. (2002), Volcanic and tectonic processes
Business Media. coinciding with glaciation and crustal rebound: an early
Sigmundsson, F., Hooper, A., Hreinsdóttir, S., Vogfjörd, K. Holocene rhyolitic eruption in the Dyngjufjöll volcanic
S., Ófeigsson, B. G., Heimisson, E. R., and Drouin, V. centre and the formation of the Askja caldera, north
(2015), Segmented lateral dyke growth in a rifting event at Iceland, Bull. Volcanol., 64(3–4), 192–205.
Bárðarbunga volcanic system, Iceland, Nature, 517(7533), Sigvaldason, G. E., Annertz, K., and Nilsson, M. (1992),
191. Effect of glacier loading/deloading on volcanism: postgla­
Sigmundsson, F., Einarsson, P., Bilham, R., and Sturkell, E. cial volcanic production rate of the Dyngjufjöll area,
(1995), Rift‐transform kinematics in south Iceland: central Iceland, Bull. Volcanol., 54(5), 385–392.
Deformation from Global Positioning System measure­ Simkin, T., Seibert, L., McClelland, W.G., Melson, D.,
ments, 1986 to 1992, J. Geophys. Res. Solid Earth, 100(B4), Bridge, C.G., Newhall, G., and Latter, J. (1981), Volcanoes
6235–6248. of the world, Hutchinson Ross, New York.
Sigmundsson, F., Vadon, H., and Massonnet, D. (1997), Simonarson, L. A. (1979). On climatic changes in Iceland,
Readjustment of the Krafla spreading segment to crustal Jökull, 29, 4446.
rifting measured by satellite radar interferometry, Geophys. Sinton, J., Grönvold, K., and Sæmundsson, K. (2005),
Res. Lett., 24(15), 1843–1846. Postglacial eruptive history of the western volcanic zone,
Sigurðsson, H. (1970), Structural origin and plate tectonics Iceland, Geochem., Geophys., Geosyst., 6(12).
of the Snæfellsnes volcanic zone, Western Iceland, Earth Skelton, A., Sturkell, E., Jakobsson, M., Einarsson, D.,
Planet. Sci. Lett., 10(1), 129–135. Tollefsen, E., and Orr, T. (2016), Dimmuborgir: a root­
Sigurdson, H. (1983), Sigurdur Thorainsson 1911–1983, less shield complex in northern Iceland, Bull. Volcanol.,
Eos, Trans. Am. Geophys. Union, 64(28), 450. 78(5), 40.
Sigurðsson, H. and Sparks, R. S. J. (1978), Rifting episode in Smellie, J. L. (2000), Subglacial eruptions, in H. Sigurðsson
north Iceland in 1874–1875 and the eruptions of Askja (ed.), Encyclopedia of volcanoes, pp. 403–418, Academic
and Sveinagja, Bull. Volcanol., 41(3), 149–167. Press, San Diego, CA.
Sigurðsson, H. and Sparks, R. S. J. (1981), Petrology of rhy­ Smellie, J. L. (2006), The relative importance of supraglacial
olitic and mixed magma ejecta from the 1875 eruption of versus subglacial meltwater escape in basaltic subglacial
Askja, Iceland, J. Petrol., 22(1), 41–84. tuya eruptions: An important unresolved conundrum,
Sigurðsson, O. (1998), Glacier variations in Iceland 1930– Earth Sci. Rev., 74, 241–268.
1995, Jökull, 45, 3–26. Smith, K. T. and Dugmore, A. J. (2006), Jökulhlaups circa
Sigurðsson, O. (2003), Jöklabreytingar 1930–1960 and Landnám: mid‐to late first millennium AD floods in south
2001–2002, Jökull, 53, 55–62. iceland and their implications for landscapes of settlement,
Sigurðsson, O., Snorrason, A, and Zophonıasson, S. (1992), Geogr. Ann. Ser. A, Phys. Geogr., 88(2), 165–176.
Jokulhlaupaannáll 1984–1988, Jökull, 42, 73–80. Smith, K. and Sigmundsson, F. (2010), Volcanogenic hazards
Sigurðsson, O., Jónsson, T., and Jóhannesson, T. (2007), and risks to road systems: Preliminary assessments at
Relation between glacier‐termini variations and summer Snæfellsjökull, Vegagerðin, Reykjavík.
temperature in Iceland since 1930, Ann. Glaciol., 46, Smithsonian Institute (2019), National Museum of Natural
170–176. History, Global Volcanism Program, Askja, https://
Sigurðsson, O., Williams, R.S. Jr., and Víkingsson, S. (2013), volcano.si.edu/ [Accessed 4 September 2019.]
Maps of glaciers in Iceland, Icelandic Meteorlogical Office, Sonnek, K. M., Mårtensson, T., Veibäck, E., Tunved, P.,
Reykjavík. Grahn, H., von Schoenberg, P., et al. (2017), The impacts
Sigurgeirsson, M., Gautason, B., Gudmundsson, Á., of a Laki‐like eruption on the present Swedish society,
Hjartarson, H., Blischke, A., Mortensen, AK,et al. (2010), Natural Hazards, 88(3), 1565–1590.
Exploration drilling in the Theistareykir high‐temperature Soosalu, H., Einarsson, P., and Jakobsdóttir, S. (2003),
field, NE‐Iceland: Stratigraphy, alteration and its relation­ Volcanic tremor related to the 1991 eruption of the Hekla
ship to temperature structure, World Geothermal Congress. volcano, Iceland, Bull. Volcanol., 65(8), 562–577.
194 REFERENCES

Soosalu, H., Jónsdóttir, K., and Einarsson, P. (2006), Stokes, C. R. and Clark, C. D. (2001), Paleo‐ice streams,
Seismicity crisis at the Katla volcano, Iceland—signs of a Quat. Sci. Rev., 20, 1437–1457.
cryptodome? J. Volcanol. Geotherm. Res., 153(3–4), Stothers, R. B. (1998), Far reach of the tenth century Eldgjá
177–186. eruption, Iceland, Climate Change, 39(4), 715–726.
Sparks, R. S. J., Wilson, L., and Sigurðsson, H. (1981), The Sturkell, E. and Sigmundsson, F. (2000), Continuous defla­
pyroclastic deposits of the 1875 eruption of Askja, Iceland, tion of the Askja caldera, Iceland, during the 1983–1998
Philos. Trans. R. Soc. Lond. A, 299(1447), 241–273. noneruptive period, J. Geophys. Res. Solid Earth, 105(B11),
Spry, A. (1962), The origin of columnar jointing, particu­ 25671–25684.
larly in basalt flows, J. Geol. Soc. Austral, 8(2), 191–216. Sturkell, E., Sigmundsson, F., and Einarsson, P. (2003),
Staines, K. E., Carrivick, J. L., Tweed, F. S., Evans, A. J., Recent unrest and magma movements at Eyjafjallajökull
Russell, A. J., Jóhannesson, T., and Roberts, M. (2015), A and Katla volcanoes, Iceland, J. Geophys. Res. Solid Earth,
multi‐dimensional analysis of pro‐glacial landscape 108(B8), 1–13.
change at Sólheimajökull, southern Iceland, Earth Surf. Sturkell, E., Sigmundsson, F., Geirsson, H., Ólafsson, H.,
Process. Landf., 40(6), 809–822. and Theodórsson, T. (2008), Multiple volcano deforma­
Stefánsson, A. 2010. The Vatnshellir Project—a first for tion sources in a post‐rifting period: 1989–2005 behaviour
Iceland, Proceedings 14th International Symposium on of Krafla, Iceland constrained by levelling, tilt and GPS
Vulcanospeleology, 115–121. observations, J. Volcanol. Geotherm. Res., 177(2),
Stefánsson, R. and Halldórsson, P. (1988), Strain release 405–417.
and strain build‐up in the south Iceland seismic zone, Sturkell, E., Einarsson, P., Sigmundsson, F., Hooper, A.,
Tectonophysics, 152(3–4), 267–276. Ófeigsson, B. G., Geirsson, H. and Ólafsson, H. (2010),
Stefánsson, R., Gudm ̄ undsson, G. B., and Roberts, M. J. Katla and Eyjafjallajökull volcanoes, in Developments in
(2006), Long‐term and short‐term earthquake warnings Quaternary Sciences, Vol. 13, pp. 5–21, Elsevier.
based on seismic information in the SISZ. Verdurstofa Sverrisdottir, G. (2007), Hybrid magma generation pre­
Islands (Vl ‐ES‐04), Rejkavik. ceding Plinian silicic eruptions at Hekla, Iceland: evidence
Stefánsson, R., Gudmundsson, G. B., and Halldorsson, P. from mineralogy and chemistry of two zoned deposits,
(2008), Tjörnes fracture zone. New and old seismic evi­ Geol. Mag., 144(4), 643–659.
dences for the link between the North Iceland rift zone Swindles, G. T., Watson, E. J., Savov, I. P., Lawson, I. T.,
and the Mid‐Atlantic ridge, Tectonophysics, 447(1), Schmidt, A., Hooper, A., et al. (2018). Climatic control on
117–126. Icelandic volcanic activity during the mid‐Holocene,
Stefansson, U. (1981), The sea around Iceland, in Natural Geology, 46, 47–50.
Islands, pp. 397–438, Almenna Bokaf6lagio, Reykjavik. Syvitski, J. P. M., Burrell, D. C., and Skei, J. M. (1987),
Stefánsson, V. (1980), Investigations on the Krafla high Fjords: processes and products, Springer‐Verlag, New
­temperature geothermal field, Natturufraedinjurinn, 50, York, 379 pp.
333–359. Syvitski, J. P., Burrell, D. C., and Skei, J. M. (2012), Fjords:
Stefánsson, V. (1981), The Krafla geothermal field, northeast processes and products. Springer Science and Business
Iceland, in L. Reybach, and L. J. P. Muffer (eds), Media.
Geothermal systems: Principles and case histories, pp. 273– Thelen, W. A., Malone, S. D., and West, M. E. (2010) Repose
294, John Wiley & Sons, Ltd., New York. time and cumulative moment magnitude: A new tool for
Steffensen, J.P., Andersen, K.K., Bigler, M., Clausen, H.B., forecasting eruptions? Geophys. Res. Lett., 37(L18301), 1–5.
Dahl‐Jensen, D.,Fischer, H., et al. (2008) High‐resolution Thórarinsson, S. (1950), Glacier outbursts in the river
Greenland ice core data show abrupt climate change hap­ Jökulsa a Fjöllum, Náttúrufræðingurinn, 20, 113–133.
pens in few years, Science, 321(5889), 680–684. Thórarinsson, S. (1953), The crater groups in Iceland, Bull.
Steindorsson, S. (1963), Ice age refugia in Iceland as indi­ Volcanol., 14(1), 3–44.
cated by the present distribution of plant species, in A. Thórarinsson, S. (1954), The eruption of Hekla 1947–1948,
Löve and D. Löve, (eds), North Atlantic biota and their The approach and beginning of the Hekla eruption,
history, pp. 355–365, Pergamon. Eyewitness accounts, pp. 1–23, Vísindafélag íslendinga
Steinthorsson, S. (1967), Two new 14C ages on the ashes from and the Museum of Natural History, Reykjavík.
Snæfellsjökull, Náttúrufrædingurinn, 37, 236–238. Thórarinsson, S. (1959), Some geological problems involved
Stevenson, J. A., Smellie, J. L., McGarvie, D. W., Gilbert, J. in the hydroelectric development of the Jokulsa a Fjöllum,
S., and Cameron, B. I. (2010), Subglacial intermediate vol­ State Electricity Authority, Reykjavík.
canism at Kerlingarfjöll, Iceland: magma–water interac­ Thórarinsson, S. (1963), Eldur í Öskju: Askja on fire,
tions beneath thick ice, J. Volcanol. Geotherm. Res., 185(4), Almenna Bókafélagið, Reykjavík.
337–351. Thorarinsson, S. (1964), Whirlwinds produced by Surtsey
Stevenson, J. A., et al. (2006), Subglacial and ice‐contact volcano, Bull. Am. Meteorol. Soc., 45(8), 440–444.
volcanism at the Öræfajökull stratovolcano, Iceland, Bull. Thórarinsson, S. (1966), The median zone of Iceland, in The
Volcanol., 68(7–8), 737–752. World Rift System, Vol. 66, p. 187.
REFERENCES 195

Thórarinsson, S. (1967), Hekla and Katla. The share of acid Thordarson, T. and Larsen, G. (2007), Volcanism in Iceland
and intermediate lava and tephra in the volcanic products in historical time: Volcano types, eruption styles and erup­
through the geological history of Iceland, Soc. Sci. tive history, J. Geodyn., 43(1), 118–152.
Islandica, 38, 190–197. Thordarson, T. and Self, S. (2003), Atmospheric and
Thórarinsson, S. (1968), The last phases of the Surtsey ­environmental effects of the 1783–1784 Laki eruption:
eruption, Natturufraedingurinn, 38, 113–135. A review and reassessment, J. Geophys. Res. Atmos.,
Thorarinsson, S. (1969), Glacier surges in Iceland, with spe­ 108(D1).
cial reference to the surges of Brúarjökull, Can. J. Earth Thordarson, T., Self, S., Oskarsson, N., and Hulsebosch, T.
Sci., 6(4), 875–888. (1996), Sulfur, chlorine, and fluorine degassing and atmo­
Thórarinsson, S. (1970), Tephrochronology and medieval spheric loading by the 1783–1784 AD Laki (Skaftár Fires)
Iceland, Scientific methods in Medieval archaeology, pp. eruption in Iceland, Bull. Volcanol., 58(2–3), 205–225.
295–328, University of California Press, Berkeley. Thordarson, T., Miller, D. J., Larsen, G., Self, S., and Sigurðsson,
Thórarinsson, S. (1971), The age of the light Hekla tephra H. (2001), New estimates of sulfur degassing and atmospheric
layers according to corrected C14‐datings, mass‐loading by the 934 AD Eldgjá eruption, Iceland,
Náttúrufræðingurinn, 41, 99–105. J. Volcanol. Geotherm. Res., 108(1–4), 33–54.
Thórarinsson, S. (1974), Vötnin stríð. Saga Skeidarárhlaupa Thoroddsen, Þ. (1892), Islands Jøkler i Fortid og Nutid,
og Grímsvatnagosa (The swift flowing rivers: the history of Geogr. Tidsskr., 11, 111–146.
Grímsvötn jökulhlaups and eruptions). Menningarsjóður, Thoroddsen, Þ. (1925), Die Geschichte der isländischen
Reykjavík. [In Icelandic.] Vulkane (nach einem hinterlassenen Manuskript), AF
Thórarinsson, S. (1979), The postglacial history of Myvtan Høst and Søn.
Area, Oikos, 32(2), 16–28. Thorolfsson, G. (2005), Maintenance history of a geothermal
Thórarinsson, S. (1981a), Greetings from Iceland: Ash‐falls plant: Svartsengi Iceland, in Proceedings of the World
and volcanic aerosols in Scandinavia, Geogr. Ann. Ser. A, Geothermal Congress, Antalya, Turkey.
Phys. Geogr., 63(3–4), 109–118. Thors, K. (1992), Bedrock, sediments, and faults in
Thórarinsson, S. (1981b), Jardeldasvædi á nútıma (Volcanic Thingvallavatn. Oikos, 64, 69–79.
areas of the Holocene), Náttúra Íslands, 2nd edn, pp, 81– Thorsteinsson, T. and Björnsson, H. (2012), Climate change
119, Almenna bókafélagid, Reykjavık. and energy systems: impacts, risks and adaptation in the
Thórarinsson, S. (1981c), The application of tephrochronol­ Nordic and Baltic countries, Nordic Council of Ministers,
ogy in Iceland, in Tephra studies, pp. 109–134, Springer, Copenhagen, 228 pp.
Dordrecht. Thy, P., Beard, J. S., and Lofgren, G. E. (1990), Experimental
Thórarinsson, S. and Grayson, D. K. (1979), On the damage constraints on the origin of Icelandic rhyolites, J. Geol.,
caused by volcanic eruptions with special reference to tephra 98(3), 417–421.
and gases, pp. 125–159, Academic Press. Tobler, D. J., Stefansson, A., and Benning, L. G. (2008), In‐
Thórarinsson, S. and Sæmundsson, K. (1979), Volcanic situ grown silica sinters in Icelandic geothermal areas,
activity in historical time, Jökull, 29, 29–32. Geobiology, 6(5), 481–502.
Thórarinsson, S. and Sigvaldason, G. E. (1972), The Hekla Tómasson, H. (1996), The jökulhlaup from Katla in 1918,
eruption of 1970, Bull. Volcanol., 36(2), 269–288. Ann. Glaciol., 22, 249–254.
Thórarinsson, S. and Sigvaldason, G. E. (1962), The Tómasson, H. A. (2002), Catastrophic floods in Iceland, in
eruption in Askja, 1961; a preliminary report, Am. J. Sci., Extremes of Extremes, IAHS Publication 271, pp. 121–
260(9), 641–651. 128, Reykjavık.
Thórarinsson, S., Einarsson, T., and Kjartansson, G. (1959), Tomlinson, E. L., Thordarson, T., Lane, C. S., Smith, V. C.,
On the geology and geomorphology of Iceland, Geogr. Manning, C. J., Müller, W., and Menzies, M. A. (2012),
Ann., 41(2–3), 135–169. Petrogenesis of the Sólheimar Ignimbrite (Katla, Iceland):
Thórarinsson, S., Einarsson, T., Sigvaldason, G. E., And implications for tephrostratigraphy, Geochim. Cosmochim.
Elisson, G. 1964. The submarine eruption off the Westman Acta, 86, 318–337.
Islands 1963‐64, Bull. Volcanol., 27, 435–445. Toomey, D. R. (2012), Plate tectonics: Piecing together rifts,
Thordarson, T. (1995), Volatile release and atmospheric Nature Geosci., 5(4), 235.
effects of basaltic fissure eruptions, Doctoral dissertation, Torfason, H. (1985), The Great Geysir, Geysir Conservation
University of Hawaii at Manoa. Committee, Reykjavík, 23 pp.
Thordarson, T. (2012), Outline of the geology of Iceland, in Trippanera, D., Ruch, J., Acocella, V., Thordarson, T., and
Champman Conference. Urbani, S. (2018), Interaction between central volcanoes
Thordarson, T. and Höskuldsson, Á. (2008), Postglacial vol­ and regional tectonics along divergent plate boundaries:
canism in Iceland, Jökull, 58(198), e228. Askja, Iceland, Bull. Volcanol., 80(1), 1.
Thordarson, T. and Höskuldsson, Á. (2014), Iceland, classic Trønnes, R. G. (2002), Geology and geodynamics of Iceland,
geology in Europe, 2nd edn, Dunedin Academic Press, pp. 1–19, Nordic Volcanological Institute, University of
Edinburgh, 256 pp. Iceland.
196 REFERENCES

Tryggvason, A., Benz, H. M., and Rognvaldsson, S. T. Walker, G. P. L. and Blake, D. H. (1966), The formation of a
(1998), Seismic travel time tomography studies of two vol­ palagonite breccia mass beneath a valley glacier in Iceland,
canoes, the Long Valley Caldera, California, and the Q. J. Geol. Soc., 122(1–4), 45–58.
Hengill Volcano, Iceland, Ann. Geophys., 16(1), 174. Wappler, T., Grímsson, F., Wang, B., Nel, A., Ólafsson, E.,
Tryggvason, E. (1964), Arrival times of P waves and upper Kotov, A. A., et al. (2014), Before the “Big Chill”: A
mantle structure, Bull. Seismol. Soc. Am., 54(2), 727–736. preliminary overview of arthropods from the middle
­
Tryggvason, T. (1965), Petrographic studies on the eruption Miocene of Iceland (Insecta, Crustacea), Palaeogeogr.
products of Hekla 1947–1948. The eruption of Hekla, Palaeoclimatol. Palaeoecol., 401, 1–12.
Soc. Sci. Islandica, 4(6), 82–133. Wastegård, S., Andersson, S., and Perkins, V. H. (2009),
Tryggvason E (1968), Result of precision leveling in Surtsey, A new mid‐Holocene tephra in central Sweden, GFF
Surtsey Res., 4, 149–158. (J. Geol. Soc. Sweden), 131(4), 293–297.
Tryggvason E (1972) Precision leveling in Surtsey. Surtsey Weber, G. and Castro, J. (2017), Experimental constraints on
Res., 6, 158–162 silicic magma storage at Hekla volcano (Iceland) and
Tryggvason, E. (1980), Observed ground deformation during potential implications for pre‐eruptive deformation, in
the Krafla eruption of March 16, 1980, Nordic EGU General Assembly Conference Abstracts, Vol. 19,
Volcanological Institute, University of Iceland. p. 5264.
Tryggvason, E. (1984), Widening of the Krafla fissure swarm Wells, G. H. (2016), Timeline reconstruction of Holocene
during the 1975–1981 volcano‐tectonic episode, Bull. jökulhlaups along the Jökulsá á Fjöllum channel, Iceland,
Volcanol., 47(1), 47–69. Doctoral dissertation, University of Texas at Austin
Tryggvason, E. (1986), Multiple magma reservoirs in a rift Weir, N. R., White, R. S., Brandsdóttir, B., Einarsson, P.,
zone volcano: ground deformation and magma transport Shimamura, H., and Shiobara, H. (2001), Crustal struc­
during the September 1984 eruption of Krafla, Iceland, ture of the northern Reykjanes Ridge and Reykjanes
J. Volcanol. Geotherm. Res., 28(1–2), 1–44. Peninsula, southwest Iceland, J. Geophys. Res. Solid Earth,
Tryggvason, E. (1989a), Ground deformation in Askja, 106(B4), 6347–6368.
Iceland: its source and possible relation to flow of the Weisenberger, D. T. (2010), Iceland, Sótt 26. 08 2011 frá DR.
mantle plume, J. Volcanol. Geotherm. Res., 39(1), 61–71. Tobias
Tryggvason, E. (1989b), Measurement of ground deformation Wastegard, S., Wohlfarth, B., Subetto, D.A., and Sapelko, T.
in Askja 1966 to 1989, Vol. 8904, Nordic Volcanological (2000), Extending the known distribution of the Younger
Institute, University of Iceland. Dryas Vedde Ash into northwestern Russia, J. Quat. Sci.,
Tryggvason, E. (1994), Surface deformation at the Krafla 15(6), 581–586.
volcano, North Iceland, 1982–1992, Bull. Volcanol., 56(2), White, J. D. and Houghton, B. (2000), Surtseyan and related
98–107. phreatomagmatic eruptions, in H. Sigurðsson (ed.),
Tuffen, H. and Castro, J. M. (2009), The emplacement of an Encyclopedia of volcanoes, pp. 495–511, Academic Press,
obsidian dyke through thin ice: Hrafntinnuhryggur, New York.
Krafla Iceland, J. Volcanol. Geotherm. Res., 185(4), White, R. and McKenzie, D. (1989), Magmatism at rift
352–366. zones: the generation of volcanic continental margins and
Tweed, F. S., Roberts, M. J., and Russell, A. J. (2005), flood basalts, J. Geophys. Res. Solid Earth, 94(B6),
Hydrologic monitoring of supercooled meltwater from 7685–7729.
Icelandic glaciers, Quat. Sci. Rev., 24(22), 2308–2318. Winpenny, B. and Maclennan, J. (2014), Short length scale
Vésteinsson, O. (2004), Icelandic farm house excavations. oxygen isotope heterogeneity in the Icelandic mantle:
Field methods and site choices, Archaeol. Islandica, 3, Evidence from plagioclase compositional zones, J. Petrol.,
71–100. 55(12), 2537–2566.
Waitt, R. B. (2002), Great Holocene floods along Jokulsa a Wittmann, M., Dorothea, C., Zwaaftink, G., Schmidt, L.S.,
Fjöllum, north Iceland, in P. I. Martini, V. R. Baker, and Guðmundsson, S., Pálsson, F., et al. (2017), Impact of
G. Garzon (eds), Flood and megaflood processes and dust deposition on the albedo of Vatnajökull ice cap,
deposits: recent and ancient examples, Special Publication Iceland, The Cyrosphere, 11, 741–754.
32, pp. 37–51, International Association of Wolfe, C. J., Bjarnason, I. T., VanDecar, J. C., and Solomon,
Sedimentologists, Blackwell Science, Oxford S. C. (1997), Seismic structure of the Iceland mantle
Walker, G. P. L. (1960), Zeolite zones and dyke distribution plume, Nature, 385(6613), 245.
in relation to the structure of the basalts in Eastern Woul, M., Hock, R., Braun, M., Thorsteinsson, T.,
Iceland, J. Geology, 68, 515–528. Johannesson, T., and Halldorsdottir, S. (2006), Firn layer
Walker, G. P. L. (1999) Part I. Basaltic volcanoes and impact on glacial runoff: a case study at Hofsjökull,
volcanic systems, in H. Sigurdsson, B. Houghton, H. Iceland, Hydrol. Proc., 20(10), 2171–2185.
Reymer, J. Stix, and S. McNutt, Encyclopedia of volcanoes, Wright, T. J., Sigmundsson, F., Pagli, C., Belachew, M.,
p. 1417, Academic Press. Hamling, I.J., Brandsdóttir, B., et al. (2012), Geophysical
REFERENCES 197

constraints on the dynamics of spreading centres from Zierenberg, R. A., Schiffman, P., Barfod, G. H., Lesher, C.
rifting episodes on land, Nature Geosci., 5, 242–250. E., Marks, N. E., Lowenstern, J. B., and Friðleifsson, G.
Ye, B., Ding, Y., Liu, F., and Liu, C. (2003), Responses of Ó. (2013), Composition and origin of rhyolite melt inter­
various‐sized alpine glaciers and runoff to climate change, sected by drilling in the Krafla geothermal field, Iceland,
Journal of Glaciology, 49(164), 1–7. Contrib. Mineral. Petrol., 165(2), 327–347.
Zakharova, O. K. and Spichak, V. V. (2012), Geothermal fields
of Hengill Volcano, Iceland, J. Volcanol. Seismol., 6(1), 1–14.
INDEX

Note: Page numbers followed by f indicate figures; t indicate tables; b indicate boxes; p indicate photos.
Aa lava flow, 50–52, 51p, 93, 113 Bølling‐Allerød Interstadial, 118, 121–22
Ablation zones, 126, 132–34, 165 Bookshelf faulting, 15, 17f, 27–28
Active fissures, 33f, 65, 66f, 77 Boreal forest vegetation, 11b, 117, 118p
Adaldalur costal plain, 34, 100 Borgarfjorður Anticline, 40–41, 41f
Aegir Ridge, 3 Braided river channels, 146, 147p, 151p
Airline travel, 67–68, 70f, 78 Breccia, 20, 55, 64
Air quality, 48b–49b, 70f Breiðamerkurjökull outlet glacier
Álftanes Stage, 121 features of, 127, 132, 134, 141t
Almannagjá tensional fracture, 21–22, 22p south region and, 146, 148, 150–52, 156
Alpha Ridge, 3 Breiðárlón basin, 151–52, 151p
AlÞing, 22 Brennisteinsfjöll rift zone, 15, 16f, 62
AMOC. See Atlantic Meridional Overturning Circulation British Isles, 8, 28, 65
Andesite, 8–9, 28t Brúarjökull outlet glacier, 134, 141t, 146, 148,
rhyolite lava, 10, 30 149t, 165
volcanics and, 36, 45, 80–81, 106 Bryophytes, 85
Anticline, 40–41, 41f Búði Stage, 121
Ar‐Ar dating, 62, 105, 137
Arctic Circle, 3, 160–61, 163 Calcite crystals, 10
Arctic Mendelev Ridge, 3 Carbon emissions, 48b, 81. See also Gases
Arȇtes, 162 Chronostratigraphy, 111f, 123
Arnarson, Ingólfur, 122 The Churn. See Strokkur
Ash dispersal, 65, 68f, 73–74, 73f, 92 Cinder cone, 50f
Askja volcanic system, 31, 33f Cirque, 125, 162, 168
eruption of 1874–1876, 91–93, 92f–93f, 92t Climate change
eruption of 1961, 93–94 isotope indicators, 118, 119b, 121b
VEI of, 88, 90t, 91 modern climate and, 123–25
volcanics and, 88–94, 89f, 90t, 91p, 92f–93f, 92t north region and, 163, 167, 169–71
Atlantic Meridional Overturning Circulation (AMOC), Reykjanes Peninsula and, 133, 140
159 south region and, 146, 154
Avalanche, 162, 165, 166f, 167b Columnar basalts, 55, 57f
Aviation zones, 67–68, 70f, 141, 157 Corifting, 29–30
Axarfjördur, 34–36 Crater rows
Lakagígar crater, 148
Bakkaflói beach, 89 Lúdentarborgir, 102
Bárðarbunga, 29 of north region, 34, 36–37, 99
glacial features, 145–48, 152, 155 Rauðhólar fissure and, 36–37
volcanics and, 49, 66f, 68–71 Reykjanes Peninsula and, 20, 62
Basaltic magma, 56, 90, 94, 112 tectonics and, 20, 52, 56, 78p
Bathymetry, 139 Crevasses, 130, 131f, 133–34, 146, 157
Bersekjahraun crater, 112 Cyclothems, 117
Birch trees, 118p, 122
Birds, 85 Dalseldar Fires, 102
Bjarnarflag field, 99, 104 Dalvík Zone (DZ), 32–34
Björnsson, Helgi, 124 Dark basalt (blágrýtismyndun), 8
Blágrýtismyndun (dark basalt), 8 Deflation, 90, 94, 96, 98f

Iceland: Tectonics,Volcanics, and Glacial Features, Geophysical Monograph 247, First Edition. Tamie J. Jovanelly.
© 2020 American Geophysical Union. Published 2020 by John Wiley & Sons, Inc.
199
200 Index

Deglaciation, 118, 133–34, 133t Hverfjall, 100–101


in north region, 159–60, 168 Krafla, 34–36, 36t
in south region, 145–46 of 1975–1984, 96–97, 97f–98f, 98p
Denmark Strait, 161f, 162 Laki, 77–78, 78p
Dense rock equivalent (DRE), 58–59 Mývatn of 1724–1729, 102–5, 103p–4p
in north region, 87–88, 92, 92t, 94 Reykjanes, 62
in south region, 70, 77 Firn layer, 143–44
Denudation rates, 127, 130t, 146 Fish, 82, 103, 123
Dettifoss waterfall, 36–37, 37f, 37p Fissure swarms
Diatreme pipes, 83–84 Grímsvötn, 145–46, 149f
Dike intrusions, 92, 94, 96, 100, 100p, 106 Hengill‐Langjökull, 64
Dimmuborgir pillars, 102 Krafla, 94–102, 95p, 95t, 97f–98f, 98p
Drangajökull ice cap, 162–66, 164f, 166f Laki, 70, 145
DRE. See Dense rock equivalent of NVZ, 31–37, 33f, 87–91, 100–104, 100p
Dryhólaey lava cave, 78–79, 79p of SISZ, 66–68, 71
Dyngjufjöll massif complex, 88–91. See also Askja volcanic tectonics of, 15, 18p, 27, 45–47
system Fjallsarlón lagoon, 151–52, 151p
Dyngjujökull outlet glacier, 134, 141t, 148, 149t Flateyjarskagi peninsula, 32–34, 33f
DZ. See Dalvík Zone Fljót, 81
Flooding. See Jökulhlaup flooding
Early Pleistocene, 8f, 9t, 11, 16, 82, 107, 118p Fluorine poisoning, 76, 81
Earthquake, 96–97, 103–4, 112, 145 Fossils, 10b–11b, 117–18, 121b
swarms, 15, 33–36, 75–76, 93, 104 Fossvogur sedimentary, 122
East Greenland Current (EGC), 123 Fremrinámar central volcano, 31, 32f, 87
East Volcanic Zone (EVZ), 7, 8f, 27–28, 31, 32f Fremrinámur fissure belt, 33f, 100
Vatnajökull glacier and, 65, 68 Froda‐Breiðavík intrusion, 107, 108f
EGC. See East Greenland Current Fumaroles, 17, 69, 92, 98
Egilsstaðir, 29p, 30
Eiríksjökull, 61 Gabbro intrusion, 63, 110
ELA. See Equilibrium line altitude Gases, 24, 81, 99, 125
Eldgjá Fires, 76–78, 78p isotope fractionation and, 121b
Eldhraun lava field, 104 volcanoes and, 48b–49b
Eldvörp‐Svartsengi rift zone, 15, 16f, 62 Gateway to Hell. See Hekla
Ellesmere Island, 3 Gauss Chron, 117
Enriched mid‐ocean ridge basalts (EMORB), 45 Geitahlid tabletop ridge, 16
Equilibrium line altitude (ELA), 132f, 164, 168–69 Geitlandsjökull tuya, 138
Esja mountain complex, 63 Geothermal energy, 17, 20, 33, 97–100
Eurasia Plate, 3, 20–21, 21p, 28, 65 for Reykjavík, 23, 23p, 24b, 64
Europe, 3, 10b, 21–22, 48b, 73 GIA. See Glacioisostatic adjustment
air travel and, 68, 70f, 157 Gigjökull outlet glacier, 123
Hekla and, 79–80 Gjástykki graben, 96
largest glacier of, 148 Glacial fjord, 128p, 161, 162f
LIA and, 122 Glacial till, 119b, 127, 130p
livestock mortality and, 77 Glaciers. See also Tectonics
Evacuations, 49, 69–70, 104, 112, 158 associated landforms and, 125–32, 126f, 127p–31p, 130t,
EVZ. See East Volcanic Zone 131f
Eyjafjallajökull volcano, 71–77, 76p denudation rates, 127, 130t, 146
eruption of, 48b, 49, 54, 67, 70, 70f ice formation and, 124–25, 124f–25f
stratovolcano, 157–58, 158p mass balance and, 132–34, 132f–34f, 133t
Eyjafjöll sandur plain, 65, 157 modern climate and, 123–24
setting of, 117–23, 118p, 120f, 122f
Fagradalsfjall rift zone, 15, 16f, 62 striations of, 109, 126, 129p, 137, 167
Faroe Islands, 8, 28, 74 volcanics and, 135
Fimmvörðuháls ridge, 157 Glaciofluvial deposits, 8, 157, 165
Fires Glacioisostatic adjustment (GIA), 118, 122, 133, 135, 146
Dalseldar, 102 Global Positioning System (GPS) data, 27–28, 137
Eldgjá, 76–78, 78p Global warming, 112, 123–24, 140, 144
Index 201

Goðabunga rise, 75 Hrafntinnuhryggur ridge, 94–95


Goðafoss waterfall, 99 Hreðavatn horizon, 41
Golden Circle Road, 127 Hreppar Formation, 82
Golden Waterfall. See Gulfoss Hreppar microplate, 3, 27–28
GOR. See Grímsey Oblique Rift Hrossadalur craters, 104
GPS. See Global Positioning System Hrútagjá shield volcano, 62
Grábók Lava Field, 113 Húsavík‐Flatey Fault (HHF), 33–34
Grænvatn, 17, 19p Hveragerði town, 24b, 28
Great Geysir. See Stóri Geysir Hverfjall Fires, 100–101
Greenland, 3, 8, 10b, 28, 73, 162 Hverfjall tuff ring, 101, 101p
Grensdalur central volcano, 22–23 Hvíhólar geothermal field, 99
Grey Basalt Formation, 8 Hvítá River, 25, 144
Grímsey Oblique Rift (GOR), 33–34 Hyalobasalt, 82
Grímsvötn volcanic system, 65–71, 66f, 71p, 111 Hyaloclastite, 89, 94, 99, 109–10, 135
eruption of, 27, 54–55, 58 deposits, 16, 78
fissure swarm, 145–46, 149f ridges, 15, 31, 36, 67, 87, 96
Gulfoss (Golden Waterfall), 23–26, 25p–26p Hyannadalshnúkur, 132
Gullborgarhraun crater, 112 Hydrogen sulfide, 20, 77–78, 82, 156

Hafragilsfoss waterfall, 36 Ice‐cored moraine, 170


Hagafell lava dome, 137 Ice‐core drilling, 119b, 121b
Hágöng dome, 94 Ice formation, 124–25, 124f–25f
Hamarshólar crater, 102 Iceland. See also North regions; South regions; Tectonics
Hásteinar mountain, 141 background geology of, 7–13, 9t, 10b–11b, 12f, 13p
Hawaiian‐type eruption, 46t, 58, 93 climate of, 8, 9t, 123–24
Hawaiite lava flow, 111 glacial features of
Hazard mitigation, 58, 111–12 associated landforms and, 125–32, 126f, 127p–31p,
Haze Famine, 77–78, 78p 130t, 131f
Heimaey Island, 82 glaciovolcanic deposits, 135
Heinrich Event, it, 119b ice formation and, 124–25, 124f–25f
Hekla (The Queen), 27, 47, 52, 53p, 58 mass balance and, 132–34, 132f–34f, 133t
ash dispersal from, 65, 68f periglacial environments, 132
eruption events of, 80–82, 80f setting of, 117–23, 118p, 120f, 122f
tephrochronology of, 69t, 79–81, 80f Meteorological Office of, 40
Heklugjá fissure system, 79 Parliament of, 99
Helgrindur volcano, 112–13 present setting of, 7, 8f, 117, 120f
3
He chronology, 54, 159 southwestern region of, 61–64, 63p
Hellisheiði Geothermal Power Station, 23, 23p, 24b, 64, 99 tectonic context of, 3, 4f
Hengill‐Langjökull rift zone, 15, 16f, 20, 24b, 62, 64 Icelandic Lúdent Period, 99–100
HHF. See Húsavík‐Flatey Fault Icelandic National Monument, 102, 103p
Hljóðaklettar fissure, 36–37 Icelandic Speleological Society, 113
Höfði, 102 Ice rafted debris (IRD), 117, 119b
Hofsjökull glacier, 121b, 167 Ignimbrites, 73–74, 106
ice cap, 140–44, 142f, 143p Inflation and deflation model, 96, 98f
Holocene Interbasaltic beds, 8–10, 11b
lava shields, 87 Interglacial lava flows, 82, 109, 112
north region and, 89, 95, 163–64, 168 IRD. See Ice rafted debris
Plio‐Pleistocene deposits, 40, 41f Irminger Oceanic Current, 122–23, 160, 162
Reykjanes Peninsula and, 137–39 Iron‐rich basalts, 67
south region and, 155, 157–58
tectonics and, 8, 9t, 11, 16, 27, 36, 39, 89 Jarðbaðshólar volcano, 34, 100–101
volcanic eruptions of, 61–62 Jarlhettur mountain range, 139
western region and, 107–9, 113 Jökulhlaup flooding, 111, 141, 161, 165
Holocene Thermal Maximum, 163–64 in south region, 145–46, 152–58, 153f, 155t
Horgsholtshraun crater, 112 tectonics and, 26, 36, 71p, 74f
Hrafnabjörg, 22 volcanics and, 69–70, 71p, 74–76, 74f
Hrafnagjá, 21 Jökulsá á Fjöllum river, 36–37
202 Index

Jökulsárgljúfur canyon, 36, 37f, 37p, 148, 155 Lapilli, 52, 83, 101–2
Jökulsárlón lagoon, 132, 148, 150p–51p, 151–52 Last Glacial Maximum (LGM), 121, 125
Jónsson, Björn, 23 of north region, 159–63, 168
Journey to the Centre of the Earth (Verne), 110 Reykjanes Peninsula and, 141, 144
of south region, 145
Katla Geopark, 49b Weichselian glacial stage, 12f, 117–18, 120f
Katla volcano, 47–49, 55, 58, 71–77, 72p, 73f–74f, 73t of western region, 159–63, 168
Kerlingarfjöll central volcano, 61–62 Lateral flow hypothesis, 87–88, 88f
Kerlingarfjöll mountains, 144 Laugarfjall mountain, 25
Ketildyngja volcano, 34, 90, 100 Lava dome, 50f, 106, 110, 137–38
Kirkjufell mountain, 162, 163p Laxárdalur Valley, 34, 100
Kistufell tabletop ridge, 16 Leggjabrjótur lava dome, 137
Kı ̄lauea, Hawaii, 102 Leirbotnar geothermal field, 98–99
Knickpoint, 36, 37p Leirhnjúkur craters, 34, 36, 96, 102f, 103–4
Kolbeinsey ridge (KR), 3, 8f, 31–33, 39 Leirufjördur moraine, 159, 160f
Kollafjörður volcanic system, 63 Leitahraun lava flow, 64
Kollóttadyngja volcano shield, 90 LGM. See Last Glacial Maximum
Kollur volcanic edifice, 88, 90 LIA. See Little Ice Age
Kötlujökull Pass, 74–75 Lichenometry, 123, 145
KR. See Kolbeinsey ridge LiDAR imagery, 102, 171
Krafla Lignite, 10b–11b, 106
fires, 34–36, 36t Lithology, 8, 10, 26, 28t, 37f
of 1975–1984, 96–97, 97f–98f, 98p log of, 92, 93f
fissure swarm, 94–99, 95p, 95t, 97f–98f, 98p Little Ice Age (LIA), 122–23, 126, 137–40
geothermal energy and, 98–99 maximum, 161, 164–66, 168–70
lava field, 96, 98p Livestock mortality, 76–77, 81
tectonics of, 33f, 34–36 Ljósufjöll volcanic system, 39, 108f–9f, 109, 112
Kröfluháls volcano, 100 Lock, William George, 88
Krýsuvík rift zone, 15, 16f, 62 Loddavötn area, 81
Krýsuvík Valley, 16–17, 18p Lofthellir lava cave, 52
Kverkfjökull volcano, 66f, 68 Lúdentarborgir crater row, 102
Kverkfjöll central volcano, 31, 32f Lúdent Cycle, 34
Kverkfjöll jökulhlaup, 155, 155t Lúdent tuff ring, 100–101
Lýsuskard volcanic system, 39, 108f, 112–13
Lagarfljót, 29p, 30
Lakagígar crater rows, 148 Maar volcano, 17, 19p, 50f
Lakes Mactra Zone, 10b–11b
Elliðavatn, 64 Maelifell dome, 110
Hvitárvatn, 138–40, 139p, 140f Mafic lava. See Aa
Jökulsárlón, 150p–51p, 151–52 MAR. See Mid‐Atlantic Ridge
Kleifarvatn, 16–17, 18p–19p Marine terrace, 168, 169p
Mývatn, 34, 35p, 97f, 100–104, 102f Markarfljót Valley, 158
Öskjuvatn, 93–94, 141 Mass balance, 132–34, 132f–34f, 133t
proglacial, 127, 139, 152, 157, 161, 164 in north region, 165–66, 168–71
Sandvatn, 34 Massif, 88–91, 125, 126f
Skoravatn, 163 Matuyama‐Brunhes transition, 107
subglacial, 69, 146, 152–54, 153f–54f, 155t Medieval Warm Period, 122
Þingvallavatn, 21–22, 140 Megafloods, 70, 155, 155t
Laki fissure swarm, 27, 70, 145 Meltwater, 89
fires, 77–78, 78p of north region, 159, 170
Lambatungajökull outlet glacier, 145 overview of, 126–27, 138–39
Landbrotshólar, 64 of Reykjanes Peninsula, 138–39
Landslides, 90, 167b, 171 of south region, 68–75, 146, 147p, 150–57
Landsvirkjun, 99 MIB. See Mid‐Iceland Belt
Langjökull, 29f Mid‐Atlantic Ridge (MAR), 3, 4f, 7, 20, 31, 45
ice cap, 137–40, 138f, 139p, 140f, 141t Mid‐Iceland Belt (MIB), 7, 8f, 27
Index 203

Miocene, 9t, 105, 117 Mývatn vent system, 99–105, 100p–101p, 102f,
fossils of, 10b–11b 103p–4p
Móberg Formation, 8–11, 62 Norðurlandsjöklar, 167–69, 167f, 169p
Möberg ridge, 12f, 13p, 34, 55, 55f, 90 Snæfellsnes Peninsula, 169–71, 170f–71f
Mohorovičić discontinuity, 39, 107 TFZ and, 87, 91, 95, 97
Molluscs, 10b–11b, 122 Tinná central volcano, 105–6, 105f
Mortality, 48b–49b, 76–77, 81 Vestfirðir Peninsula, 162–68, 163p, 164f, 166f, 167b
Móskarðshnúka, 63 waterfalls of, 36–37, 37f, 37p
Mounts North Volcanic Zone (NVZ)
Esja, 63 fissure swarms and, 87–91, 94–97, 95t, 100–104
Hafnafjall, 41 of northeast region, 31–32, 32f–33f
Hengill, 63p, 64 tectonics of, 7, 8f
Hofsfjall, 106 Vatnajökull and, 65, 67
Húsavíkurfjall, 33 Nunataks, 141, 160, 162, 163p
St Helens, 110 NVZ. See North Volcanic Zone
Skjaldbreiður, 22 Nýjahraun lava field, 91
Vindbelgur, 34
Múlajökull lobe, 141–43 Obsidian, 56, 94
Mýrdalsjökull, 71–72, 74 Ocean Drilling Project (ODP), 106
ice cap, 155–57, 156f Olafsvík, 111
Mýrdalssandur coast, 74–75, 155, 156f Older Laxá Lava (OLL), 100
Mýrdalur sandur plain, 74f Oligocene, 3, 9t, 11b
Mývatn Fires, 34–36, 95–96 Olivine basalt, 28t, 39, 45, 82
of 1724–1729, 102–5, 103p–4p OLL. See Older Laxá Lava
volcanic features and, 99–102, 100p–101p, 102f Öræfajökull
glacial lobe, 145–46, 148–50, 149t
Námafjall geothermal field, 33, 98–100 stratovolcano, 49, 50f
National Power Company (Landsvirkjun), 99 Öskjuvatn caldera, 90–94
Neogene volcanoes, 105–6 Outwash plain, 62, 123, 127, 127p, 146, 154–55
Neovolcanic zones, 7, 13, 65, 135. See also East Volcanic Oxygen isotopes, 118, 119b, 121b, 125
Zone; North Volcanic Zone; South Iceland Seismic
Zone; West Volcanic Zone Pahoehoe flows, 50–52, 51p, 52p, 56, 113
Nesjavellir geothermal field, 23, 24b, 64 Peralkalic rhyolites, 111
Nonsurge‐type glacier, 148, 169, 171f Permafrost, 123, 132, 143, 168
Norðurlandsjöklar, 167–69, 167f, 169p Phenocrysts, 82, 91, 93, 111, 159
North America, 3, 10b, 77 Phreatic eruptions, 55–56, 63–64, 82, 146
tectonics and, 20–21, 21p, 28, 45 Phreatomagmatic eruption
North Atlantic Ocean, 3–4, 4f, 20, 73, 82 overview of, 45, 54–55
currents of, 160–61, 161f of south region, 69, 74, 82
deep‐sea cores from, 117 of western region, 100–101, 105
IOC and, 122–23 Pillow lava, 12f, 20, 84f
oxygen isotopes and, 118, 119b, 121b glacial features and, 117, 135
Snæfellsnes Peninsula and, 169–70 ridges, 89
thermohaline circulation of, 154, 156 volcanics and, 55, 56p, 62, 64, 89, 110
North Atlantic Oscillation index, 124 Plagioclase, 82, 91, 96
North regions Plants, 11b, 85, 122–23
Askja volcano, 88–94, 89f, 90t, 91p, 92f–93f, 92t Pleistocene, 54, 55, 62, 72, 117–18
Drangajökull ice cap, 162–66, 164f, 166f deposits, 109–10
Dynáufjöll massif complex, 88–94, 89f, 90t, 91p, 92f–93f, Early, 8f, 9t, 11, 16, 82, 107, 118p
92t möberg ridge, 90
Krafla fires and, 34–36, 35p, 36t Reykjanes Peninsula and, 135, 137, 139, 143
Krafla fissure swarm, 94–99, 95p, 95t, 97f–98f, 98p rhyolitic, 110
lateral flow hypothesis and, 87–88, 88f Upper, 8, 9t, 32
LGM, 159–63, 168 volcanotectonic line, 109
mass balance and, 165–66, 168–71 Plinian eruption, 45, 73–74, 80–81, 106
MOR regions of, 87–88, 88f at Askja, 88, 91–92, 92t
204 Index

Pliocene, 63, 105, 117 Hvítá, 25, 144


fossils of, 10b–11b, 11, 13 Jökulsá, 157
Plio‐Pleistocene deposits, 9f, 11, 40, 41f, 79, 107 Jökulsá á Brú, 146
Plume‐origin hypothesis, 3–4 Jökulsá á Fjöllum, 100, 154–55
Polar Front, 118, 124, 160 Skeiðará, 69–70
Pompeii of Iceland. See Þjórsárdalur Valley Skjálfandafljót, 99
Porphyritic basalt, 8, 28t Þjórsá, 82, 144
Proglacial lake, 127, 139, 152, 157, 161, 164 Þórisdalur, 138
Proto‐Lake Mývatn, 101 Rootless cone (pseudocrater), 34, 35p, 54p, 56, 63–64,
Pseudocrater. See Rootless cone 77, 101
Pseudo‐islands, 83–84 RR. See Reykjanes Ridge
Pyroclastic debris, 83–84, 111 RVB. See Reykjanes Volcanic Belt
flow, 28t, 49, 50f, 59, 73, 92
Pyroxene, 45, 91 Saalian glaciation, 118, 120f
Sandfellshæð, 62
Quaternary period, 9t, 112, 119b Sandur plain, 127, 131p, 146, 147p, 170
glaciations, 123–24 Scandinavia, 68, 92. See also Viking settlements
ice sheet growth, 12f Scoria cone, 52, 62, 77, 101–5
The Queen. See Hekla Scotland, 68
Selfoss waterfall, 36–37
Radiocarbon dating, 88, 111, 118, 122–23, 160 Seljavellir farm, 77
Radio echo measurement, 66, 68, 71, 145, 151–52, 171 Sellandafjall, 34
Rauðhólar fissure, 36–37, 100, 100p Seltún drill site, 17
Rauðhólar rootless cone complex, 63–64 Shield volcanoes, 49–52, 50f, 51p, 62
Repose time, 58, 69t Sideromelane, 82, 84
Respiratory hazards, 48b–49b Siðujökull lobe, 134
Reydarfjördur, 28f, 29–30 Silica magma, 45, 65, 80–81, 111
Reykjahíð town, 34, 104p Silica tephra, 65, 77, 79–81, 80f
Reykjanes Fires, 62 SILK deposits, 73
Reykjanes Peninsula SISZ. See South Iceland Seismic Zone
Blue Lagoon, 17–20, 19p Skaðsá town, 23
Hofsjökull ice cap, 140–44, 142f, 143p Skagafjörður, 81
Kerlingarfjöll mountains, 144 fjord, 160f, 167
Krýsuvík Valley, 16–17, 18p–19p valley, 34, 105–6, 105f
Lake Hvitárvatn, 138–40, 139p, 140f Skati Dome, 105f, 106
Langjökull ice cap, 137–40, 138f, 139p, 140f, 141t Skeiðará River, 69–70
MAR in, 4f, 20 bridge of, 71p, 146, 147p
Stóri Geysir, 23–26, 25p–26p Skeiðarárjökull
tectonics of, 15–26, 16f–17f, 21p–23p outlet glacier, 69, 134f, 141t, 148
volcanism of, 61–64, 63p valley, 127, 149f, 149t
Þingvellir National Park, 9t, 20–23, 23p, 24b Skeiðarársandur plain, 127, 131p, 146, 148, 150–54
Reykjanes Ridge (RR), 3, 4f, 7, 8f, 15–16, 20 Skjaldbreiður shield volcano, 22, 49, 51p
Reykjanesviti, 20 Skjálfandi Bay, 34
Reykjanes Volcanic Belt (RVB), 8f, 15, 16f, 46, 61–62 Skógar tephra, 73
Reykjavík city, 24b, 39, 63–64, 169 Skridan dome, 110
Reykjavík Energy, 24b Skutustaðagigar craters, 102, 103p
Reynisfjara Beach, 79 Snæfellsjökull volcanic system, 39–40, 40p
Rhyolite lavas, 8–10, 28t, 30, 106 glacial features of, 109–12, 110p, 111f
magma and, 65, 80, 89, 96 Snæfellsnes Peninsula, 39, 40p
Rifting episodes, 31, 36, 39 active volcanos, 15, 16f
in north region, 87, 91, 95, 95t glacial features of, 159, 160f, 162, 163p, 169–71,
in southwest region, 62, 66–67 170f–71f
Ring Road, 70, 71p volcanics of, 107–9, 108f–9f
Rivers Snæfellsnes Syncline, 40–41, 41f
channel and, 146, 147p, 151p Snæfellsnes‐Vatnsnes Rift Zone, 40–41, 41f
Elliðaár, 64 Snæfellsnes Volcanic Belt (SVB), 7, 8f, 39, 41f, 46, 47f,
Geitá, 138 107–13
Index 205

Snagahraun lava field, 96 S‐wave shadowing, 96


Snow avalanche, 162, 165, 166f, 167b Syncline, 40–41, 41f, 107, 108f
Snow pack, 125f
Solfataras, 20, 93 Tabletop mountains, 13, 13p, 55, 117–18, 120f
Sólheimar ignimbrite, 73 Tapes Zone, 10b
South Iceland Seismic Zone (SISZ) Tectonics
fissure swarms and, 65–68, 71 subglacial eruptions and, 54–56, 55f, 57f
tectonics of, 7, 8f, 16–17, 20, 27–32 tephrochronology and, 46t, 56–58
volcanics of, 61, 65–67, 107, 112 VEI and, 45, 46t, 58–59
western region and, 107, 112 volcanic morphology and, 49–54, 50f, 51p–55p
South regions volcanic repose, 58
Bárðarbunga volcano, 66f, 68–71 volcanic setting and, 45–49, 46f–47f, 46t, 48b–49b
Breiðárlón basin, 151–52, 151p Tephra, 30, 80–84
Dryhólaey, 78–79, 79p from Askja, 92–93, 92t
Eyjafjallajökull stratovolcano, 157–58, 158p deposits, 65, 67f, 73–74, 80, 106, 119b
Eyjafjallajökull volcano, 71–77, 76p dispersal, 57, 59, 68, 76
fires of, 77–78, 78p from Dyngjufjöll, 89–90
Grímsvötn volcano, 65, 66f, 67–71, 71p silica and, 65, 73, 77, 79–81, 80f
Hekla, 79–81, 80f volume, 46t, 69t, 73t, 92, 92t
Jökulsárlón lake, 150p–51p, 151–52 Tephrochronology, 46t, 56–58
Katla volcano, 71–77, 72p, 73f–74f, 73t Terminal moraine, 117, 126, 129p
meltwater in, 68–75, 146, 147p, 150–57 of northern region, 164–65, 168
Mýrdasjökull ice cap, 155–57, 156f of Reykjanes Peninsula, 137, 144
subglacial lakes of, 152–55, 153f, 154f, 155t of south region, 145
Surtsey volcanic island, 82–85, 83f–85f Tertiary Basalt Formation, 8–11, 28, 29f, 40
tectonics of, 145–46, 152–58, 153f, 155t Tertiary period, 10–11
Vatnajökull National Park, 146–50, 149f, faults of, 20f–30f, 28–30, 28t, 29p
149t, 150p glaciations of, 123–24
Þjórsárdalur Valley, 82–83 ice sheet growth of, 12f
Spatter cone, 53p, 62, 77 TFZ. See Tjörnes Fracture Zone
Stardalur volcanic system, 63 Tholeiite, 8, 62, 67, 106
Steam vents, 20, 61 basalts, 28t, 45, 46f, 90, 108–9
Steinholtsjökull outlet glacier, 123 Tindars, 41, 55, 55f, 61, 110, 137
Stöng farmstead, 81–82 Tinná central volcano, 105–6, 105f
Stóra‐Laxá volcano, 82 Tinnárdalur Valley, 106
Stóri Geysir, 23–26, 25p–26p Tjörnes Formation, 10b
Strath terrace, 36, 37p Tjörnes Fracture Zone (TFZ)
Stratovolcanoes northeast region and, 31–35, 87, 91, 95, 97
Eyjafjallajökull, 157–58, 158p tectonics and, 7, 8f
Öræfajökull, 49, 50f Tjörnes peninsula, 10, 32–34, 33f, 117
Snæfellsjökull, 39–40, 40p, 110p Tombolo, 78–79, 79p
of south region, 75, 76p Torfajökull central volcano, 65, 79, 111–12
tectonics and, 49–50, 52 Trachybasalt, 111
tuya hybrid, 111 Trandilsson, Gaukur, 81–82
Strokkur (The Churn), 25, 25p Trölladyngja volcano shield, 90
Subglacial eruptions, 54–56, 56p, 57f Tröllaskagi Peninsula, 160f, 162, 167–69
Subglacial lakes, 69, 146, 152–54, 153f–54f, Tuff cone, 77–78, 82
155t Tuff ring, 50f, 100–101
Suðurhídar geothermal field, 99 Tuya, 55, 55f, 61–62, 110–11
Sulfur fog, 77–78
Surge‐type glaciers, 133, 134f, 156, 165, 166f UNESCO. See United Nations Educational, Scientific, and
Surtsey Research Progress Reports, 82 Cultural Organization
Surtsey volcanic island, 78, 82–85, 83f–85f United Kingdom, 48b, 68
Svalbard glaciers, 165 United Nations Educational, Scientific, and Cultural
Svartsengi geothermal plant, 17, 20 Organization (UNESCO), 82
SVB. See Snæfellsnes Volcanic Belt Upper Pleistocene, 8, 9t, 32
Sveinagja graben, 91 U‐shaped valley, 117, 126, 128p, 161
206 Index

Vascular plants, 11b, 85, 122–23 Gulfoss, 23–26, 25p–26p


Vatnajökull glacier, 45, 58, 87 Hafragilsfoss, 36
EVZ and, 65, 68 Water phase diagram, 125f
features of, 146–50, 149f, 149t, 150p Weichselian glacial stage, 9t, 11, 13, 22, 163, 164f
mass balance of, 132–34, 134f ice sheet, 117, 120f, 159, 163
subglacial volcanoes and, 65–68, 66f LGM and, 12f, 117, 120f
Vatnajökull ice cap, 3, 88 Western region, 39–41, 40p, 41f
Vatnshellir Cave, 113, 113f deglaciation in, 159–60, 168
Vedde Ash, 73–74 Drangajökull ice cap, 162–66, 164f, 166f
VEI. See Volcanic explosivity index Helgrindur volcano, 112–13
Veiðivötn fissue swarm, 145 Holocene, 107–9, 113
Verne, Jules, 110 LGM in, 159–63, 168
Vertebrate fossils, 11b Ljósufjöll volcano, 112
Vestfirðir Peninsula, 40, 159–68, 163p, 164f, 166f, 167b Norðurlandsjöklar, 167–69, 167f, 169p
Vestmannæyjar Islands, 65, 82, 156 Snæfellsjökull volcano, 109–12, 110p, 111f
Viking settlements, 47–49, 48b–49b, 121b, 122 Snæfellsnes Peninsula, 107–9, 108f–9f, 169–71,
of Reykjanes Peninsula, 143, 143p 170f–71f
in south region, 74, 81–82, 145, 155 Vatnshellir Cave, 113, 113f
Viti crater, 34, 91–92, 103 Westman Islands, 82, 156
Vogar lava field, 34, 100 West Volcanic Zone (WVZ), 7, 8f, 27, 31
Volcanic explosivity index (VEI) Reykjanes Peninsula and, 15, 16f, 20, 61–64, 63p
of Askja, 88, 90t World Heritage Site, 82
of Eyjafjallajökull, 76, 146 WVZ. See West Volcanic Zone
of Hekla, 69t, 79–80
of Katla, 73t Xenoliths, 102
Mývatn fires and, 103
overview of, 45, 46t, 58–59 YLL. See Younger Laxá Lava
Volcanoes. See also Stratovolcanoes Younger Dryas Stadial, 9t, 26, 73
background on, 7, 8f glacial features and, 121–22, 122f, 137, 165, 168
maar, 17, 19p, 50f Younger Laxá Lava (YLL), 101–2
morphology and, 49–54, 50f, 51p–55p Younger Stampar cone row, 62
neogene, 105–6 Yti Bugur, 111
repose, 58 Ytriflói Basin, 103
setting and, 45–49, 46f–47f, 46t, 48b–49b
shield, 49–52, 50f, 51p, 62 Zeolite zones, 10, 16, 106
tectonic line and, 109 Þeistareykir central volcano, 31, 32f
Vulcanian‐type eruption, 46t, 58 Þingvellir National Park, 64, 107, 108f
tectonics of, 20–23, 21p–23p, 24b
Water contamination, 81, 103 Þjórsárdalur Valley, 82–83
Waterfalls Þorarinsson, Sigurður, 81–82
Dettifoss, 36–37, 37f, 37p Þráinskjöldur shield volcano, 62
Goðafoss, 99 Þrengslaborgir‐Lúdentsborgir cone craters, 101

You might also like