Download as pdf or txt
Download as pdf or txt
You are on page 1of 59

Caltech Enrichment Trip Iceland

Iceland Field Guide

August 16-29, 2014


Caltech Enrichment Trip Iceland

Table of Contents
Logistics………………………………………………………………………………………...2

Icelandic Language…………………………………………………………………………...3

Daily Itinerary………………………………………………………………………………..…4

Þingvellir (Cultural History & Geologic Setting)…………………………………………5

Geysir and Gulfoss………………………………………………………...........................10

Rhyolitic Volcanism in Landmannalaugar…………………………………………...….12

Icelandic Highlands……………………………………………………………....………….16

Columnar Basalts: Morphology and Processes……………………………….…….....19

Chronology and Characteristics of the Hverfjall Eruptive Fissure……………..…..25

Geothermal Power and Hydrothermal Systems……………………………………......29

Faults, Fissures, and Catastrophic Floods………………………………………..........34

Askja Volcano and Surrounding Area……………………………………………….......38

Eastern Iceland: Geology, Culture, and Industry…………….………………………...40

Icelandic Mineralogy and Geochemistry…………………………….……………….….41


.
Glaciers and Jökulhaups…………………………………………………………….......…44
 
Recent Volcanism………………………………………………………………………..…..53

Reykjavik.......................................................................................................................55

List of Maps……………………………………………………………………………...……58

  1
Caltech Enrichment Trip Iceland

Logistics
Flights
AS = Alaska Airlines; FI = Icelandair; LAX = Los Angeles; SEA = Seattle; KEF = Keflavik

Outbound (August 16, 2014)


•AS439 LAX (depart 11:00 PST) à SEA (arrive 13:30 PST)
•FI680 SEA (depart 16:30 PST) à KEF (arrive 06:45 GMT August 17)

Inbound (August 29, 2014)


•FI681 KEF (depart 17:00 GMT) à SEA (arrive 17:45 PST)
•AS456 SEA (depart 19:55 PST) à LAX (arrive 22:19 PST)
•AS458 SEA (depart 20:30 PST) à LAX (arrive 23:00 PST)

Iceland Basics
Emergency Numer: 112
USA embassy: 595-2200

•Currency: Icelandic króna (IKR)


•Exchange Rate (12-month range): 1 USD = 112—123 IKR
•Time zone: GMT (but no DST, so summers are 1 hour behind London, 7 hours ahead of LA)
•Credit Cards: Accepted at most locations, even in rural areas. VISA and MasterCard are
ubiquitous, AmEx less so
•ATMs: Available in all towns
•Electricity: 220 V/50 Hz, 2-pronged (CEE-type)
•Tipping: None
•Etiquette: Smoking is illegal in enclosed public spaces, bars, and restaurants. Remove shoes
when entering homes. Shower thoroughly before entering spas and hot springs
•August weather and climate: daytime highs around 55˚F (12˚C) and daytime lows around 45˚F
(7˚C); ~15 hours of daylight (06:00–21:00), clear skies 15% of the time, 75% chance of drizzle
each day, 50% chance of rain each day, 70-90% daily relative humidity

Trip Participants
Faculty: Mark Simons, Rob Clayton, Bethany Ehlmann
Students: Toby Bischoff, Jennifer Buz, David Case, Jen Caseres, Peter Gao, Josh Kammer,
Masha Klescheva, Semechah Lui, Miki Nakajima, Stephen Perry, Bryan Riel, Kirsten Siebach,
Sarah Slotznick, Natalia Solomatova, and Robb Wills

Acknowledgments
We would like to thank Shirley A. Kliegel for her generous sponsorship of this trip. David Case
edited this field guide and Peter Gao compiled the maps. Existing field guides from two previous
academic trips to Iceland were helpful in putting ours together: The 2009 Harvard University
field guide (edited by G. Sterenborg, J. Crowley, and E. Kiser), and the 2010 Columbia
University field guide (edited by J. Jweda and M. Reitz). The cover photo is by Orsolya and
Erlend Haarberg from nationalgeographic.com and shows Hverfjall and Lake Mývatn in winter.

  2
Caltech Enrichment Trip Iceland

Icelandic Language
Descended from Old Norse, Icelandic is a Germanic language which retains several phonetic
characters that have fallen out of use in modern English. Icelandic is the national language of
Iceland, but citizens learn English in primary school and generally speak it well. As is always the
case when travelling internationally, a little knowledge and effort in the local native language will
go a long way, even if the conversation could easily be accomplished in English.

Below is a list of common characters, and how they are pronounced1:


Character Pronunciation
Áá ow (as in “how”)
Ðð dh (as “th” in “that”)
Éé ye (as in “yet”)
Íí ee (as in “see”)
Óó oh (as “note”)
Úú oo (as in “too”)
Ýý ee (as in “see”)
Þþ th (as in “think”)
Ææ ai (as in “aisle”)
Öö eu (as “u” in “nurse”)

Below is a list of common and useful expressions1,2:


Hello Halló
How are you? Hvað segir þú gott?
Excuse me Afsakið
Yes Já
No Nei
Please Takk
I’m fine Allt fínt
Can you show me on the map? Geturðu sýnt mér á kortinu?
What’s your name? Hvað heitir þu?
My name is… Ég heiti…
Where are you from? Hvaðan ertu?
I’m from… Ég er frá
Good morning/afternoon Góðan daginn
Goodbye Bless
Cheers! Skál!
How much is this? Hvað kostar þetta?
Sorry Fyrirgefðu
Thank you Takk fyrir (or just takk)
How do you say … in Icelandic? Hvernig segir maður … á íslensku?
Do you have vegetarian food? Hafið þið grænmetisrétta?

1
Lonely Planet: Iceland. Presser, B., Bain, C., and Parnell, F., 2013.
2
Harvard EPS Graduate Student Field Trip to Iceland. Sterenborg, G., Crowley, J., Kiser, E.,
2009.

  3
Caltech Enrichment Trip Iceland

Itinerary
DAY OF DAY OF ACTIVITIES EVENING
WEEK MONTH Geology/Culture DESTINATION
Sat 16 Leave Los Angeles midday
Arrive Keflavik early AM, shop for food
Þingvellir (rift/columnar basalt, 1st parliament) Sjóva house
Sun 17
Geysir Ásólfsstaðir in
Gulfoss Þjórsárdalur (801
Landmannalaugar (~3-4hrs, with hike) Selfoss)
Mon 18
Reykholt swimming pool (open 10:00-22:00)
Deildartunguhver Thermal Spring (30min stop)
Tue 19 Nyidalur Hut
Stong (old mud house)
Aldeyjarfoss
Wed 20
Ódáðahraun "desert" lava field
Grjótagjá cave
Mývatn rhyolitic cone, view of phreatovolcanics
Hverir, Námafjall springs Mývatn
Thu 21
Krafla Power Plant (University of
Krafla Volcano (if time) Iceland Science
Mývatn Nature Baths (12:00-22:00) Institute house in
Husavik Reykjahlid)
Tjornes fracture zone
Fri 22 Triple junction at þeistareykir
Åsbyrgi
Dettifoss
Sat 23 Pillow lava stop (if time)
Thorsteinsskali in
Askja, including hike into caldera and to crater lake
Sun 24 Herdubreidalindir
Vatnajoekull outwash plains (64.9N 16.75W)
Karahnukur dam and powerplant
Visit a farm (if time)
Eiðar in
Mon 25 Berufjordur zeolites #1, 64.8057N 14.5582W
Egilsstaðir
Berufjordur dike, 64.7912N 14.5153W
Breiðdalsvík, George P.L: Walker geology center
Djupivogur
Teigarhorn, Zeolite type locality
Tue 26 Horgsland hotel
Rock museum in Stodvarfjordur
Fish factory (if time)
Jokulsarlon glacial lagoon, black sand beach Þykkvabaejarklau
Skaftajell glacier (if time) stur
Wed 27
Skeidararjokull – Gigjukvisl, view of outwash plain http://kiddasiggi.is
Laki, Svartifoss (if time) /node/234
Coastal columnar basalt at Dyrhólaey, Vik
Solheimarjokull glacier hike Loft Hostel,
Thu 28 þorvaldseyri visitor center, Eyjafjallajokull eruption Bankastræti 7,
Seljalandsfoss Reykjavik
Grindavik lighthouse (if time)
Fri 29 Visit Reykjavik in the morning, leave for afternoon flight; arrive late night LA

  4
Caltech Enrichment Trip Iceland

Þingvellir: Cultural History


Contributed by Toby Bischoff

Relevant Icelandic vocabulary:


Þing – Thing / Assembly
Þingvellir - Thing fields / Parliament plains
Alþingi – All-thing / General assembly
Lögberg - Law Rock

Short summary:
At Þingvellir, which literally stands
for assembly field or parliament
plains, the general assembly of
the nation of Iceland is located.
The assembly first came together
at Þingvellir in 930 AD and
continued to meet there until 1798.
Þingvellir National Park was
founded in 1930 and remains a
property of the Icelandic nation.
Today it is one of the most popular
tourist destinations because of its
historical, cultural, and geological
importance.
Figure 1. Þingvellir from the information center overlook
(from Wikipedia).
Long Summary:
After settlement of Iceland
in 874 AD, people of Norse (Vikings/people form northern Europe speaking the old Norse
language) and Celtic (another ethno-linguistic group of tribal societies in Europe) origin had
populated the island to the extent that small regional assemblies were no longer sufficient to
deal with political matters and a general assembly was necessary. The Þingvellir region was
chosen due to it’s relatively convenient location, minimizing travel time and effort to and from the
assembly location, and because the land was declared public around the time (conveniently the
previous owner was found guilty of murder). The longest travel time was ~17 days for the tribal
chief that was in charge of the easternmost region of Iceland. Historians regard the foundation
of the general assembly in 930 AD at Þingvellir as the founding moment of the Icelandic nation
because it brought the population on a path of common cultural heritage and ultimately to
national identity.
The Alþingi is the name of the general assembly of Iceland and remained at Þingvellir
until 1271 as Iceland's supreme legislative and judicial authority. The so-called Lögberg was the
focal point of the Alþingi and a platform for holding speeches. The Lawspeaker, who was the
president of the assembly, was elected to three year terms. His job was to recite the law of the
land, which before written-down laws meant orally reciting the laws from memory over the three
summers during the Lawspeaker’s term. Additionally, he presided over assembly procedures
that took place every summer at the Lögberg. Important announcements concerning the nation
were made there and anyone attending the assembly was allowed to present his case at the
Lögberg.

  5
Caltech Enrichment Trip Iceland

In addition to inauguration and dissolution


of the assembly, new rulings by the Law Council
were announced at the Lögberg. The Law Council
was the legislative assembly and thus a sub-
institution of the Alþingi. The work of the Law
Council was not limited to passing new laws only,
but it was responsible for settling disputes as well.
It thus functioned as a legislative and judicial
institution at the same time. Unlike the Alþingi, the
Law Council was a closed body in which only
certain people enjoyed full rights: chieftains (48 of
them) who held the office of "goði", their "Þingmen"
(advisors) and later also bishops. However,
everyone at the assembly was entitled to watch
and listen to the Law Council at work.

th
Figure 2: 19 century rendering of the
Lögberg (from Wikipedia).

References & Further Reading


http://www.thingvellir.is/
http://en.wikipedia.org/wiki/%C3%9Eingvellir
http://en.wikipedia.org/wiki/Althing
http://en.wikipedia.org/wiki/L%C3%B6gberg
http://www.personal.utulsa.edu/~Marc-Carlson/history/grontime.html
http://www.thingvellir.is/history/the-law-council.aspx

Þingvellir: Geologic Setting


Contributed by Sarah Slotznick
Þingvellir: Rifting
Þingvellir is located to the northwest of the shield volcano Hengill that is the location of
one of the triple junctions that forms the Southern Iceland microplate (Einarsson 2008). In this
location, the Reykjanes Peninsula Ridge (RPR) meets the east-west striking South Iceland
Seismic Zone (SISZ) and the Western Volcanic Zone (WVZ) striking to the northwest (Figure 1).
Thingvellir thus records rifting activity along the WVZ, which was the primary zone of spreading
from ~6Ma until 2Ma when the Eastern Volcanic Zone (EVZ) formed via propagation from the
Northern Volcanic Zone (Ivarsson 1992). Based on GPS measurements, today only 20-30% of
the total spreading across southern Iceland is accommodated by rifting on the WVZ; the rest is
on the EVZ (Sinton et al. 2005). Based on 9 years of measurements, the spreading rate across
the boundary is 3mm/yr in the NE to 7mm/yr in the SW, which classifies the WVZ as an ultra-
slow spreading ridge (LaFemina et al. 2005, Sinton 2005).

  6
Caltech Enrichment Trip Iceland

Figure 1: Map of the main


fault structures and volcanic
belts that form the plate
boundaries in Iceland. Plate
boundary segments are
labeled: RPR Reykjanes
Peninsula Rift, WVZ Western
Volcanic Zone, SISZ South
Iceland Seismic Zone, EVZ
Eastern Volcanic Zone, SIVZ
South Iceland Volcanic Zone
(propagation of the Eastern
Volcanic Zone), CIVZ Central
Iceland Volcanic Zone, NVZ
Northern Volcanic Zone, GOR
Grímsey Oblique Rift and HFZ
Húsavík-Flatey Zone (Part of
the Tjörnes Fracture Zone),
ER Eyjafjardaráll Rift, DZ
Dalvík Zone. Kr, Ka, H, L, V
mark the volcanoes of Krafla,
Katla, Hengill, Langjökull, and
Vestmannaeyjar. (from
Einarsson 2008).

Figure 2: Geologic
map of the postglacial Z
lavas and main faults
in the northern part of
the WVZ. Location of
Thingvellir,
Thingvallavatn, and Z’
Hrafnabjörg table
mountain (H) are
marked. Topography
contours areat 100m.
(from Sinton et al.
2005).

  7
Caltech Enrichment Trip Iceland

a)

A, Almannagja; S, Sledaasgja; AF; Armannsfell; LT, Lake


Thingvallavatn; AR, Arnarfell; L, Litlagja; H, Hrafnagja;
Gildruholtsgja; HE, Heidargja. 1, normal fault (only those with
throws of several meters are thus indicated); 2, tectonic fissure
(with a small or no throw); 3, deformation zone; 4, Pleistocene
rocks (basaltic lavas or hyaloclastites) or alluvium.
b)

Figure 3: a) Map of the Thingvellir Fissure Swarm. (from Gudmundsson 1987) b) Cross section of
Thingvellir Graben near Thingvellier, exact location can be seen in Figure 2 (Z to Z’). msal = meters
above sealevel. (from Saemundsson 1992)

Geologic mapping paired with 14C-dating has identified 44 eruptive units along the WVZ
in the past 11,000 years although there has been a decline in volcanism over that period of time
(Figure 2, Sinton et al. 2005). 85% of these post-glacial volcanic eruptions have been from
volcanic centers of lava shields and cones instead of fissures, which are the majority of
eruptions in the other Icelandic volcanic zones. The EVZ lacks any shield volcanoes, while
>90% of volcanic production in the WVZ is from these shield volcanoes (Sinton et al. 2005).
The lavas seen in Þingvellir are about 10,000 years old, and the lack of recent resurfacing by
younger lavas allows us to see the most striking structural feature of the WVZ: the Þingvellir
Graben. This tectonic depression, 25km wide in the NW to 10km in the SW, is bound by several
successive layers of faults (Figure 3). The large inner graben tectonism occurred around 9500-
8000 years ago (Sinton et al. 2005), bound on the east by the normal fault Almannagjá and
other minor faults with large throws up 40m (at Þingvellir) and on the west by the Hrafnagjá fault
system with smaller vertical offsets (Gudmundsson 1987). Some faults cutting older hundred-
thousand-year-old units have larger vertical offsets up to 400m indicating a longer history and
gradual development of the Þingvellir Graben, the deepest graben with the largest single fault
throws in Iceland (Saemundsson 1992).
There has been recent rifting activity on the WVZ. In early June of 1789, following
previous events on the EVZ in 1783 and SISZ in 1784, 10 days of significant earthquake activity
were felt and recorded by the vicar of Þingvellir along with subsidence and flooding of the lake

  8
Caltech Enrichment Trip Iceland

waters in the center of the graben (~1.4-1.5m) and elevation at the edge of the graben with
wells running dry and trails across faults becoming impassable (vertical throws up to 2m
reported to the south). To the south of Hengill (in the RPR), a new hot spring appeared but no
volcanic eruption occurred, suggesting the event was associated with crustal dike propagation
(Saemundsson 2006).
The Þingvellir Graben with its extraordinary deepness, fast subsidence, and large width
has been focal point in understanding the nature of rifting along the WVZ since this zone is very
different from the other rift zones in Iceland. Its ultra-slow spreading rates, a decline in
volcanism, and the deep graben structure at Þingvellir have led many to suggest that the WVZ
is a failed or dying rift and all motion will eventually be transferred to the EVZ as it continues its
southward propagates (e.g. Pálmason 1981, Einersson 1991). Both in theory and through
models it has been shown that the deep graben is a sign of a magma-starved rift with plate
divergence being accommodated by crustal stretching and normal faulting instead of volcanism
(Saemundsson 1992, Karlsson and Sigmundsson 2008, Sturkell et al. 2013). However, several
pieces of contradictory evidence have led to debate about the WVZ in the past several years.
Although there has been a decrease in volcanism, production is similar to that seen all over
Iceland post-glacially and is constant along the WVZ, whereas a dying rift should be
progressively dying from north to south (Sinton et al. 2005). In a counterargument, Sonette et al.
2010 find a decline in tectonic activity along the WVZ (via decreased fissure zone growth rates)
and suggest if there is steady-state volcanism still present, it could be due to a decoupling of the
Icelandic hotspot and the spreading center/ridge as it moves to the west. Other studies note that
this zone is still very seismically active with more earthquakes than the EVZ in the past several
decades (Einarsson 1991).

References & Further Reading


Einarsson, P. (1991). Earthquakes and present-day tectonism in
Iceland.Tectonophysics, 189(1), 261-279.
Einarsson, P. (2008). Plate boundaries, rifts and transforms in Iceland. Jökull,58, 35-58.
Gudmundsson, A. (1987). Tectonics of the Thingvellir fissure swarm, SW Iceland. Journal of
Structural Geology, 9(1), 61-69.
Ivarsson, G. (1992). Geology and petrochemistry of the Torfajokull central volcano in central
south Iceland, in association with the Icelandic hot spot and rift zones (Doctoral
dissertation).
Karlsson, E.& Sigmundsson, F.(2008). How do grabens form: the influence of plate spreading
on topography in a magma starved rift. In Geophysical Research Abstracts, Vol. 10,
EGU2008-A-03158
Pálmason, G. (1981). Crustal rifting, and related thermo-mechanical processes in the
lithosphere beneath Iceland. Geologische Rundschau, 70(1), 244-260.
Saemundsson, K. (1992). Geology of the Thingvallavatn area. Oikos, 40-68.
Sæmundsson, K. (2006, December). The 1789 rifting event in the Hengill volcanic system, SW-
Iceland. In AGU Fall Meeting Abstracts (Vol. 1, p. 1568).
Sinton, J., Grönvold, K., & Sæmundsson, K. (2005). Postglacial eruptive history of the western
volcanic zone, Iceland. Geochemistry, Geophysics, Geosystems, 6(12).
Sonnette, L., Angelier, J., Villemin, T., & Bergerat, F. (2010). Faulting and fissuring in active
oceanic rift: Surface expression, distribution and tectonic–volcanic interaction in the
Thingvellir Fissure Swarm, Iceland. Journal of Structural Geology, 32(4), 40
 
 

  9
Caltech Enrichment Trip Iceland

Geysir and Gulfoss


Contributed by Toby Bischoff
 
Relevant Icelandic vocabulary:
Geysir – Namesake of the English “geyser” (from the Icelandic “to gush”)
Gullfoss – Golden Falls
Hvítá – White

Short summary:
Geysir (The Great Geysir) is a geyser in Haukadalur, in southwest Iceland. It is famous
because it lent its name to all other erupting geothermal springs worldwide. Gullfoss is a large
waterfall in southwest Iceland. It is located on the Hvítá River. The waterfall has two stages,
11m and 21m high, and thus a total height of 32m (105ft; cf. Niagara Falls: 51m). The average
flow rate is 140 m3s-1 (cf. Niagara Falls: 2,400 m3s-1). Gullfoss is approximately 20m (60ft) wide
and one of the most popular attractions for tourists on the Golden Circle.

Long summary:
The Great Geysir
(named from the Icelandic “to
gush”) is located in the
Haukadalur Valleys, which it
shares with other geysers
such as Strokkur (up to 40m
high). The hot spring was first
mentioned in 1294 when a
strong earthquake changed
the geological activity of the
region. The first mentioning of
the name “Geysir” dates to
1647 by Bishop Brynjólfur
Sveinsson and it was
recognized that the activity of
the hot spring is connected to
the occurrence of earthquakes Figure 1: The Great Geysir (from Wikipedia).
in the region. Geysir’s activity,
like all natural hot springs, is
related to surface water that seeps through the ground until it reaches rocky material that is
heated by magma from below. The water then rises back to the surface by convection through
cracks and porous rocks.
The main difference between a conventional hot spring and a geyser lies in the
subsurface structure. Geysers usually consist of a small hole that is connected to one or
multiple narrow tubes below the surface. The dynamics behind the geyser mechanism were first
explained by Robert Bunsen in 1846 (cf. Bunsen burner), who was based in Marburg at the
time. The water at the top of a geyser can easily cool off over time, but exchange with the
warmer water below is difficult due to the small width of the tubes below the surface. This
prevents convective cooling of the water deeper below the surface. The heat flux through the
rock at the bottom heats the water until it reaches the boiling point, at which point gas bubbles

  10
Caltech Enrichment Trip Iceland

develop and ultimately reach the surface


(cf. pressure cooker). This reduces the
pressure on the superheated water
reservoir at the bottom, which leads to
instant vaporization of the bottom water
and the geyser erupts violently. The time
scale of eruption depends on the heat flux
that warms the water below and on the
length of the tubes below the ground. The
longer the tubes, the higher the water
pressure at the bottom of the geyser will
be. This increases the boiling temperature
of the water at the bottom and thus
lengthens the eruption time scale.
Similarly, a larger heat flux from below can
Figure 2: The Great Geysir erupts (from Wikipedia). lead to shorter eruption periods because
the water reaches its boiling point faster.
The Great Geysir in Iceland was
dormant between 1915 and 1935 when it became active for a few years. The most recent period
of natural activity began in June 2000 after an earthquake had occurred in the region, but
eruptions remain sporadic. Before, the Great Geysir was reactivated by human intervention
only. The Great Geysir has shown
some very high eruptions in the
past, some of which have reached
100m and more.
The Gullfoss waterfalls are
located close the geyser field
surrounding the Great Geysir.
They are located along the Hvítá
(Ölfusá) river. The Gullfoss
consists of two waterfalls that
have a combined height of 32m.
As the Hvítá flows southward it
turns sharply to the West about
1km (0.62 mi) before the falls and
rushes down into a wide, curved
three-step "staircase" before it
abruptly plunges in two steps
(11m and 21m, respectively) into a Figure 3: Gullfoss waterfall in September (from Wikipedia).
crevice 32 m (105 ft) deep. The
crevice, about 20m (60 ft) wide,
and 2.5 km in length, extends perpendicular to the flow of the river. The name of the waterfalls
“Gullfoss” translates to golden falls. This is because the river carries glacial sedimentary
materials that can give the falls a golden appearance under the right sunlight conditions.

References & Further Reading


http://en.wikipedia.org/wiki/Geysir
http://en.wikipedia.org/wiki/Gullfoss
http://www.gullfoss.is/

  11
Caltech Enrichment Trip Iceland

Rhyolitic Volcanism in Landmannalaugar


Contributed by Semechah Lui
 
Overview
Landmannalaugar is an area in south central Iceland with centuries of volcanic activity. It
is situated in the large, rhyolitic Torfajökull Caldera, the largest caldera in Iceland. The 18-km
diameter caldera is home to the country’s largest geothermal area. Torfajökull is a rhyolitic
stratovolcano and complex of subglacial volcanoes located north of Mýrdalsjökull and south
of Þórisvatn Lake. It is located at the intersection of a spreading zone to the northeast, the
South Iceland Seismic Zone to the west and a non-rifting flank zone to the south. Its last
eruption was in 1477 and consists of the largest area of silicic extrusive rocks in Iceland. The
region is anomalous because of its rhyolite to basalt ratio of 4:1, in contrast to the 1:5 seen at
the other central volcanoes around the island. The following includes a general section on
rhyolitic volcanism, followed by a description on the volcanic features, geology and petrology of
the region.
Rhyolitic volcanism
Rhyolite is characterized by its felsic (silica-rich) composition (65 – 70% SiO2). It is high
in K and Na+ but low in Fe2+, Mg2+ and Ca2+. The mineral assemblage is mostly quartz, sanidine
+

and plagioclase. Biotite and hornblende are common accessory minerals. Texture varies from
glassy to aphanitic to porphyritic. In general, temperature, viscosity, and gas content are three
major factors governing the behavior of magma. The temperature of typical rhyolitic magma is
650 – 800°C, which is lower than basaltic and andesitic magma, hence the higher viscosity.
Felsic magmas also tend to have higher gas contents than mafic magmas. These give rhyoltic
magma an explosive eruptive behavior. As rhyolites reach the surface of the earth, those that
cool too quickly to grow crystals form a natural glass or vitrophyre (obsidian). Slower cooling
forms microscopic crystals in the lava and results in textures such as flow foliations, spherulitics,
nodular, and lithophysal structures. Some rhyolites are highly vesicular pumice. Since many
eruptions of rhyolite are highly explosive and pyroclastic, the deposits may consist of fallout
tephra or of ignimbrites.
The physical and chemical characteristics of rhyolitic volcanic ash are primarily
controlled by the style of volcanic eruption. Ash is formed under different processes.
Fragmentation is one that generates very fine ash without the addition of water. During
explosive eruptions, magma decompresses as it rises, allowing dissolved volatiles (dominantly
water and carbon dioxide) to exsolve into gas bubbles, which nucleate a foam that decreases
the density of the magma, accelerating it up the conduit. Fragmentation occurs when bubbles
occupy approximately 70 – 80 volume % of the erupting mixture, during which expanding
bubbles tear the magma apart into fragments and they are then ejected into the atmosphere
where they solidify into ash particles (< 2 mm). Ash is also produced during phreatomagmatic
eruption when there is direct coupling of the cold water and hot magma. Heat transfer leads to
the rapid expansion of water and fragmentation. The most abundant ionic species found in fresh
ash leachates are Na+, K+, Ca2+, Mg2+, Cl-, F- and SO42-. High-silica ash consists of pulverized
products of pumice, individual phenocrysts and some lithic fragments. Ash generated during
phreatic eruptions primarily consists of hydrothermally altered lithic and mineral fragments,
commonly in a clay matrix. The morphology of ash particles is mostly dependent on the shape
of vesicles in the rising magma before disintegration, or during on the stresses within the chilled
magma during phreatomagmatic eruption.

  12
Caltech Enrichment Trip Iceland

Typical volcanic landform associated with rhyolitic volcanism is stratovolcano (or


sometimes called composite volcano), which exhibits inter-layering of lava flows as well as
pyroclasts. Stratovolcanoes are usually concave-shaped and have steeper slopes than other
types of volcanoes. The steep slope near the summit is due to thick viscous lava flows that do
not travel far downhill from the vent. The gentler slopes near the base are due to accumulations
of material eroded from the volcano and to the accumulation of pyroclastic materials. Some
stratovolcanoes consist of craters at the summit that are formed by explosive ejection of
material from a central vent. Sometimes the craters can be filled by lava flows/domes, glacial
ice, or water.
Landscape / volcanic features of the region
The Torfajökull volcanic system forms a massive highland elevated 300–600 m above
the surroundings. Its eastern half is heavily eroded contrary to the western half where Holocene
lavas have been erupted. The eastern area offers the best exposures and allows the
development of the volcano to be traced far back. Bright colored rhyolite is almost the sole rock
type there with remnants of hyaloclastite (basaltic) overlying it in places with NE-SW-feeder
dykes cutting up through the
rhyolite below. The western half is
undulating high ground where
hyaloclastite, obsidian flows and
dark grey ash shed over it from
neighboring volcanoes lend the
landscape a blackish hue. In
Landmannalaugar, one can see
the striking landscape of elongated
sharply crested hills and V-shaped
valleys, which is formed by rapid
erosion of the soft, rhyolitic lava
that were recently deposited there.
The mountains are split with
gullies and gorges, one of which,
the Jökulgil, is about 13 km long.
Rhyolites in the area have
subdued, earth colors, fluctuating
from green to yellow-beige and
brick red. A recent outpouring of
obsidian, a glassy and fluid variety
of lava, can be seen meandering
across the rhyolite hills like a black
snake.
Geothermal activity is
widespread, mostly as fumaroles
inside the caldera. Warm and hot
springs are also common around
the periphery of young lavas and
other permeable formations. CO2-
Figure 1: Outline geologic map of the Torfajökull central rich hot springs occur in the east of
volcano. The curved line represents the ring structure (from the caldera and to the SW of it.
Gretar Ivarsson, 1992).
Alteration is pervasive in the older
.

  13
Caltech Enrichment Trip Iceland

rock units. Hot ground is extensive around the fumarole fields.

Geology / Petrology of the area


Volcanic activity in this area is traced back to the last interglacial period and appears to
have been most vigorous during the last glaciation (Weichsel) when large volumes of
peralkaline rhyolites were produced. The silicic rocks at Torfajökull can be divided into four main
structural series (Figure 1). These are, in order of decreasing age: the Brandsgil Series
(115,000 – 130,000 YBP), the Jökulgil Series (65,000 – 115,000 YBP), the Blahnukur Series
(10,000 – 65,000 YBP) and the Postglacial Series (last 10,000 years).
During the Brandsgil Series, there was formation of thick rhyolitic airfall welded tuff,
followed by the eruption of subaerial rhyolitic lava flows found in Brandsgil, Jökulgil and in low
exposures in Sudumamur and Barmur, as well as eruptions of similar units in the south and
west. The boundary between the Brandsgil Series and the Jökulgil Series is marked by an
abrupt change in the chemistry of the erupted rhyolites. This change resulted in the
establishment of a highly peralkaline rhyolitic top in the magma chamber and the doming of the
central region, accompanied by rhyolitic eruptions on short NW-SE trending fissures within the
outside the central area.
The transition from the Jökulgil Series to the Blahnukur Series is reflected by the end of
further production of highly peralkaline rhyolites, following a large silicic eruption through arcuate
fissures, which effectively removed this magma from the chamber. A peralkaline tephra layer
has been found in piston cores south of Iceland. It is 65,000 years old and has an estimated
volume of 7 to 8 km3. As Torfajökull is the only volcano in the North-Atlantic known to have
produced peralkaline magmas in the last 100,000 years, it is likely that a large plinian eruption
took place there. Connecting this plinian eruption with the extrusion of the peralkaline rhyolites
forming the ring structure and that it marks the end of the Jökulgil Series period, gives the
possible age of 65,000 years for the transition into the Blahnukur Series period. During the
Blahnukur Series, through NE-SW trending fissures, there is increased and wildly distributed
eruptions of transitional basalt within and around the ring structure. It is also believed to be the
period of re-establishment of a rhyolitic magma body at Torfajökull. Rifting events in the Eastern
Rift Zone has caused eruptions of rhyolites outside the ring structure. Tholeiite was found for the
first time as inclusion in a rhyolitic host.
Lastly, the transition into the Postglacial Series is based on the disappearance of the
icesheet, but there is no apparent abrupt changes in the eruptive behavior or geochemisty.
During the Postglacial Series, there is a mixed and hybridized rhyolitic and tholeiitic basaltic
activity in the NE, and rhyolitic activity in the central region. In particular, the Postglacial Series
consists of 10 separate eruptive events, mostly metaluminous rhyolites, transitional basalts and
intermediate rocks, tholeiites and various hybrid compositions. Some of the eruptions were
actually triggered by tholeiitic injections from the northeast. The oldest occurred about 9000
years ago, the three youngest about 2000, 1100 and 500 years ago. These Postglacial
volcanics cover 55 km2 and have a volume of 0.9 km3 (0.8 km3 of rhyolites and 0.1 km3 of
basalts and intermediate rocks). The production rate was much lower than previously
experienced at Torfajökull. Some of them had an initial plinian phase which produced
widespread air fall ash layers (Figure 2). The volcanics have all formed on NE-SW trending
fissures and some are related to simultaneous tholeiitic eruptions in the Veidivötn fissure swarm.
The lava fields are located in three distinct areas with Torfajökull, one on the northeastern
perimeter, one in the western central region, and one in the far southwest. Additionally, a
number of lava flows, many of which are hybrids representing mixing between Torfajökull
magmas and tholeiites from the Veidivötn fissure swarm, occur on the NE margin of Torfajökull.

  14
Torfajökull 5
Caltech Enrichment Trip Iceland

Figure 2: Postglacial
lavas flow of the
Torfajökull central
volcano and plinian
phase fall-out of rhyolitic
ash which was erupted
in the initial stage of the
eruptions (brown
arrows). Rhyolite lavas
are in red. Basaltic and
andesitic lavas are blue.
The yellow dotted
contour marks the
rhyolite area (from
Kristján Saemundsson,
2009.)

There are also distinct differences observed between the basalt in the northern part and the
FIGURE 4:
central and southwestern Holocene
part lavasInoftheTorfajökull
of Torfajökull. volcano
north, tholeiites and plinian
are found phase
along with fall-out of rhyoli
various
was erupted
hybrid compositions, in the
indicating initial
mixing stage ofwith
of tholeiite therhyolite
eruptions.
and Rhyolite
transitional(mixed) lavas
icelandite. are red. Basaltic
In the
lavas areablue.
central region and the far southwest, rangeThe yellow, stippled
of transitional contour
compositions marks thebutrhyolite
is observed not area
tholeiites.
In summary, more than 90% of volcanics erupted within the large ring structure are
rhyolites, while units of basaltic composition are mostly restricted to the outer margins and as
inclusions. Rhyolitic activity was most intense during the Brandsgil and Jökulgil Series and has
decreases with time. Based on geochemistry and morphology, one can speculate that the large
ring structure was formed in a single large eruption. Increased basaltic activity is observed in the
area following the eruption.

References & Further Reading


1. Dave McGarvie, "Rhyolitic volcano–ice interactions in Iceland." Journal of Volcanology and
Geothermal Research 185.4 (2009): 367-389.
2. Gretar Ivarsson, “Geology and petrochemistry of the Torfajokull central volcano in central
south Iceland, in association with the Icelandic hot spot and rift zones.” Dissertation (1992).
3. Kristján Saemundsson, “Torfajökull, iceland – a rhyolite volcano and its geothermal resource”
Exploration for Geothermal Resources (2009).
4. http://www.tulane.edu/~sanelson/eens1110/volcanoes.htm

  15
Caltech Enrichment Trip Iceland

Icelandic Highlands
Contributed by Robb Wills
The Icelandic highlands aren’t very high, predominantly 500-1000 meters, but the 65°N
latitude and volcanic nature of the soil means a nearly complete lack of vegetation except along
the many rivers draining the major glaciers in the area. The terrain is primarily volcanic sand /
ash and lava fields, and has been reworked by glaciation during the ice ages, and presently by
fluvial and aeolean erosion. A soil map shows the distribution of different soil types and amount
of volcanic and glacial reworking throughout Iceland (see Maps section at the end of this field
guide). Types that indicate the presence of vegetation are 1-13 or green, yellow, and pink.

Figure 1: (A) Regional mean annual air temperature (MAAT)-based map of mountain permafrost
distribution on Iceland. The contours indicate the lower limit of potential mountain permafrost. The
shaded areas show the distribution of predicted permafrost based on topography extending to
elevations with MAAT <38°C. The enclosed areas are sporadic permafrost according to Brown et al.
(1995) and Priesnitz and Schunke (1978), mostly related to palsa and organic-rich soils. The hatched
ellipses denote areas of active rock glaciers and their elevation range according to Águst
Guðmundsson (unpublished MSc thesis, 2000). The triangles indicate the location of rock glacier
velocity measurements east for Hjaltadalur (Wangensteen et al., 2006) and Nautadalur (Whalley et al.,
1995a,1995b). (B) The inset map shows the original MAAT map based on Tveito et al. (2000). (C)
Topographic profile lines with proposed lower limit of mountain permafrost. The profile lines are
indicated as solid lines in (A). [From Etzelmüller et al. 2007, references contained therein].
  16
Caltech Enrichment Trip Iceland

Much of the
highlands are covered
in permafrost, where
some portion of the
soil is permanently
frozen. Permafrost is
able to persist in
regions where mean
annual air
temperature is below -
2 or -3 °C. In these
regions, an active
layer, which melts and
refreezes seasonally,
covers a deep
permanently frozen
layer. This melting
and refreezing
Figure 2: The Sprengisandsleið (F26), our route across the Icelandic process transports
Highlands. [From Wikipedia: Route F26 (Iceland)] sediment and causes
the arrangement of
often nearly-hexagonal patterns. Figure 1 shows the mean annual air temperature that can be
expected across Iceland and the resulting distribution of permafrost, glaciers, and rock glaciers.2
Our route will take us across profile A, through a few of the areas of active permafrost, and near
the area of sporadic permafrost or palsa, where ice lenses can form seasonally or for several
years when there is plenty of soil moisture but little insulating snow cover. Overall, permafrost
covers about 8% of the surface area of Iceland.2 Transport and burial of carbon by microbes in
permafrost are important processes in mediating global atmospheric CO2 and methane,
particularly in areas that aren’t as small as Iceland, such as Arctic Canada and Siberia.
There are no paved roads across the highlands, but it is traversed by a network of gravel
roads, most of which require 4 wheel drive vehicles capable of fording rivers. We will be taking
the F26 or Sprengisandsleið, the longest of these roads. The name is based on the words
‘sanssleið’ or sand path and ‘sprengja’ which means “to ride a horse to death or to be on the
point of bursting after running for too long”,3 and was so named because of the extended
distance one must travel without any horse fodder or shelter. This name originates from the
period 900-1200 AD during the Icelandic Free State when this pass was used to connect
villages in Northern Iceland to the yearly summer parliament at Þingvellir.3 Most medieval
Icelanders however, chose to avoid this path and take the longer paths along the inhabited
coasts, both because of the dangers involved and because they thought it to be inhabited by
ghosts. We will cover the road’s more than 200 km in 2 days.
We will be travelling through Sprengisandur, the highland plateau that extends between
Vatnajökull and Hofsjökull glaciers. The immediate surroundings of the road will mostly be a
desert of volcanic sand, punctuated by a few lakes. The word desert here has more to do with
the lack of vegetation than the lack of precipitation. Though dry for Iceland, some areas of
Sprengisandur still receive more than 1 meter per year of rainfall.4 The road will provide long-
distance views of Vatnajökull and Hofsjökull, the biggest and 3rd biggest glaciers in Iceland,
respectively. Our destination is Mývatn, a volcanic lake in the north of Iceland, with a few
villages of a few hundred people each. There are several stops or potential stops along the way
of geologic or other interest:

  17
Caltech Enrichment Trip Iceland

Stöng. Þjóðveldisbærinn Stöng is a reconstruction of a farmstead as would be found at this


location 1000 years ago in the Viking era. Archeological evidence suggests that it was buried by
volcanic ash during the 1104 eruption of Hekla.5

Veidivötn. A series of crater lakes near the beginning of our journey across the
Sprengisandsleið. These craters are believed to have formed associated with the phreatic
phase of the eruption of Bárðarbunga around 1477.6 A phreatic eruption is the explosive release
of steam associated with the heating of groundwater by a nearby magma chamber.

Hágöngulón and Vatnajökull outlet. This lake is situated between the F26 and Vatnajökull. It
is fed by meltwater from Vatnajökull, bounded on the southern side by the Hágönguhraun lava
field, and surrounded by a group of small cinder cones. With time to take the 10 km detour from
the main road, this would be a great spot to see the combined volcanic and glacial drivers of the
landscape.

Kvislavatn and Hofsjökull outlet. Down the road past Kvislavatn, Google Earth reveals a
fantastic braided river network draining the Hofsjökull glacier through volcanic sand. This is a
great vantage point of Hofsjökull and a good spot to talk about river dynamics in this heavily
transport limited river network near the source of the river Þjórsá, the longest river in Iceland.7

Tungnafellsjökull. A glacier-capped volcano off the road near Nyidalur Hut where we will
spend the night. This could be a good spot for a hike to get up close to this glacier and to get a
vantage point on the middle of Iceland.

Aldeyjarfoss. A fall of the Skjálfandafljót river where it drops 20 meters through the
Suðurárhraun lava field.8 The falls is made spectacular by the surrounding basalt columns.

References & Further Reading


1. Nygard, Iver J., 1959. Jardvegskort af Islandi (soil map of Iceland). Scale – 1:750,000. Map
adapted by university Research Institute Reykjavik, Iceland. Printed by US. Geological Survey.
2. Etzelmüller, Bernd, Herman Farbrot, Águst Guðmundsson, Ole Humlum, Ole Einar Tveito,
and Helgi Björnsson, 2007. The regional distribution of mountain permafrost in Iceland.
Permafrost and Periglacial Processes
3. Wikipedia: Sprengisandur. http://en.wikipedia.org/wiki/Sprengisandur <accessed July 25,
2014>
4. Crochet, P., T. Jóhannesson, T. Jónsson, O. Sigurðsson, H. Björnsson, F. Pálsson and I.
Barstad, 2007. Estimating the spatial distribution of precipitation in Iceland using a linear model
of orographic precipitation. J. of Hydrometeorol.
5. Wikipedia: Stöng. http://en.wikipedia.org/wiki/Stöng <accessed July 25, 2014>
6. Wikipedia: Veiðivötn. http://de.wikipedia.org/wiki/Veiðivötn <accessed July 25, 2014>
7. Wikipedia: Þjórsá. http://en.wikipedia.org/wiki/Þjórsá <accessed July 25, 2014>
8. Wikipedia: Aldeyjarfoss. http://en.wikipedia.org/wiki/Aldeyjarfoss <accessed July 25, 2014>

  18
Caltech Enrichment Trip Iceland

Columnar Basalts: Morphology and Processes


Contributed by Sarah Slotznick
Columnar Basalts: Summary
Columnar basalts are (a) 50(b) 200

found on every continent and m


LEGEND
Flow (single-unit)
338
cm

ne
90 cm Flow (multiple-unit)
8
33

:1 li
Lava lake

Igneous body thickness H (m)


40
throughout Iceland such as in Dome

100
Intrusion, sub-surface

Mean joint-length at a site (cm)


150 Dyke

Hrepphólar/Hreppar area (Mattson 30

et al. 2011, Bosshard et al. 2012, 20


100

Almqvist et al. 2012, Forbes et al. LEGEND

2014), Hjálparfoss and Gjáin (Lyle 10 Flow (single-unit)


Flow (multiple-unit) 50
line Lava lake
10:1 Dome
Intrusion, sub-surface

2000, Forbes et al. 2014), 0


0 50 100 150
Dyke

200
Mean joint-length at a site (cm)
Dverghamrar, Gerðuberg, 42 44 46 48 50 52 54
Lava SiO2-content [wt.%]
56 58 60

Hljóðaklettar, Kirkjugólfíð,
Reynishverfi (Hetényi et al. 2012) Figure 1: a) Igneous body thickness versus the average side
and Svartifoss (Guy 2010, Hetényi length of columns at 50 different columnar jointing sites in 3
et al. 2012, Tanner 2013). In countries. Note that erosion and partial exposure might have
addition to tiers of long equal-sized reduced the thickness of H and the thickness of the dike site
parallel columns and regular was divided by 1.5 to compare to free flows. Symbols filled
polygonal fracture patterns, these with grey are sites where the flow type could not be readily
sites exhibit smaller surface established and guesses were taken. b) Mean side length of
morphologies such as horizontal a column at each side and Lava SiO2 content in wt%. Upper
limits for intrusions and lava lakes is shown with dashed lines
striations, plumose hackles,
of corresponding colors. The long-dashed red lines are the
inscribed circles, and concentric best fit correlations between jointing side length and SiO2
ring features. Based on the content. (from Hetényi et al. 2012)
morphological observations,
petrography, in situ observation,
experiments, and numerical modeling, there are four models for the formation of columnar
basalts and their internal structures: 1) thermal
contraction potentially with a) water
interactions for entablature and/or b) pressure
and crystallization-induced melt migration with
viscous fingering 2) double diffusive
convection or constitutional super-cooling.

Columnar Basalts: Morphology and


Processes
Columnar joints consist of long
colonnades/columns with locally parallel axes
and regular polygonal fracturing. They
regularly occur in subaerial basalt flows, but
also can occur in lavas of other chemistries.
The columns can range in scale from
centimeters to meters in diameter and can
Figure 2: Sketch detauling hypothesis to explain extend through an entire flow unit up to ~ 30
equal-size distribution of columns within a flow meters, although usually they are in two or
unit. (from Budkewitsch and Robin 1994) more tiers of jointed columns up to many 10s
of meters (DeGraff and Aydin 1987,

  19
Caltech Enrichment Trip Iceland

Grossenbacker and McDuffle


1995). The basic model for
understanding how these
columns form is through thermal
volume contraction as the basalt
cools from its two contact
surfaces. Stress will accumulate
as the temperature falls below
that of elastic behavior. When
this stress exceeds the tensile
strength, tensional cracks will
form at the margins of these
flows perpendicular to their
boundaries and propagating
inward (DeGraff and Aydin 1987,
Budkewitsch and Robin 1994). Figure 3: Distribution of number of sides on polygonal cross
Isotropy of the crack pattern sections of worldwide columnar basalts. (from Budkewitsch
suggests that contraction occurs and Robin 1994)
equally in all directions during the
jointing (Budkewitsch and Robin
1994).
Most studies suggest the diameter of the columns is proportional to the cooling rate, i.e.
large thick columns have slower cooling times (Ryan and Sammis 1978, Long and Wood 1986,
Grossenbacker and McDuffle 1995). However, recently the column diameter/aspect ratio was
linked instead to the geology setting and chemistry of the columnar bodies. When geology
constrains the geometry emplaced body such as in dikes and intrusive bodies, columns are
thinner; in unconstrained geometries such as lava flows, chemistry becomes important with
increased SiO2 content creating thicker columns (Figure 1, Hetényi et al. 2012). The diameter of
the columns is fairly regular throughout a flow; one hypothesis is that while the cracks propagate
downward, they deviate parallel to the highest thermal gradient to maintain and achieve an
overall uniform size (Figure 2, Budkewitsch and Robin 1994).
Although normally pentagonal or hexagonal in cross section, the columns can have three
to eight sides. Columnar basalts around the world contain slightly different, but overall similar,
distributions of number of sides (Figure 3, Budkewitsch and Robin 1994, Hetényi et al. 2012).
The most common joint intersections are those of Ts, Ys, and Xs; immature flows contain more
T and X junctions – thus quadrilateral polygons. More mature flows contain more Y junctions –
thus increasing the sides on the polygons (hexagonal being the maximum if all the Ys were
perfect 120°) (Gray et al. 1976, Aydin and DeGraaf 1988). This joint evolution can be modeled
by Voronoi tessellation around anticlustered centers (space-filling polygons with distant
centers), crack propogration, and thermal gradients to show that any initial crack pattern will
mature into a quasi-hexagonal pattern (Budkewitsch and Robin 1994). Although approaching
perfect hexagons, the patterns do not mature enough to reach this stage (with average number
of sides <6). The polygon column cross-sections also have smaller areas than regular polygons
at all side numbers (aka their centroids are not the Voronoi center) (Figure 4). Based on
numerical modeling, it is suggested that joint evolution slows down after a certain point and an
increasing large number of cracks are needed to form a more regular hexagons, reaching a
natural limit based on cooling rates (Budkewitsch and Robin 1994). It has also been suggested
that the jointing process isn’t in equilibrium and inertia, heterogeneities, or environmental

  20
Caltech Enrichment Trip Iceland

constraints (such as cooling rate) prevent a perfect hexagonal network from forming (Hetényi et
al. 2012).
Understanding the tiering pattern found in

6
6

N=

N=

N=
300
6 columnar basalt outcrops has also been of interest to

5
scientists. As discussed before, singly tiered lavas can

N=
6
Area of a column section A (cm²)

250 8 7 be up to 30m thick, normally any flows larger become


7 66
7 7
6 6 5 multi-tiered systems. These systems often consist of
6
7
777
6
6 6 6 an upper colonnade, entablature, lower colonnade and
200 66
66 6
77 6 6 6 7 5 sometimes a basal pillow zone or breccia at the bottom
7 7 66
66 6 55
5
8 666
6 66 5
555 N = 4 (Figure 5, Long and Wood 1986). Entablature is a zone
150 7 666 6 55 5
6
66
66 5
55
4 of irregularly oriented columnar joints, curved
6 6 6 55
5 5
6 5 555
55 fracturing, and small cube-jointing. It is suggested that
7
7 66666
100 7 66
6 6
6
555
5
5 5 the interaction of columnar joints and pseudopillow
5 4
666 5 5 55 fracture systems (with a master fracture and several
6 5 5 5
55 5
50
6
55 55 Ideal N-sided polygon smaller ones) creates the jointing seen in entablature
5 Best fit to Somosk
4
outcrop data
(Forbes et al 2014). Some systems only have two tiers
6 8 10 12
with colonnades that do not match up, and it is noted
Mean joint-length of a column L (cm) that the downward growing joints meet the lower
colonnade well below the middle of the tier. This rapid
Figure 4: Cross sectional area of a cooling of the upper portions compared to the lower
column versus the mean side length of portions could be due to the convection of water from
a column from one site. Each point is the surface through the columnar joints to aid in cooling
labeled by a number to indicate the (DeGraff and Aydin 1987). Several studies on the
number of sides on a column. Solid petrography of entablature show textural signs of
black curves show the theoretical curve quenching compared to the lower colonnade. In multi-
for different n-sided regular polygons, tiered columnar basalts around the world, evidence of
while dotted grey curves show the best paleo-river valleys, damming of paleo-drainages, high
fit curves to the data (anchored at the
rainfall/evidence of surface water, and associated
origin). (from Hetényi et al. 2012)
lacustrine or fluvial sediments suggest that surface
flooding of the cooling lava flow due to displaced drainages is the cause behind this multi-tiered
architecture (Long and Wood 1986, Lyle 2000, Forbes 2014). Most of the water enters along the
master fractures in the pseudopillow fracture system, and cube-jointing likely forms when more
water enters the lava (Forbes et al. 2014).
In addition to these large-scale features, scientists have noted several other surface
features on the faces and parting surfaces of Lava flow, single-unit (SUF) Lava flow, multiple-unit (MUF)

columns to help understand the details of the


~x10-100 m
~x1-x10 m

heat heat

processes forming columnar basalts. flow


heat
flow

Horizontal striations or banding are often


seen around the entire perimeter of the Upper breccia Upper breccia

columns, although sometimes they are (Upper colonnade)

vertically offset between faces (Figure 6a,c). Colonnade

The striations are on the order of cms to 10s Entablature


Basal breccia
of cms. Two different mechanisms have been
used to explain striation thickness: 1)
Striation height varies inversely with the (Lower) colonnade

thermal gradient developed during cooling


Basal breccia
(Grossenbacher and McDuffle 1995), and 2)
striation thickness correlates inversely with Figure 5: Diagram showing columnar basalt at
the velocity of the cooling front (Goehring et the outcrop in single unit flows/single tiers and in
multiple-unit flows. (from Hetényi et al. 2012).

  21
Caltech Enrichment Trip Iceland

al. 2009). The horizontal striations are interpreted to be a stepwise propagation of the polygonal
fractures (Ryan and Sammis 1978, Budkewitsch and Robin 1994). Often each striation contains
a smooth and a rough surface which is suggested to be formed by alternating brittle elastic and
non-elastic incremental failure as the crack propagates into hotter regions which are more
ductile but which then cool due to the new crack exposure (Ryan and Sammis 1978).
Within the horizontal striations, but rarely visible in the “rough” part, are crescent hackles
also called a plumose structure (e.g. Figure 6b), which commonly reverses direction from one
striation to the next. Originally
interpreted to be relicts of
rotational shear during thermal
contraction (Ryan and Sammis
1978), DeGraff and Aydin (1987)
reinterpreted them as having
formed during crack propogation,
starting at a point of weakness for
the fracture origin and radiating
away in the direction of
propagation. Thus these plumose
structures can also be used to aid
Figure 6: Interpretations of columnar joint side morphology. understand the direction of growth
Photo (a), planar sketch (b), and cross section sketch (c) in basalt columns.
show horizontal striations, each with one plumose structure. On the flat cross-sections
Dots indicate crack origination at each striation, arrows show of the columns, there is often an
propagation directions, and numbers indicate order of crack inscribed circle with relief of a few
formation. Scale is in inches. (from DeGraff and Aydin 1987) mm above or below the rest of the
parting surface often only a small
rim. Radiating hackles are seen within the circle ring is marked by an oxidizing weathering of external origin).
coming from a central point (Figure 7, Tanner
2013, Guy 2010). Tanner (2013) suggests that
these hackles are similar to the plumose
structures described earlier formed by tensile
stresses in the column and that the circle and
periphery are due to differences in tensile
strengths of the early crystallized outer column
and slower cooling interior of the column. Guy
(2010) doesn’t believe that these hackles come
from thermal stress due to their central starting
point (fractures could start on a corner) and
termination in a perfect circle. He instead thinks Figure 7: Example of inscribed circles with
that they are from a directional growth of minerals radiating hackles from Svartifoss columns.
guided by the geometry of solid fingers within the Watch for scale. (from Guy 2010)
sample as it cools. This is one piece of evidence
used to support his hypothesis that columns form
by constitutional supercooling instead of by thermal contraction. In this model, a slight
heterogeneity in the melt will cause the formation of non-planar conditions between the crystals
and the melt. Gilman (2009) suggested that early minerals with high melting points will
crystallize first slowly moving to crystals with low melting points. Guy (2010) thought basaltic
melts were fairly homogeneous in composition, but that increased H2O content of the melt as
solidification occurs could drive constitutional supercooling, based on rings of small bubble

  22
Caltech Enrichment Trip Iceland

circles on the rims of the columns. This


a b c new model for columnar jointing could
also explain multi-tiered flows and
entablature as being due to a higher
degree of supercooling with poorly

Strike of column axis


defined smaller thermal gradients in the
middle of these larger units.
Other columns have concentric
ring features of alternating dark and light
bands resulting from slightly different
proportions of the main minerals (e.g.
plagioclase versus olivine). In polished
d
cut slabs, one can see that these bands
e f
are pseudohexagonal on the rims
becoming more circular toward a central
point (Mattson et al. 2011, Guy et al.
2010). Cuts parallel to the column axis
show that these rings are part of larger
Figure 8: Example of the internal compositional rings elongate finger shapes of slight
and fingers found in basalt columns from photos and compositional difference within the basalts
sketches of the photos. Scale bar is b) 15mm, c) (Figure 8). These rings are used to
5mm, f) 52cm. (from Mattson et al. 2011) support the idea of constitutional
supercooling (Guy 2010). Another
explanation for this fingering is double a
Convective heat transfer in the air
diffusive convection in which thermal or Convective heat transfer (air)
compositional variations within the lava within column-bounding fractures
Conductive heat transfer
could result in instabilities forming basalt Solid
fingers which eventually develop into 3
layered structures with strong convection Partially crystallized material
in the center and fingers on the edges
Crack advance point
(Kantha 1981). Although similar to
Molten interior
constitutional supercooling where a
compositional difference (or increasing b Original level of lava surface
gas bubbles) should be seen from the
center to edge of a column, this Loading due to weight of solidified crust
additionally suggests differences should
be seen from the top to the bottom of a
columnar basalt flow. Testing these
hypotheses, Bosshard et al. (2012) found
that similar crystallization temperatures of
magnetic minerals, similar compositions Partially crystallized material

of plagioclase, and lack of downwarped Molten interior

material did not suggest convective Figure 9: Diagram of how crystallization and
motions or constitutional supercooling. pressure loading can force melt-migration in the
Instead a new model was proposed in interior of columns. resulting in viscous fingers in the
which steep isotherms are created inside partially crystallized columns. (from Mattson et al.
the columns when cooling on crack and 2011)
joint surfaces becomes locally dominate

  23
Caltech Enrichment Trip Iceland

over heat transfer to the air at the top of the flow. Crystallization of titanomagnetite will cause a
volume decrease (15% is observed) and sinking of the solidified upper portions forcing
convection on the interior of columns and flows (Figure 9, Mattson et al. 2011). Systematic
variations of plagioclase lath orientation/size across the column diameter and of the anisotropy
of magnetic susceptibility (measuring orientation of titanomagnetite and paramagnetic grains)
support the idea of vertical melt migration in the interior of the columns (Bosshard et al. 2012,
Almqvist et al. 2012).

References & Further Reading


Almqvist, B. S., Bosshard, S. A., Hirt, A. M., Mattsson, H. B., & Hetényi, G. (2012). Internal flow structures
in columnar jointed basalt from Hrepphólar, Iceland: II. Magnetic anisotropy and rock magnetic
properties. Bulletin of volcanology, 74(7), 1667-1681.
Aydin, A., & DeGraff, J. M. (1988). Evoluton of polygonal fracture patterns in lava
flows. Science, 239(4839), 471-476.
Bosshard, S. A., Mattsson, H. B., & Hetényi, G. (2012). Origin of internal flow structures in columnar-
jointed basalt from Hrepphólar, Iceland: I. Textural and geochemical characterization. Bulletin of
volcanology, 74(7), 1645-1666.
Budkewitsch, P., & Robin, P. Y. (1994). Modelling the evolution of columnar joints. Journal of Volcanology
and Geothermal Research, 59(3), 219-239.
DeGraff, J. M., & Aydin, A. (1987). Surface morphology of columnar joints and its significance to
mechanics and direction of joint growth. Geological Society of America Bulletin, 99(5), 605-617.
Forbes, A. E. S., Blake, S., & Tuffen, H. (2014). Entablature: fracture types and mechanisms. Bulletin of
Volcanology, 76(5), 1-13.
Gilman, J. J. (2009). Basalt columns: Large scale constitutional supercooling?.Journal of Volcanology and
Geothermal Research, 184(3), 347-350.
Goehring, L., Mahadevan, L., & Morris, S. W. (2009). Nonequilibrium scale selection mechanism for
columnar jointing. Proceedings of the National Academy of Sciences, 106(2), 387-392.
Gray, N. H., Anderson, J. B., Devine, J. D., & Kwasnik, J. M. (1976). Topological properties of random
crack networks. Journal of the international association for mathematical geology, 8(6), 617-626.
Grossenbacher, K. A., & McDuffie, S. M. (1995). Conductive cooling of lava: columnar joint diameter and
stria width as functions of cooling rate and thermal gradient. Journal of volcanology and
geothermal research, 69(1), 95-103.
Guy, B. (2010). Comments on “Basalt columns: Large scale constitutional supercooling? by John Gilman
(JVGR, 2009) and presentation of some new data [J. Volcanol. Geotherm. Res. 184 (2009), 347–
350]. Journal of Volcanology and Geothermal Research, 194(1), 69-73.
Hetényi, G., Taisne, B., Garel, F., Médard, É., Bosshard, S., & Mattsson, H. B. (2012). Scales of columnar
jointing in igneous rocks: field measurements and controlling factors. Bulletin of
volcanology, 74(2), 457-482.
Kantha, L. H. (1981). ‘Basalt fingers’–origin of columnar joints?. Geological Magazine, 118(03), 251-264.
Long, P. E., & Wood, B. J. (1986). Structures, textures, and cooling histories of Columbia River basalt
flows. Geological Society of America Bulletin, 97(9), 1144-1155.
Lyle, P. (2000). The eruption environment of multi-tiered columnar basalt lava flows. Journal of the
Geological Society, 157(4), 715-722.
Mattsson, H. B., Caricchi, L., Almqvist, B. S., Caddick, M. J., Bosshard, S. A., Hetényi, G., & Hirt, A. M.
(2011). Melt migration in basalt columns driven by crystallization-induced pressure
gradients. Nature communications, 2, 299.
Ryan, M. P., & Sammis, C. G. (1978). Cyclic fracture mechanisms in cooling basalt. Geological Society of
America Bulletin, 89(9), 1295-1308.
Tanner, L. H. (2013). Surface Morphology of Basalt Columns at Svartifoss, VatnajökulsÞjóðgarður,
Southern Iceland. Journal of Geological Research,2013.

  24
Caltech Enrichment Trip Iceland

Chronology and Characteristics of the Hverfjall Eruptive Fissure


Contributed by Bryan Riel
Summary
The Hverfjall tuff ring /
crater lies east of Lake Mývatn
and was created approximately
2500 years ago due to
magmatic activity along the
southern extension of the
Krafla fissure swarm. At that
period, water from the proto-
Lake Mývatn covered the vents
in the Hverfjall area. During the
initial phase of the eruption, hot
Figure 1: Cross-section of the Hverfjall tuff ring.
magma from a single vent
interacted with the water from
the shallow lake, leading to explosive phreatomagmatic fragmentation, which formed the tuff
ring. The tuff ring itself has a maximum relative height of about 150 m, and the interior crater
has a maximum diameter of 1040 m (Figure 1). The date of the Hverfjall eruptions has been
estimated to be 2500 BP via dating of tephra layers. By analyzing soil profiles acquired
throughout the Mývatn area, the Hverfjall eruptions can be chronologically sorted into three
stages: an initial Hverfjall fallout via phreatomagmatic eruptions, lava flow at the Jarðbaðshólar
vents, and base surge of magma fragments and fluids expanding outward from the Hverfjall
vents.

Geologic Setting
Postglacial volcanism in northern Iceland is concentrated along the Northern Volcanic
Zone (NVZ), which forms a northward continuation of southern Iceland’s Eastern Volcanic Zone.
The NVZ consists of five separate volcanic systems that are arranged en-echelon, and in which
the active volcanism and plate divergence in northern Iceland is currently confined. One of these
volcanic systems is the Krafla volcanic system (KVS), which consists of a central caldera and a
transecting fissure swarm. With a length of approximately 100 km and a width that varies from 5
to 8 km, the Krafla fissure swarm is oriented in a SSW-NNE direction. The swarm forms a set of
graben structures, of which the central graben stretches north of Hverfjall up through the Krafla
caldera. The composition of erupted magmas is predominantly tholeiitic, but more evolved
quartz tholeiites, basaltic andesites, andesite, dacites, and rhyolites occur within the Krafla
caldera.
At the beginning of the Younger Dryas period, about 11000 years ago, inland ice in the
Mývatn area began advancing northwards until its front became stagnant around the location of
the current Lake Mývatn. There, the front remained stagnant for hundreds of years, forming
glacio-fluvial deposits referred to as the Reykjahlíd moraines. Before the northward
advancement of ice, subaerial volcanism had already started on a small scale in the Mývatn
area, which would set the stage for the subsequent phreatomagmatic eruptions.

Phreatomagmatic Eruptions
Phreatomagmatic explosions result from interaction of magma of any chemical
composition with ground or surface water. If magma and water mix mechanically, the

  25
Caltech Enrichment Trip Iceland

temperature contrast between the hot and cool liquids results in formation of insulating vapor
films (“Leidenfrost effect”). The collapse of the vapor films in combination with an extremely fast
increase in surface area of the melt results in rapid heat transfer from the magma to the water.
The water then gets superheated and transformed into steam. Since the heat transfer from
magma to water happens at a faster rate than the vaporization, the water vaporizes completely
and homogeneously, which results in an explosive expansion of highly pressurized steam to
ambient pressure. Explosive interactions of large volumes of melt and water (in scales of cubic
meters) may escalate to thermal detonations resulting in large-scale explosions.
Generally, only a portion of the magma melt interacts with the water for rapid heat
transfer. Most of the melt is passively ejected by and after the thermal explosion. The amount of
“interactive” melt is proportional to the intensity of the explosion. Ejected particles participating in
the rapid heat transfer are often characterized by angular to subrounded shapes.

Lake Mývatn and the Phreatomagmatic


Hverfjall Eruption
The area surrounding the Hverfjall
tuff ring shows extensive fracturing by
normal faults and numerous open
fissures. Combined with previous studies
suggesting a larger extent of the proto-
Lake Mývatn, it is likely that the lake
covered the area around the main
Hverfjall vent and thus promoted
phreatomagmatic fragmentation. The
presence of diatoms (phytoplankton)
mixed with the initial Hverfjall fall deposits
and the absence of tree molds within a
radius of 2 km from the vent support this
interpretation. Magma from the central
magma chamber beneath the Krafla
caldera to the north migrated laterally to
the south. Prior to reaching the Hverfjall
vents, the magma passed through the
Jarðbaðshólar system and dried up the
shallow groundwater table, inhibiting
phreatomagmatic activity there.
The age of the Hverfjall eruption
has been estimated using
tephrochronology to approximately 2500
BP. This estimate is based on a light-
colored tephra layer, called the Hekla-3
layer, beneat the Hverfjall deposits. The
Hekla-3 layer was formed circa 1000 B.C.
from a massive eruption of the Hekla
Figure 2: Map showing the deposits of the Hverfjall stratovolcano in southern Iceland. The
fissure eruption together with the main faults. Hekla-3 and Hverfjall layers are separated
Contouring interval is 20 m. The tentative eruptive by 1-3 cm of soil, and 14C dating of the
fissure is drawn as a dashed line.
older Hekla-3 layer yielded an age of
approximately 2800 years. The volcanic

  26
Caltech Enrichment Trip Iceland

episode in which the tuff ring formed is referred to as the Hverfjall Fires. The Hverfjall Fires
comprise eruptions from several vents within the Krafla fissure swarm, which formed both
pyroclastic deposits and emplaced lava flows. The vents that were active during this volcanic
episode are confined to the east by the Grjótagjá fault and to the west by the Beinihryggur and
Krummaskard faults (Figure 2), forming a 1-1.5 km wide graben-like structure. The amount of
vertical offset in this graben ranges from a few tens of centimeters to several meters with the
largest offsets in the northern part west of Námafjall.

Figure 3: Photographs illustrating the main characteristics of deposits


generated during the three stages of the eruption. (Left) Basal sequence
of laterally continuous, planar-stratified, deposits from the Hverfjall fallout
underlying the Jarðbaðshólar scoria deposits. (Right) Coarse-grained
pyroclastic deposits of the second stage overlayed by the clear bedding of
the stage III ash deposits.

Three-Stage Eruption
The Hverfjall fissure eruption can be subdivided into three main phases depending on vent
location and overall characteristics of deposits sampled around the Mývatn area (Figure 3):
I) Hverfjall fallout: During the first stage, a relatively low eruption plume deposited fallout,
forming relatively well-sorted, planar-stratified deposits. These deposits commonly
overlie between 1 and 3 cm of brownish soil and the rhyolitic Hekla-3 ash layer. The
thickness of the deposits decreases away from the Hverfjall vent. These deposits are
attributed to phreatomagmatic fragmentation due to sparse vesicles and the abundance
of smaller particles adhering to the outer surface of large particles. The eruption rate was
high and continuous in this stage.
II) Jarðbaðshólar lava flows: The second stage can be characterized by Jarðbaðshólar
deposits, which represent activity at the northernmost part of the eruptive fissure and the
two partially overlapping scoria cones of Jarðbaðshólar. The activity at the
Jarðbaðshólar vents also produced lava flows, which were emplaced mainly towards the
west and covered an area extending to the shoreline of the present-day Lake Mývatn.
The eruption of the scoria cones produced coarse-grained, vesicular scoria deposits,

  27
Caltech Enrichment Trip Iceland

which overly the Hverfjall deposits at soil profiles obtained south of the cones. As new
vents at Jarðbaðshólar started to draw magma from the same reservoir as Hverfjall, the
eruption rate at Hverfjall dropped significantly, resulting in more extensive magma-water
interactions there.
III) Hverfjall base surges: The third stage of the eruption was dominated by base surges.
Stage III deposits display structures indicative of lateral transport, i.e. presence of
ripples, dunes, low-angle cross bedding, plastering against obstacles, etc. Surge
deposits are found as far away from the vent as Námafjall, more than 5 km away from
the Hverfjall vent and approximately 100 m uphill. At this location, the total thickness of
surge deposits are less than 20 cm, but still preserve clear flow-structures. Closer to the
vent, the surge deposits are thicker and are indicative of a high-energy depositional
environment. Additionally, the deposits closer to the vent show evidence of greater water
concentrations with systematic “drying up” with distance away from the vent (stronger
sorting of deposits with distance). Thus, wet surges were most likely channelized inside
preexisting graben structures while drier surges were able to escape the graben and
travel farther distances.
 
References & Further Reading

Mattsson, H.B. and Höskuldsson, Á. (2011), Contemporaneous phreatomagmatic and effusive


activity along the Hverfjall eruptive fissure, north Iceland: Eruption chronology and resulting
deposits, Journal of Volcanology and Geothermal Research, 201, p. 241-252.

Saemundsson, K. (1991). Geology of the Krafla system. Nattura Mývatns, Hid Islenska
Natturufraedifelag, Reykjavik, 25-95.

Thorarinsson, S. (1979), The Postglacial history of the Mývatn area, Oikos, 32, p. 16-28.

Zimanowski, B., Frölich, G., and Lorenz, V. (1991), Quantitative experiments on


phreatomagmatic explosions, Journal of Volcanology and Geothermal Research, 48, p. 342-
358ˆ.

  28
Caltech Enrichment Trip Iceland

Geothermal Power and Hydrothermal Systems


Contributed by Kirsten Siebach
Overview
Iceland’s unique geological origin has created a remarkably high geothermal gradient,
enabling Iceland to derive about 65% of their total national energy consumption from geothermal
heat [1]. For comparison, the worldwide average geothermal gradient is 25°C/km [2], whereas in
Iceland there are at least 20 areas where water temperatures reach 200°C within a kilometer of
the surface (high-temperature zones created by near-surface magma), and over 250 areas have
water temperatures approaching 150°C at the same depths (low-temperature geothermal zones
created by flow through high geothermal gradients) (Figure 1) [3, 4]. This geothermal heat is
converted to electricity in geothermal power plants as well as being used directly to provide hot
water and heat buildings (in 2011, about 90% of homes were heated by direct geothermal heat)
[6]. Additionally, the geothermal energy is important for heating ~90% of Iceland’s popular open-
air swimming pools and creating the conditions for natural hot spring recreation areas
throughout the island [6]. Geothermal heat is therefore a critical resource for this small island
nation.

Figure 1. Map of known geothermal fields in Iceland, plotted with bedrock ages. High temperature
fields are defined by subsurface temperatures in excess of 200°C within the top kilometer under the
surface, whereas low temperature fields have temperatures approaching 150°C. Image from [3].

  29
Caltech Enrichment Trip Iceland

Geologic Setting and Source of Heat


Iceland is a geologically young island, estimated to be less than 20 million years old, and
it sits on top of both the Mid-Atlantic spreading ridge and a mantle hotspot. This unique
geological setting means that there is abundant heat trapped in the crust below Iceland, creating
an estimated energy current of approximately 30 GW (24 GW carried in magma and 6 GW from
heat conduction). As this energy approaches the surface, 7 GW are dispersed through volcanic
activity (there is an average of one eruption every four years), 8 GW are dispersed in water and
steam energy, and 15 GW are conducted to surface. Of this energy, about 7 GW of energy are
harnessable by current techniques [3]. One of the reasons that this energy is so readily
available is because the relatively young rocks that make up Iceland have not been buried and
retain relatively high permeability, allowing water to easily flow through the rocks and absorb
heat [3]. The Icelandic National Energy Authority is focused on finding sustainable ways to
utilize this energy resource through both direct utilization and electricity production [6].

Geologic Implications
of Geothermal Heat
One of the
interesting uses of
Icelandic geothermal
heat is as a rare
geologic case where
enough variables can
be considered
“constant” to directly
test the results when
only one variable is
changed, in this case,
temperature. The rocks
in Iceland are relatively
similar basalts,
chemically and with
respect to porosity and
permeability, especially
within a given volcanic
field. Furthermore, the
chemistry of the
meteoric water that is Figure 2. Basaltic alteration zones defined in Iceland based on basalt
heated in most of the alteration within different geothermal temperature regimes. From [4].
geothermal systems is
relatively homogeneous. This means that (after some corrections for porosity and vein
abundance) the amount and type of chemical alteration of a given basaltic rock is directly
related to the temperature of rock alteration, which in Iceland is well-mapped and includes a
significant range of temperatures due to the high geothermal gradient [4]. Iceland has therefore
been used as the type location for determining basalt chemical weathering and alteration
patterns with temperature (Figure 2) [4]. Broadly speaking, basaltic rock alteration proceeds
through alteration of the glass and minerals within the rock (typically in the opposite order from
mineral crystallization) to form clays, until temperatures above about 250°C are reached and
metamorphism occurs, forming amphiboles (Figure 2) [4]. Both modern and ancient Icelandic
geothermal zones have been used to begin to constrain the amount of alteration that occurs at

  30
Caltech Enrichment Trip Iceland

given temperatures, enabling plots like Figure 2 [4]. This has implications for reconstructing past
climates or geothermal environments on both Earth and Mars [4, 5].

Geothermal Electricity Generation


In 2012, geothermal power
plants in Iceland generated 4,600
GWh of electricity (661 MW), which
represented 24.5% of the nation’s total
electricity production. Seven major
power plants produce this energy,
including the Hellisheiði Power Station
(303 MW capacity), Reykjanes Power
Station (150 MW capacity), Nesjavellir
Geothermal Power Station (120 MW
capacity), Svartsengi Power Station
(76.5 MW capacity), Krafla Power
Station (60 MW capacity), Bjarnaflag
Power Station (3.2 MW capacity), and
Húsavík Power Station (2 MW
capacity) [6]. Each of these
Figure 3. Schematic of a flash steam type of power plant. geothermal plants is designed to
Hot water (>180°C) is retrieved from the subsurface and maximize the potential of the specific
moved into a low-pressure “flash tank,” where it vaporizes geothermal resource while taking into
(in a flash), driving a turbine that produces the energy. account the needs of the local
The water is then condensed and returned to the community [7]. Most of the power
subsurface to minimize impact. Image from [9]. plants in Iceland are “flash steam”
power plants, which use a borehole
drilled deep (~2.5 km) into the subsurface to collect high-pressure water that is at >180°C, and
then move this water into a low-pressure tank, where it vaporizes rapidly, driving a turbine that
runs a generator. The water is the condensed and returned to the subsurface through an
injection well (Figure 3) [8].

Krafla Power Station


The Krafla flash steam power station was developed in northeast Iceland near Lake
Mývatn under the direction of the government to meet urgent power needs in the north of
Iceland in the 1970s. It is still run by the government-owned energy company Landsvirkjun
[9,10]. The first test boreholes were drilled in 1974 and these showed that the geothermal
potential was promising, but volcanic eruptions only 2 km away in 1975 threatened the power
plant development. However, despite the volcano and related seismic activity, the plant survived
and the first turbine started in August 1977. The plant began producing about 30 MW of energy
with the single steam turbine in February 1978, and by 1984 the seismic and volcanic activity
had died down [6,7]. In 1996, additional drilling commenced and a second steam turbine was
installed, producing an additional 30 MW of energy under the direction of the Mannvit energy
corporation. Mannvit reports that: “In total, 33 boreholes were drilled, including 17 high pressure
production wells and 5 low-pressure production wells. The plant uses 110kg/second of 7.7 bar
saturated high-pressure steam and 36 kg/sec of 2.2 bar saturated low-pressure steam and has
been in operation at 60MW since 1999” [7].

  31
Caltech Enrichment Trip Iceland

Direct Utilization of Geothermal


Energy
While geothermal power
plants produce about 25% of the
electricity in Iceland, over 50% of
the total primary energy usage in
Iceland is geothermal. The
difference is made up by direct
geothermal energy usage (Figure
4). In 2013, 47% of the
geothermal energy used in Iceland
was to heat buildings, and 89% of
all homes were heated with
geothermal energy, enabling
Iceland to have the lowest
household heating costs of all the
Nordic countries [3, 11]. Direct Figure 4. Plot showing how geothermal energy was used in
geothermal energy is also used to Iceland in 2013. Image from [6].
heat 138 of the 169 swimming
pools in Iceland, in addition to the popular recreational hot springs swimming areas like the Blue
Lagoon. Other uses of direct geothermal energy, which involves pumping hot water from the
subsurface directly to the heating site, include: greenhouse warming, warming of pools for fish
farming, heating the surface under roads, sidewalks, and parking areas to melt snow in the
winter, and some industrial uses, including drying fish in heated warehouses [3]. As indicated by
the variety of uses for geothermal heat, the utilization of this resource has significantly impacted
the economy and culture of Iceland.

Grjótagjá cave
The Grjótagjá cave is a small volcanic cave that was supposedly a hideout and bathing
site for the outlaw Jón Markússon in the early 18th century. It was a popular bathing site until
volcanic eruptions between 1975 and 1984 made the water too hot for bathing [12,13]. Today,
according to internet rumors, there is still a sign warning that the water is too hot for bathing
temperatures (regulation: bathing water is at or below 42°C), but the water is apparently around
43-46°C and so some locals and various tourists take a dip for short periods of time despite the
heat [13,14].

Hverir, Námafjall springs


At the base of the Námafjall volcano, there are hot springs called Hveraröndor Hverir
that produce several fumeroles, mud volcanoes, and mud pools. These sulfurous and very
active features attract tourists because of the variety of colorful minerals deposited and the
remarkable scale of the mud craters and springs [15].

Mývatn Nature Baths


The Mývatn Nature Baths are a popular tourist destination where a large (5000 m2) man-
made lagoon with a sand and gravel base is kept at a constant temperature of 48-50°C. The
water is supplied from a manmade hot spring, the National Power Company’s borehole from the
Bjarnarflag Power Station, where ~130°C water from a depth of 2.5 km is brought to the surface.
The area was developed into a bathing facility in 2004 and there are saunas, shower, hot tubs,
etc in addition to the lagoon [16].

  32
Caltech Enrichment Trip Iceland

References & Further Reading


[1] http://askjaenergy.org/iceland-introduction/iceland-energy-sector/
[2] http://www.ipcc.ch/pdf/supporting-material/proc-renewables-lubeck.pdf
[3] Björnsson, S. (Ed.) (2010) Geothermal Development and Research in Iceland. Orkustofnun,
Reykjavik, Iceland. Available online at http://www.nea.is/media/utgafa/GD_loka.pdf
[4] Franzson, H., Zierenberg, R., & Schiffman, P. (2008). Chemical transport in geothermal
systems in Iceland: evidence from hydrothermal alteration. Journal of Volcanology and
Geothermal Research, 173(3), 217-229.
[5] Ehlmann, B. L., Mustard, J. F., Murchie, S. L., Bibring, J. P., Meunier, A., Fraeman, A. A., &
Langevin, Y. (2011). Subsurface water and clay mineral formation during the early history of
Mars. Nature, 479(7371), 53-60.
[6] http://www.nea.is/geothermal/direct-utilization/
[7] http://www.mannvit.com/GeothermalEnergy/GeothermalPowerPlants/
[8] http://energyalmanac.ca.gov/renewables/geothermal/types.html
[9] http://www.landsvirkjun.com/company/powerstations/kraflapowerstation/
[10] http://askjaenergy.org/iceland-introduction/iceland-energy-sector/
[11] http://www.samorka.is/doc/1841/%C3%93lafur+G.+Fl%C3%B3venz,+afm%C3%A6lisdag
skr%C3%A1+hitaveitu,+sept+08.pdf
[12] http://en.wikipedia.org/wiki/Grj%C3%B3tagj%C3%A1
[13] http://travel.stackexchange.com/questions/25609/is-it-allowed-to- bathe-at-
grj%C3%B3tagj%C3%A1
[14] http://jdombstravels.com/bathing-grjotagj-thermal-spring/
[15] http://www.northiceland.is/whattoseedo/viewattraction/namafjall
[16] http://www.jardbodin.is/en/

  33
Caltech Enrichment Trip Iceland

Faults, Fissures, and Catastrophic Floods


Contributed by Natalia Solomatova
Húsavík-Flatey Fault
The Húsavík-Flatey fault (HFF), part of the Tjörnes
Fracture zone (TFZ), is a system of right-lateral transform
faults oriented NW. It runs 1-3 km north of the Flateyjarskagi
peninsula (passing between the peninsula and Flatey island)
and through Húsavík, the second largest town in North
Iceland with 2,300 people. The HFF bends at Húsavík,
resulting in two sag ponds which align with the surface fault Figure 1
traces. In the east, the HFF exhibits NE trending graben in
the Skjalfandi region (Bergerat et al., 2000). However, most of the HFF is located offshore. The
motion on the HFF is 40% of the total transform motion across the TFZ (Metzger et al., 2011)
and the slip rate on is estimated to be 5-7 mm/yr (Arnadottir et al., 2009). Although there have
been four large earthquakes on the HFF within the last 200 years, no major earthquakes on the
HFF have occurred since two M6.3 earthquakes in 1872 (close to Flatey and Húsavík). In 1755
and 1838, M7 and M6.5 earthquakes occurred in Sjalfandi bay, respectively. The HFF has a
significant dip-slip component, as indicated by a fault scarp of ~200 m in several locations and a
maximum fault scarp of 1400 m! The right-lateral displacement is 5-10 km (Gudmundsson et al.,
1993).

Tjörnes Fracture Zone


The Tjörnes Fracture zone
(TFZ) is mainly composed of the
Húsavík-Flatey fault (HFF),
Guðfinnugjá (a large normal fault)
and the volcanically-active
Grimsey Oblique Rift (GOR),
connecting the northern volcanic
zone (NVZ) with the North Atlantic
Kolbeinsey Ridge (KR). The TFZ
has a 120 km-offset between two
segments of the Mid-Atlantic
Ridge and it accommodates the
total spreading rate between the
North American plate and the
Eurasian plate of 18 mm/yr
(Metzger et al., 2013). The largest
earthquakes in Iceland occur on
the TFZ and more than 90% of the
earthquakes are shallow, occurring
Figure 2: The Tjörnes Fracture Zone
shallower than 10 km depth
(Einarsson, 2008).
The HFF meets the normal Guðfinnugjá fault at an angle of 60 degrees (Húsavík strikes
N55W while Guðfinnugjá strikes N5E) (Gudmundsson et al., 1993) while GOR runs offshore and
is characterized by a set of steep N-S-oriented en echelon faults. Both normal and strike-slip

  34
Caltech Enrichment Trip Iceland

faults occur on GOR with a geometry that can be described as bookshelf faulting (Metzger et al.,
2011).

Fissures and Fractures


The North Volcanic Zone (NVZ) is
50 km wide and 200 km long, consisting
of a number of volcanoes and fissure
swarms. The orientation of the fractures is
roughly perpendicular to the spreading
direction, indicating that plate spreading
affects their orientation. The density of
eruptive fissures and fractures is greatest
20-30 km from the central volcano, and
depends on the age of the lava flow
(oldest lava has the most fissures). The
density of fractures is lowest near eruptive
fissures due to the fresh lava and highest
at the junction of transform zones and
fissure swarms (for example, at the
junction of HFF and the
Theistareykir/Krafla fissure swarms).
The mechanical and physical
junction of Tjörnes Fracture Zone (TFZ),
the Grimsoy Oblique Rift (GOR), Húsavík
Flatoy Fault (HFF) and the Guðfinnugjá
fault has resulted in the formation of the
shield volcano, Theistareykjarbunga.
Theistareykjarbunga erupted 12,000,
9,000 and 2,500 years ago and although
there hasn’t been any recent volcanic
activity, it uplifted ~3 cm between the
summer of 2007 to 2008 due to inflation of
the magma reservoir located at a depth of
8.5 km (Metzger et al., 2011). The most
recent eruption 2,500 years ago formed
the ‘Theistareykjahraun’ lava flow
between Theistareykjarbunga and HFF.
There have been no eruptive fissures near
Theistareykir, although the density of
fractures is very high. Additionally, the
orientation of fractures is anomalous at
the junction of the Theistareykir fissure
swarm and the Húsavík Transform Zone
Figure 3: The North Volcanic Zone. Source: Hjartardottir
(they bend from N-S orientation to NW-SE (2013).
orientation) (Hjartardottir, 2013).

Åsbyrgi
Åsbyrgi is a horseshoe-shaped canyon that is 3.5 km long and 1.1 km wide. The vertical
walls of Åsbyrgi reveal 25 stacked lava flows, each 1-3 meters thick (the top two have pahoehoe

  35
Caltech Enrichment Trip Iceland

Figure 4. Source: Alho et al. (2005).

texture). It was formed as a result of two catastrophic glacial floods (as well as 16 moderate
floods) of the Jökulsá á Fjöllum river, one 8,000-10,000 years ago and the other 2,000-3,000
years ago due to the eruption of a volcano underneath the Vatnajökull ice cap. Vatnajökull is the
largest glacier in Europe with an area of 8,100 km2 and thickness of 400-600 m, covering
volcanoes Bárðarbunga and Grímsvötn and currently feeding glacial rivers that actively create
today's waterfalls and canyons (Waitt, 2009). Interestingly, Norse mythology explains the shape
of the Åsbyrgi canyon as a result of Odin’s eight-legged horse, Sleipnir, touching one of his feet
to the ground. Legends also say that hidden people called Huldufolk (similar to elves) live there.

Dettifoss
Dettifoss, located on the Jökulsá á Fjöllum river, is the largest waterfall in Europe by
volume with an average flow of 193 m3/s. The waterfall is 100 meters wide and 45 meters tall.

Catastrophic Floods
Glacial outburst floods (jokulhlaups) have resulted in scablands and canyons throughout
Iceland due to the failure of glacier dams following volcanic eruptions under the ice, earthquakes
and when large pieces of glacier break off. The Vatnajökull icecap feeds the Jökulsá á Fjöllum
river, which is the largest glacial river in Iceland, flowing northward into the sea. The most
catastrophic flood during the Holocene, which contributed to the Åsbyrgi formation, probably
occurred ~2,500 years ago with a peak flow of 0.7-0.9 million m3/s. In comparison to other
floods, the Altai flood had a peak discharge of ~18x106 m3/s with 20 times a higher power per

  36
Caltech Enrichment Trip Iceland

area while the Lake Missoula flood had 10 times higher discharge with 2 times higher power per
area. The most recent great flooding event occurred in November 1996, releasing 3.8 km3 of
water across Skeijararsandur within 48 hours (with a peak flow of ~50,000 m3/s) (Alho et al.,
2005; Waitt, 2009).

References & Further Reading


Alho, P., Russell, A. J., Carrivick, J. L., & Käyhkö, J. (2005). Reconstruction of the largest
Holocene jökulhlaup within Jökulsá á Fjöllum, NE Iceland. Quaternary Science Reviews,
24(22), 2319-2334.
Árnadóttir, T., Lund, B., Jiang, W., Geirsson, H., Björnsson, H., Einarsson, P., & Sigurdsson, T.
(2009). Glacial rebound and plate spreading: results from the first countrywide GPS
observations in Iceland. Geophysical Journal International, 177(2), 691-716.
Bergerat, F., Angelier, J., & Homberg, C. (2000). Tectonic analysis of the Húsavík‐Flatey Fault
(northern Iceland) and mechanisms of an oceanic transform zone, the Tjörnes Fracture
Zone. Tectonics, 19(6), 1161-1177.
Einarsson, P. (2008). Plate boundaries, rifts and transforms in Iceland. Jökull,58, 35-58.
Gudmundsson, A., S. Brynjolfsson and M. Th. Jonsson, 1993. Structural analysis of a transform
fault-rift zone junction in North Iceland. Tectonophysics, 220, 205-221.
Hjartardottir, A.R. (2013). Fissure swarms of the Northern Volcanic Rift Zone, Iceland. (Doctoral
dissertation). Faculty of Earth Sciences, University of Iceland.
Metzger, S., Jónsson, S., & Geirsson, H. (2011). Locking depth and slip-rate of the Húsavík
Flatey fault, North Iceland, derived from continuous GPS data 2006-2010. Geophysical
Journal International, 187(2), 564-576.
Metzger, S., Jónsson, S., Danielsen, G., Hreinsdóttir, S., Jouanne, F., Giardini, D., & Villemin, T.
(2013). Present kinematics of the Tjörnes Fracture Zone, North Iceland, from campaign
and continuous GPS measurements.Geophysical Journal International, 192(2), 441-455.
Waitt, R. B. (2009). Great Holocene floods along Jökulsá á Fjöllum, north Iceland. Flood and
Megaflood Processes and Deposits: Recent and Ancient Examples, International
Association of Sedimentologists special publication, 32, 37-51.

Websites
http://en.wikipedia.org/wiki/%C3%81sbyrgi
http://en.wikipedia.org/wiki/Dettifoss
http://www.nat.is/travelguideeng/plofin_tjornes_more.htm
http://www.norvol.hi.is/html_i/geol_i/trip4_i.html
http://www.vatnajokulsthjodgardur.is/english/education/asbyrgi/
http://www.visithusavik.com/attractions/asbyrgi-canyon/

  37
Caltech Enrichment Trip Iceland

Askja Volcano and Surrounding Area


Contributed by Stephen Perry
Askja is a stratovolcano located in the central highlands of Iceland. The name “Askja”
actually refers to a nested caldera system situated on top of the volcano – askja meaning box or
caldera in Icelandic. Askja was not a notable volcano until it erupted violently in 1875. This
event created the smaller caldera, which has since filled with water and now forms Öskjuvatn
Lake. You can see the two calderas in the aerial photo shown in Figure 1. The outer caldera
was formed in an event about 11,000 years ago, and the darker, inner region is Öskjuvatn Lake.
Another notable feature of Askja is its Viti crater located on the northeast shore of Öskjuvatn.
This 150 meter wide crater, also formed in the eruption of 1875, is filled with a geothermal lake
of mineral-rich, sulphurous water.3

Figure 1: Nested Askja Caldera System - The larger, outer caldera was formed in a prehistoric
eruption and the inner, water-filled caldera was formed in the large 1875 eruption.

Askja is most famous for it’s enormous volcanic eruption, which occurred on March 29,
1875. It was part of a series of regional volcanic and tectonic events that took place in the
northern rift zone in 1874 and 1875. This explosion produced about 3 km3 of ryolitic tephra.5 In
fact, the ashfall was thick enough that it covered easthern Iceland, killing livestock and choking
out crops, and causing a significant emigration for the following couple of years.3 Some of the
particulate matter was even blown all the way to parts of Sweden and Norway. The map in
Figure 2 shows the extent of the ejecta from the 1875 eruption.5
During the 1920s, four small eruptions produced basalt lava in the caldera. The most
recent eruption was in 1961, and was of the Hawaiian type, with lava fountains reaching up to
500m high. The eruption lasted about 5 weeks and produced aa lava during the first week,
followed by pahoehoe for the remaining 4 weeks.4 The two lava types differ in their mineralogy:
the first characterized by relatively few phenocrysts of plagioclase and pyroxene and with no
olivine present.4 The second type contains relatively more phenocrysts as well as olivine.

  38
Caltech Enrichment Trip Iceland

Figure 2: Left - Askja and the surrounding region. The dotted lines show the extent of the 1875
ejecta. Right- Lava flows around the Askja caldera labeled by their matching eruption dates.

This caldera is inherently an unstable region, and further illustrating this point was a
massive landslide on July 21st, 2014 (about 1 month before our scheduled arrival at the site).1
The collapsed region is about 1 km wide and flowed directly down into Öskjuvatn.2 This created
a tsunami, or seiche wave (so-called because it was a standing wave inside the fully bounded
lake), inside the lake that reached about 50 meters in height; high enough to overflow the Viti
crater.2 The total displacement raised the water level in the lake by about 2 meters after
everything had settled. As of the end of July, the hiking trails surrounding the Askja caldera
have been closed until that the area is confirmed to be safe.

References & Further Reading

1. http://www.wired.com/2014/07/massive-landslide-inside-the-askja-caldera/
2. http://www.jonfr.com/volcano/?p=4633
3. http://en.wikipedia.org/wiki/Askja
4. Thorarinsson, S., and G. E. Sigvaldason. "The Eruption in Askja, 1961; a Preliminary
Report." American Journal of Science 260.9 (1962): 641-51. Web.
5. Sparks, R. S. J., L. Wilson, and H. Sigurdsson. "The Pyroclastic Deposits of the 1875
Eruption of Askja, Iceland." Philosophical Transactions of the Royal Society A: Mathematical,
Physical and Engineering Sciences 299.1447 (1981): 241-73. Web.
 

  39
Caltech Enrichment Trip Iceland

Eastern Iceland: Geology, Culture, and Industry


Contributed by Josh Kammer
General Geology
Basaltic lavas predominate the geology of eastern Iceland, most noticeably forming
impressively steep fjords along the coastline. Nearly 15,000 feet of mostly basalt lavas can be
found in the Reyðarfjörður area in the middle of a large Tertiary outcrop. Much of the lava was
created from eruptions on dry land, and many can be found in hundreds of dike formations in
the region. Olivine basalts, porphyritic basalts, and tholeiites make up most of the lava piles in
the area. Periods of acid volcanism have also shaped the geology of the region, forming acid
and intermediate lavas and pyroclastic rocks, much of which were sourced from the Thingmuli
central volcano. Zeolites can be found around Teigarhorn, and many colorful examples of
jasper, onyx, opal, agate, and amethyst are also prevalent in the region. Reputedly the world's
largest private collection of rocks can be found in Stöðvarfjörður, in the gardens of local collector
Petra Sveinsdottir, who has been collecting them since 1946.

Culture
Eastern Iceland is full of history, with museums, ruins, or both scattered in almost every
fjord, farm, or village. The Djúpivogur historical center provides a look at the earliest arrivals by
the Greeks and Romans, as well as Irish monks. Roman coins dating back to the first century
AD were found near here. Later in about the 8th century, however, the Vikings arrived, and much
of Iceland's history has been dominated by their influence. Much of the local culture has been
influenced by these settlers, too. Eastern Iceland folklore is full of tales of the supernatural,
ranging from a great wyrm that lives in the lake Lagarfljót, to many stories of trolls and elves.
Current and antique pieces of art often reflect the stories of eastern Iceland's past, both
historical and mythological. Modern day shops often still carry quality handcrafted art ranging
from paintings and ceramics, to reindeer leather and antlers, and carvings from wood. Many
examples of modern work can be found at the Slaughterhouse Culture Center in Egilsstadir.

Industry
Historically, eastern Iceland prospered from the sea – both herring fishing and the
whaling industry dominated during the 19th century, and for a time the world's largest whaling
station was based in the region. International communications have been important here ever
since the first telegraph cable connecting Iceland to mainland Europe was installed in 1906. In
2007, an aluminum smelter has been operating in the region, producing nearly 1000 tons of
aluminum per day; the Kárahnjúkar hydroelectric power plant was completed in 2009 to provide
4600 Gwh annually to the smelter. Tourism also continues to be a significant factor for job
growth. Fishing is still the main source of income, however, but agriculture is also prevalent in
the area, especially based on sheep but also involving barley and other grain products.

References & Further Reading


Walker, G. P. L., 1957, “Geology of the Reydarfjordur area, eastern Iceland”, Quarterly Journal
of the Geological Society, 144, 367
Benedikz, B. S., 1973, “Basic Themes in Icelandic Folklore”, Folklore, 84, 1
East Iceland: The Official Tourist Guide (http://en.east.is/)

  40
Caltech Enrichment Trip Iceland

Icelandic Mineralogy and Geochemistry


Contributed by Jen Caseres
Volcanic rocks cover about 8% of the world’s surface, and about half of that is basalt.
Areas like central Iceland, which has a lot of exposed basalt and few inhabitants, are ideal to
study the chemical weathering of Ca-Mg silicates. The weathering of Ca-Mg silicates is
important for studying the CO2 budget, and the weathering of basalts is important to
understanding early Earth environments (Gisalson & Armannsson 1994).
Mars, like early Earth, is covered in volcanic rocks. The signature red dust that coats the
surface of Mars is likely to be an alteration product of basalt, not rhyolite. The palagonitization of
basalt has been suggested as a means to produce the red soils of Mars (Bishop et al. 2002).
Palagonite is the product of alteration of basalt by water. Palagonitized tuff is very common on
Iceland as a subglacial deposit. It is found as a resin-like red orange coating surrounding
basaltic glass (Stroncik & Schmincke 2002).  
The large amount of palagonite-like dust of Mars suggests that it may be widely altered,
and its presence is used to argue that large amounts of water were present on Mars, although
the relative lack of clay minerals suggests that water-rock interactions may have been limited
(Michalski et al. 2005).
Another famous alteration product is found in eastern Iceland. Zeolites are tectosilicate

Figure 1: From Neuhoff et al., 1999

  41
Caltech Enrichment Trip Iceland

minerals with hydrated


aluminosilicate frameworks
that are loosely bonded to
alkali and alkali earth
cations. The water
molecules and large cations
occupy the large channels
and cages present in all
zeolite structures. Zeolites
are a highly variable group
of minerals, as up to 50% of
silicon and aluminum can be
replaced by phosphorus and
beryllium. Exchange of
cations and dehydration can
be induced at relatively low
temperatures (below 100˚C
for exchange and between
250˚C and 400˚C for
dehydration) (Tschernich
1994). Once dehydrated,
zeolites will absorb water at
room temperature. This
property is extensively
utilized as an industrial
desiccant.
Zeolites are formed
in a range of environments,
from dry lakebeds to cooling

Figure 2: From Neuhoff et al., 1999 basalt flows. They are also
formed as a result of low-
grade metamorphism. In
Iceland, they are formed as the result of regional hydrothermal metamorphism and burial
metamorphism (Tschernich 1994, Neuhoff et al. 1999). Often, basalt flows in Iceland are subject
to multiple stages and types of alteration. At Teigarhorn, coastal erosion has exposed famous
zeolite rich outcrops produced by a multistage process.
If you examine the amygdules at Teigarhorn, you may see green and white rims around
the edges. This is the first alteration product, celadonite (a green layer silicate) and quartz. No
zeolites have formed at this stage yet. As the flow is buried by another basalt flow, chlorite and
smectite forms, then zeolites precipitate. The type of zeolite produced is very sensitive to depth,
and the zeolites here can be used a barometer (Neuhoff 1999).
The large, euhedral zeolites for which Teigarhorn is famous were formed by
hydrothermal alteration caused by the Álftafjörður mafic dike swarm after burial. The most
common large crystals found here are stilbite, heulandite, mordenite, scolecite and laumontite.
Iceland spar calcite and euhedral quartz crystals may also be found (Neuhoff et al. 1999).
Zeolites formed by these dikes are found throughout most of the area where the dikes are, in a
swath from Álftafjörður to Breiðdalsvík. Exposures along the coast of Berufjörður are particularly
rich in zeolites (Lesiow 2008).

  42
Caltech Enrichment Trip Iceland

Figure 3: Heulandite-Ca, Teigarhorn, Berufjörður, Figure 4: Scolecite and stilbite. Teigarhorn,


Suður-Múlasýsla, Eastern Region, Iceland. This is an Berufjörður, Suður-Múlasýsla, Eastern Region,
important old specimen probably at least 125 years Iceland. “A 6.3 x 5.2 x 4.3 cm cluster of prismatic
old, if not more. Before India, these Icelandic zeolites colorless translucent scolecite crystals measuring
set the standard and were considered both superb to 3.6 cm in size. Some stilbite crystals to 1.9 cm
specimens and a mark of the sophisticated collector. are present. Excellent for the species from this
With lustrous doubly-terminated crystals to 7.5 cm, locality! Ex Steve Pullman collection with a
isolated on basalt matrix, this is still a superb previous label from the W.J. Rodekohr
heulandite even compared to modern Indian material collection.”
and yet it is priced about the same, although its a http://www.mindat.org/photo-337003.html
major rarity and classic. 21 x 12.1 x 5 cm
http://www.mindat.org/photo-53065.html

Zeolites are often quite distinct in the field, although they are almost impossible to
distinguish from one another. Their crystal habit, which ranges from fans and blades to fine
needles to blocky, is quite distinct from quartz and calcite. But, each mineral species may be a
variety of colors and forms, and the overlapping physical properties make it difficult to tell which
zeolite is which without further analysis.

References and Further Reading


Bishop, J. L., Schiffman, P., & Southard, R. (2002). Geochemical and mineralogical analyses of
palagonitic tufts and altered rinds of pillow basalts in Iceland and applications to Mars. In J. L. Smellie &
M. G. Chapman (Eds.), Volcano-Ice Interaction on Mars (pp. 371–392). Geological Society of London.
Gislason, S. R., & Armannsson, H. (1994). Present chemical weathering of basalt in Iceland. Mineralog.
Mag. Goldschmidt Conf. A, 333–334.
Lesiow, D. (2008). Iceland’s zeolites. http://www.mindat.org/article.php/383/Iceland's+zeolites. Web.
Michalski, J., Kraft, M., Sharp, T., & Christensen, P. (2005). Palagonite-like alteration products on the
Earth and Mars I: Spectroscopy(0.4-25 microns) of weathered basalts and silicate alteration products.
36th Annual Lunar …. Retrieved from http://adsabs.harvard.edu/abs/2005LPI....36.1188M
Neuhoff, P. S., Fridriksson, T., Arnorsson, S., & Bird, D. K. (1999). Porosity evolution and mineral
paragenesis during low-grade metamorphism of basaltic lavas at Teigarhorn, Eastern Iceland. American
Journal of Science, 299, 467–501.
Sigmarsson, O., & Steinthórsson, S. (2007). Origin of Icelandic basalts: A review of their petrology and
geochemistry. Journal of Geodynamics, 43(1), 87–100. doi:10.1016/j.jog.2006.09.016
Stroncik, N. a., & Schmincke, H.-U. (2002). Palagonite – a review. International Journal of Earth Sciences,
91(4), 680–697. doi:10.1007/s00531-001-0238-7
Tschernich, R. W. (1994). Zeolites of the World.

  43
Caltech Enrichment Trip Iceland

Glaciers and Jökulhaups


Contributed by Jennifer Buz
Glaciers
After fresh snow is
deposited on the ground it
is subjected to either
sublimation or melting.
These processes cause
the porous snow crystals
(porosity ~ 95%) to
become more compacted
and rounded, creating old
snow (porosity ~ 50%).
The densification of snow
continues as more snow is
added to the pile.
Eventually the old snow
becomes firn, from the
German word meaning
“last year snow”. Individual
crystals, up to 2mm in
diameter, are created
through intergranular
movements, plastic
deformation, crushing, and
recrystallization. Although
firn has a low porosity, it is
still permeable. Firn
eventually leads to new Figure 1: Figure from Martini, Brookfield, Sadura textbook.
glacier ice, which is
relatively impermeable and
contains even larger crystals. As these large crystals deform within a glacier due to the weight
of the glacier itself, temperature variations, and external forces they become old glacier ice. The
transformation of new snow to old glacier ice is illustrated in Figure 1a.
Glaciers are defined as bodies of ice capable of moving under their own weight. Glaciers
occur in areas where there is a net accumulation of snow in the average year (i.e. more snow is
deposited than is melted/sublimated away). The accumulation zone of a glacier typically extends
from the head of the glacier to the firn line. The ablation zone represents the region of the
glacier where more material is lost than gained in a given year. The exception to this is
superimposed ice which is formed from re-frozen melt water from the accumulation zone. A
schematic cross section showing the stratigraphy of a glacier is shown in Figure 1b. Many types
of glaciers exist; they are typically classified based on their extent, location, and shape.
Glaciers move to achieve a gravitational equilibrium as material is moved from the
accumulation zone to the ablation zone. The speed at which the glacier moves depends on the
local environment. Glaciers which incorporate more rock typically move slower than pure ice
glaciers. Additionally, glaciers increase their speed during the winter as the accumulation
increases. Some glaciers may surge, moving up to 40 m/day, while others may be nearer 0.3-

  44
Caltech Enrichment Trip Iceland

Figure 2: Figure from Benn and Evans textbook.

0.5 cm/day (Martini, 2001). Friction at the walls and floor bounding a glacier slow the movement
of the ice in these regions causing a parabolic velocity profile. During a surge the ice near the
edges of the glacier may shear, causing a more box-like velocity profile. There are four main
hypotheses for how glaciers move: rotation of ice crystals past one another, movement of water
down slope in the pore spaces, slippage of the glacier over rock, and internal shearing in the
glacier.

  45
Caltech Enrichment Trip Iceland

Glaciers cause many unique landscape features both large and small. For example,
striae, are linear features formed when rocks embedded in the glacier bottom press and scratch
against the bedrock below them. Examples of striae are found in Figure 2 (Benn and Evans,
1998). Medium scale erosional features include m-scale grooves carved by large boulders
gouging out softer material below. Roche moutonnées are elongate hills (1-50 m in height, 10-
1000 m long) where the up-flow end is smoothed via abrasion, while the down-flow end is
jagged from plucking by the glacier. Crag and tail features are found where a resistant bedrock
knob shields material leeward from it. Sometimes semilunate troughs can form around the crag.
Drumlins are streamlined hills with a tapered end leeward. They are usually composed of till.
The origin of drumlins is currently under debate. Flutes are elongate streamlined ridges (cm to
m high and wide) or clusters of ridges aligned parallel to glacial flow, they are believed to be the
result of subglacial sediment deformation down flow of an obstruction. Moraines are
accumulations of glacial debris formed under actively moving parts of glaciers and at glacial
margins. Many types of moraines exist depending on their location (inside or outside) relative to
the glacier. Examples include ground, end, interlobate, push, and lateral moraines. Eskers are
elongate ridges of glacial sand and gravel which are the infillings of ice-walled river channels.
Kames are hills of glacial material that were originally deposited in glacial depressions and
remain after the glaciers melted away. Kames and eskers are illustrated in Figure 3, which

Figure 3: Figure from Martini, Brookfield, Sadura textbook.

  46
Caltech Enrichment Trip Iceland

shows how the sediments on a glacier may be transposed onto the surface after the glacier
melts. Many more glacial features exist, too many to include in this field guide.

Jökulhaups (Glacial Outburst Floods)


Jökulhaups are catastrophic flood events that occur when a glacial lake suddenly drains.
Iceland is a very volcanically active region, and some of the volcanoes are overlain by glaciers.
Eruption of these volcanoes can cause the glaciers to melt and accumulate huge amounts of
water. Lakes are capable of forming beside glaciers, either dammed by ice or by moraines.
Lakes can also form below glaciers. In the case of moraine-dammed lakes, as soon as the
water level breaches the surface of the moraine, which may occur as the result of a rock fall for
example, rapid incision into the sediment ensues as moraines are easily eroded by the focused
water flow. Ice-dammed lakes and subglacial cupolas on the other hand, are usually breached
near the bed as the pressure condition changes between the water reservoir and the
surrounding ice (Benn and Evans, 1998). Nye (1976) has developed a mathematical model for
jökulhlaups initiated in subglacial lakes. When a lake reaches 90% of the thickness of the ice
dam, the drainage potential outside of the reservoir will be lower, water will then raise the dam
and travel along this potential gradient. In geothermal areas, the water in the lake will likely be
well above the melting point of ice. The movement of this water will therefore further melt the
channel in which it escapes, increasing the flow. Discharge observations are consistent with this
tunnel-widening hypothesis because discharges increase rapidly over the course of a few to
several days after which they drop substantially, likely representing the emptying of the lake or
creep closure (Björnsson, 1992). An alternative to the ice-dam flotation hypothesis has been
proposed by Fowler and Ng (1996). In this scenario, the catastrophic rupture occurs when the
subglacial hydrology switches from porewater flow to a distributed canal-system.

Jökulsárlón (Glacial River Lagoon)


Jökulsárlón is a proglacial lake at the
mouth of Breiðamerkurjökull, an outlet of the larger
Vatnajökull. Jökulsárlón may have formed recently;
it was first reported in 1932 (Todtmann, 1960). No
reports of lakes predate this initial account, though
the lack of report does not necessitate the lack of
lake. Two maps produced of the Vatnajökull region
in the 1700s indicate that the glacier drained
directly into the sea through a series of
anastamosing channels. Subsequent reports
indicated that the glacier advanced and retreated
throughout the 1800s, at one point being only 1 km
from the ocean (Thorarinsson, 1943). At present,
the lake has a 1 km long outlet to the Atlantic
ocean through the Jökulska river.
Soundings of the lake depth and ice
elevation revealed that a shelf of ice extends away
from the glacier into the water (Howarth and Price,
1969). Temperature measurements of Jökulsárlón
rarely rise above 1°C, whereas surface
temperatures exceed melting. The ice ablation
during the summer months averages 7.5 cm/day.
This suggests that ice removal above water is Figure 4: Figure from Howarth and Price.

  47
Caltech Enrichment Trip Iceland

faster than that below water, causing the shelf (Howarth and Price, 1969). Eventually the
density contrast between the ice and water becomes significant enough to raise the submerged
ice shelf, creating crevasses parallel to the ice-water interface and eventually leading to calving
(Howarth and Price, 1969). This process is illustrated in the Figure 4 (Howarth and Price, 1969).
The process by which proglacial lakes form is not known. Benches at the bottom of the
proglacial lake are proposed to have formed from deltaic deposits from streams at the surface or
under hydrostatic pressure (Howarth and Price, 1969). An alternative hypothesis is that the lake
basins formed gradually from downcutting and lateral migration of subglacial rivers (Jonsson,
1955).
The temperature and salinity profile of Jökulsárlón was studied at multiple times during
the year by Harris (1976). For most of the year, the salinity of the water increases sharply at the
lake surface and also has a subtle increase at the basin floor. The exception to this is during the
summer when melt water flushes the saline water from the surface but leaves the saline water
at the floor. A schematic diagram is shown in Figure 5 (Harris, 1977).

Figure 5: Figure from Harris, 1977.

Skaftafell Glacier
Detailed stratigraphic columns of the Skaftafell region were constructed with the use of
K-Ar dating and paleomagnetic techniques (Halgason and Duncan 2001). By correlating two
stratigraphic columns in the region, 16 glacial and interglacial intervals were accounted for since
5 Ma. Glacial intervals were marked by discontinuous pillow basalt ridges and hyalocastic
sedimentary rocks and breccias. Interglacial intervals consisted of continuous, subaerial lava
flows and thin volcaniclastic sediments. Halgason and Duncan (2001) hypothesize that the
topography seen in the region today originated along with the glacial interglacial periods of the
past 5 Ma. They construct the following chronology of events, also illustrated in Figure 6.

  48
Caltech Enrichment Trip Iceland

Preglacial period (> 4 Ma)- No striations


present or thinning of lava beds due to
erosion. The first striated surface appears ~
4 Ma. There is less than 10 m relief
Early landscape-forming period (3 – 4 Ma)-
Minor subglacial volcanic activity present. A
90-m thick lava flow is eroded away between
two cliff profiles, indicating relief ~100 Ma.
Main landscape-forming period (0.8 – 3 Ma)-
A subglacial volcanic ridge from 2.8 Ma is
found with a relief of 200 m, subsequent
volcanic ridges have reliefs up to 400 m.
This marks the beginning of the deepening
of the valleys and massive erosion of
underlying volcanics.
Present Period (0- 0.8 Ma)- The valleys
formed during the main landscape forming
period became so deep that the interglacial
period could no longer fill them with lavas.
As a result these valleys have become
permanent channels, each year further
deepening.The sharp elevation changes
contribute to the frequency and intensity of
glaciations in the region.

Skeiðarárjökull and Gígjukvísl


Skeiðarárjökull is an outlet glacier
from the larger Vatnajökull. Underlying
Vatnajökull is Grímsvötn, Iceland’s most
active volcano. The volcano and associated
hydrothermal field create a depression in
Skeiðarárjökull that is 10-km wide, below
which lies Iceland’s largest subglacial lake,
situated in the Grímsvötn caldera. Currently,
the lake level rises 10-15 m/yr until it
reaches a height of 80-110 m, after which it
produces a jökulhlaup (Bjornsson, 2003). In
addition to this periodicity, ice quakes, a
depression in the overlying ice, and a sulfur
smell in the drainage rivers indicate a
jökulhlaup has begun its 50 km journey from
Grímsvötn to the closest outlet at
Skeiðarársandur. The relative periodicity and
easy identification of the jökulhlaups has
facilitated thorough studies of the mysterious
events.
The jökulhlaup of 1996 was a
particularly well-studied event and many
Figure 6: Figure from Halgason and Duncan, publications followed it. The overwhelming
2001.

  49
Caltech Enrichment Trip Iceland

Figure 7: Figure from Jóhannesson


opinion on the jökulhlaup was that it was not a typical jökulhlaup. In traditional jökulhlaups,
discharges increase exponentially until a peak discharge is reached in a matter of a few to
several days, after which the discharge decreases slowly over one to a few weeks. This is well
explained by the jökulhlaup initiation hypothesis by Nye (1976) in which a positive feedback
mechanism exists where the friction of the water in the ice tunnels enlarges the tunnels until the
water pressure in the reservoir drops below the ice overburden pressure, after which ice creep
slowly closes the tunnel. In the 1996 jökulhlaup, the increase in discharge was so large, peaking
in just 16 hours, that friction in an ice tunnel would not be substantial enough to account for the
volume required (Bjornsson, 2002). As the jökulhlaup burst through the glacier terminus, it took
with it several hundred meters of ice, suggesting that the water pressure was significantly higher

  50
Caltech Enrichment Trip Iceland

Figure 8: Figure from Russel, 2001.


than the ice overburden pressure, which would not be the case for an ice tunnel. Crevasses
were observed along much of the path from Grímsvötn to Skeiðarársandur, suggesting that the
glacier was deformed during the jökulhlaup (Bjornsson, 2002, Jóhannesson, 2002). The final
unusual observation during the 1996 jökulhlaup is that an ice crystal slurry was observed at the
surface of the floodwaters. The presence of these crystals only after their exit from the glacier
further suggests that the pressures in the glacier were significantly higher than atmospheric
pressure, enough to supercool the fluid. Skeiðarárjökull is a temperate glacier and therefore
would not otherwise be capable of freezing the water-sediment mixture. All of these
observations have led to an alternative hypothesis for initiation of some jökulhlaups, such as the
one in 1996. Bjornsson (2002) and Jóhannesson (2002) propose a scenario in which the glacier
is momentarily lifted above the flow. This allows for the extreme increases in discharge as a
huge volume of water is able to pass at a time. As the glacier falls back down, the drop in
discharge is also expected to occur quickly. This hypothesis also explains the supercooling of
the sediment-water mixture as well as the explosive exit the waters made at Skeiðarársandur. A
schematic diagram illustrating how the glacier is lifted as the pressure wave exits is shown in
Figure 7 (Jóhannesson).
Gígjukvísl is one of three main outlet rivers from Skeiðarárjökull (Figure 8; Russell,
2001). The glacier has retreated significantly from the proximal zone of Skeiðarársandur, leaving
a depression in its absence. This depression is pivotal in regulating the flow of water during the
jökulhlaup (Gomez, 2002). During the 1996 jökulhlaup which was truly enormous, this river
drained approximately 60% of the total flood water volume (Russell et al. 1999). Damage was
significant but would have been worse without the ponding effect of the depression. Less than
25% of the Skeiðarársandur proximal zone was inundated (Gomez, 2002). The Icelandic ring
road was partially washed away because of the event, but nobody was hurt due to far advanced

  51
Caltech Enrichment Trip Iceland

warning of the coming jökulhlaup. The decoupling of the glacier from the sandur, caused by
glacial retreat, therefore determines the erosive and destructive powers of jökulhlaups.

Refernces & Further Reading

Benn, Douglas I., and David JA Evans. "Glaciers and glaciation: London." Edward Arnold
(1998).
Björnsson, Helgi. "Subglacial lakes and jökulhlaups in Iceland." Global and Planetary Change
35.3 (2003): 255-271.
Fowler, A. C., and F. S. L. Ng. "The role of sediment transport in the mechanics of jökulhlaups."
Annals of Glaciology 22 (1996): 255-259.
Gomez, Basil, et al. "Erosion and deposition in the proglacial zone: the 1996 jökulhlaup on
Skeioarársandur, southeast Iceland." The Extremes of the Extremes: Extraordinary
Floods 271 (2002): 217.
Harris, P. W. V. "The seasonal temperature-salinity structure of a glacial lake: Jökulsarlon,
south-east Iceland." Geografiska Annaler. Series A. Physical Geography (1976): 329-
336.
Harris, P. W. V. Sedimentation in a Proglacial Lake: Jokulsarlon, South-East Iceland. Diss.
University of East Anglia, 1977.
Helgason, Johann, and Robert A. Duncan. "Glacial-interglacial history of the Skaftafell region,
southeast Iceland, 0–5 Ma." Geology 29.2 (2001): 179-182.
Howarth, P. J., and R. J. Price. "The proglacial lakes of BreiDamerkurjökull and Fjallsjökull,
Iceland." Geographical Journal (1969): 573-581.
Johannesson, T. Ô. M. A. S. "The initiation of the 1996 jokulhlaup from Lake Grimsvotn,
Jonsson, J. "On the formation of frontal glacial lakes." Geogr. Annlr37 (1955): 3-4.
Martini, Ireneo Peter, Michael E. Brookfield, and Steven Sadura. Principles of glacial
geomorphology and geology. Pearson College Div, 2001.
Þórarinsson, Sigurður, and Hans Wilhelmsson Ahlmann. Vatnajökull: scientific results of the
Swedish-Icelandic investigations 1936-37. Geografiska annaler, 1943.
Nye, J. F. (1976) Water flow in glaciers: jôkulhlaups, tunnels and veins. J. Glaciol. 17(76), 181-
207.
Russell, Andrew J., and Óskar Knudsen. "An ice-contact rhythmite (turbidite) succession
deposited during the November 1996 catastrophic outburst flood (jökulhlaup),
Skeiðarárjökull, Iceland." Sedimentary Geology 127.1 (1999): 1-10.
Russell, A. J., et al. "Morphology and sedimentology of a giant supraglacial, ice-walled,
jökulhlaup channel, Skeiðarárjökull, Iceland: implications for esker genesis." Global and
Planetary Change 28.1 (2001): 193-216.Vatnajokull, Iceland." IAHS Publication 271
(2002): 57-64.
Todtmann, E. M., (1960), Gletscherforschungen auf Island (Vatnajokull) Abhandlung aus dem
Gebeit der Auslandskunde, Hamburg. Bd. 65. Rh. c, Bd. 19.95pp.
 
 
 
 
 
 
 

  52
Caltech Enrichment Trip Iceland

Recent Volcanism
Contributed by Masha Klescheva
Laki Eruption
In the west area of Vatnajökull
National Park, we can visit the site of the
deadliest eruption in historical times.
Disturbed earth extends out around Mt Laki
in parallel fissures for 27 km, now impotent
deformed earth, but at one time bringing
annihilation to the entire Northern
hemisphere. The Laki Eruption lasted over
an 8 month period between 1783 and 1784.
Each of the ten episodes of the eruption
ensued by the formation of a fissure, then
short lived explosive phreatomagmatic
activity (basalt magma interacting with
groundwater), then continued gentle flow from Figure 1: Laki, photo by Ulrich Latzenhofer
the vents, Hawaiian style. Over the 8 month
period, 14 km3 of basalt lava poured out from the system (that’s enough to pave over Boston 63
meters deep in basalt!)
The real killer, however, creeped out simultaneously with the enraged antics of the lava.
Acid haze, clouds of hydrofluoric acid and sulfur dioxide, proved to be much more deadly. Eight
million tons of HF and 120 million tons of SO2 were released into the atmosphere alongside the
explosion. 50% of Icelandic livestock died of dental and skeletal fluorosis, causing a famine
which annihilated a quarter of the Icelandic population. These difficult times in Icelandic history
are known as the Mist Hardships.
The Mist, as it were, did not
remain in Iceland. Of the 120 million
tons of SO2, 95 million tons made it to
the upper troposphere and lower
stratosphere, where the jet stream
circulated the haze around the entire
northern hemisphere. The resulting
drop in global temperatures caused
crop failures in Europe, and was
noticed as far as Egypt, India, and
North America. The Laki Eruption is
estimated to have killed over 6 million
people worldwide.

Reynisdrangar Basalt Columns


There was no hope. All night the
crew struggled against the trolls, but
their human ingenuity was no match Figure 2: Reynisdrangar basalt columns, photo from
for the trolls’ inexhaustible www.virtualtourist.com.
constitutions. The ship captain
defiantly yelled orders into the now dawn, but he heard the ship’s bottom scrape the black

  53
Caltech Enrichment Trip Iceland

pebbles of the beach: the trolls had pulled them on shore. It was certain death, until… the
morning rays struck the creatures, and they instantly turned to stone: columns of stacked basalt,
right on the black beach in Southern Iceland.
So tells us Icelandic Folklore. Reynisfjara is a black pebble beach next to Vik, the
southernmost village in Iceland of only 350 people. Reynisdrangar is a cliff of basalt columns on
the beach, home to a variety of bird-life. Surrounding the cliffs are caves from twisted basalt,
from which puffin chicks belly-flop into the ocean.

Eyjafjallajokull Eruption, þorvaldseyri visitor center


The eruption of Ey-ya-fyad-la-ou-couddle made the news in 2010 not only because none
the news anchors could pronounce the name, but also because of the level of travel disturbance
it created, the highest since World War II. As a volcanic eruption it was not the most impressive,
but it managed to disrupt human travel to such a degree for several reasons. First, it sits directly
underneath the jet stream, which is able to swiftly redistribute massive amounts of ash directly
into Northern Europe. Second, the eruption happened underneath 200 meters of glacial ice. The
melted water flowed right back
into the erupting volcano,
rapidly vaporized, and
elevated the explosive power
of the eruption. In addition, the
lava cooled very fast, which
created clouds of glass-rich
ash that shot straight into the
jet stream.
The þorvaldseyri Visiter
Center opened at the foot of
the volcano "when the ash
settled." There is a wall
showcasing the history of
Icelandinc volcanism, and a Figure 3: Eyjafjallakokull eruption, photo from
short film about a family farm in http://scienceblogs.com/eruptions/2010/04/16/eyjafjallajokull-
þorvaldseyri. "It’s NOT a eruption-cont/
typical home-movie, but a
dramatic 20-minute film."

References and Further Reading

http://en.wikipedia.org/wiki/Laki
Th Thordarson 1993
http://www.wired.com/2013/06/local-and-global-impacts-1793-laki-eruption-iceland/
http://www.guidetoiceland.is/travel-iceland/attraction/reynisfjara
Lonely Planet, Iceland
http://scienceblogs.com/eruptions/2010/04/16/eyjafjallajokull-eruption-cont/
http://en.wikipedia.org/wiki/2010_eruptions_of_Eyjafjallaj%C3%B6kull
http://www.icelanderupts.is/en/

  54
Caltech Enrichment Trip Iceland

Reykjavik
Contributed by Miki Nakajima
History
Reykjavik is the capital of Iceland and the largest city in Iceland. Its
population is around 120,000, which is ~60% of the Icelandic populations.
Reykjavik means “Smokey Bay”, which is named after steam rising from
geothermal vents [2]. The first permanent Icelander is believed to be
Ingólfur Arnarson (AD 871). He decided to live in this location based on a
Viking tradition: throwing his high-seat pillars into the ocean when he saw
the coastline and settled wherever the pillars came to shore. Until the
18th century, Reykjavik was just a small farmland. In 1752, the king of
Denmark donated Reykjavik to Innréttingar Corporation. This movement
was led by Skúli Magnússon, as known as “Father of Reykjavik”. He
started wool factories, which became the major industry in Iceland. The
Danish crown abolished the monopoly trading in 1786 and this date is
recorded as the foundation of Reykjavik. Reykjavik boomed during Figure 1: Flag of
Reykjavik [1]
World War II when British and American soldiers built camps there. The
city kept growing until the financial crisis in 2008.

Geography
During the Ice Age, this region was partly covered by a large glacier and partly by sea water.
At the end of the Ice Age, some hills in Reykjavik existed as islands. The sea level during this
period could have been 43m (141 ft) higher than the current sea level as indicated by clam
shells found in sediments.

Weather
Reykjavik is warm for its high latitude due to the Gulf Stream and Westerlies. The
temperature in winter rarely goes below -15˚C (5˚F). In summer, it is between 10-15˚C (50-59˚F)
(Figure 2). The length of the day can be as short as 4 hours in winter and as long as 21 hours in
summer. On average, precipitation occurs 148 days per year.

Figure 2: Weather in Reykjavik [1]

  55
Caltech Enrichment Trip Iceland

Energy
Reykjavik is one of the greenest cities in the
Figure 2:
world. Space heating is provided completely by
Hydrogen
geothermal energy. Some buses in Reykjavik use fuel station
public hydrogen fuelling stations (The Ecological [3]
City Transport System, ECTOS, project, Figure
2).  

Places to visit
•The Blue Lagoon geothermal spa – the largest outdoor spa
(5000 m2) and one of the most visited attractions in Iceland,
located 20 minutes from Keflavik airport and 40 minutes from
Reykjavik by car. It is an artificial lagoon and is fed by the water
output of the geothermal power plant, Svartsengi. The water is
rich in minerals (e.g. silica and sulfur) and the water
temperature is controlled to 37-39˚C (98-102˚F). Facility hours:
9am-9pm, price: 35 EUR. Most crowded between 10am-2pm.
Figure 3: The Blue Lagoon [4]
•Perlan – The building has been used to store hot water. It has
large space for exhibition/dining/shopping. The restaurant on
the highest floor has a great view of the city and rotates every
two hours. The Saga museum exhibits the early history of
Iceland.

•Hallgrimskirkja – a Lutheran parish church named after a poet


in 17’s century, Hallgrímur Pétursson. This is the largest church
in Iceland (73 meters, 244 ft). State Architect Guðjón
Samúelsson's designed this building in 1937 inspired by Figure 4: Perlan [5]
columnar jointed basalt. The statute is Leif Erikson, who found the North American continent
500 years before Christopher Columbus. The elevator inside of the church brings you up to top
of the building where you can enjoy the great view of the city.

•Höfði – a house initially built for


the French consul Jean-Paul
Brillouin in 1909. It is best known
as the location for the 1986
Reykjavík Summit meeting of
presidents Ronald Reagan and
Mikhail Gorbachev to take a step
to end the Cold War. Figure 6: Höfð i[7]

•Bæjarins Beztu Pylsur (4 stores in/near Reykjavik) – a small chain


of popular hot dog stands in Reykjavik. The British newspaper The
Guardian selected this chain as the best hot dog stand in Europe
Figure 5: Hallgrimskirkja in 2006. A number of celebrities have visited, including Bill Clinton,
[6] James Hetfield, and Charlie Sheen.

  56
Caltech Enrichment Trip Iceland

•More – Tjörnin (lake), Sun Voyager (Viking monument based on the myth), Laugardalur (spa)
[9]

Figure 7: Detailed map of Reykjavik sites.

References & Further Reading

[1] Reykjavik
http://en.wikipedia.org/wiki/Reykjav%C3%ADk
[2] Iceland, Lonely planet, 7th edition
[3] ECTOS, project
http://www.global-hydrogen-bus-
platform.com/InformationCentre/PhotoGallery
[4] Blue Lagoon
http://en.wikipedia.org/wiki/Blue_Lagoon_(geothermal
_spa)
[5] Perlan
http://www.barth.com/iceland/reykjavik/pages/dsc_306
8.htm
[6] Hallgrimskirkja
http://en.wikipedia.org/wiki/Hallgr%C3%ADmskirkja
Figure 8: Map of Reykjavik area.
[7] Höfði
http://en.wikipedia.org/wiki/H%C3%B6f%C3%B0i
[8] Bæjarins Beztu Pylsur
http://en.wikipedia.org/wiki/B%C3%A6jarins_Beztu_Pylsur
[9] Sun Voyage http://pilgrimito.com/sun-voyager-reykjavik
[10] Laugardalur http://www.holidaym.ru/iceland/gid_laugardalur_valley.php

  57
Caltech Enrichment Trip Iceland

List of Maps
Iceland Soil Map
•see “Icelandic Highlands” on page 16

Overview map
•Locations of our 40 geologic and cultural stops
•Locations of our 8 accommodations
•Overlays of 21 regional maps

Regional Maps:

Regional Map Number Region Detailed Map Type


1 SW Iceland Geologic
2 SE Iceland Geologic
3 N-Central Iceland Geologic
4 NE Iceland Geologic
5 E Iceland Geologic
6 Þingvellir Geologic, Satellite
7 Geysir, Gulfoss, Reykholt Pool, Geologic, Satellite
Deildartunguhver Spring, Sjova House
8 Landmannalaugar Topographic, Satellite
9 Sjova House, Deildartunguhver Spring, Stong Topographic, Satellite
10 Aldeyjarfoss, Ódáðahraun Lava Field Topographic, Satellite
11 Reykjahlid (Mývatn), Grjótagjá Cave, Mývatn Topographic, Satellite
Rhyolitic Cone, Hverir/Námafjall Springs,
Krafla Power Plant, Krafla Volcano, Mývatn
Nature Baths
12 Husavik Topographic, Satellite
13 Tjornes Fracture Zone Satellite
14 Þeistareykir Triple Junction, Äsbyrgi, Dettifoss Topographic, Satellite
15 Askja Topographic, Satellite
16 Askja, Vatnajoekull Outwash Plaines, Satellite
Karahnukur Dam
17 Berufjordur Zeolites, Berufjodur Dike, Satellite
Breiðdalsvik, Djupivogur, Teigarhorn,
Stodvarfjordur
18 Jokulsarlon Glacial Lagoon, Skaftajell Glacier, Topographic, Satellite
Laki
19 Laki, Eldgja Topographic, Satellite
20 Landmannalaugar, Eldgja, Dyrhólaey, Geologic, Satellite
Solheimajokull, Eyjafjallajokull, Þorvaldseyri
Museum, Seljalandsfoss,
Þykkvabaejarklaustur
21 Grindavik, Loft Hostel Geologic, Satellite

  58

You might also like