Elsevier Editorial System(tm) for Earth Science
Reviews
Manuscript Draft
Manuscript Number: EARTH1381
Title: Subaerial salt extrusions in Iran as analogues of ice sheets, streams and
glaciers
Article Type: Review Article
Keywords: glaciers; salt sheets; namakiers; structures and fabrics; material
properties.
Corresponding Author: Dr. Christopher Talbot, PhD
Corresponding Author's Institution: Uppsala University
First Author: Christopher J Talbot, PhD
Order of Authors: Christopher J Talbot, PhD; Veijo
Pohjola , PhD
Abstract: Ice (H20) and salt (halite, NaCl) share many physical properties and
resemble each other in hand specimens and subaerial gravity-driven flows.
However, while most significant bodies of ice accumulate in cold highlands and
gravity spread where and soon after they form, most significant bodies of salt
accumulate in tropical marine basins and have to be buried by >1 km of other
rocks before they flow. Buried salt is driven by differential loading into
various categories of piercing structures known as diapirs. Many diapirs extrude
onto the surface as sheets of allochthonous (out of place) salt. Thousands of
sheets of allochthonous salt have been interpreted in over 35 basins worldwide
in the last 25 years, mainly in the toes of passive continental margins and in
orogenic belts where some are >10E3 km2 in area. Most former salt sheets are now
submarine or subsurface and have to be studied by seismic profiling ± drilling
but several active examples are beautifully exposed in Iran. These were compared
to ice glaciers soon after they were introduced to western science, a comparison
that has been neglected since. Here we update this analogy and use modern
understanding of flowing ice and salt to examine the similarities and
differences that might be mutually beneficial to both fields of study as well as
extraterrestrial scientists.
The profiles, internal structures and fabrics in flowing bodies of ice and
salt are sensitive gauges of the histories of their budgets of supply and loss.
However, whereas ice caps merely compact where snow accumulates, salt sheets are
fed by already deformed salt from below. When salt diapirs first emerge on land
they extrude domes that mature to the profiles of viscous fountains that often
feed glacier-like flows known as namakiers. After locally exhausting their deep
source layers, salt fountains spread to the profiles of viscous droplets normal
for ice caps.
Ice typically deforms at >80% (usually > 90%) of its absolute melting
temperature while most salt deforms at <50% of its homologous temperature; as a
result, grain shape fabrics in salt are clearer and have longer strain memories
than in ice. Foliations in deformed salt map streamlines that aid understanding
of how internal folds develop. Salt sheets seldom erode their channels like
flowing ice and internal debris accumulates on their tops rather than their
bases. Ice sheets float on water but, as salt sheets are twice as dense as ice,
rain that falls onto the top surface of namakiers tends to stay there. Both
glaciers and namkiers surge but the association between surges and changes in
boundary conditions are much clearer for namakiers than glaciers. Because the
rate of delivery of land ice to the oceans is such an important control on sea
level, we end by considering how the implications of surging salt might help
studies of surging ice and converge on recent glaciological findings that pont
to changes in boundary conditions other than their bases.
Suggested Reviewers: Ian Alsop
gia@st-andrews.ac.uk
Expert in salt tectonics
Wendy Lawson
wendy.lawson@canterbury.ac.nz
Expert structural glaciologist
Garry KC Clarke
clarke@eos.ubc.ca
International expert in glaciology
JK Warren
jwarren@brunet.bn
International expert in Salt
Author Agreement
Cover letter,
Dear Editors,
Subaerial salt extrusions in Iran as analogues of ice sheets, streams and glaciers
On behalf of both authors, I enclose an entirely re-written version of a script that was
submitted as EARTH1192 with a slightly different name over a year ago. We were asked by
our reviewers and editor of that script for a major reworking in what was a reasonable time
were it not for the fact that I retired and moved my household from Sweden to England soon
after. I apologise for the delays beyond the original and subsequently-negotiated deadlines
and fully understand that you wish to treat this as a new script. However, we wished to
produce a paper we would be proud of rather than to rush to meet your schedule.
This script is much longer than the original version because every one of the reviewers’
and editor’s comments on the previous version have been taken into account. We have taken
to heart their encouragement to add more coverage on ice and have added two significant
sections, one on micro-organisms in ice and salt, the other summarising the hot topic of
surging ice (up to September 2008) and relating it to surging salt. We now have a clearly
stated aim and theme. We have also added to and improved out Figures and cited primary
references (as well as a few text books).
I have had great difficulty uploading the Figure files to your web site and submit all the
figures and plates in black and white jpegs to save space. If the script is accepted we would
like the figures on the web to be in colour.
Sincerely,
Christopher Talbot
* Manuscript
Click here to download Manuscript: Talbot-ice-final.doc
Click here to view linked References
1
1
Subaerial salt extrusions in Iran
2
as analogues of ice sheets, streams and glaciers
3
Christopher J. Talbot*, Veijo Pohjola
4
Department of Earth Sciences, Uppsala University, Villavägen 16, 752 36 UPPSALA, Sweden.
5
Received February 2007; returned November 07 Resubmitted 15 September 2008 Accepted
6
7
Abstract: Ice (H20) and salt (halite, NaCl) share many physical properties and resemble each
8
other in hand specimens and subaerial gravity-driven flows. However, while most significant bodies
9
of ice accumulate in cold highlands and gravity spread where and soon after they form, most
10
significant bodies of salt accumulate in tropical marine basins and have to be buried by >1 km of
11
other rocks before they flow. Buried salt is driven by differential loading into various categories of
12
piercing structures known as diapirs. Many diapirs extrude onto the surface as sheets of
13
allochthonous (out of place) salt. Thousands of sheets of allochthonous salt have been interpreted
14
in over 35 basins worldwide in the last 25 years, mainly in the toes of passive continental margins
15
and in orogenic belts where some are >103 km2 in area. Most former salt sheets are now submarine
16
or subsurface and have to be studied by seismic profiling ± drilling but several active examples are
17
beautifully exposed in Iran. These were compared to ice glaciers soon after they were introduced to
18
western science, a comparison that has been neglected since. Here we update this analogy and
19
use modern understanding of flowing ice and salt to examine the similarities and differences that
20
might be mutually beneficial to both fields of study as well as extraterrestrial scientists.
21
The profiles, internal structures and fabrics in flowing bodies of ice and salt are sensitive
22
gauges of the histories of their budgets of supply and loss. However, whereas ice caps merely
23
compact where snow accumulates, salt sheets are fed by already deformed salt from below. When
24
salt diapirs first emerge on land they extrude domes that mature to the profiles of viscous fountains
25
that often feed glacier-like flows known as namakiers. After locally exhausting their deep source
26
layers, salt fountains spread to the profiles of viscous droplets normal for ice caps.
27
Ice typically deforms at >80% (usually > 90%) of its absolute melting temperature while most
28
salt deforms at <50% of its homologous temperature; as a result, grain shape fabrics in salt are
29
clearer and have longer strain memories than in ice. Foliations in deformed salt map streamlines
30
that aid understanding of how internal folds develop. Salt sheets seldom erode their channels like
2
31
flowing ice and internal debris accumulates on their tops rather than their bases. Ice sheets float on
32
water but, as salt sheets are twice as dense as ice, rain that falls onto the top surface of namakiers
33
tends to stay there. Both glaciers and namkiers surge but the association between surges and
34
changes in boundary conditions are much clearer for namakiers than glaciers. Because the rate of
35
delivery of land ice to the oceans is such an important control on sea level, we end by considering
36
how the implications of surging salt might help studies of surging ice and converge on recent
37
glaciological findings that pont to changes in boundary conditions other than their bases. (485
38
words)
39
40
Key words: Ice sheets, glaciers, salt sheets, namakiers, structures and fabrics, material properties.
41
* Corresponding author: Now at 14 Dinglederry, Olney, Bucks, 46MK 5ES, UK (Tel: +44 1234 714
42
43
44
45
46
47
48
140)
E-mail addresses: Christopher.Talbot@geo.uu.se and Veijo.Pohjola@geo.uu.se
1. Introduction
Water ice and salt (consisting mainly of halite, NaCl) can resemble each other in both hand
49
specimen and in surface bodies (Figs 1 & 2) despite forming by entirely different processes in
50
different and mutually exclusive environments. Most significant bodies of ice form at high latitudes
51
near to maritime air masses and soon gravity spread from their high inland accumulation areas as
52
lobes and outlet glaciers to more continental and drier areas. By contrast, most significant bodies of
53
salt crystallise by the evaporation of seawater in marine basins between latitudes 10-40º N & S.
54
Autochthonous (in place) salt has to be buried to depths of km and squeezed back to the surface in
55
intrusive bodies known as diapirs before some extrude* sheets of allochthonous (out of place) salt
56
that gravity spreads over the surface in smaller analogues of ice caps and glaciers (Fig. 3). Large
57
-------------------------------------------------------------------------------------------------------------------------
58
*Only footnote: Except where signalled in section 5.6, we use extrusion here in the geological
59
sense of an elasto-plastic solid being driven upward through vents in other rocks, not in the
60
glaciological sense (Demorest, 1942; Gudmundsson, 1997a) of an increase in horizontal flow
61
velocity with depth. -----
3
62
bodies of both ice and salt on the surface are temporary because they are gravitationally unstable
63
and vulnerable to the climate. Significant bodies of ice seldom survive for more than a million years
64
and, while extruded sheets of allochthonous salt are also short-lived, the salt in them can be 5-500
65
million years old.
66
Here we update early comparisons between glaciers of ice and salt (e.g. Harrison, 1930) and
67
explore how the comparisons and contrasts between flows of ice and salt might be of mutual
68
benefit to studies of both materials on Earth and analogue materials on extra-terrestrial bodies.
69
Bodies of both water ice and salt are being recognised or suspected on increasing numbers of
70
extraterrestrial bodies where the different environments may lead to ice flowing rather like salt on
71
Earth (Schenk and Jackson, 1993; Pappalardo and Barr, 2004, and Schenk and Pappalardo, 2004)
72
or methane flowing like water on Titan (Lunine and Atreya, 2008). To understand the processes
73
that give rise to extraterrestrial landforms, it is becoming increasingly important to study their
74
analogues on Earth. It is also advantageous to be able to distinguish by remote means such
75
phenomena as snow fields from salt playas, cryo-volcanoes from salt diapirs, ice sheets from salt
76
sheets and glaciers from namakiers. However, here we will focus mainly on Earth-bound studies as
77
structural glaciologist and geologists.
78
Early hominids probably used ice as “land bridges” to spread through high latitudes and must
79
have learned to cope with the harsh climates of the ice age. They probably also learned that salt is
80
a necessary for their health and have probably “mined” it for many millennia. The first known
81
solution mining was in Poland at least ~3000 years ago and salt mining began in the middle east
82
~2000 years before science began to take an interest in salt (for a historical review of salt tectonics,
83
deformation involving salt flow, see Jackson, 1995).
84
Glaciology began slowly when glaciers invaded valleys farmed in the Alps during the little ice
85
age in the 17th century (Walker and Waddington, 1988). However, the subject really took off after
86
1837 when Louis Agassiz, accepting earlier suggestions that the huge erratic boulders of granite
87
high on limestones in the Jura Mountains were emplaced by ice rather than the “Deluge”, went on
88
to suggest that Earth had been subject to a past ice age.
89
Svein Pálsson, an Icelander, first described in his diary at the end of the 18th century how
90
glaciers flowed under gravity like viscous tar, a theory proved in 1846 by James Davis Forbes
91
using experiments on the Mer de Glace in the French Alps (Walker and Waddington, 1988). We
4
92
guess that Pálsson had in mind a recent Icelandic lava eruption as another analogue of viscous
93
flow in rocks and ice, an approach we follow in this review.
94
After a century of relatively slow progress in theoretical glaciology during the exploration of
95
polar geography, glaciology kicked off again after WWII, boosted by the Habbakuk project – one of
96
Churchill’s pet projects attempting to build aircraft carriers using ice and wood pulp (Perutz, 1948).
97
This project inspired British ice physicists to start post-war glaciological field projects to test their
98
theories about glacial dynamics (Clarke, 1987). In fact, exploration of the ice sheets of Antarctica
99
and Greenland had begun in the 1930s (Robin and Swithinbank, 1988) but blossomed as a result of
100
the large geophysical campaigns and >50 Antarctic stations installed for the International
101
Geophysical Year (IGY) of 1957-58 by the 12 original signatories of the Antarctic treaty that came
102
into force in 1961.
103
The literature that has accumulated on glaciology since the work of Agassiz in mid 19th century
104
is rather larger than that on salt extrusions since Ville (1856) first described a mountain salt of salt
105
extruding geyser-like in the Sahara Atlas of Algeria. This imbalance reflects snow and ice covering
106
>10% of the Earth’s surface whereas analogous extruded salt covers much less (but see maps in
107
Hudec and Jackson, 2006 for regions underlain by salt).
108
Shumskii (1964) is a definitive text on structural glaciology and Maltman et al., (2000) reviewed
109
modern developments. These authors summarised how structural glaciologists developed new
110
insights by applying structural geological concepts, first to the small individual valley glaciers and
111
ice caps that form over centuries, and then to the larger ice sheets that form over millennia
112
(Maltman et al, 2000; Hambrey and Lawson, 2000). Such studies accelerated in 1988-1998 after
113
the structures and deformation histories of surge-type glaciers and ice streams (fast flowing rivers
114
of ice that drain most of an ice sheet) were distinguished from those subject to slow penetrative
115
creeping flow.
116
Glaciology advanced further when short-wave surface radar identified new categories of folds
117
well above the beds of ice streams (Jacobel et al., 1993) and when high-resolution satellite images
118
became available (Merry and Whillans, 1993; Bindschadler, 1993, 1998; Unwin and Wingham,
119
1997). It surged again when ice cores found unexpected complications in near-basal ice strata
120
beneath the current divides of ice sheets (e.g. Souchez et al., 2000). Since then the recognition of
5
121
how the drainage of subglacial lakes relate to ice streams promises another significant advance
122
that will be considered later (section 9).
123
Hudec and Jackson (2007) summarized current ideas about the mechanics of salt flow, the
124
processes of diapir growth (Fig. 4), and the ways that these processes interact with regional
125
deformation to produce salt structures with a huge variety of shapes (Fig.5). They correct the past
126
emphasis on the buoyancy of structures of viscous salt in ductile surroundings to the current picture
127
where the location and shape of most salt bodies depend on how the stronger surrounding rocks
128
deformed. There has always been discussion about the relative importance of gravity and lateral
129
tectonic forces as drives for particular suites of salt structures, a tradition that we will continue here.
130
Hudec & Jackson (2006) reviewed the literature on the thousands of sheets of allochthonous salt
131
now known in orogens and passive continental margins and distinguished them into 4 different
132
categories. We will focus here on one of these, the subaerial salt extrusions in Iran that are
133
relatively accessible and provide the closest analogues of flowing ice.
134
Although past studies of ice have been funded by the hydro-electric power industry keen to
135
exploit melt water, today’s focus is on the concern about climate change and how the rate of
136
delivery of land ice to the oceans affects sea-level changes. By contrast, the data and funding for
137
studies of salt have come largely from industry. In the 1930s, hundreds of mines in the salt domes
138
of central Europe were extracting potash and sodium salts, gypsum, sulphur, borates, nitrates, and
139
zeolites as the feedstocks for the chemical and construction industries (Warren, 1999). In the 1970s
140
and early 80s, salt, was a target for nuclear waste isolation because excavating or dissolving
141
underground space in it is comparatively easy; by the late 1980s deformed salt was recognised as
142
too mobile for this use (the WIPP depository for military nuclear waste in New Mexico is in bedded
143
salt). From its beginning, the oil industry has always been interested in salt because salt affects
144
most aspects of hydrocarbon systems by creating structural traps, influencing reservoir distribution
145
and blocking fluid flow. Within 5 years of the first hydrocarbon blow-out in Texas in 1901, oil or gas
146
were being produced from sands in the flanks of ~50 salt diapirs. Something like 60% of
147
hydrocarbon fields are associated with salt structures (Halbouty, 1979). Many of the world's great
148
hydrocarbon provinces lie in such salt basins as the Gulf of Mexico (Nelson and Fairchild, 1989;
149
Worrall and Snelson, 1989, Wu et al., 1990, Diegel et al., 1995; Peel et al.,1995; Schuster, 1995,
150
Rowan et al. 1999), Persian Gulf (Kent, 1979; Talbot and Alavi, 1996; Sherkati and Letouzey, Eo,
6
151
Volgeo (ELS)2004), North Sea (Coward and Stewart ,1995; Kockel, 1998), Lower Congo Basin
152
(Rouby et al., 2002) the Campos Basin off Brazil (Cobbold et al., 1995) and the Pricaspian Basin
153
(e.g. Volozh et al., 2003).
154
Like water, salt is vital to most life forms. Just as a little salt flavours larger volumes of food,
155
a little salt exaggerates so many geological deformation processes (Talbot, 1992) that, as for ice,
156
salt structures can be treated as natural analogues models of many deformation processes in more
157
usual rocks Thus thin layers of salt can detach overlying rocks in regional extension and shortening
158
and simulate decoupling by shear along weak zones in and beneath the lithosphere. Salt often
159
plays a role in sedimentary basins similar to that of weak and buoyant granitoids in the crystalline
160
continental crust (Talbot, 1992).
161
This comparative review explores some of the analogies between ice and salt. As processes
162
involving salt are generally less well known than glacial processes, we will focus more on salt than
163
ice. Section 2 starts by contrasting how ice sheets and analogue salt bodies form. Section 3 then
164
explains how salt is buried and expelled back to the surface to form salt diapirs, some of which
165
extrude analogues of ice sheets. We then compare the shapes and rates of salt extrusions with ice
166
before discussing the dynamics of salt diapirs and extrusions. Hudec & Jackson (2006) recently
167
suggested that the emplacement of salt sheets require lateral compression but we will describe
168
evidence that questions this. Section 4 describes the layering within ice and salt and the debris they
169
generate, carry and deposit. Section 5 discusses their grain-scale shape fabrics and deformation
170
mechanisms before distinguishing their macroscopic internal structures into kinematic and dynamic.
171
We then outline kinematic folds in ice and salt before considering dynamic folds near the bottom
172
and top boundaries. Section 6 contrasts the fractured carapaces of ice and salt bodies before
173
section 7 dicusses microscopic their fluid inclusions and biological contents. Section 8 considers
174
the mechanical effects of water on ice and salt and leads to section 9 discussing current
175
understanding of surging flows in not only ice and salt but also of laboratory experiments on the
176
gravity spreading of viscous fluids. The conclusions highlight the areas where comparative studies
177
of bodies of flowing ice and salt might be mutually beneficial.
178
179
180
7
181
182
2. Formation and deformation of ice and salt
It usually takes decades to centuries for snow sedimented on topographic highs in cold climates
183
to be buried to depths where firn compacts to crystalline ice that can gravity spread or glide down-
184
slope to levels where it ablates (Fig. 3A). By comparison, rock salt is a comparatively rare rock that
185
precipitates wherever seawater evaporates beyond 11 times its general salinity in restricted basins
186
between latitudes 10 & 40º N and S with or without additional input from volcanic brines (Fig. 3B)
187
(Warren, 1999; Hardie, 1991; Hovland et al., 2006). Unlike ice and most other rocks that compact
188
on burial, rock salt precipitates as a crystalline rock with a grain size of cm that typically loses any
189
initial porosity by 40 –70 m burial (often over millions of years) as halite precipitated from
190
subsurface brines infills the pores (Warren 1999, p. 22). By comparison, compaction and
191
crystallisation (e.g. to cm grains) at depths of 10-100 cm in a snow pack involve vapours rather than
192
solutions.
193
We will follow a classical approach and consider both glacier ice and rock salt as essentially
194
mono-minerallic metamorphic rocks that deform comparativelt rapidly at low stresses and low
195
temperatures and therefore model more common rocks deformed more slowly at much higher
196
stresses and temperatures. Table 1 lists some of the properties of water ice and halite.
197
The strengths of ice crystals are strongly anisotropic so that dislocation glide on the basal
198
plane is comparatively easy until hindered by kinks (Shumskii, 1964; Wilson and Marmo, 2000)
199
Compaction with depth increasingly rotates the initially random c-axes of snow through broad
200
clusters of c-axis about the vertical in firn that tightens over 103-4 years to a narrow girdle in ice at
201
an ice depth ~1.5 km (Wang et al., 2002). Compaction therefore progressively aligns the basal
202
shear planes in ice to the orientations where they are increasingly vulnerable to the bed-normal
203
layer thinning, along-bed extension and bed-parallel shear that increases toward the bed. Shear
204
along the margins of ice streams and glaciers imposes narrow girdles much faster (e.g. ~102 years).
205
Large strains are aided by diffusion, grain rotation and boundary translation processes (Alley et al.,
206
1997; Wilson and Marmo, 2000) that may lead to dynamic recrystallisation. The crystallography of
207
halite, is very different from ice and its face-centred cubic symmetry with 3 orthogonal potential slip
208
planes oblique to the faces ensures that, in bulk, salt is close to mechanically isotropic.
209
210
The cycle from snow to glacier ice through gravity-driven compaction and flow to water by
melting or ablation usually takes <<1 Ma on Earth today. The equivalent cycle, where brine
8
211
evaporation forms bodies of crystalline rock salt that are buried and squeezed back to the surface
212
where they gravity spread in flows that share visual and mechanical similarities to ice sheets, caps,
213
streams and glaciers usually lies somewhere between 5 and 500 Ma.
214
Significant salt sequences often begin accumulating rather slowly in cyclic beds of other
215
evaporites (carbonates and calcium sulphates) by evaporation of ephemeral brines in rift valleys
216
(Warren, 1999). The brines in rift valleys are likely to have high components of volcanic input and
217
Hovland et al., (2006a, b) have recently suggested that brines heated to supercriticality by volcanic
218
heat may deposit sizeable salt deposits. If the rift valleys open to narrow seaways and then oceans,
219
any initial cyclic sequence of subaerial evaporites can be buried by purer salt that can crystallise
220
much more rapidly (cm a-1) by epitaxial growth on pre-existing halitic substrates beneath permanent
221
brines (Talbot et al., 1996; Sage et al., 2005). Because salt is usually one of the first sediments to
222
form in a new basin it is usually buried by rocks of other origins (clastic sediments ± volcanic rocks).
223
Such basal salt beds may be >1 km thick over Mediterranean-like areas the but are still
224
volumetrically small compared to current and past ice sheets.
225
Old dirty ice with strong grain-shape fabrics are comparatively weak and can decouple cleaner,
226
younger and stronger overlying ice (Hooke and Hanson, 1986). Similarly, basal salt sequences are
227
so weak that their ductile flow can mechanically decouple the overlying cover rocks from their
228
crystalline basement during subsequent “thin skinned” tectonics Thus cover sequences tens of km
229
thick may undergo enormous lateral extension when they spread and/or slide off “passive”
230
continental margins (such as the Gulf of Mexico, Brazil, West Africa, etc: Duval et al., 1992; Hudec
231
and Jackson 2004) or enormous lateral shortening in foreland basins and thrust-fold belts beside
232
mountain chains where continents converge after intervening oceans have closed. About 120 of the
233
deformed basins with surviving salt tectonics are shown on maps by Hudec &Jackson (2007, figs.
234
5a & b).
235
236
237
3. Salt diapirs and extrusions
Ice and salt can be considered some of the weakest rocks (Table 1) so that substantial bodies
238
of both deform as crystalline elasto-viscous fluids as they flow “down” load gradients induced
239
mainly by gravity.
9
240
The closest salt comes to emulating ice and flowing in the place it first forms is where clastic
241
sediments accumulate and sink into thick, recently formed, salt layers. Although less dense than
242
salt until compacted at depths >~1km, these clastic sediments sink into the weak underlying salt as
243
they accumulate and drive the salt along lateral pressure gradients into adjacent highs. These form
244
passive cylindrical or wall-like diapirs with crests that stay near the depositional surface in a
245
process known as downbuilding (Fig. 3A, Barton, 1933; Jackson et al., 1994; Hudec and Jackson
246
2006).
247
The Zechstein (250 Ma) salt sequence in Germany and the North Sea and is known to have
248
developed downbuilt diapirs and extruded stacked namakiers in Late Triassic (Mohr et al., 2007)
249
and Cretaceous times (Zirngast, 1996). Similarly, the Jurassic Louanne salt sequence began
250
moving at shallow depths in the Gulf coast interior basins (Rosenkrans and Marr, 1967) and was
251
buried to as deep as14 km by younger clastic sediments that built out into the Gulf of Mexico and
252
drove most of the salt basinward in several generations of diapirs that spawned the hundreds of
253
submarine salt sheets now extruding on its floor (Wu et al., 1990, Gemmer et al., 2004).
254
Other salt sequences have been buried by overburdens of constant thickness faster than the
255
salt could rise by its (rather weak) inherent buoyancy. However, some diapirs actively upbuild
256
through overburdens that are only slightly denser but ductile (Fig. 4B, e.g. Jackson et al., 1990).
257
More usually, differential loading of a salt source layer by the mass of the overburden drives
258
reactive salt diapirs upward where faults have thinned and weakened stronger brittle overburdens
259
(Fig. 4C and Jackson et al., 1994, Hudec and Jackson, 2006a).
260
The crests of passive salt diapirs remain near the surface as their source layer deepens- and
261
the crests of many active or reactive diapirs rise to the surface (at 0.1 to 3 mm a-1: Jackson and
262
Talbot, 1986) from their deep source layer(s). What happens then depends on the surface
263
environment into which the salt extrudes. High rainfalls onshore can dissolve salt as fast as it
264
extrudes so that the insoluble residues in the extruding sequence accumulate atop the salt as
265
downward thickening cap rocks. These develop as hard cemented caps in reducing conditions
266
below the water table in the US gulf coast (e.g. Seni, 1987, Posey et al., 1987, Warren, 1997), but
267
as loose soils on salt emerging in oxidizing conditions in Iran (Talbot, 1998; Bruthans et al., 2007).
268
Equivalent diapirs rising into seawater probably dissolve more slowly in less aggressive seawater
269
and can extrude large volumes of salt that flow downslope with or without the burial that can lead to
10
270
further cycles of intrusion and extrusion (Jackson et al., 1994). The first geologists to encounter
271
such subaerial salt extrusions in the south of Iran referred to them as salt glaciers (Lees 1927, Busk
272
1929, DeBockh et al., 1929, Harrison, 1930). However, the term salt glacier involves a bilingual
273
redundancy, and because “river of salt” and “saltier” have other meanings, we use here another
274
bilingual term for sheets of allochthonous salt extruding over the former surface from salt diapirs:
275
namakier (from namak, Farsi for salt; Talbot and Jarvis, 1984).
276
Ice analogues of emergent salt diapirs are pingoes, where overpressured groundwater rises
277
through ground icing on fine-grained flats in permafrost regions and grow large “frost-blisters”.
278
These bulge the surface on scales of hundreds of metres in terrestrial structures and are
279
sometimes referred to as cryo-volcanoes (Washburn, 1979). Larger cryovolcanos exist on the icy
280
moons around the outer planets (Lucchitta, 2001; Pearle, 2003).
281
Salt diapirs are more common and usually larger than pingoes on Earth, being typically a few km
282
across in their shallow levels and 3-12 km from crest to root on wavelengths between 6 and 26 km.
283
However, the subaerial namakiers that simulate bodies of comparatively young glaciers are smaller,
284
fewer and usually extrude long after the salt in them formed. Such disparities in the ages of ice and
285
salt bodies indicate that ice on the surface is more vulnerable to degradation than buried salt. Being
286
denser than water, most salt extrudes over the seafloor offshore. The focus here is on subaerial
287
extrusions of salt over rocks and soils onshore that evolve through similar shapes in various arid
288
regions of Iran despite consisting of salts of two very different ages.
289
290
291
3.1 Shapes of salt extrusions
Although all viscous and strain-rate softening fluids (like ice and salt) flow downhill beneath
292
their top free boundaries, all valley glaciers, most outlet glaciers and several of the known ice
293
streams also generally flow downslope over their bottom boundaries (i.e. they are steered by the
294
topographies of their substrates).
295
Subaerial salt extrusions are also shaped by the rate at which distal salt advances over a
296
particular substrate from a vent of a particular shape. Thus axisymmetric extrusions spreading by
297
pure shear over plains slow as they expand circular sheets of allochthonous salt that can be ~ 100
298
m thick with surface velocities calculated at between 4.10-4 and 20 m a-1 (Wenkert, 1979). Faster
299
and thinner (e.g. 5 x 2 x 0.3 km) monoclinic namakiers spreading by simple shear down the dip
11
-1
300
slopes of anticlines in the surrounding limestones (at velocities measured at up to 17.6 m a : Talbot
301
et al., 2000) do not widen downslope, even where they advance over a plain (Fig. 1SB). Current
302
subaerial namakiers in Iran do not coalesce (Fig. 1SB). However, the bulbs of some groups of
303
diapirs have merged to form “salt canopies”, the nearest saline equivalents of coalesced piedmont
304
glaciers. Twelve diapirs contribute to an onshore salt canopy (38 x13 km) in central Iran (e.g.
305
Jackson et al., 1990), perhaps 30 form an incomplete 140 x 40 km canopy off the shore of Yemen
306
(Heaton et al., 1995) and hundreds have merged beneath the Gulf of Mexico (Hudec and Jackson,
307
2006). Most offshore salt extrusions are reactivated into later, smaller diapirs after being buried by
308
younger clastic sediments.
309
Although still-flowing salt is known to bypass now static salt along steep strike-slip faults
310
(Talbot, 1979), few individual salt sheets are sufficiently well known to recognise possible “salt
311
streams” (but see fig 21 in Hudec & Jackson, 2007). The closest equivalents to ice streams
312
currently known in salt sheets are where salt flows faster between partially sunken roof blocks (e.g.
313
fig. 6 in Hudec & Jackson 2006).
2
314
The largest source-fed thrust sheets of allochthonous salt are > 1000 km in area (Hudec &
315
Jackson, 2006) but the subaerial salt extrusions in Iran are much smaller (~ 25 km2) and flow more
316
slowly than most ice streams and glaciers.
317
Domes and namakiers of extruding salt emulate many more than visual aspects of ice sheets
318
and glaciers (Fig. 1 & 2) – with intriguing differences. The mass budgets of many active ice bodies
319
depend mainly on climatic and topographic influences on the supply of snow and the loss of ice. By
320
comparison, the mass budgets of active salt extrusions in Iran can be more complicated (and more
321
like those of ice streams) because their supply depends on the rates that already crystalline
322
material is fed into them (at mm to m a-1) from small vents in their country rocks (Fig. 6) and the
323
loss of extruded salt depends on the rate of its gravity spreading as well as and local river erosion
324
and general surface dissolution (Talbot and Jarvis, 1984) in the local rainfall (e.g. <1500 mm a in
325
Iran: Bruthans et al., 2007).
326
-1
The salt extrusions in Iran are studied as accesiblle versions of submarine salt extrusions that
327
can trap hydrocarbons, and as smaller and faster natural models of steady-state mountains of other
328
metamorphic rocks extruding from near the suture zones where continents converge (e.g. Talbot
329
and Jackson 1987; Talbot, 1998) Piles of nappes of salt and other metamorphic rocks can have
12
330
steady-state topographies and fluxes of mass and heat if they degrade by gravity spreading and
331
climatic processes as fast as they extrude from sub-horizontal channels in the crust (e.g. Pazzaglia
332
and Knueper, 2001). Thus the Himalaya extrude silicate gneisses at mm a on length scales of 10
333
km for 107 years (Beaumont et al., 2001) but the salt mountains of Iran are easier to study as they
334
extrude at 0.1-1 m a on scales of 10 m for 10 years. By comparison, ice characteristically flows
335
1m per 0.1-0.01 yr for <<106 years and a typical cycle from snow to melt is six orders of magnitude
336
faster than the Himalaya orogen (Hambrey and Lawson, 2000).
-1
-1
3
3
5
337
Extrusions of different salt sequences in different regions of Iran appear to evolve through
338
similar geometries although on different time and length scales (Fig. 6). Thus perhaps a hundred
339
diapirs of Oligocene to Miocene salt (20-5 Ma) emerge on the central plateau of Iran and something
340
like 150 diapirs of Hormoz (Neoproterozoic to Cambrian in age, ~500 Ma) salt extrude in the Zagros
341
Mountains and their foreland basin (Kent, 1970). Most such extrusions spread from rhombic vents a
342
few km across where salt has been expelled from depths of 3-12 km up the stems of reactive
343
diapirs pulled apart at releasing bends along strike-slip faults (Fig. 7 and fig. 29 in Hudec &
344
Jackson, 2007). Whatever its age, the salt first extrudes a rounded topographic dome that rises a
345
few hundred metres above its vent, probably in millenia, whether this is on a plain near sea level or
346
atop an anticlinal ridge >1 km high (like diapir 1 in Fig. 1SA). These domes of salt cannot support
347
their own weight in the surrounding air and so their lower slopes gravity spread short wide aprons
348
and/or narrow namakiers up to 8 km over the surrounding scenery (diapir 2 in Fig. 1SA and 7A).
349
While supplied by salt expelled from depth, extrusions of salt develop the distinctive smooth profiles
350
of viscous fountains (Lister and Kerr, 1989) that contrast with the angularity of the surrounding
351
mountains of other rock types (Figs. 6 & 7).
352
A high ratio of rate of extrusion to rate of spreading raises a summit dome well above the apron
353
spreading from the flanks of a salt fountain. After something like 60,000 years (Talbot et al., 2000,
354
fig. 7), a salt fountain locally exhausts its deep source layer, its ratio of extrusion- to spreading-rates
355
falls, and the summit dome subsides into the salt apron. No longer fed locally from below, the
356
former viscous fountain then gravity spreads to the profile of a viscous droplet (dashed in Fig. 6B
357
after Huppert, 1982) that characterise most ice caps with their comparatively widespread supplies
358
of snow. Unless salt extrusion is re-invigorated by lateral forces or further burial, salt erosion
359
eventually outpaces extrusion and the droplet profile degrades to conical piles of residual cap soils
13
360
growing atop the dissolving salt as insoluble components in the extruded sequence accumulate.
361
Eventually only a hollow of residual soils above a chimney of mega-breccia marks a former salt
362
extrusion (Fig. 6B).
363
Annual surveys of markers on ten mountains of Hormoz salt in the Zagros Mountains in the
364
1990s found that five were extruding faster that they degraded while five were wasting (Talbot et al,
365
2000, fig.1). Modelling such results indicate that the extruding salt rises out of its vent at over 1 m
366
a with a surface bulge due to budget irregularities travelling faster downslope than the salt (Talbot
367
et al, 2000). Fountains fed by 0lgocene-Miocene salt (~5-20 Ma) 3 km deep in central Iran are
368
smaller than those in the Zagros and an example reaches only 315 m above the surrounding plain
369
because it has been extruding more slowly e.g. 82 mm a-1 in an average annual rainfall near 23 mm
370
a for the last 42,000 years (Talbot and Aftabi, 2004).
-1
-1
371
The stratigraphy of a flowing body of ice only accumulates strain as it gravity spreads from
372
beneath a visible upslope accumulation zone to its ablation zone downslope. By contrast, the salt
373
equivalent of the glaciologists’ accumulation zone is out of sight in the extrusive vent (Fig. 1). The
374
salt surfacing on the crest of a salt extrusion is already strongly deformed having flowed many km
375
along the deep source layer and then up the diapir. It has also undergone strong lateral extension
376
where it passed through a flow separation zone just beneath the crest of a dome (starred on
377
profiles on Fig. 7). Once exposed on the surface, the salt can be dissolved by rain falling on any
378
part of it. The resulting surface relief in flowing salt allows study of the initiation and development of
379
internal deformation structures and fabrics more easily than in flowing ice where relief tends to be
380
lower, particularly in accumulation zones.
381
382
383
3.2 Dynamics of salt diapirs and extrusion
Hudec & Jackson (2007) detailed the forces driving and hindering salt flow. Differential loading
384
provides the dominant driving stress which is opposed by the strength of the overlying rocks and
385
friction along the boundaries of the sal body. Salt flows as soon as the driving forces overcome the
386
resisting forces, otherwise, it can remain static for hundreds of millions of years. Three types of
387
loading can drive salt flow: gravitational loading, displacement loading, and thermal loading.
388
Because salt behaves as a fluid over geologic time scales, Hudec & Jackson (2007) simplified the
389
effects of gravitational loading by using the fluid statics concept of hydraulic head (Kehle, 1988) and
14
390
distinguished hydraulic head gradients into two components: elevation head and pressure head. An
391
elevation head gradient is between two particles of fluid at different elevations above an arbitrary
392
horizontal datum. A pressure head is the height of a fluid column that could be supported by the
393
pressure exerted by the overlying rock. A pressure head gradient is induced by lateral variations in
394
the load on salt layers induced by erosion, sedimentation or deformation. The geometry of the
395
bottom boundary of the fluid layer does not influence the head gradient although it may influence
396
the geometry of flow once it begins.
397
Hudec and Jackson (2007 p.29) defined displacement loading as the “results from the forced
398
displacement of one boundary of a rock body relative to another” (e.g., Suppe, 1985). Although they
399
went on to state that “in salt tectonics, this type of loading occurs when the flanks of a salt body
400
move toward or away from one another during regional shortening or extension” we find the first
401
sentence obscure as to whether the boundaries of the salt body are moved by gravity or lateral
402
tectonic forces. We will therefore refer to lateral tectonic forces rather than displacement loading.
403
Subhorizontal shears are known to be very important along gently dipping salt layers in thin- and
404
thick-skinned tectonics but the flanks of salt allochthons have such small areas that the normal
405
stresses induced by lateral tectonic forces may only become significant when applied to salt
406
structures that already have significant relief.
407
Theoretically, the thermal expansivity of salt is sufficiently higher than its compressibility that
-1
408
geothermal gradients quite normal in sedimentary basins (e.g. 30ºC km ) could drive thermal
409
convection within salt layers with a viscosity of 1016 Pa s and thickness >2.6 km (Jackson et al.,
410
1990, fig 3.8). Such internal circulation could be either immediately beneath the surface (e.g. in
411
Afar: Talbot, 1978, 2008) and offer an accessible analogue of thermal convection in the mantle, or
412
beneath insulating layers of other rocks; however, no such movements have yet been
413
demonstrated. The vortex coils in the mushroom shaped diapirs of Iran are difficult to explain
414
without invoking a component of thermal convection (Jackson et al., 1990) but thermal loading
415
probably usually merely plays an additional subsidiary role in most other salt movements.
416
Ice is too compressible to convect on Earth. However, the salt mushrooms in central Iran were
417
quoted as analogues of the cellular pattern of gravity overturns of three layers of ice in the ~30 km
418
thick crust of Triton (Schenk & Jackson, 1993). Similarly, a space filling pattern of pits, spots and
419
domes (5- to 30-km-in diameter) on the sparsely cratered ice-rich surface of the Jovian satellite
15
420
Europa have been attributed to solid-state thermal convection in the ice shell as a result of tidal
421
heating (Pappalardo et al., 1998), a concept since partially tested by numerical models (Showman
422
& Han, 2004).
423
Largely to counteract pre-1988 ideas, Hudec & Jackson (2007) argued that buoyancy is
424
unimportant for initiating salt diapirism and, in effect, only converts reactive diapirs of sufficient relief
425
into active diapirs. They also argue that “Most salt diapirs in the world pierced during extension
426
events” (Hudec & Jackson, 2007, p.26). However, no large scale diapirs feed terrestrial namakiers
427
in modern extension settings (Mohr et al., 2007) which is why Fig. 4B illustrates a reactive diapir in
428
the locally oblique extensional setting that is usual in Iran which is subject to regional N-S
429
shortening.
430
431
3.3 How important are lateral forces?
432
Two forces are usually invoked as drives for the extrusion of emergent salt diapirs: gravity and
433
lateral tectonic forces. However, noting that most salt sheets form in compressed orogens and the
434
compressed toes of passive continental margins, Hudec & Jackson (2006) went further and
435
suggested that salt sheet emplacement requires lateral compression.
436
The effects of gravity on salt sheets can only vary at the rates that sedimentation can thicken
437
the overlying rocks or erosion can thin either them or the salt. Typical sedimentation and erosion
438
rates are a few m Ma , about the rate that subsurface salt diapirs rise. One might expect, therefore,
439
that lateral tectonic forces vary faster than the effects of gravity. However, such effects are not
440
visible in the 3 best-documented salt extrusion rates in the literature.
-1
441
Arguing that the heights of salt mountains are potential tectonic strain gauges, Talbot & Alavi
442
(1996, fig. 9) used them to map active faults in part of the Zagros Mountains laterally shortening
443
where Arabia and Asia converge. However, most of the salt extrusions in the Zagros are no higher
444
than the highest peak of country rocks nearby- implying that gravity is their main drive (Talbot
445
1998). The only Zagros salt extrusions that rise higher than their surroundings are the offshore salt
446
islands close to the Zagros deformation front but even these could be supported by dense
447
carbonates loading their deep salt source.
448
The heights of salt fountains as tectonic gauges are complicated by the climatic signal (Talbot
449
and Jarvis, 1984). More direct gauges of the forces extruding them are emergent salt diapirs that
16
450
have had their extrusive salt removed or bevelled by sea-level at known times. This happened to 3
451
salt diapirs emerging close to two of the boundaries of the lithospheric plate capped by the Zagros
452
Mountains.
453
Mount Sedom, a diapiric salt wall extruding from the Dead Sea oblique-extension fault zone, was
454
turned into such a simple tectonic gauge when the precursor to the adjoining Dead Sea removed
455
any earlier extruded sat and trimmed its crest to a smooth flat surface at ~14 ka. Using geological
456
markers, precise ground leveling and Synthetic Aperture Radar interferometry, Weinberger et al.,
457
(2006) were able to constrain the rates of rise of different parts of Mount Sedom to between 5 and 9
458
mm a over the last 14 ka. After carefully modeling the Sedom architecture, these workers
459
attributed the steady salt extrusion solely to gravity without any need for additional lateral forces.
460
Similarly, a small group of diapirs of a much older salt in and beside the Hormoz straits on the
-1
461
southern contracting boundary of the same lithospheric plate were also converted into simple
462
tectonic gauges when they were partially truncated by the Persian Gulf at 9 ka BP. Using the
463
current altitudes of dated oysters and bivalves that lived close to sea level in the sediments
464
deposited on these now-tilted Holocene marine terrace, Bruthans et al., (2006) were able to
465
constrain the rates of rise of the centres of two such diapirs to 7 ±1 mm a-1 over the last 9 ka BP.
466
Despite involving salt of two different ages, all the three above diapirs are reported to have
467
extruded at the same steady rates (near 7 mm a-1) for several millennia. This is particularly
468
remarkable for Mount Sedom that extrudes from the Dead Sea oblique-extension fault zone where
469
major earthquakes associated with surface faulting were common in the two millennia before 1894
470
but have been absent since (Ambraseys and Jackson, 1998, cf. figs 4 & 5). This is one of the few
471
documented cases where the historical record indicates a possible change in either the direction or
472
values of lateral tectonic forces. The steady extrusion of Mount Sedom from a fault along which
473
major earthquakes stopped ~200 years ago implies that gravity is its main driver. Hereafter we will
474
assume that the influence of lateral tectonic forces and their changes on salt extrusions are
475
sufficiently small that they can be neglected. We also consider it unlikely that lateral forces in their
476
bedrocks are likely to be transmitted into ice bodies.
477
478
479
17
480
4. Compositional layering, inclusions and moraines
481
Apart from falls of volcanic tephra that sometimes add spectacular competent marker layers to
482
ice masses, most primary stratification in ice is picked out by differences in grain-size and bubble-
483
content surviving from layering in the initial snow pack and such diagenetic processes as refreezing
484
of ice due to raditative and advective heat events (Hudleston and Hooke, 1980). The resulting
485
layering is comparatively subtle compared to stratification in salt, largely because fewer impurities
486
accumulate with snow than commonly accumulate with salt.
487
While many salt sequences are interlayered with other sediments or igneous rocks that can
488
range in thickness from mm to km, beds of relatively pure, clear and transparent salt often reach
489
thicknesses of hundreds of m. Stratification, layering or bedding in such pure salt is defined by
490
halite of different grain size and colours due to various concentrations of different impurities (e.g.,
491
magenta clay, buff-coloured anhydrite, black iron or green volcanic dust). The colours in extrusive
492
salt sheets generally intensify down their exposed lengths as dissolution concentrates the insoluble
493
components. Even in salt that may have flowed 10 km along the source layer, perhaps 5 km up the
494
diapir, and 5 km downslope over the surface, compositional layering in distal salt is generally
495
accepted as inherited from the initial bedding and can look like simple planar bedding until colour
496
repetitions and rare isoclinal fold hinges attest to the huge strains the salt has undergone (Talbot,
497
2004).
498
Like ice, exposed salt bodies develop nearly every known category of karst feature although
499
subsurface stream channels erode down to base drainage levels rather than the base of salt (e.g.
500
see Bruthans et al., 2007)
501
Unlike ice, salt extrusions rarely pluck, quarry or incorporate their surrounding strata so most
502
inclusions (that can range in siz from km to microscopic) are part of the initial salt sequence
503
disrupted and dispersed by salt flow in the diapir and are equivalent to en-glacial moraines
504
(Weinberg, 1993). Denser than salt, even the largest solid inclusions (e.g. 3 x 2 x 03 km- Kent,
505
1958) carried >3 km up active diapirs in Iran can be stranded on the modern surface or choke the
506
vent when the salt dissolves (as in the Flinders Range of, Australia: Dyson, 2004). Large dense
507
inclusions lifted high in an active diapir can sink back through otherwise inactive salt diapirs (Koyi,
508
1991; Chemia et al., 2008).
18
509
By contrast, submarine salt extrusions are often buried by beds of younger rocks that can be
510
disrupted by salt flow to simulate supra-glacial moraines that are often subsequently shed as
511
complex slumps equivalent to side or terminal ice moraines (see Fig. 1SA). Inclusions concentrated
512
at the sutures of submarine salt canopies help to delineate where salt sheets have coelesced
513
(Rowan, 2003).
514
Push moraine-type structures with simple thrusts and/or imbricate thrust complexes are
515
common in the soft sediments in or beneath the fronts of salt sheets in the Gulf of Mexico (Jackson
516
and Hudec, 2004). However, subaerial salt sheets do not erode U-shaped channels or quarry and
517
abrade such distinct bed-forms as roches moutonée or striations like ice because they attach to
518
their substrates along frictional contacts. Both ice and salt are too weak to abrade hard substrates
519
significantly but, more importantly, because most extrusive salt is dissolved by rain falling from
520
above, potential abrasive tools concentrate on the tops of namakiers rather than their bases (Fig.
521
1SA). This means that the only saline equivalents of sub-glacial moraines are shed and overridden
522
en-and supra-glacial moraines. Saline equivalents of eskers are likely to exist locally but have not
523
yet been recognised as such.
524
There also appear to be very few saline equivalents of the glaciotectonic structures in weak sub-
525
glacial substrates or moraines (Boulton, 1996; Maltman et al., 2000, p.4). What appear to be
526
ground moraines (Fig. 1SA) are equivalent to “melt-out” tills that remain where supra-salt moraines
527
that cloak large areas of most salt extrusions are dropped after eventual salt dissolution (Fig. 1SA,
528
Talbot, 1979, 1998). Conical mounds of loose residual cap soils-and-blocks accumulating on top of
529
dissolving salt commonly become badlands with steep slopes scarred by landslides.
530
Former moraines and ice-sculpted features are the main records of the many past ice ages. The
531
equivalent of erratic blocks and melt-out tills remaining after the complete dissolution of past
532
extrusions of salt in Iran are recognisable as black erratic blocks ± red soils spread over the plains
533
(Fig. 1SA) or interbedded among the surrounding buff coloured carbonates Such insoluble residues
534
in the stratigraphy of subsequently downbuilt sediments show that many of the salt diapirs currently
535
active in the Zagros Mountains first began extruding as long ago as the Triassic, 220 Ma ago
536
(Player, 1969; Kent, 1970; 1979).
537
538
There is evidence that evaporites formed in Archaean times (Warren, 1997). However, because
salt is so weak and soluble, it is usually recycled back to the world’s oceans as the surrounding
19
539
rocks pass through diagenetic into metamorphic facies. The salt is dissolved while it is still in place
540
or after it has been squeezed to the surface by the early stages of orogeny. Just as there were
541
many former ice ages, many former salt sequences have already been recognised from their
542
metamorphosed remnants or their mega-breccias after dissolution (see Warren, 1997; Jackson et
543
al., 2003).
544
545
5. Internal fabrics, structures and deformation mechanisms
546
547
Deformation structures and fabrics on all scales in glaciers and namakiers resemble those in
548
more usual metamorphic rocks and naturally model them at rates that can be monitored over a few
549
years.
550
5.1 Grain-scale fabrics
551
The hexagonal crystal lattices in a grain of ice are mechanically anisotropic, while the cubic
552
lattice of a halite grain is comparatively isotopic. Initially isotropic materials (like snow) develop
553
anisotropic (mainly subhorizontal) fabrics (as in new glacier ice) by homogeneous compaction and
554
planar markers (like stratification) can develop structures (like folds) by inhomogeneous strains.
555
Markers are strain active where they involve materials that deform at different rates (across
556
contacts that refract foliations) or strain passive where the strain rates of any distinguishable
557
components are negligible (e.g., colour bands).
558
Grain boundaries in flowing ice and salt can readily dynamically recrystallise and anneal (Fig.8).
559
As a result, grain shapes in deformed ice and salt (Table 1) are said to have short strain memories
560
because they record only the latest stages in their strain history (Talbot and Jackson, 1987). The
561
strain memories of grain fabrics in ice are even shorter than those in salt. This means that both ice
562
and salt that can look much the same however far it has flowed. Thus drill cores of salt (~170 Ma)
563
that has probably flowed distances approaching 100 km near the floor of the Gulf of Mexico can still
564
display horizontal beds of uniform thickness made up of isotropic grains cm across (Talbot, 2004)
565
just as distal and proximal ice in an Antarctic ice sheet can be indistinguishable.
566
Fabrics in salt deformed in different environments show different dominant deformation
567
mechanisms (Fig. 8) so their rheological behaviour should be described by different flow laws
568
(Schleder & Urai, 2008). Thus for still-allochthonous (Zechstein) salt, the subgrain sizes indicate
20
569
stresses between 0.6 and 1 MPa at depths implying deformation temperatures <50ºC (Schleder &
570
Urai, 2008). In thin sections, patches of primary crystallisation structures full of dispersed primary
571
brine inclusions are surrounded by areas of recrystallised grains with evidence of grain boundaries
572
that migrated to collect fluid inclusions and become very mobile by solution-precipitation creep.
573
Solution-precipitation creep and dislocation creep contribute almost equally to the total strain rate
574
and the equation describing the salt rheology should reflect this.
575
Diapiric salt tends to consist of grains of uniform size (e.g. 0.5-5 cm) but irregular isotropic
576
shapes implying that, like ice, it can flow large distances with a steady-state dynamic recrystal-
577
lisation fabric involving mainly dislocation creep and fluid assisted grain boundary migration. Every
578
grain is competent with respect to its neighbouring grain boundaries and so undergoes the
579
oscillating super-shear strains of stiffer inclusions rotating with vorticity higher than the bulk flow
580
(Ramberg, 1975; Ramsay and Huber, 1983, Talbot and Jackson, 1987; Talbot, 1992, 2004). The
581
application of experimentally derived flow laws relating the steady-state size of sub-grains with flow
582
stress indicates that diapiric salt deforms by climb-controlled dislocation processes at rates
583
between 10
584
40-150ºC (Schleder and Urai 2006, 2008).
585
-11 to 16
s
–1
at differential stresses 1-5 MPa at depths implying deformation temperatures
Diapiric salt with its typical course uniform grain size is too strong to flow in subaerial surface
586
conditions. Extrusive salt in Iran exhibits the necessary weakening process in fabrics that have
587
bimodal grain sizes with an average grain size that decreases downslope (Fig. 8SD-SE). Such
588
bimodal grain size fabrics are rare in ice and render shape and orientation grain-fabrics in salt much
589
more obvious than most secondary foliations picked out by discontinuous layers of different grain
590
size and bubble content in ice (Hambrey and Lawson, 2000). The bimodal grain sizes of extrusive
591
salt means that foliations can appear with new orientations (e.g. axial planar to new folds) over
592
distances as short as a few dm (Talbot, 1979). As in all rocks, foliations develop parallel to the
593
longest and intermediate axes of the cumulative strain ellipse that fomed them (Hambrey &
594
Lawson, 2000).
595
Salt extruding from a diapiric vent surfaces with both a strong a horizontal stratification and
596
gneissose foliation as a result of strong lateral extension and vertical flattening in the crest of the
597
summit dome (Fig. 7). Halite megacrysts (cm across, e.g. Figs. 8SA-SE) surviving from the diapir
598
exhibit very small subgrains indicating dislocation climb at stresses 3-5 MPa in the diapir. These
21
599
porphyroclasts decrease in size and number downslope as they disperse within increasing volumes
600
of smaller (< a few mm) daughter grains shed from their cores. The lack of subgrains in the small
601
daughter grains indicates flow by pressure solution at differential stresses of <<1 MPa by pressure-
602
solution, at temps from 20-40ºC implying effective viscosity values as low as 1014 Pas (Schleder &
603
Urai, 2006, 2008). Dynamic recrystalisation decreases the grain size of both salt (Fig. 8 SE) and ice
604
in shear zones (Figs. ID & IE, Hudleston and Hooke, 1980; Wilson and Zhang, 1996).
605
Deformation fabrics in extrusive salt can thus be defined by the shapes of both the large
606
porphyroclasts and the orientations of the surrounding smaller grains of halite ± spicules of
607
anhydrite (CaS04) or specularite (Fe2SiO3). Shape and orientation fabrics in the porphyroclasts are
608
more obvious (Fig. 8SD) but the fabrics in the finer grained groundmass are often stronger.
609
Nevertheless, grain shape fabrics are close to isotropic over most bodies of flowing salt (Fig. 8SC)
610
and become noticeably anisotropic only locally. Halite grains with length/width ratios of <1.5:1
611
characterise zones of salt gneisses tens to hundreds of metres wide (Fig. 8SB-E), and smaller
612
halite grains (<~0.2 mm) with length/width ratios up to 5:1 characterise salt mylonites that may be
613
hundreds of metres long but are seldom >1 dm wide (Fig. 8SE, Talbot 1979, 1981). Lacking
614
crystallographic preferred orientations, the micro-fabrics in halitic mylonites indicate that they
615
deform by solution precipitation creep involving non-conservative grain boundary migration and
616
grain boundary sliding (Schleder and Urai, 2006). Sub-grain sizes indicate that such mylonites are
617
capable of the rapid strain rates of 10
618
temperatures of ~10 to 40ºC found by field measurements of markers on namakiers after recent
619
rain (e.g. Talbot and Rogers, 1980; Talbot et al., 2000).
620
-10
-1
s at differential stresses of 0.4 MPa and surface
The simple listing of physical properties (Table 1) obscures some of the complexities that must
621
be borne in mind when considering the deformation of natural bodies of ice or salt. Thus, both ice
622
(Hambrey & Lawson 2000) and salt (Critescu & Hunsche, 1998) deform with bulk properties that
623
depend, on their purity, grain size, water content, and fabric as well as the PT conditions; scale also
624
becomes important if strain inhomogeneities such as shear or mylonite zones are included within
625
the volume of interest.
626
On scales >1 km, the smooth outlines of both intrusive and extrusive salt structures (Figs. 5-6),
627
together with the parabolic profiles of Holocene terraces on diapirs truncated by marine erosion
628
(Bruthans et al, 2006), indicate that confined by loads >~1 Mpa salt deforms as a linear viscous
22
629
crystalline fluid (n=1) consistent with model formulae indicating effective viscosities in the range
630
1016-18 Pa s (Critescu & Hunsche, 1998). Future formulae for the mechanics of rock salt flowing
631
from the conditions of alllocthponous source depths through diapirs to namakiers will have to
632
extend the range of effective viscosities to at least 1014-19 Pa s (Schleder & Urai, 2008).
633
Because glaciologists usually treat ice as a strain-rate-softening power-law crystalline fluid with
634
a stress exponent near 3 since Nye (1952) they are wary of quoting even effective viscosities.
635
Shumskii (1964) wrote of 10
10-11
14
Pa s for “warm” ice and 10
Pa s for “cold” ice.
636
The not-quite parabolic arcs of insoluble moraines on the tops of namakiers (Fig. 2A) can be
637
compared to plug flows along channels. The best-fit between such arcs and the non-dimensional
638
velocity curves for channel flow (Turcotte and Schubert, 1982) implies that gneissose salt on scales
639
of hundreds of m flows as a strain-rate-softening power-law fluid with a strain exponent near 3
640
(Talbot and Jarvis, 1984), as commonly taken for ice (Nye, 1965; Walker and Waddington, 1988;
641
Hambrey & Lawson, 2000 but see Harper et al., 1998). On still smaller length scales (<dm), the
642
power-law exponent of dilated salt increases to account for the n= 6-8 measured in mine, borehole
643
and laboratory tests (Cristescu and Hunsche, 1998). We will return to strain rates when considering
644
surging flows in section 9.
645
646
5.2 Macroscopic structures
647
From now on we will attempt to distinguish the deformation structures and fabrics that develop
648
within visco-elastic fluids flowing down gravity-driven pressure head gradients into three categories
649
on the basis of their origins:
650
Kinematic deformation structures and fabrics are those that develop where passive markers
651
(like stratification and foliation) develop new structures where they adapt to changes in the
652
geometries of their external boundaries they flow past. Examples are where passive markers and
653
flow lines in ice or salt remain parallel but develop forced or drape folds over irregularities in their
654
basal boundaries or follow bends in the rigid boundaries of a channel.
655
Dynamic deformation structures (± fabrics) are the result of inhomogeneous strains imposed on
656
strain-active markers by differential stresses. Examples are buckles in competent markers
657
shortened along their length, pinches or boudins developed in competent markers extended along
23
658
their length, mullions in incompetent markers shortening along their length and “inverse folds” in
659
incompetent markers extened along their length.
660
Lateral tectonic forces are more likely to strain salt bodies (e.g. by changing channel gradient)
661
than shorter-lived ice bodies. Nevertheles, regional stress fields (dominated by gravity) generally
662
impose longitudinal extension in the upper reaches and longitudinal compression in the lower
663
reaches of both glaciers (Hambray & Lawson, 2000) and sheets of allochthonous salt.
664
An important sub-category of dynamic structures (± fabrics) can develop where local changes
665
in velocity or volume induce flow lines to cross passive markers. The most obvious way to change
666
local volumes or velocities is to invoke local changes in the budget of supply relative to loss of the
667
flowing fluid. As a result, such structures will be called here budget change induced structures (±
668
fabrics). Examples are recumbent flow folds induced in ice or salt by budget changes (Hudleston,
669
1976).
670
Dynamic structures (± fabrics) in another sub-category develop where flowing fluids encounter
671
(or are subject to) changes in frictional resistance along parts of their external boundaries. Such
672
change in the forces retarding flow can be local iin time or space. Examples of such changes in
673
space are where flowing ice detaches from a rock substrate and thins as it accelerates over a free-
674
slip base with water; another is where salt in high frictional contact with the surrounding rocks
675
extrudes into air. Examples of changes of boundary conditions in time are where batches of surface
676
melt-water hydraulically fracture their way to the base of a glacier or ice sheet (Sohn et al., 1998) or
677
where rain weakens the top boundary of a namakier. We refer to these as boundary induced
678
budget change structures.
679
In contrast to the short strain memories of grain shape fabrics in deformed ice and salt,
680
macroscopic markers like stratification and veins, have indefinite strain memories and can record
681
large cumulative strains that can be lost by ablation, melting or dissolution, but are unlikely to
682
anneal (e.g. for salt see: Talbot and Jackson 1987). Most macroscopic strain markers in ice are
683
more or less strain active because compaction and diagenetic layering increasingly orientates the
684
bulk mechanical anisotropy of their constituent ice crystals. Thus stratifications with conformable ice
685
lenses are slightly strain active, crevasses, crevasse traces or infilled-crevasses are moderately
686
strain active and dirt bands, silt beds and tuff layers are strongly strain active.
687
24
688
689
5.3 Macroscopic kinematic structures
Stratification in both ice and salt form horizontal and usually first deform by flowing down nearly
690
horizontal load pressure gradients. Clear 3D grain shape and orientation fabrics with short strain
691
memories map the flow lines. They generally define a planar foliation ± a lineation although the
692
lineation can dominate in parts of some salt structures (e.g. constricted in the stem of a diapir or
693
rolled around the hinges of folds in extruded salt: Talbot and Jackson 1987). In steady penetrative
694
flows, flow lines and the foliation parallel stratification and their nearest external boundaries and
695
remain parallel as they sweep together around changes they encounter in their external
696
boundaries, even right angle corners (e.g. in planform for glaciers; from source-to-diapir and diapir-
697
to-extrusion in profiles of salt). The top boundary is generally a smoother version of the often-
698
irregular bottom boundary in both namakiers and glaciers. The infills of moulins and crevasses in
699
ice can develop vertical foliations (Hooke and Hudleston, 1978).
700
The stratification in ice converging on a narrow channel from several accumulation basins, or
701
downstream from nunataks (bedrock highs exposed through ice), tends to develop upright
702
symmetrical kinematic folds with subhorizontal axes and a vertical axial planar foliation (Fig. 2ID-E,
703
Hambrey et al., 1999; Iizuka et al., 2001). Equivalent kinematic folds and foliations develop in the
704
deep source layer where flowing salt converges radially on the base of the diapir stem (Ramberg,
705
1981; Talbot and Aftabi, 2004). The radial and subhorizontal axes of these folds turn through a
706
kinematic bend to become vertical up the diapir where they are known as curtain folds (Talbot and
707
Jackson, 1987). These curtain folds emerge built into the salt that first surfaces on the crest of a
708
salt extrusion. Rather than open in the divergent radial flow within the summit dome, the steep axial
709
surfaces of curtain folds turn through another kinematic bend (related to the divergent flow points
710
starred in Fig. 7) to parallel the top free surface of the extruding salt (Talbot and Aftabi, 2004). As a
711
result, unlike ice that usually only accumulates strain in its compaction and downslope flow, most
712
salt is extruded from depth reaches the surface with mature folds already incorporated into the
713
stratification. Indications of these folds can persist into the most distal salt as local layers with steep
714
along-flow strikes.
715
716
717
25
718
719
5.4 The kinematic tank-track fold
Analogue experiments (Ramberg, 1981, pp. 219-226) studying how gravity spreads rectangular
720
blocks of elasto-viscous silicone putty with different length/ height ratios over frictional and no-slip
721
basal boundaries are relevant to the effect of gravity spreading bodies of both salt and ice. Where
722
blocks of silicone putty spread over a horizontal no-slip basal boundary, their initially vertical sides
723
bulged and then advanced by rolling their tops over the front and attaching to the substrate (Fig.
724
9A1-3). As a result, initially-horizontal passive marker layers near the distal edges of experimental
725
viscous flows advance like lavas and caterpillar tank-tracks. The top surface overtakes the toe, rolls
726
down a steep flow front (and over any debris spilled from it) and attaches to its substrate along a
727
frictional basal boundary that lengthens with further advance. Passive stratification in the flow rolls
728
over the hinge of a kinematic tank track fold with an axial surface that climbs slightly as it elongates
729
behind an axis parallel to the the advancing flow front (Figs. 7 and 9 and Talbot, 1998).
730
Stratification that faces upward in the upper limb (thinned to invisibility in Figs. 9B-F) is inverted to
731
face downward in their strongly sheared lower limbs.
732
In extrusion experiments (Talbot and Aftabi, 2004), the upper limbs of tank-track folds are
733
simple if they advance steadily over planar horizontal substrates (Fig. 9B) but develop flow folds
734
with geometries that relate to changes in the relative rates of extrusion and gravity spreading (Fig.
735
9C-F). Despite significant dissolution from their top surfaces, every subaerial salt extrusion studied
736
in Iran involves a fundamental recumbent sheath-like tank-track fold. In marked contrast, major
737
recumbent folds have been reported from very few distal profiles of glaciers and ice caps and those
738
have been attributed merely “to ice flowing fastest at mid-levels” (Shumskii, 1962).
739
Hudleston (1976) described discontinuous examples of recumbent folds near the base of the
740
Barnes ice cap, listed a few other examples in the literature, and expected such folds to be more
741
common than then generally recognised- but very few additional examples have been added since.
742
Chinn (1989) attributed single folds in short dry-based glaciers to their advance- but invoked basal
743
buckling and thrusting of a sub-glacial till rather than the upper limb of a supra-glacial till rolling over
744
the advancing front and then being truncated by ablation. Recently several authors have
745
investigated the potential for folding near the base of ice caps in order to understand the deposition
746
and deformation history in ice cores taken from ice caps (e.g. Jacobson and Waddington,
747
2005).
26
748
A clue for this absence comes from blocks of silicone putty spreading over free slip basal
749
contacts in the laboratory. These spread rapidly over mercury and thinned uniformly by almost
750
uniform pure shear so that passive horizontal markers curl upward in distally as a result of their toes
751
spreading faster than their tops (Figs. 9A4-6). Whereas stratifications curling upward distally are so
752
far unkown in subaerial namakiers that have advanced over frictonal substrates of rocks and soils,
753
they appear to be the norm for profiles of glaciers and ice sheets (Hambrey and Lawson, 2000)
754
suggesting that they have low basal friction (upstream of a periphery with high basal friction in cold
755
examples). Seismic profiles rarely image stratification within submarine salt sheets but the most
756
distal few km of their bases curl upward above simple thrusts and/or imbricate thrust complexes in
757
some of the frontal sediments beneath the Gulf of Mexico (Jackson and Hudec, 2004). Submarine
758
salt sheets probably advance faster and further over saturated sediments than subaerial namakiers
759
usually flow over dry rocks and soils. In summary, we infer that stratifications curling downward
760
distally signal ductile flow over high-frictional basal contacts, whereas stratifications curling upward
761
distally signal flows over low-frictional basal contacts.
762
763
5.5 Kinematic structures associated with irregular bedforms
764
In addition to their fundametal tank-track folds, namakiers with irregular beds have fairly well
765
developed basal zones of folds ± thrusts like many glaciers (Shumskii, 1964 p. 444). Most of these
766
structures are dynamic and treated in the following section (5.6); here we mention kinematic
767
undulations where the fluid adapts to irregularities in its bottom boundaries. Both ice, salt and
768
viscous analogues develop internal upright kinematic similar-type folds (class 2 folds of Ramsay
769
1967) wherever gravity drapes their internal structures over irregularities in their beds (Figs. 7C &
770
9F). Such kinematic “undulations” are characterised in vertical profiles by upright folds with
771
amplitudes that diminish upward from the base but can reach the top free surface as ogives. Even
772
the stratification of near-basal ice beneath the current divides of ice sheets develops such upright
773
folds over bed relief (Souchez et al., 2000; Waddington et al., 2001; Thorsteinsson and
774
Waddington. 2002; Jacobson and Waddington. 2005). Equivalent structures in other rocks are
775
known as compaction, drape or forced folds (Cosgrove and Ameen, 2000). The dynamic folds of
776
the next section can appear upright in profiles perpendicular to the direction of flow but are
777
asymmetric in other sections.
27
778
5.6 Budget change induced flow folds and their axial planar foliations
779
Axial surfaces to drape folds that are vertical in near-basal ice can curve upward to become
780
axial planar to asymmetric reclined, even recumbent folds, where carried downstream. Anatomising
781
stripes, planes of ice grains with c-axes approximately in the dip direction of the planes, are
782
commonly axial planar to both deep symmetric upright folds and less deep asymmetric folds (Alley
783
et al., 1997). Such asymmetric folds in ice are dynamic in the sense that they record variations in
784
flow velocity in either space or time. Hudleston’s (1976) explanation for how such folds develop in
785
glaciers when the flow is unsteady on a wide variety of time scales is equally applicable to
786
namakiers. Folds develop wherever and whenever flow trajectories cross passive markers. Minor
787
advances and retreats change the thickness and/or surface slope of glaciers in time so that
788
markers already draped over an irregular base deform into increasingly asymmetric folds as they
789
are carried downstream. Complex budget changes over complex basal irregularities can lead to
790
multiple recumbent folds in distal ice and salt (Fig. 10).
791
It is not clear whether all folds verging downstream in ice are deformed upright folds or if any
792
can initiate as such (Alley et al., 1997). However, many folds within namakiers certainly initiate with
793
asymmetries, perhaps because the bimodal grain size of extrusive salt allows the generation of new
794
foliations to record a new set of stream lines over only a few dm (Talbot, 1979).
795
Our understanding of how dynamic flow folds generate near the base of flowing fluids benefited
796
considerably from two linked papers by Gudmundsson (1997a & b). In this paragraph (only) we use
797
the term “extrusive flow” in the usual glaciological sense of referring to increases in horizontal flow-
798
velocities with depth. Gudmundsson (1997a) introduced a form of Local Extrusion Flow (LEF)
799
where the vertical expansion and contraction of the ice changes laterally close to an undulating
800
free-slip bed so that LEF above the trough dominates that above the ridge. Rather than adding to
801
ice flux as considered normal for the general extrusive flow of the preceding glaciological literature,
802
Gudmundsson's (1997a) LEF instead decreases ice flux by the general flow separating from, and
803
flowing beyond, ice stagnant in troughs. The effect of sinusoidal bed relief is largest a distance
804
above the bed that relates to the depth/width ratio of its relief. In appropriate conditions, a high-
805
pressure zone can develop above the bed and superpose a Poiselle flow on the gravity driven
806
plane flow. Gudmundsson (1997b) used numerical models to show that LEF increases with the
807
power law exponent of stress. (When and where the depth of a trough is >1.5 times its along-flow
28
808
width, the main downslope flow to the right will separate from, and flow over ice that remains in the
809
trough and is either stagnant or slowly circulating in a clockwise corner eddy). Multiple flow folds
810
can develop where flow perturbations have higher harmonics than the bedrock topography inducing
811
them, as found empirically by Hooke et al., (1987) and experimentally by Reeh (1987).
812
Similar phenomena have been described from a different viewpoint when relating trains of
813
asymmetric folds and slides in a namakier to step-like (rather than sinusoidal) irregularities in its no-
814
slip bed (Fig. 10 from Talbot, 1981). Rather than eroding obstructions, namakiers instead smooth
815
their channels by infilling irregularities with static salt and then detaching from and flowing beyond
816
the infills alnong near-basal mylonite zones.
817
To surmount a steep upstream-facing step in its bottom boundary, a namakier slows and
818
thickens. The gneissose grain shape fabric mapping the flow surfaces diverge in the decelerating
819
salt and cross the passive colour layers to become axial planar to a suite of new flow folds. These
820
similar-type folds amplify and tighten as the flowing salt detaches along mylonite shear zones and
821
flow up and over static veined salt immediately upstream of the stationary step (Fig. 10B). The
822
stratification curls upward distally as though the near-basal mylonite zones act as low frictuional
823
detachments between the ice above and below them. Downstream of the obstruction, the foliation
824
converges in the accelerating salt, the folds tighten to isoclinal, and the mylonite zones that acted
825
as thrusts upstream, turn to slides (extending detachments) downstream (Fig. 10B). A steep strike-
826
slip shear zones marks where an active salt stream by-passes salt now damed by bedrock ridge in
827
response to budget changes (Talbot 1979).
828
Although folds of most categories in salt structures have been simulated in laboratory
829
experiments with viscous ductile fluids, trains of dynamic folds relating to irregularities in the bottom
830
boundary of namakiers (Talbot, 1979) have not (Talbot & Aftabi, 2004 p. 331).
831
832
833
5.7 Budget induced dynamic folds near top boundary
Mapping the vergence of minor folds and how axial planar foliations relate to stratification implies
834
that major similar-type folds face outward in the upper slopes of salt fountains. All subaerial salt
835
extrusions studied in Iran are characterised by these major flow folds in the upper limbs of their
836
tank-track folds (Fig. 7A-D). Absent from the crest of the dome, these folds first appear as gentle
837
and open reclined asymmetric flow folds in layering on the steep (>15°) upper slopes of the dome.
29
838
Axial surfaces dip steeply upslope where they first develop but flatten in successive folds
839
downstream as they amplify to isoclinal recumbent folds in distal salt (Fig. 7C-D). Their axes
840
parallel the flow front and therefore sub-parallel smoothed topographic contours on the salt. These
841
similar-type flow folds are salt equivalents to the ductile fold-thrust nappes that characterise the
842
metamorphosed cores of orogenic belts of all ages and, like Helvetic-type nappes, their lower limbs
843
often detach along mylonitic shear zones or slides (e.g. Pfiffner, 1993). No equivalent folds appear
844
to have been reported from ice.
845
It is easy to simulate viscous extrusions in the laboratory by loading a horizontal wooden board
846
(representing stiff overlying rocks) floating in a pan of viscous fluid (e.g. polydimethylsiloxane GM-
847
26, a transparent silicone polymer with a viscosity of ~ 5.105 Pa s in Talbot and Aftabi, 2004). The
848
sinking board displaces the viscous fluid (representing the salt) up around its edges (from where it
849
is removed) and up a central hole that simulates the diapiric vent of a salt extrusion. Thin horizontal
850
colour bands built into the source layer represent passive stratification in the salt.
851
If suitably small loads are applied to the board, the fluid rises slowly up the rhomboidal vent and
852
extrudes over the board with the profile of a droplet (Huppert, 1982) or a miniature ice cap (dashed
853
in Fig. 6B). Simulating realistic loads on salt layers buried beneath overburdens a few km thick
854
lead to the extrusion of a dome that spreads as an axisymmetric viscous fountain with a diameter of
855
about a dm after 6 hours and 2 dm after 24 hrs.
856
In the laboratory, it is easier to change the rate of viscous extrusion than to change the rate at
857
which it gravity spreads or is removed by erosion. Thus the viscous fluid can be extruded in pulses
858
separated by brief intervals without a load when the spreading rate of the extrusion changes
859
comparatively little (Talbot and Aftabi, 2004). The number of outward facing folds in the upper limb
860
of the basic tank-track fold is a direct measure of the number of loading phases (Fig. 9B-F). This is
861
because each renewal of the load significantly increases the extrusion rate so that resurgent flow
862
trajectories cross the passive markers re-orientated by the gravity spreading that continued during
863
the pause in loading. The resultant flow folds therefore have budget-induced dynamic origins. The
864
relative lengths of their lower limbs reflect intervals of increasing extrusion rate (or slowing gravity
865
spreading in nature) and the lengths of their upper limbs reflect intervals of falling extrusion rate (or
866
accelerating gravity spreading in nature). Early flow folds rolled over the axis of the tank-track fold
867
are effectively unfolded by strong shear along the bottom boundary.
30
868
In essence, the laboratory loading history of an experimental viscous extrusion can be read
869
from the shapes of folds in the upper limbs of its tank-track fold (Talbot and Aftabi, 2004).
870
Generating experimental dynamic folds by varying the extrusion rate relative to the spreading rate
871
does not rule out the effects other mechanisms in nature. Indeed, the spreading rates of natural salt
872
extrusions depend on the climate and are thus likely to change much faster than changes in the
873
extruding forces. If the extrusion rates of emergent salt diapirs are essentially steady because they
874
are driven mainly by gravity (section 4.3), then the major dynamic flow folds within subaerial salt
875
extrusions in Iran are more likely to record significant (millennia-scale) changes in their rates of
876
gravity spreading than any slower changes in their extrusion rate. Because warm salt flows faster
877
than cold salt and, even more significant, wet salt flows faster than dry salt (Urai et al., 1986),
878
changes in the rate of gravity spreading of extruded salt depend mainly on the climate. Each of the
879
major folds within Iranian salt extrusions therefore presumably record wet intervals when the salt
880
gravity spread faster relative to its almost steady extrusion rate. The structures within salt
881
extrusions therefore offer proxy records of climate changes, particularly rainfall. The simplest test of
882
this hypothesis is that salt extruded from different vents in the same region over the same time
883
interval should have developed similar sequences of flow folds in the upper limbs of their tank-track
884
folds.
885
It is generally accepted that the rise of mountain chains can result in orographic rainfall and that
886
the resultant increase in erosion rate can feed back to influence the orogenic structures (e.g.
887
Beaumont et al., 2001). However, it is still debateable whether the dynamic flow folds represented
888
by piles of orogenic nappes record fluctuations in climate and/or tectonic forces.
889
Hudleston (1976) envisaged flow of ice developing folds where and when the flow lines cross
890
old stratification perturbed by relief in the bottom boundary. However, the clearer folds in salt flows
891
suggest that dynamic flow folds near the top boundary of flowing ice may be even clearer records of
892
budget changes due to fluctuations in climate. The potential for folds in ice to record climate change
893
has been recognised in ice streams (see e.g. Bindschadler, 1993; Jacobel et al., 1993; Merry and
894
Whillans, 1993; and Fahnestock, et al., 2000) but woefully neglected in ice sheets (Alley et al.,
895
1997) and glaciers (Hambrey and Lawson, 2000).
896
897
31
898
899
6. Fractured carapaces in ice and salt
Every exposed rock mass develops a dilated and fractured outer layer. This elastic carapace
900
is ~10-70 m thick in ice and ~5-10 m thick in salt. Most fractures in both materials initiate with
901
apertures less than a few mm: however, whereas many fractures in extending or shearing ice
902
develop significant gapes as crevasses, very few open in exposed salt (Talbot 1998).
903
Complex crevasse patterns that record strain histories in flowing ice (Nye, 1952) can be
904
emphasised by seracs or ice clefts sculpted by rapid ice sublimation. Dissolution by rain, snow melt
905
and streams (Bruthans et al, 2007) widens a few joints in salt (see photos in Talbot, 1998) but
906
generally frets a jagged salt relief of pinnacles that is largely independent of fractures (Fig. 1SB-E).
907
The topographic relief weathered into extruded salt in Iran increases down-slope from a minimum
908
(e.g., 5 m) near the summit to a maximum (e.g. 40 m) near the terminus (inserts on Fig.7B).
909
Bergschrunds are frequent forms in glacial settings, where they develop between the ice
910
body and the head wall in a combination of thermal and ice dynamics (e.g. Mair, and Kuhn. 1994).
911
Bergschrunds also gape behind a few local salt cliffs <10 m high that shed falls of m-scale blocks
912
every few hours or days (Talbot, 1979) but, in general, the only fractures that gape in salt define
913
top-surface-parallel slabs in about the top 10 m (Fig. 11). These define top-surface-parallel slabs in
914
the outermost ~10 m of all subaerial salt extrusions. When dry, these salt carapaces strain
915
elastically in immediate response to temeperature changes and fracture noisily during rapid
916
temperature drops but accumulate permanent ductile strains silently within minutes of light rainfall
917
(Talbot and Rogers 1980). Talbot (1998) assumed that the carapaces of brittle dilated dry salt
918
encasing Iranian salt extrusions are only veneers as thick as the local salt relief. However, master
919
fractures over a 100 m long suggest that regional stress fields have broken dry salt deep within at
920
least one salt fountain at unknown times in the past (e.g. Talbot and Aftabi 2004).
921
922
6.1 Fracture dynamics
923
Factures develop in elastic solids when and where stresses in them exceed their tensile or shear
924
strength. In ice, such stresses are usually generated by differential flow rates. In salt they, these
925
fractures are attributed to salt extruded from depths of km de-stressing to the free surface on
926
exposure with the aid of weathering processes (Talbot and Aftabi, 2004).
32
927
Monitoring the movements of markers on salt extrusions in Iran (e.g. Talbot and Rogers,1980;
928
Talbot et al, 2000; Talbot and Aftabi, 2004) finds them to move back and forth on time scales from
929
minutes to months or even (locally) years. The (unusually high) standard value of the linear thermal
930
expansivity of salt (Table 1) has been confirmed to an accuracy of a few % by translating to strains
931
the changes in the distance between two wooden stakes 2 m apart and relating them to the
932
temperature changes on the intervening dry ground over a few hours (Aftabi, unpublished MSc
933
thesis, 2000). These elastic strains indicate that the salt expands on heating and contracts on
934
cooling as expected; however salt masses also expand on wetting and shrink on drying. Marker
935
movements are therefore complicated because salt temperatures fall when it rains and rise when
936
sunshine dries the salt. However, in general, thermal expansion and contraction dominates the
937
swelling and shrinkage with wetting and drying. Large as they are, elastic strains are eventually
938
exceeded by ductile strains that accumulate when the salt is damp.
939
The ambient conditions of salt are usually warmer than ice at all depths but surficial salt is
940
commonly subject to larger diurnal and annual temperature changes than ice (30-50ºC). The
941
thermal expansivity of salt is an order of magnitude higher than most other rocks (including ice:
942
Table 1). Calculating the thermal skin depth (τκ/π)1/2 (equation 4.90, Turcotte and Schubert, 1982)
943
for ice and salt with a thermal diffusivity of κ, subject to regular temperature fluctuations with cycles,
944
τ indicate that the daily and annual temperature cycles reach depths of 19cm, and 3.66 m in ice and
945
32 cm, and 6 m in salt respectively.
946
Thermal strains and shocks induced in surficial salt by temperature cycles of about a year
947
therefore account for the outer 10 m or so of subaerial salt bodies in Iran being broken into brittle
948
elastic carapaces. When they are dry during winter sunsets, such salt carapaces sound like a
949
small-arms battles when new fractures develop as a result of rapid temperature falls. Surficial
950
glacier ice also snaps and cracks under thermal stresses but more impressive is the singing of lake
951
ice as it expands near noon in the warm Spring-time sun. Whereas the 10-70 m thick rigid surficial
952
crusts carried by flowing ice are there because they are cold, the 10 m thick crusts usual on salt
953
flows are fractured by falling temperatures but their yield strengths depend on water content. On the
954
way to considering the effects of water on ice and salt we consider their fluid inclusions.
955
956
33
957
7 Fluid inclusions
958
Most if not all mineral grains contain fluid inclusions that trap samples of any volatiles surrounding
959
them when they crystallised. In ice, these chemical and physical impurities are usually ejected from
960
the ice matrix by crystal growth, although some chemical species do dissolve with water in the
961
molecular structure (e.g. Legrand and Mayewski, 1997). Since the first ice coring experiments (e.g.
962
Langway, 1967), impurities in the ice matrix have been used as proxies of climatic and
963
environmental changes, Such studies have grown exponentially in the last two decades (e.g., Petit
964
et al. 1999, Alley, 2002) and a recent ice core from Antarctica documented the record of more than
965
eight glacial cycles over 900 ky (EPICA, 2004).
966
Primary and diagenetic textures in undeformed crystalline salt are distinctive of the environment
967
in which they formed and also trap fluid inclusions that sample the surrounding atmosphere or the
968
brines they evaporated from (Lowenstein and Hardie, 1985). Halite crystals that accumulate in
969
episodically desiccating playas typically grow vertically from horizontal substrates. Cubic inclusions
970
of the brines in which they crystallised are commonly incorporated in distinctive face-parallel
971
chevron-shaped growth zones that point upward where they are not interrupted by wavy dissolution
972
zones (stylolites) recording episodic additions of fresher water. In permanent brines, chevron
973
crystals of halite plus inclusions radiate from reef-like substrates and can grow more or less
974
continuously and very rapidly (Talbot et al., 1996).
975
Growth and cementation fabrics inherited from the initial crystallisation of salt can survive
976
indefinitely in salt confined by burial without flow. Studies of paleoclimatic records in salt lag behind
977
such studies in ice. Nevertheless, a 186 m core of salt interbedded with muds from Death Valley in
978
California documented a 200 ky paleoclimate record (Lowenstein et al. 1999). Similarly, a 106 ky
979
paleoclimate record has been recovered from a 100 m salt core from northern Chile and
980
demonstrated that wet conditions in tropical South America were coeval with cold sea surface
981
temperatures in the high-latitude North Atlantic on both millennial and orbital (20,000-year)
982
timescales (Bobst et al., 2001; Baker et al., 2001).
983
Paleoclimatic records from salt can be much longer than ice because salt can often survive
984
dissolution longer than most ice survives ablation or melting. Lowenstein and Demicco (2006) used
985
the co-precipitation of nahcolite (NaHCO3) and halite (NaCl) preserved in evaporites of the western
986
United States to infer that atmospheric CO2 was >1125 ppm (four times pre-industrial
34
987
concentrations) between ~56 and 49 Ma and to confirm that the highest prolonged global
988
temperatures of the past 65 million years were in the early Eocene.
989
In addition to paleoclimate studies, primary fluid inclusions in marine halites allow studies of
990
systematic changes in the chemistry of evaporated seawater going back to the late Precambrian
991
(~500 Ma; Lowenstein et al., 2001). These fluctuations are in phase with oscillations in seafloor
992
spreading rates, volcanism, global sea level, and the primary mineralogies of marine limestones and
993
evaporites.
994
995
996
7.1 Biological materials
It has slowly become clear in the last decade or so that micro-organisms are present in virtually
997
all rock samples examined for biological properties. Anomalies in the concentration and isotope
998
ratio of biogenic trace gases (e.g., CH4, N2O) in ice (Abyzov, 1993, Priscu et al., 2005) indicate that
999
some of the micoorganisms (that include fungi, bacteria, and viruses: Castello and Rogers, 2005) in
1000
solid ice and subglacial water to depths >3.6 km are living and reproducing (Sowers, 2001, Campen
1001
et al., 2003). The difficulties in sampling such extreme life-forms in ice and salt provide useful
1002
practise for sampling potential equivalents in the hundreds of lakes beneath the Antarctic ice and
1003
on extraterrestrial bodies. The subaerial biota on ice usually resides in small melt cups on bare ice
1004
called cryocontite (Säwström et al., 2004). The dark bacterial communities in these cryoconite
1005
holes can reach of 6 area % of glaciers in Svalbard and, by lowering their albedo, increase their
1006
rate of melting and so become a negative component of their mass balance.
1007
Rather than increase with the age of the surrounding ice iin a core, the numbers of recoverable
1008
bacteria correlate with the climate during deposition of the initial snow. Cores from low-latitude
1009
glaciers contain more bacteria than polar ice core but the same genera and species occur in both
1010
(Priscu and Christner, 2004). The fact that particular strains can be revived and sub-cultured from
1011
ice everywhere points to them having adapted to the extreme conditions of deep ice. Priscu and
1012
Christner (2004) estimated that the polar ice sheets contain a total organic carbon pool almost
1013
equivalent to the bacterial carbon contained in all the freshwaters on the Earth s surface. The trace
1014
gasses that these microbes consume or liberate have to be taken into account when changes in the
1015
paleo-climate are read from paleo-gas records in ice cores (Campen et al. 2003, Sowers 2001).
35
1016
Conversely, layers within ice repositories dated accurately for paleoclimate studies provide
1017
important stratigraphic ages for studies of how these microbial populations have evolved.
1018
Rogers et al. (2004) speculated that the annual release of 10
17
21
to 10
viable microbes by
1019
melting ice up to 105 to 106 years old mixes the genotypes of ancient and modern microbial
1020
populations and therefore affect their mutation rates, fitness, survival, and pathogenecity. Smith et
1021
al. (2004) even speculated that releases of such ancient microbiota may explain cyclic calicivirus
1022
events and decades-long disappearances and reappearances of influenza-A subtypes.
1023
Organic components can also be trapped in brine inclusions within crystallising halite and their
1024
degradation can generate methane so that many salt mines face an explosion risk. Microbes
1025
trapped in crystallising salt go much further backs in time than the ~1 Ma record in ice. This was
1026
dramatically demonstrated by encysted halophytic bacteria carefully recovered from an undisturbed
1027
fluid inclusion in Permian salt (~ 250 Ma) preserving primary and diagenetic textures at the WIPP
1028
repository for isolating military nuclear waste in New Mexico (Vreeland et al, 2000). These bacteria
1029
divided after being released and “revived” with primitive sugar 1 that no longer exists in nature.
1030
Nowadays, before publication, these authors should have had their results replicated by an
1031
independent laboratory (where possible contaminants would be unlikely to be the same). Since
1032
then, Sattefield et al., (2005) confirmed that the brine inclusions from the same layer of salt that
1033
housed the revived bacillus are indeed ~250 Ma old. Gragg et al., (2006) reported that they are
1034
studying a 220 m long salt core from the Salar de Uyuni in Bolivia with its >276 ka record to
1035
acertain whether the preservation of viable microorganisms trapped in fluid inclusions relates to the
1036
age of the halite, or (as in ice) the environment of organism entrapment, or both.
1037
Fluid inclusions in salt can also lead to unexpected physical effects. Experiments to study the
1038
effects of thermal gradients imposed by isolating nuclear waste in engineered salt cavities found
1039
that, because the solubility of halite in water increases with temperature, 3-phase fluid inclusions
1040
can migrate up thermal gradients through the salt (by salt dissolving on the warm wall, diffusing
1041
through the brine and precipitating on the cool wall), vent the gas bubble at the engineered surface,
1042
and then migrate back into the rock mass (Roedder and Belkin, 1980). Similar phenomena are
1043
involved in the migration of solid particles and fluid inclusions and films to wam interfaces during
1044
thermal regelation in ice (Worster & Wettlaufer, 1999).
1045
36
1046
8 Ice, Salt and Water
1047
Water and its ice are different phases of the same chemistry so that changes in P and T can
1048
result in each spontaneously appearing in the other by melting or crystallisation. By contrast, salt
1049
and water have different chemistries so that, although increasing pressure increases the density of
1050
halite, there is no equivalent in salt of pressure melting regelation or basal lubrication by water, as
1051
in ice.
1052
1053
8.1 Solubility
1054
Apart from ice which is infinitely soluble, salt is one of the most soluble rocks and 100 g of pure
1055
water can dissolve 36 g of NaCl at 25°C (Langer & O ffermann, 1982) so that 1 cm of rain can
1056
dissolve 0.1667 cm of salt (Lockner, 1995) implying that the potential dissolution rate of salt can
1057
reach 17% of the annual precipitation.
1058
Talbot (1998) and Talbot and Aftabi (2004) anticipated that much less salt is usually dissolved
1059
than predicted by these values. This is because most salt extrusions are topographic highs so that
1060
most precipitation drains off them without carrying its potential load of dissolved salt, and as springs
1061
issuing from the base of Iranian namakiers are very rare. Although secondary salt deposits on and
1062
around them demonstrate that subaerial salt extrusions do lose salt via dissolution, most
1063
precipitation probably evaporates on their surfaces (Talbot, 1998). The first direct measurements of
1064
denudation rates of salt exposed in an arid climate (Frumkin, 1996) reinforced this impression. Thus
1065
99% of current denudation of Mount Sedom in Israel is by dissolution estimated at 0.5 to 0.75 mm
1066
a (which is 1 to 1.5 % of the annual rainfall of 50 mm/y). However most of that dissolution is
1067
restricted to a system of subsurface conduits only metres wide that entrench at rates between 4
1068
and 25 mm a-1.
1069
-1
In strong contrast, salt erosion rates measured on Cardona diapir in Spain were 20 mm per 100
1070
mm rainfall per year (Mottershead et al., 2005), greater than the theoretical potential. Similarly,
1071
Bruthans et al., (2007) used 26 plastic pegs implanted on a variety of slopes with different aspects
1072
on 3 salt extrusions in Iran from 2000-2005 and standardised their direct measurements of
1073
denudation rates of salt surfaces and their cap-soils to horizontal surfaces subject to long term
1074
mean precipitation rates. Long term denudation rates for exposed salt were 30-40 mm a-1 for two
1075
coastal diapirs (Namakdan and Hormoz island) where the rainfall averaged 103 mm a-1 for the
37
-1
1076
relevant 5 years (60% of the 30 year average) and 120 mm a for an inland diapir (Jahani) where
1077
the average rainfall was 295 mm a-1 for the same 5 years. Thin residual cap soils cut the
1078
denudation rate on coastal diapirs to 3.5 mm a , ~1/10 that for salt surfaces exposed nearby.
1079
Erosion rates for exposed salt were found to relate to rainfall directly and to the slope inversely,
1080
being lowest on vertical faces and highest on horizontal surfaces. Salt slopes steeper than 70º were
1081
estimated to retreat ~10 m/1000 years on coastal diapirs and ~30 m/1000 years on inland Jahani.
1082
Bruthans et al (2007) attributed the mean calculated values of denudation rate of exposed salt
1083
surfaces being between 29 and 39% higher than the maximum dissolution capacity of the
1084
corresponding rainfall to the cap soils retaining water like a sponge. However, halophytic
1085
microorganisms might also be involved (Castanier, et al., 1999).
-1
th
1086
1087
1088
8.2 Strength
The rheologies of both ice and salt depend strongly on their liquid water content. A liquid water
1089
content of ~1 volume % in ice (Murray et al., 2005) and of ~0.01 volume % in salt (Cristescu and
1090
Hunsche, 1998) increases the strain rate by ~400%. In glacier ice, this weakening is mainly due to
1091
free water at triple-grain contacts (e.g. Duval, 1977). Experiments on thin sections of natural salt
1092
filmed under the microscope have demonstrated that thumb pressure can distort initially isolated
1093
cubic brine inclusions through amoeboid shapes that connect into continuous films of brine along
1094
new, very mobile, grain boundaries (Urai et al., 1986). Such distortion of dispersed brine inclusions
1095
into continuous brine-filled grain boundaries affectively changes the stress exponent, n, of salt
1096
deforming as a power law fluid from the usual n=3 to 8 in the laboratory (Cristescu and Hunsche,
1097
1998) to an n=1 viscous fluid (Urai et al., 1986).
1098
Water is denser than ice but less dense than salt. As a result, water tends to migrate to different
1099
boundaries in flowing ice and salt. Water forms at, or migrates to, the bottom boundary of ice
1100
masses, either as gravitational flow through porous firn, or by drainage via structural weakness
1101
zones forming arcuate en- and subglacial networks (Fountain and Walder, 1998). Conversely,
1102
extruded salt is wet by rain that falls on the top surface before draining off or evaporating (Talbot,
1103
1998). Ice with surface temperatures well below 0ºC and without crevasses (as in the snouts of
1104
smaller and less dynamic polar or sub-polar glaciers) is usually regarded as impermeable, and thus
1105
an analogue of salt).
38
1106
Salt diapirs were being considered as repositories for radioactive waste in the 1960s and 70s
1107
(Gera, 1972) despite the salt mechanics of the time indicating that salt could not flow over the
1108
surface as in Iran (Heard, 1972). Gussow (1968) suggested that the Iranian salt extrusions were
1109
emplaced rapidly by lava-like, hot and temporary flow; a claim contested by Kent (1966) who
1110
argued that the local inhabitants often heard major rock falls on salt mountains that would not
1111
survive for long in the local rainfall if they were not actively supplied from depth.
1112
Assuming steady state flow and dissolution, Wenkert (1979) calculated the flow rates for 5 of
1113
the salt extrusions in the Zagros to be ~108 faster than allowed by contemporary salt mechanical
1114
measurements in the laboratory (Heard, 1972) and suggested that rainfall softened salt by allowing
1115
intragranular liquid diffusion. Fifty years after they were first suggested by Bailey (1931), the first
1116
direct measurements of salt movement were made in late 70s (Talbot and Rogers 1980) and
1117
confirmed that salt flow was still active. These first measurements found that markers on a
1118
namakier oscillated back and forth when the salt was dry but moved downslope almost 1 m in two
1119
days after a rain storm. Fibres of halite <1 mm thick and a few cm long are brittle when dry but
1120
some flex like a human hair when help under running water (Odé, 1968, p. 543). This, the Joffée
1121
effect, is attributed to water dissolving gasses poisoning the grain boundaries. A similar but
1122
macroscopic effect explains how a namakier that had been deforming noisily as an elastic solid for
1123
most of the preceding year was turned into a silent ductile fluid by only 20 minutes of light drizzle
1124
that precipitated 7 mm in the next 24 hrs (Talbot and Rogers, 1980).
1125
Inspired by these field studies (Urai et al .,1986 and Spiers et al .,1990) found that damp
1126
evaporites, halite, bischofite, MgCl2·6(H2O), and) carnallite KMgCl3·6(H2O), do indeed flow much
1127
faster at lower stresses by solution-transfer creep than the dislocation creep of dry evaporites. Later
1128
laboratory tests suggest that fine-grained rock salt flows about an order of magnitude faster than
1129
coarse-grained salt and that damp dilated salt flows about 40 times faster than dry confined salt
1130
(Cristescu and Hunsche, 1998). Even an increase in humidity >67% accelerates the creep of
1131
dilated salt by a factor of 12 (Cristescu and Hunsche, 1998).
1132
1133
9 Surging flows in ice and salt
1134
Both glaciers and namakiers can temporaily surge and flow hundreds of times faster than usual.
1135
Such surges can be local in time and space and we will treat ice streams as longer-lived spatial
39
1136
surges in ice sheets (e.g. Engelhardt, 2004)- like ocean currents flowing rapidly through the slow-
1137
flowing surrounding oceans or as rivers, fed by sluggish Darcian flow, in the surrounding ground
1138
water zones).
1139
Maximum values for strain rates in both compressed and extended ice are typically in the range
1140
of 0.1-0.2 a (=3 to 6.10 s ) for normal ice flow and >100 a (=3.10 s ) in surging flows (table 2
1141
in Hambrey & Larson 2000). By comparison, subsurface salt with coarse uniform grain-size flows at
1142
strain rates between ~3.10 to 3.10 a (=10
1143
sizes flows at velocities between cm to m a-1 at strain rates of 3.10-6 to 3.10-4 a-1 (10-13 to 10-11 s-1)
1144
over years before being recycled by dissolution. Namakiers surge for a few days at ~dm day at
1145
3.10-3 to 0.3 a-1 (=10-10 to 10-8 s-1) (Jackson and Talbot, 1986).
-1
-9
-1
-8
-1
-6
-1
-15
to 10
-13
-6
-1
-1
s ) and extruded salt with bi-modal grain
-1
1146
1147
9.1 surging flows in ice
1148
The ice activated in a surging glacier may result in the terminus advancing by km in a few
1149
months (Clarke, 1991; Kotlyakov et al., 1997). Most glacial surges last only 2-3 years but some on
1150
Svalbard last longer (12 years) and run-out further (4 km) more slowly, at ~ 2 m day (Lørnne,
1151
2006). Surging glaciers are more likely on weak sedimentary substrates which may explain why
1152
they cluster in particular areas (Clarke, 1991; Jiskoot et al., 1998). In longitudinal profiles, bulges of
1153
surging ice undergoing plug flow (± marginal and internal anastomosing shear zones) tend to fold
1154
against or override slower ice still frozen to the bed downstream along one or more thrusts (e.g.
1155
Lawson et al., 1994). This process results in stratifications above basal contacts with low frictional
1156
contact curling upward distally where friction increases along the base (cf. Fig. 9A4-6). In plan views,
1157
surges result in distinctive tear-drop or bulbous folds in moraines with axes parallel to valley sides
1158
that are clearest where glaciers converge (Hambrey and Lawson, 2000 p.72). Of particular note is
1159
that the plug-like flows of surging glaciers superpose cross-cutting longitudinal and transverse
1160
crevasses on the distinctive crevasse patterns associated with steady-state flows (Lawson, 1996).
1161
Furthermore, crevasse patterns in normal and surging glaciers are very different from each other
1162
and from crevasse patterns in both ice streams and ice sheets (Mayer and Herzfield, 2000).
-1
1163
The generally accepted explanation of surges in glaciers (Meier and Post, 1969) is that the
1164
accumulation of the ice mass upstream reaches a critical normal pressure where above-normal ice
1165
pressures trigger a collapse of an arcuate-tunnel-based subglacial drainage system into a cavity-
40
1166
type drainage system. This smears the subglacial water along a larger area of the ice bed and
1167
decreases basal friction so that the surge occurs by enhanced basal sliding (Kamb, 1987; Raymond
1168
and Harrison, 1988; Eisen et al., 2001). Glaciologists also tend to contrast fast and slow ice flow in
1169
terms of water-lubricated basal ice and ice that is “still frozen” to the bed. The salt of namakiers is of
1170
course “frozen” but temperature does not play a significant role in whether it flows slowly or rapidly.
1171
Even if the transition from normal to surging ice flow does involve basal melting due to increased
1172
friction and increased buoyancy, Ramberg’s experiments on basal traction (Fig. 9A), suggest that
1173
the most relevant aspect of melting or adding water to the basal ice contact is that it decreases the
1174
basal friction. We therefore follow Fowler et al., (2001) and equate the basal sliding of surging ice
1175
(± the often associated processes of shearing or plastic yielding of lubricated underlying soft
1176
sediments ± shearing above the basal contact) as equivalent to what mathematician refer to as a
1177
free slip basal boundary. At its simplest, focussing on the bottom boundary effect suggests that
1178
glaciers surge when water reduces friction along their soles and ice streams appears where water
1179
reduces friction along their soles. However, Joughin’s (2008) study of the lower reaches of ice
1180
streams and outlet glaciers point to friction at the margins of a glacier as a control area accelerating
1181
ice flow. These findings agree with the implications of namakiers surging when water dampens their
1182
bulk mass and their upper levels rather than merely their basal contacts.
1183
More than 150 subglacial lakes beneath the ice of Antarctica have been found by ground
1184
penetrating, flown or satellite radar since the 1960s (Kohler, 2007; Siegert, 2007). Study of these
1185
has helped understanding of the link between smaller surging glacier systems and ice streams. The
1186
subglacial lakes are associated with smooth -floored depressions in the surface topography of the
1187
ice sheet above the lake due to acceleration of the ice associated with longitudinal stretching where
1188
ice detaches from its solid substrate. Surface melt is negligible in inland Antarctica so subglacial
1189
lakes there cannot be supplied by surface melt waters fracturing or draining their way through the
1190
ice as in Greenland (Sohn, 1998; van de Wal et al., 2008). Instead geothermal and frictional heat
1191
can bring large areas of basal ice to its melting point.
1192
Subglacial lakes were assumed to be stagnant until satellite radar from East Antarctica showed
1193
that the ice surface dropped suddenly while rising simultaneously hundreds of km away (Clarke,
1194
2006; Wingham et al., 2006). The surrounding ice appears to buttress most subglacial lakes
1195
(Kohler, 2007) but Belll et al., (2007) provided evidence for an all-important link between subglacial
41
1196
water and the dynamics of the overlying ice. Discharges comparable to the river Thames have been
1197
documented between subglacial lakes >300 km apart over 16 months (Bell et al., 2007). Repeat-
1198
track satellite altimetry by laser also demonstrated several lake drainage events as water
1199
pressurised to near the ice load was squeezed between different parts of the lower reaches of the
1200
West Antarctic ice streams (Fricker et al., 2007). One large lake near the grounding line drained ~2
1201
cubic km water over ~3 years.
1202
Bell et al., (2007) suggested that subglacial hydrology could be responsible for the initiation of
1203
an ice stream by changing the basal thermal regime of the ice sheet and/or by introducing water
1204
and reducing friction at the ice/bed interface. Thus lake water freezing onto the underside of the
1205
floating ice adds thermal energy and softens a zone of basal ice that has a lower frictional contact
1206
with downstream bedrock. Water spilling downstream of the lake also reduces basal friction and
1207
speeds ice flow. Processes like these likely account for sudden accelerations of ice streams in
1208
Greenland (Joughin et al., 2004, 2008; Luckman et al., 2006) while the rerouting of former
1209
subglacial water flows could account for the drastic slowing (“shutdown”) of the Kamb ice stream in
1210
Antarctica (Zwally et al., 2006).
1211
One problem with blaming the initiation of ice streams on subglacial lakes is that this approach
1212
neglects the thermodynamics behind the onset of streaming ice. However it develops, a pool of
1213
water beneath cold ice is insufficient on its own to initiate an ice stream. If the base of the ice is
1214
cold, the water from the subglacial lake will freeze onto the base, and prevent sliding, so there is no
1215
ice stream downstream of Lake Vostok (Bell et al., 2002; Pattyn et al., 2004). The possibility that
1216
frictional heating due to ice motion and high local geothermal heat flux could initiate ice streams has
1217
been discussed and modelled extensively (e.g. Clarke et al., 1977). Although the suggestion that a
1218
thermal run-away could lead to a surge of the Antarctic ice sheet (Yuen, et al., 1986) was later
1219
discounted, numerical modelling has demonstrated that strain heating could initiate streaming flow
1220
(Payne and Dongelmans, 1997) and satisfactorily explain modern (Payne, 1999) and past ice
1221
streams (Payne and Baldwin, 1999) as well as find the zone of initiation of fast glacial flow (Pohjola
1222
and Hedfors, 2003). The recent recognition of hydraulic connections between subglacial lakes
1223
revitalised this debate. In order for the lakes to be where they are, the geothermal heat flow must
1224
be higher than previously estimated, basal strain heat is better conserved than earlier believed, or
1225
heat conducts through the ice more slowly than expected.
42
1226
Bell (2008) distinguished how subglacial water affects the mass balances of the 3 largest
1227
current ice sheets. In the warmest, Greenland, (with a volume equivalent to 7.3 m sea-level rise)
1228
basal meltwater produced in regions of elevated heat flow is joined by surface meltwater that drains
1229
through the ice sheet on elevated crust and reaches the ocean through fjords that pierce the
1230
mountain rim. The West Antarctica ice sheet, similar in area to Greenland, has a volume equivalent
1231
to 5.5 m sea-level rise. Basal melt develops here beneath the interior ice with a base largely
1232
beneath sea level and the ice-stream tributaries (particularly in regions of elevated geothermal
1233
heatflow); water stored in subglacial lakes episodically drain into the ocean beneath the ice streams
1234
that are still largely buttressed by two large ice shelves. In East Antarctica, the largest current ice
1235
sheet, with a volume equivalent to ~65 m sea-level rise, large deep subglacial lakes in the interior
1236
are connected by intermittent subglacial flooding events that transmit water along the axes of ice
1237
streams and outlet glaciers.
1238
Monitoring large areas of the Greenlandic ice sheet found that surface meltwater reaching the
1239
basal contact almost doubled the velocity of the local ice sheet (Joughin et al., 2008). However,
1240
very little of this acceleration reached the relevant outlet glaciers, probably because of friction along
1241
their walls (Joughin et al., 2008). Water at the ice bed has been known to accelerate the overlying
1242
ice since the 1980s (e.g. Iken and Bindschadler, 1986). Pressurised water backing up behind
1243
overwhelmed subglacial plumbing systems lubricates the bottom boundary locally but can soon be
1244
dissipated by hydrodynamic changes in the subglacial sheets and channels. Recent work by
1245
Bartholomaus et al, 2008 adds to the knowledge of water decreasing the strength of the ice mass
1246
and friction along boundaries other than the bases as playing roles in the surging of flowing ice.
1247
1248
9.2 Surging flows in salt
1249
1250
Comparisons of Synthetic Aperture Radar interferograms with ground measurements of marker
1251
arrays dispersed over subaerial extrusions of salt of different ages and at different stages of
1252
evolution in Iran (Fielding and Talbot, unpublished) find that steady extrusion from the unseen vent
1253
continuously inflates the upper reaches of salt extrusions during dry weather. Such swellings of
1254
extruded salt are stored at high structural levels by elastic carapaces that are strong during dry
43
1255
years but are dissipated by accelerated gravity spreading when the carapaces are weakened by
1256
rain in wet years.
1257
An obvious working model for the surges of subaerial salt extrusions assumes that salt
1258
extrudes steadily from its nearly constant loading at depth and then gravity spreads downslope at
1259
rates that depend on the surface temperature but mainly on the local rainfall. Salt extruded steadily
1260
through the subsurface vent can be stored high in the mountain by its outer carapace or corset of
1261
broken elastic salt that case-hardens every salt mountain during the usual dry weather (Fig. 12A).
1262
When it is dry, the extruded salt flows at a few cm to m a-1 and generates a tank track fold as it
1263
advances and attaches to its high-friction basal contact. Flow folds deforming the upper limb of this
1264
fold probably record changes in the rate of gravity-spreading controlled by variations in rainfall over
1265
millennia. Constraint by the elastic corset is lost when the salt is saturated and weakened by
1266
rainstorms a few days per decade. This allowis the potential energy stored in the salt to be released
1267
by surges at dm day for a few days (Fig. 12B).
1268
-1
The closest water comes to migrating along the basal contact of a namakier is probably when
1269
unsaturated brines fed by storm-water dampen near-basal mylonitic shear zones <1 dm thick in the
1270
generally gneissose salt of bimodal grain size. These detach the overlying salt and curl upward
1271
distally allowing the overlying salt to shear over slower or static salt (Fig. 9, Talbot, 1998). Such
1272
halitic mylonites have fine grain sizes of <2 mm (Fig. 8SE) that indicate strain rates near 0.3 a-1 and
1273
micro-fabrics that imply that they deform by solution-precipitation creep and grain-boundary sliding
1274
(Schleder and Urai, 2006). Their sub-grain sizes indicate differential stresses of <0.4 MPa at
1275
temperatures <40ºC consistent with field observations (Talbot and Rogers, 1980). Storm-water
1276
draining through near-basal mylonitic detachment zones in namakiers that only surge for a day or
1277
two a decade take the place of the wet basal contacts that allow much faster and longer surges of
1278
glaciers and ice streams. However, the main control of surges of namakiers over their substrates of
1279
rocks and sediments appear to be weakening of the salt mass and its top boundary (Fig. 12). This
1280
working model for surging salt converges on recent suggestions that surges in the lower reaches of
1281
ice streams and outlet glaciers are controlled more by changes in strength of the ice mass and its
1282
outer boundaries. This argument suggests that salt extrusions in the damper mountains of Iran
1283
surge more than those along the drier coast.
44
1284
Individual salt sheets beneath the Gulf of Mexico have been calculated to have advanced at
1285
rates between 1.5 and 275 mm a-1 during Miocene and Pleistocene times (Wu et al., 1990).
1286
Submarine salt sheets therefore advance further and faster over saturated sediments than
1287
subaerial equivalents advance over dry substrates of rocks and sediments. Although it is not yet
1288
known whether submarine salt sheets surge, we consider this unlikley because their outer zones
1289
may undergo minor changes in strength during relatively small and slow temperature changes but,
1290
being surrounded by a perpetual bath of seawater, are unlikely to undergo the sudden drops in
1291
strength that allow surges in subaerial in desert climates.
1292
1293
10 CONCLUSIONS
1294
Despite differences in origin, chemistry, mineralogy, physical properties and supply processes
1295
etc, subaerial salt extrusions have visual and mechanical similarities to ice caps and glaciers (see
1296
Figs.1, 2, 8 and Table 2). Both ice and extruded salt can gravity-spread to the profiles of viscous
1297
fountains or droplets with profiles that may approximate steady-states.
1298
In bodies consisting of polygonal grains of ice with its hexagonal crystallography, primary
1299
stratification, diagenetic ice lenses and foliations picked out by discontinuous layers of different
1300
grain size and bubble-content are slightly strain active with a subhorizontal planar anisotropy. As a
1301
result, kinematic drape or forced folds (± axial planar stripes) form upright in ice above irregular
1302
bottom boundaries and probably have to be carried increasing distances downstream to be sheared
1303
to inclined or recumbent geometries.
1304
By contrast, in bodies of polygonal grains of cubic halite, primary stratification and secondary
1305
foliations are much more obvious in extrusive salt with its bimodal grain-size but are essentially
1306
strain passive. The grain shape fabrics in salt are clearer and more informative than grain-size
1307
foliations in ice and have longer strain memories. Nevertheless the strain memories of deformed
1308
salt fabrics are sufficiently short for foliations to map new, locally developed flow lines, most of
1309
which sub-parallel both the stratification and external boundaries. If this were true in flowing ice it
1310
would be difficult to distinguish secondary foliation from primary stratification (Hambrey and Muller,
1311
1978; Hooke and Hudleston, 1978). Passive markers like stratification and foliation in salt can
1312
remain parallel where they flow steadily round right-angled corners from source to diapir, and from
45
1313
diapir to extrusion; they can also drape over bedrock obstructions (Fig. 9G) in kinematic equivalents
1314
of upright folds in near-basal ice.
1315
The top free surface of a gravity-spreading fluid moves faster under air than where it is in
1316
frictional contact with a solid substrate (Talbot, 1998). As a result, the stratification rolls over the
1317
hinge of a kinematic tank track fold that grows behind every active namakier and inverts in its lower
1318
limb. A kinematic tank-track fold is inevitable in and behind the fronts of all viscous flows that have
1319
advanced over fictional substrates unless it is truncated by loss of material from their top surface.
1320
Their lack of recognition in glaciology may be because their upper limbs are truncated by ablation
1321
or because the stratification is so rarely exposed in cold ice flows.
1322
By comparison with laboratory experiments (Fig. 9) we have attributed the tank track folds that
1323
characterise the profiles of subaerial namakiers of Iran to their advance over high-frictional basal
1324
contacts. This would be confirmed if future studies find that tank track folds are absent from
1325
submarine salt extrusions that have advanced further and faster but essentially steadily over low
1326
friction basal contacts with weak saturated sediments.
1327
Both Ice and salt are supplied to significant surface bodies at rates that are close to steady over
1328
decades. By contrast, the rates at which ice (Shumskii 1964, p.. 326) and extruded salt gravity
1329
spread downslope are noticeably irregular on a large range of time scales. Accepting that the
1330
extrusion rates of individual salt diapirs are close to steady, flow lines marked by a foliation cross
1331
the (mainly) passive bedding and impose asymmetric dynamic flow folds on the upper limb of the
1332
kinematic tank-track fold when the ratio of their rates of extrusion and spreading changes. As
1333
rainfall is likely to change faster than the tectonic forces affecting subaerial salt extrusions, such
1334
budget-induced dynamic flow folds within subaerial salt extrusions offer a potential record of climate
1335
change over tens of millennia. Any flow folds in submarine salt sheets might record long term
1336
temperature changes in the surrounding seawater. Indivivual flow folds in the upper limbs of tank
1337
track folds also offer natural analogue models of individual nappes within piles of nappes of
1338
extruded infrastructures in orogens like the Himalaya. They suggest that individual nappes within
1339
orogenic nappe-piles might reflect changes in the ratios of their rates of tectonic extrusion relative
1340
to their rates of gravity spreading. Whether orogenic nappes spread ar rates affected by the
1341
climates is debateable. A great advantage of studying rock deformation structures and fabrics
1342
within bodies of ice or salt is that internal structures, such as folds and shear zones of different
46
1343
categories, could be instrumented and monitored and treated monominerallic analogues of other
1344
metamorphic rocks.
1345
Hudleston (1976) envisaged flowing ice bodies developing folds where and when the flow lines
1346
crosses old stratification perturbed by relief in the bottom boundary. However, the clearer folds in
1347
namakiers suggest that dynamic flow folds near the top boundary of flowing ice may be even
1348
clearer records of budget changes due to fluctuations in climate. The potential for folds in ice to
1349
record climate change has been recognised in ice streams (see e.g. Bindschadler, 1993; Jacobel et
1350
al., 1993; Merry and Whillans, 1993; and Fahnestock, et al., 2000).
1351
Fluctuations in supply and loss (in ice) and gravity spreading and loss (in salt) result in budget-
1352
induced dynamic flow folds in both glaciers and namakiers. However, while both glaciers and
1353
namakier undergo slow penetrative creeping flow that depend mainly on their thickness, slope,
1354
grain size, purity and temperature, both may surge at velocities hundreds of times faster than usual
1355
when water weakens these rock masses and/or sufficient areas of their external boundaries. The
1356
factors controlling surges in namakiers are more obvious than those controlling surges in ice. The
1357
possibility that surging ice is influenced by water weakening the bulk ice and its outer boundaries
1358
(as well as its base) has been recognised only recently.
1359
As for all exposed rock masses, the outer zones of bodies of ice and salt salt readily dilate to
1360
brittle elastic carapaces aided by destressing and weathering or ablation. The resulting fractures
1361
and micro-fractures render the outer zones of both ice and salt masses sensitive to weather-
1362
induced changes. That this applies to ice tends to be overlooked because variations in wet and dry
1363
basal conditions are more obvious. Ice has long been recognised as capable of detaching and
1364
sliding over wet substrates because most water associated with ice forms or migrates to the bottom
1365
boundary. Although salt can detach along damp near-basal mylonite zones and slide over the
1366
underlying salt, the water that has most effect on the downslope flow rate of namakiers falls as rain
1367
on their top surfaces and seldom reaches the bottom boundary of denser salt.
1368
Detailing the records of changes in past climates from fluid inclusions in ice go back ~1 Ma whereas
1369
salt deposits potentially offer a discontinuous record of changes in climate and oceanic chemistry
1370
back to at least ~500 Ma; on Mars and some of the icy moons of the outer planets, equivalent
1371
records may extend to billions of years in both ice and/or salt. Microbiota up to ~1 Ma may be
1372
released by old ice ablating or melting, and some of the ~500 Ma microbiota released by salt
47
1373
dissolution may yet turn out to be viable although the foodstocks they originally fed on are no longer
1374
freely available on the modern Earth. The implications of mixing such ancient and modern genome
1375
pools of cryophile and halophiltic micro-biota bears continued examination.
1376
1377
Acknowledgments
1378
Various versions of this manuscript benefited considerably from constructive reviews by Ian
1379
Davison, J.K. Warren, Peter Hudleston, Michael Hudec, and Martin Jackson; any mistakes are ours
1380
rather than theirs. Both authors thank innumerable colleagues who have helped them in the field
1381
throughout their careers. Most of their travels to and from the field have been funded by The
1382
Swedish Natural Science Foundation (now VR). CT would also takes this opportunity to thank Dr
1383
Korehei, former General Director of the Geological Survey of Iran and his colleagues, for their
1384
enthusiastic encouragement and considerable logistic support for the 1993-2006 project on the
1385
kinematics of Iranian salt.
1386
1387
Talbot and Pohjola References
1388
Abyzov, S. S., 1993. Microorganisms in the Antarctic ice, In: Friedmann, E. I. (Ed.), Antarctic
1389
Microbiology John Wiley & Sons, Inc., New York., pp. 265-295.
1390
Alkattan, M., Oelkers, E.H., Dandruran J-L., Schott, J., 1997. Experimental studies of saltrock
1391
dissolution kinematics, 1: the effects of saturation state and the presence of trace metals.
1392
Chemical Geology 137. 201-219.
1393
1394
1395
Alley, R.B., 2002. The Two-Mile Time Machine: Ice Cores, Abrupt Climate Change, and Our
Future. Princeton University Press, Princeton, New Jersey, USA, 240 pp.
Alley, R.B., Gow, A.J., Meese, D.A., Fitzpatrick, J. J., Wadddington, E. D., Blozan, J. F., 1997.
1396
Grain scale processes, folding and stratigraphic disturbances in the GISP2 ice core. Journal of
1397
Geophysical Research 102, C12, 26,819-26,830.
1398
1399
1400
Ambraseys, N. N., Jackson, J. A., 1998. Faulting associated with historical and recent earthquakes
in the eastern Mediterranean region. Geophysical Journal International 133, 390-406.
Bailey, E.B., 1931.Salt plugs. Geological Magazine 68, 335-339.
48
1401
Baker, P.A., Rigsby, C.A., Seltzer, G.O., Fritz, S.C., Lowenstein, T.K., Bacher, N.P., Veliz, C., 2001.
1402
Tropical climate changes at millennial and orbital timescales on the Bolivian Altiplano. Nature
1403
409, 698-701.
1404
1405
1406
1407
Barton, D. C., 1933. Mechanics of formation of salt domes with special reference to Gulf Coast salt
domes of Texas and Louisiana. AAPG Bulletin 17, 1025-1083.
Bartholomaus, T. C., Anderson, R.S., Anderson, S.P., 2008. Response of glacier basal motion to
transient water storage. Nature geoscience 1, 33-37.
1408
Beaumont, C, Jamieson, R. A., Nguyen, M. H., Lee, B., 2001. Himalayan tectonics explained by
1409
extrusion of a low-viscosity crustal channel coupled to focussed surface denudation. Nature
1410
414, 738-742.
1411
1412
Bell, R. E., 2008. The role of subglacial water in ice-sheet mass balance. Nature geoscience 1,
297– 304.
1413
Bell, R.E., Studinger, M., Tikku, A.A., Clarke G.K.C., Gutner, M.M., Meertens. C., 2002. Origin and
1414
fate of Lake Vostok water frozen to the base of the East Antarctic ice sheet. Nature 416, 307-
1415
310.
1416
1417
1418
Bell, R. E., Studinger, M., Suman, C. A., Fahnestock, M. A., Joughin, I., 2007. Large subglacial
lakes in East Antarctica at the onset of fast flowing streams. Nature, 445, 904-907.
Bindschadler, R., 1993. Siple Coast Project research of Crary Ice Rise and the mouths of Ice
1419
Streams B and C, West Antarctica: review and new perspectives. J. Glaciol., 39, 538-552.
1420
Bindschadler, R. 1998. Monitoring ice sheet behavior from space. Reviews of Geophysics 36, 79-
1421
1422
104.
Bobst , A.L., Lowenstein T.K., Jordan T.E., Godfrey, L.V., Ku T.L., Luo, S., 2001. A 106ka
1423
paleoclimate record from drill core of the Salar de Atacama, northern Chile .Palaeogeography,
1424
Palaeoclimatology, Palaeoecology 173, 21-42.
1425
1426
Boulton, G. S., 1996. Theory of glacial erosion, transport and deposition as a consequence of
subglacial sediment deformation. J. Glaciol., 42, 43-62.
1427
Bruthans, J., Filippi, M., Gersl, M., Zare, M., Melkova, J., Pazdur, A., Bosak, P., 2006. Holocene
1428
marine terraces on two salt diapirs in Persian Gulf (Iran): age, depositional history and uplift
1429
rates. Journal of Quaternary Science 2, 843-857.
49
1430
1431
1432
Bruthans, J., Asadi, N., Filippi, M., Wilhelm, Z., Zare, M., 2007. A study of erosion rates on salt
diapir surfaces in the Zagros Mountains, SE Iran. Environmental Geology 53, 1079-1089.
Busk, H. G., 1929, Earth flexures; their geometry and their representation and analysis in geological
1433
section with special reference to the problem of oil finding: Cambridge University Press,
1434
Cambridge, U.K., 106 pp.
1435
1436
Campen, K.R., Sowers, T., Alley, R.B., 2003. Evidence for microbial consortia metabolizing within a
low latitude mountain glacier. Geology 31, 231–234.
1437
Castanier, S., Perthuisot, J-P., Matrat. M., Morvan, J-Y., 1999. The salt ooids of Berre salt works
1438
(Bouches du Rhône, France): the role of bacteria in salt crystallisation, Sedimentary Geology
1439
125, 9-21.
1440
1441
1442
1443
1444
1445
1446
1447
1448
1449
Castello, J.D., Rogers, S.O., 2005. Life in Ancient Ice. Princeton University Press. Princeton, New
Jersey, USA. 328 pp.
Chemia Z., Koyi H., Schmeling H., 2008. Numerical modelling of rise and fall of a dense layer in salt
diapirs. Geophysical Journal International, 172, 798-816.
Chinn, T. J. H., 1989. Single folds at the margins of dry-based glaciers as indicators of glacial
advance. Annals Glaciology 12, 23-30.
Clarke, G.K.C., 1987. A short history of scientific investigations on glaciers. J. Glaciol., Special
Issue, 4-24.
Clarke, G.K.C., 1991. Length, width and slope influences on glacier surging. J. Glaciol., 37, 236246.
1450
Clarke, G.K.C., 2006. Glaciology: Ice-sheet plumbing in Antarctica. Nature 440, 1000-1001.
1451
Clarke, G.K.C., Nitsan, U., Paterson, W.S.B., 1977. Strain heating and creep instability in glaciers
1452
1453
and ice sheets. Reviews of Geophysics and Space Physics 15, 235-247.
Cobbold, P.R., Szatmari, P., Demercian, L.S., Celho, D., Rossello, E.A.., 1995. Seismic and
1454
experimental evidence for thin-skinned horizontal shortening by convergent gliding on
1455
evaporites, deep-water Santos Basin, Brazil. In: Jackson, M.P.A., Roberts, D.G., Snelson, S.,
1456
(Eds.) Salt tectonics; a global perspective; AAPG Memoir 65, 305-321.
1457
Cosgrove, J.W., Ameen, M.S., 2000. A comparison of the geometry, spatial organisation and
1458
fracture patterns associated with forced folds and buckle folds. In: Cosgrove, J. W., Ameen, M.
1459
S., (Eds.) Forced folds & fractures: Geological Society London Special Publications 169, 7-21.
50
1460
Coward, M., Stewart, S., 1995. Salt-influenced structures in the Mesozoic–Tertiary cover of the
1461
southern North Sea, U.K. In: Jackson, M.P.A., Roberts, D.G., Snelson, S. (Eds.) Salt Tectonics:
1462
a Global Perspective. AAPG Memoir 65, 229–250.
1463
1464
1465
1466
1467
1468
1469
Cristescu, N., Hunsche, U. 1998. Time effects in Rock mechanics- Series: Materials, modelling and
Computation. John Wiley and Sons, Chichester. 342 pp.
De Böckh, H., Lees, G.M., Richardson, F.D.S., 1929. Contribution to the stratigraphy and tectonics
to the Iranian ranges. In Gregory, J.W. (Ed) The structure of Asia. Methuen, London, 58-176.
Demorest, M., 1942. Glacier regimens and ice movements within glaciers. American Journal
Science 240, 31-66.
Diegel, F.A., Karlo, J.F., Scuster, D.C., Shoup, R.C., Tauvers, P.R., 1995. Cenozoic structural
1470
evolution and tectonostratigraphic framework of the Northern Gulf Coast continental margin. In
1471
Jackson, M.P:A., Roberts, D.G., Snelson, S., (Eds.) Salt Tectonics: a global perspective, AAPG
1472
Memoir 65, 109-151.
1473
Drewry, D., 1986. Glacial geological processes. Edward Arnold, London, 276 pp.
1474
Durham, W. B., Abey, A.E., Trimmer, D.A., 1981. Thermal Conductivity and diffusivity and
1475
expansion of Avery Island salt at pressure and temperature. Technical Report University
1476
California, Lawrence Berkeley Laboratory, Livermore, and CA. USA.
1477
1478
1479
1480
1481
Duval, P., 1977. The role of water content on the creep rate of polycrystalline ice. IUGG General
Assembly of Grenoble, 1975. IAHS Publication 118, 29-33.
Duval, B., Cramez, C., Jackson M.P.A., 1992. Raft tectonics in the Kwanza Basin, Angola. Marine
and Petroleum Geology 9, 389-404.
Dyson, D.A., 2004, Christmas tree diapirs and the development of hydrocarbon reservoirs; a model
1482
from the Adelaide Geosyncline, South Australia. In Post, P. et al., (Eds.) Salt sediment
1483
interactions and hydrocarbon prospectivity: concepts, applications and case studies for the 21st
1484
century: 24th Annual Gulf Coast Section SEPM Foundation Bob F. Perkins Research
1485
Conference, Houston Texas pp. 79-96.
1486
Eisen, O., Harrison, W.D.M., Harrison, C.F., 2001. The surges of Variegated Glacier, Alaska,
1487
U.S.A., and their connection to climate and mass balance. J. Glaciol., 47, 351-358.
1488
Engelhardt, H. 2004. Thermal regime and dynamics of the West Antarctic ice sheet. Annals
1489
Glaciology 39, 85-92.
51
1490
1491
1492
1493
1494
1495
1496
1497
1498
1499
1500
1501
1502
EPICA community members (56 authors). 2004. Eight glacial cycles from an Antarctic ice core.
Nature 429, 623-629.
Fahnestock, M.A., Scambos, T.A., Bindschadler R.A., Kvaran. G., 2000. A millennium of variable
ice flow recorded by the Ross Ice Shelf, Antarctica. J. Glaciol., 46, 652-664.
Fricker, H. A., Scambos, T., Bindschadler, R., Padman, L., 2007. An active subglacial water system
in West Antarctica mapped from space. Science 315, 1544-1548.
Fountain, A.G., Walder, J.S., 1998. Water flow through temperate glaciers. Reviews in Geophysics
36, 299–328.
Fowler, A.C., Murray, T., Ng, F.S.L. 2001. Thermally controlled glacier surging. J. Glaciol., 47, 527538.
Frumkin, A., 1996. Determining the exposure age of a karst landscape. Quaternary Research 46,
99-106.
Gemmer, L., Ings, S. J., Medvedev, S., Beaumont, C., 2004. Salt tectonics driven by differential
1503
sediment loading: Stability analysis and finite-element experiments. Basin Research 16, 199-
1504
218.
1505
1506
1507
1508
1509
Gera, F. 1972. Review of salt tectonics in relation to the disposal of radioactive wastes in salt
formations. Bulletin of the Geological Society of America 83, 3551-3574.
Gevantman, L. H. (Ed.), 1981, Physical Properties Data for Rock Salt; US Department of
Commerce NBS Monograph 167, p. 214.
Gragg, K, E.. Lowenstein, T. K., Timofeeff, M. N., Schubert, B. A., Parker, M., 2006, Long-term
1510
survival of microorganisms preserved in halites, Salar De Uyuni salt core, Bolivia. Abstracts of
1511
Philadelphia Annual Meeting (22–25 October 2006).
1512
1513
1514
1515
1516
1517
1518
1519
Gudmundsson, G.H., 1997a. Basal flow characteristics of a linear medium sliding frictionless over
small bedrock undulations. J. Glaciol .43, 71-79.
Gudmundsson, G.H., 1997b. Basal flow characteristics of a non-linear flow sliding frictionless over
strongly undulating bedrock. J. Glaciol 43, 80-89.
Gussow, W.C. 1966. Salt temperature: a fundamental factor in salt dome intrusion. Nature 210,
518-519.
Halbouty, M.T., 1979. Salt domes, Gulf region, United States and Mexico (2nd edition). Gulf
Publishing Company, Houston, Texas, 561 pp.
52
1520
1521
1522
1523
1524
Hambrey, M.J., Muller, F., 1978. Structures and ice deformation in the White Glacier. J. Glaciol. 20,
41-66.
Hambrey, M.J., Bennett, M.R., Dowdeswell, J.A., Glasser, N.F., Huddart, D., 1999. Debris
entrainment and transfer in polythermal valley glaciers. J. Glaciol., 45, 69-86.
Hambrey M. J., Lawson, W., 2000. Structural styles and deformation fields in glaciers: a review. In:
1525
Maltman A. J, Hubbard, B., Hambrey, M. (Eds.) Deformation of Glacial materials. Geological
1526
Society of London, Special publication 176, 59-83.
1527
1528
1529
1530
1531
1532
1533
1534
1535
Hardie, L. A., 1991. On the significance of evaporites, Annual Reviews of Earth and Planetary
Science 19, 131-168.
Harper, J.T, Humphrey, N.F., Pfeffer, W.T., 1998. Three-dimensional deformation measured in an
Alaskan glacier. Science 281, 1340-1342.
Harrison, J.V., 1930. The geology of some salt-plugs in Laristan, southern Persia; Geological
Society of London Quarterly Journal 86, 463-522.
Heard, H.C, 1972. Steady-state flow of polycrystalline halite at pressure of 2 kilobars, American
Geophysical Union Geophysical Monograph 16, 191-209.
Heaton, R.C., Jackson, M.P.A., Bamahmoud, M., Nani, A.S.O., 1995. Superposed Neogene
1536
extension, contraction and salt canopy emplacement in the Yemini Red Sea. In: Jackson,
1537
M.P.A., AAPG Memoir 65, 333-367.
1538
Hooke, R. LeB., Hudleston, P., 1978. Origin of foliation in glaciers. J. Glaciol., 20, 285-299.
1539
Hooke, R. LeB., Hanson. B., 1986. Borehole deformation experiments, Barnes Ice Cap, Canada.
1540
1541
1542
1543
Cold Regions Science and Technology 12, 261-276.
Hooke, R.LeB., Holmlund, P., Iverson. N. R., 1987. Extrusion flow demonstrated by borehole
deformation measurements over a riegel, Storglaciären, Sweden. J. Glaciol. 33, 72-78.
Hovland, M., Fichler, C., Rueslåtten, H., Johnsen, H.K., 2006. Deep rooted piercement structures in
1544
deep sedimentary basins- manifestations of supercritical water generation at depth. Journal of
1545
Geochemical Exploration 89, 157–160.
1546
Hovland, M., Kuznetsova, T., Rueslåtten, H., Kvamme, B., Johnsen, H.K., Fladmark, G.E., Hebach,
1547
A., 2006. Subsurface precipitation of salts in supercritical seawater. Basin Research 18, 221–
1548
230.
53
1549
1550
1551
1552
1553
1554
1555
1556
1557
1558
Hudec, M.R., Jackson, M.P.A. 2006, Advance of allochthonous salt sheets in passive margins and
orogens. AAPG Bulletin 90, 1535-1564.
Hudec, M.R., Jackson, M.P.A. 2007. Terra Infirma: understanding salt tectonics. Earth Science
Reviews 82, 1-28.
Hudleston, P., 1976. Recumbent folding in the base of the Barnes Ice Cap, Baffin Island, Northwest
Territories, Canada. Geological Society of America Bulletin 87, 1684-1692.
Hudlestone, P., Hooke, R.LeB., 1980. Cumulative deformation in the Barnes Ice Cap and
implications for the development of foliation. Tectonophysics 66, 127-146.
Huppert, H. E., 1982. The propagation of two-dimensional and axisymmetric viscous gravity
currents over a rigid horizontal surface. Journal of Fluid Mechanics 121, 43-58.
1559
Iken, A., Bindschadler, R.A., 1986. Combined measurements of subglacial water pressure and
1560
surface velocity of Findelengletscher, Switzerland: conclusions about drainage system and
1561
sliding mechanism. J. Glaciol., 32, 101-119.
1562
Iizuka, Y., Satake, H., Shiraiwa, T., Narus R., 2001. Formation processes of basal ice at Hamna
1563
Glacier, Sôya Coast, E. Antarctica, inferred by detailed co-isotopic analyses. J. Glaciol., 47,
1564
223-231.
1565
1566
1567
1568
1569
1570
1571
1572
1573
1574
Jackson, M.P.A., 1995. Retrospective salt tectonics, In: Jackson, M.P.A., Roberts, D.G., Snelson,
S., (Eds.) Salt tectonics; a global perspective; AAPG Memoir 65, 1-28.
Jackson, M.P.A., Talbot, C. J., 1986. External shapes, strain rates and dynamics of salt structures:
Geological Society of America Bulletin 97, 305-328.
Jackson, M.P.A., Cornelius, R.R., Craig, C.R., Gansser, A., Stöcklin, J., Talbot C.J., 1990. Salt
diapirs of the Great Kavir, central Iran. Geological Society of America Memoir 177.
Jackson, M.P.A., Talbot, C.J., 1994. Advances in salt tectonics. In: Hancock, P.L, (Ed) Continental
deformation: Pergamon Press and IUGS, Oxford, UK, 159-179 pp.
Jackson, M.P.A., Vendeville, B.C, Schultz-Ela, D.D., 1994. Structural dynamics of salt systems.
Annual Reviews of Earth and Planetary Sciences 22, 93-117.
1575
Jackson, M.P.A., Warin, O.N., Woad, G.M., Hudec, M.R., 2003. Neoproterozoic allochthonous salt
1576
tectonics during the Lufilian orogeny in the Katangan Copperbelt, central Africa: Geological
1577
Society of America Bulletin 115, 314–330.
54
1578
Jackson, M.P.A., Hudec, M.R., 2004. A new mechanism for advance of allochthonous salt sheets,
1579
in Post, P.J., Olson, D.L., Lyons, K.T., Palmes, S.L., Harrison, P.F., Rosen, N.C., (Eds.) Salt-
1580
sediment interactions and hydrocarbon prospectivity: concepts, applications, and case studies
1581
for the 21st century: 24th Annual GCSSEPM Foundation Bob F. Perkins Research Conference,
1582
pp. 220-242.
1583
Jacobel, R.W., Gades, A. M., Gottschling, D.L., Hodge, S.M., Wright, D. L., 1993. Interpretation of
1584
radar-detected internal layer folding in West Antarctica ice steams. J. Glaciol. 39, 528-537.
1585
1586
1587
1588
1589
1590
1591
1592
1593
1594
1595
Jacobson, H.P. Waddington, E.D., 2005. Recumbent folding of divide arches in response to
unsteady ice-divide migration. J. Glaciol., 51, 201-209.
Jiskoot, H., Boyle, P., Murray, T., 1998. The incidence of glacier surging in Svalbard: evidence
from multivariate statistics. Computers and Geosciences 24, 387-399.
Joughin, I. Abdalati, W., Fahnestock, M., 2004, Large fluctuations in speed on Greenland's
Jakobshavn Isbræ glacier. Nature 432, 608-610.
Joughin, I., Das, S. B., King, M. A., Smith, B. E., Howat, I. M., Moon, T., 2008. Seasonal Speedup
along the Western Flank of the Greenland Ice Sheet. Science 320, 781 – 783.
Kamb W. B., 1987.Glacier surge mechanism based on linked cavity configuration of the basal water
conduit system. Journal Geophysical Research 92 (B9): pp. 9083-9100.
Kawamoto, E., Shimamoto, T., 1997. Mechanical behavior of halite and calcite shear zones from
1596
brittle to fully plastic deformation and a revised fault model. In Yadong, Z., Davis, G.A., Yin, A
1597
(Eds) Structural Geology and Geomechanics, Proceedings of the 30th International Geological
1598
Congress, Beijing, China, 4-14 August 1996, Volume 14, pp 89-105.
1599
1600
Kehle, R.O., 1988. The origin of salt structures. In: Schreiber, B.C. (Ed.), Evaporites and
Hydrocarbons. Columbia University Press, New York, pp. 345–404.
1601
Kent, P.E., 1958. Recent studies of South Persian salt plugs. AAPG Bulletin 42, 2951-2979.
1602
Kent, P.E., 1966. Salt temperature: a fundamental factor in salt dome intrusion: discussion. Nature
1603
1604
1605
1606
1607
211, 1387-1388.
Kent, P.E., 1970. The salt plugs of the Persian Gulf region. Leicester Literary and Philosophical
Society Transactions 64, 56-88.
Kent, P.E., 1979. The emergent Hormuz salt plugs of southern Iran. Journal of Petroleum Geology
2, 117-144.
55
1608
Kockel, F., 1998. Salt problems in Northwest Germany and the German North Sea sector. In
1609
Kockel, F., Marschall, R. (Eds.) Geology and Geophysics of Salt Structures. Journal of Seismic
1610
Exploration 7, 219–235.
1611
Kohler, J., 2007. Glaciology: lubricating lakes. Nature 445, 830-831.
1612
Kotlyakov, V.M., Osipova., G.B., Tsvetkov D.G., 1997. Fluctuations of unstable mountain glaciers:
1613
scale and character. Annals Glaciology 24, 338-343.
1614
Koyi, H., 1991. Mushroom diapirs penetrating into high viscous overburden. Geology 19, 1229-123.
1615
Langer, H., Offermann, H., 1982. On the solubility of sodium chloride in water. Journal of Crystal
1616
1617
1618
1619
1620
1621
1622
Growth 60, 389-392.
Langway, C.C., 1967. Stratigraphic analysis of a deep ice core from Greenland. Research Report
Cold Regions Research and Engineering Laboratory 77, 130 pp.
Lawson, W., 1996. The relative strengths of debris-laden basal ice and clean glacier ice: some
evidence from Taylor Glacier, Antarctica. Annals Glaciology 23, 270-276.
Lawson, W.J., Sharp, M.J., Hambrey, M.J., 1994. The structural geology of a surge-type glacier.
Journal of Structural Geology 16, 1447-1462.
1623
Lees, G.M., 1927, Salzgletcher in Persien: Mitt der deutschen Gesellschaft, Wien 20, 29– 34.
1624
Legrand, M., Mayewski, P., 1997. Glaciochemistry of polar ice cores: a review. Reviews of
1625
1626
1627
1628
1629
1630
1631
1632
1633
1634
1635
1636
1637
Geophysics 35, 219-243.
Lister, J.R., Kerr, R.C, 1989. The propagation of two-dimensional and axisymmetric viscous gravity
currents at a fluid interface. Journal of Fluid Mechanics 203, 215-249.
Lockner, D.A., 1995. Rock failure. In Ahrens, T.J. (Ed). Rock Physics and Phase Relations. A
Handbook of physical constants, AGU Reference Shelf 3, AGU, 127-147.
Lowenstein, T.K., Hardie, L.A., 1985. Criteria for the recognition of salt-pan evaporites.
Sedimentology 32, 627-644.
Lowenstein, T.K., Demicco, R.V. 2006. Elevated Eocene atmospheric CO2 and its subsequent
decline. Science 313. 1928.
Lowenstein, T.K., Li, J., Brown, C., Roberts, S.M., Ku, T-L., Luo, S., Yang, W., 1999. 200 ky
paleoclimate record from Death Valley salt core. Geology 27, 3-6.
Lowenstein, T.K., Timofeeff, M. N. Brennan, S.T., Hardie, L.A., Demicco, R.V., 2001. Oscillations in
Phanerozoic Seawater Chemistry: Evidence from Fluid Inclusions. Science 294, 1086 – 1088.
56
1638
1639
1640
1641
1642
Lørnne, I., 2006. Low-velocity glacial surges- processes unlocked by modern surge on Svalbard.
Geology 34, 663-556.
Lucchitta, B.K., 2001. Antarctic ice streams and outflow deposits on Mars. Geophysical Research
Letters 28, 403-406.
Luckman, A., Murray, T., de Lange, R., Hanna, E., 2006. Rapid and synchronous ice-dynamic
1643
changes in East Greenland, Geophysical Research Letters 33, L03503,
1644
doi:10.1029/2005GL025428.
1645
Lunine, J.I. Atreya, S.K., 2008. The methane cycle on Titan. Nature Geoscience 1, 159-164
1646
Mair, R., Kuhn M., 1994. Temperature and movement measurements at a bergschrund. J. Glaciol.,
1647
1648
1649
1650
40, 561-565.
Maltman A.J, Hubbard, B., Hambrey, M. 2000. Deformation of glacial materials. Geological Society
of London Special Publication 176, 1-9.
Mayer, H., Herzfeld, U.C., 2000. Structural glaciology of the fast-moving Jakobshavn Isbræ,
1651
Greenland, compared to the surging Bering Glacier, Alaska, U.S.A. Annals Glaciology 30, 243-
1652
249.
1653
1654
1655
1656
1657
Meier, M.E., Post, A., 1969. What are glacial surges? Canadian Journal of Earth Sciences 6, 807817.
Merry, C. J., Whillans. I.M., 1993. Ice-flow features on Ice Stream B, Antarctica, revealed by SPOT
HRV imagery. J. Glaciol., 39, 515-527.
Mohr, M., Warren, J.K., Kukla, P.A., Urai, J.L., Irmen, A., 2007.Subsurface record of salt glaciers in
1658
an extensional intracontinental setting (Late Triassic of northwestern Germany). Geology 35,
1659
963-966.
1660
Mottershead, D.N., Duane, W., Inkpen, R.J. Wright, J.S., 2005. Subaerial karstic erosionof small-
1661
scale saltrock terranes. In Gutierrez, F., Guitiierrez, M., Desir, G. Guerro, J. Lucha, P. Marin,
1662
C., GuarciaRuiz, J. M., (Eds.) 6th International conference on Geomorphology Zaragosa,
1663
Cardona, Spain. Abstract volume, p. 453.
1664
Murray, T., Rippin. D, Barrett, B., Kilner, M., 2005. Thermal regime and ice-water content of a
1665
glacier in the European Alps using borehole and surface radar. Eos, Trans. AGU, 86(52), Fall
1666
Meet. Suppl., Abstract C51B-0280.
57
1667
Nelson, T.H., Fairchild, L., 1989. Emplacement and evolution of salt sills in the northern Gulf of
1668
Mexico (abs.). Houston Geological Society Bulletin 32, 6–7.
1669
Nye, J. F., 1952. The mechanics of glacial flow. J. Glaciol., 2, 82-93.
1670
Nye, J.F. 1965. The flow of a glacier in a channel of rectangular, elliptic or parabolic cross-section.
1671
1672
1673
1674
J. Glaciol., 5, 661-690.
Odé, H., 1963, Mechanical properties of salt. Geological Society of America Special Paper 88, 120.
Pappalardo, R.T., Barr, A.C., 2004. The origin of domes on Europa: the role of thermally induced
1675
compositional diapirism. Geophysical Research Letters 31, L01701,
1676
doi:10.1029/2003GL019202.
1677
Pappalardo, R.T., Head, J.W., Greeley, R., Sullivan, R.J., Pilcher, C., Schubert, G., Moore, W.B.,
1678
Carr M.H., Moore, J.M., Belton, M.J., Goldsby, D.L., 1998. Geological evidence for solid-state
1679
convection in Europa's ice shell. Nature 391, 365-368.
1680
Paterson, W.S.B., 1994. Physics of glaciers. Third edition, Pergamon Press, Oxford, 390 pp.
1681
Pattyn, F., De Smedt, B., Souchez. R., 2004. Influence of subglacial Vostok lake on the regional
1682
1683
1684
1685
1686
1687
1688
ice dynamics of the Antarctic ice sheet: a model study. J. Glaciol., 50, 583-589.
Payne, A.J. 1999. A thermomechanical model of ice flow in West Antarctica. Clim.Dyn., 15, 115125.
Payne, A.J., Dongelmans, P.W., 1997. Self-organization in the thermomechanical flow of ice
sheets. Journal of Geophysical Research 102(B6), 12219-12233.
Payne, A.J. Baldwin, D.J., 1999. Thermomechanical modelling of the Scandinavian ice sheet:
implications for ice-stream formation. Annals Glaciology 28, 83-89.
1689
Pazzaglia, F.J., Knueper, P.L.K., (Eds.), 2001. The steady-state orogen: concepts, field
1690
observations, and models; American Journal of Science 302, Issues Nos. 4 & 5.
1691
1692
1693
Peale, S.J., 2003. Tidally induced volcanism. Celestial Mechanics and Dynamical Astronomy 87,
129–155.
Peel, F.J., Travis, C.J., Hossack, J.R., 1995. Genetic structural provinces and salt tectonics of the
1694
Cenozoic offshore U.S. Gulf or Mexico: a preliminary analysis. In: Jackson, M.P.A., Roberts,
1695
D.G., Snelson, S., (Eds.) Salt tectonics; a global perspective; AAPG Memoir 65, 153-175.
58
1696
Perutz, M. F., 1948. A Description of the Iceberg Aircraft Carrier and the Bearing of the Mechanical
1697
Properties of Frozen Wood Pulp upon Some Problems of Glacier Flow. J. Glaciol., 1, 95–104.
1698
Petit, J-R., Jouzel, J. Raynaud, D. Barkov, N.I. Barnola, J.M. Basile, I. Benders, M. Chappellaz, J.
1699
Davis, M. Delaygue, G. Delmotte, M. Dotlyakov, V.M. Legrand, M. Lipendoc, V.Y. Lorius, C.,
1700
Pepin, L. Ritz, C., Saltzman, E., Stievenard, M., 1999. Climate and atmospheric history of the
1701
past 420,000 years from the Vostok ice core, Antarctica. Nature 399, 429-436.
1702
1703
1704
1705
1706
Pfiffner, O.A., 1993. The structure of the Helvetic nappes and its relation to the mechanical
stratigraphy. Journal of Structural Geology 15, 511-521.
Player, R.A., 1969, Salt plugs study. Iranian Oil Operating Companies, Geological and Exploration
Division, Tehran, Report No. 1146.
Pohjola, V.A., Hedfors. J. 2003. Studying the effects of strain heating on glacial flow within outlet
1707
glaciers from the Heimefrontfjella Range, Dronning Maud Land, Antarctica. Annals Glaciology
1708
37, 134-142.
1709
Posey, H.R., Price, P.E., Kyle, J.R., 1987. Mixed carbon sources for calcite cap rocks of gulf coast
1710
salt domes. In Lerche, I., O’Brien, J.J., (Eds.) Dynamical Geology of salt and related structures.
1711
Academic Press, Orlando, Florida, 593-630.
1712
1713
1714
1715
1716
1717
1718
1719
Priscu, J.C., Christner, B., 2004. Earth’s icy biosphere. In: Bull A. T. (Ed), Microbial Diversity and
Bioprospecting. American Society for Microbiology, Washington, D.C. p. 130-145.
Priscu, J.C, Christner, B.C, Foreman, C.M., Royston-Bishop, G., 2005. Biological Material in Ice
Cores. Encyclopedia of Quaternary Sciences Submitted 6 October 2005.
Ramberg, H., 1975. Particle paths, displacement, and progressive strain applicable to rocks.
Tectonophysics 28, 1-37.
Ramberg, H., 1981. Gravity, Deformation and the Earth’s crust (second edition), Academic Press,
London. 452 pp.
1720
Ramsay, J.G., 1967. Folding and fracturing of rocks, McGraw Hill, New York. 568pp.
1721
Ramsay, J.G., Huber, M.I., 1983. The techniques of modern structural Geology, Vol 1: Strain
1722
1723
1724
analysis, Academic Press, London, 307 pp
Raymond, C. F., Harrison, W. D. 1988, Evolution of Variegated Glacier, Alaska, U.S.A., prior to its
surge. J. Glaciol., 34, 154-169.
59
1725
1726
1727
Reeh, N., 1987. Steady-state three-dimensional ice flow over an undulating base: first-order theory
with linear ice rheology. J. Glaciol. 33, 177-185.
Reyss, J. L., Pirazzoli, P. A., Haghipor, A., Hatté, C., Fontugne, M., 1998. Quaternary marine
1728
terraces and tectonic uplift rates on the south coast of Iran. Geological Society London Special
1729
Publications 146; p. 225-237
1730
1731
Robin, G.De Q., Swithinbank, C., 1988. Fifty years of progress in understanding ice sheets. J.
Glaciol., Special Issue, 33-48.
1732
Roedder, E., Belkin, H.E., 1980. Thermal gradient migration of fluid inclusions in single crystals of
1733
salt from the Waste Isolation Pilot Plant site (WIPP), In Northrup, C.J.M., (Ed), Scientific Basis
1734
for Nuclear Waste Management: Volume 2, Plenum Press, New York City, New York, p. 453-
1735
464.
1736
1737
1738
1739
1740
Rogers, S.O, Starmer, W.T., Castello, J.D. 2004. Recycling of pathogenic microbes through
survival in ice. Medical Hypotheses 63, 773-777.
Rosenkrans, R.R., Marr, D.J., 1967. Modern seismic exploration of the Gulf coast Smackover trend.
Geophysics 32, 184-206.
Rouby, D., Raillard, S., Guillocheau, F., Bouroullec, R., Nalpas, T., 2002. Kinematics of a
1741
growth/raft system on the W African margin using 3-D restoration. Journal of Structural Geology
1742
24, 783–796.
1743
Rowan, M.G., 2003, Stratigraphy and biostratigraphy in and immediately adjacent to salt bodies, an
1744
Overview AAPG Annual Convention Abstracts May 11-14, 2003, Salt Lake City, Utah (no page
1745
#).
1746
1747
1748
Rowan, M.G., Jackson, M.P.A., Trudgill, B.D., 1999. Salt-related fault families and fault welds in the
northern Gulf of Mexico. AAPG Bulletin 83, 1454–1484.
Sage, S.F., Gudela Von Gronefeld, G.V., Déverchère, J., Gaullier, V., Maillard, V., Gorini, C., 2005.
1749
Seismic evidence for Messinian detrital deposits at the western Sardinia margin, northwestern
1750
Mediterranean Marine and Petroleum Geology 22, 757-773.
1751
Satterfield, C.L., Lowenstein, T.K., Vreeland, R.H., Rosenzweig, W.D., Powers, D.W., 2005. New
1752
evidence for 250 Ma age of halotolerant bacterium from Permian salt crystal: Geology 33, 265–
1753
269.
60
1754
Säwström, C., Mumford, P., Marshall, W., Hodson, A., Laybourn-Parry. J., 2004. The microbial
1755
communities and primary productivity of cryoconite holes in an Arctic glacier (Svalbard 79°N).
1756
Polar Biology 25, 591-596.
1757
1758
Schenk, P., Jackson, M. P. A., 1993. Diapirism on Triton: a record of crustal layering and instability.
Geology 21, 299–302.
1759
Schenk, P.M., Pappalardo, R.T., 2004. Topographic variations in chaos on Europa: implication for
1760
diapiric formation. Geophysical Research Letters 31, L16703, doi:10.1029/2004GL019978.
1761
Schleder, Z., Urai, J.L. 2006. Deformation and recrystallization mechanisms in mylonitic shear
1762
zones in naturally deformed extrusive Eocene-Oligocene rocksalt from Eyvanekey plateau and
1763
Garmsar hills (central Iran), Journal of Structural Geology 29, 2241-255.
1764
Schleder, Z., Urai, J.L. 2008.
1765
http://www.ged.rwthaachen.de/Ww/people/alumni/zsolt/SchlederUrai_SPP04_v08/
1766
Halite_microstructures_SPP.htm
1767
Schuster, D.C., 1995. Deformation of allochthonous salt and evolution of related salt-structural
1768
systems, Eastern Louisiana Gulf Coast. In: Jackson, M.P.A., Roberts, D.G., Snelson, S., (Eds.),
1769
Salt tectonics; a global perspective; AAPG Memoir 65, 177-198.
1770
Seni, S.J., 1987. Evolution of Boling dome cap rock with emphasis on terrgigenous clastics, Fort
1771
Bend and Wharton Counties, Texas. In Lerche, I., O´Brien J.J., (Eds.), Dynamical Geology of
1772
salt and related structures. Academic Press, Orlando, Florida, 543-591.
1773
1774
Sherkati, S, Letouzey, J., 2004. Variation of structural style and basin evolution in the central
Zagros (Izeh zone and Dezful Embayment), Iran. Marine and Petroleum Geology 21, 5535-554.
1775
Showman, A.P., Han, L., 2004. Numerical simulations of convection in Europa's ice shell:
1776
implications for surface features Journal of Geophysical Research 109, E01010,
1777
doi:10.1029/2003JE002103.
1778
Shumskii, P.A., 1964. Principles of structural glaciology. Dover Publications, New York, 497 pp.
1779
Siegert, M.J. 2007. Sunless Seas. Geoscientist 17, 20-23.
1780
Smith, A.W., Skilling, D.E. Castello, J.D. Rogers. S.O., 2004. Ice as a reservoir for pathogenic
1781
human viruses: Specifically calciviruses, influenza viruses, and enteroviruses. Medical
1782
Hypotheses 63, 560–566.
61
1783
1784
Sohn, H.G., Jezek, K., van der Veen C., 1998. Jakobshavn Glacier, West Greenland: 30 years of
spaceborne observations, Geophysical Research Letters 25, 2699-2702.
1785
Souchez, R., Vandenschrick, G., Lorrain, R., Tison, J.L., 2000 Basal ice formation and deformation
1786
in central Greenland; a review of existing and new core data. In: Maltman A.J, Hubbard, B.,
1787
Hambrey, M., (Eds.) Deformation of Glacial materials. Geological Society London Special
1788
Publications 176. 13-22.
1789
1790
1791
Sowers, T., 2001. The N2O record spanning the penultimate deglaciation from the Vostok ice core.
Journal Geographical Research 106, 31903-31914.
Spiers, C.J, Schutjens, P.M.T.M, Brzesowsky, R.H, Peach, C.J., Liezenberg, J.L, Zwart, H.J., 1990.
1792
Experimental determination of constitutive parameters governing creep of rocksalt by pressure
1793
solution. In Knipe, R.J., Rutter, E.H., (Eds.), Deformation mechanisms, rheology and tectonics.
1794
Geological Society of London Special Publications 54, 509-522.
1795
1796
Suppe, J., 1985. Principles of Structural Geology. Prentice-Hall, Englewood Cliffs, New Jersey. 537
pp.
1797
Talbot, C. J., 1978. Halokinesis and thermal convection. Nature, 173, 739-741.
1798
Talbot, C.J., 1979. Fold trains in a glacier of salt in southern Iran. Journal of Structural Geology 1,
1799
1800
1801
1802
1803
1804
1805
1806
1807
1808
1809
1810
1811
5-18.
Talbot, C.J., 1981. Sliding and other deformation mechanisms in a salt glacier, Iran. Geological
Society of London Special Publication 9, 173-183.
Talbot, C.J., 1992. Quo vadis tectonophysics? With a pinch of salt! Journal of Geodynamics 16, 120.
Talbot, C.J., 1998, Extrusions of Hormuz salt in Iran, In: Blundell, D.J., Scott, A. C., (Eds.),
Geological Society of London Special Publication 143, 315-334.
Talbot, C.J., 2004, Extensional evolution of the Gulf of Mexico basin and deposition of Tertiary
evaporite: Discussion. Journal of Petroleum Geology 27, 95-104.
Talbot, C.J., 2008. Hydrothermal salt- but how much? Discussion. Marine & Petroleum Geology
25, 191-202.
Talbot, C.J., Rogers, E., 1980. Seasonal movements in an Iranian salt glacier. Science, Wash. 208,
395-397.
62
1812
1813
1814
1815
1816
1817
1818
1819
Talbot, C.J., Jarvis, R. J., 1984: Age, budget and dynamics of an active salt extrusion in Iran.
Journal of Structural Geology 6, 521-533.
Talbot, C.J., Jackson, M. P. A.., 1987. Internal kinematics of salt diapirs. AAPG Bulletin 71, 10681098.
Talbot, C.J., Alavi, M., 1996. The past of a future syntaxis across the Zagros. Geological Society
London Special Publications, 100; 89-109.
Talbot, C.J., Stanley, W., Soub, R., A-Sadoun, N., 1996. Epitaxial salt reefs and mushrooms in the
southern Dead Sea. Sedimentology 43, 1025-1047.
1820
Talbot, C.J., Medvedev, S., Alavi, M., Shahrivar, H., Heidari, E., 2000. Salt extrusion rates at Kuh-e-
1821
Jahani, Iran: June 1994 to November 1997, Geological. Society London Special Publications
1822
174, 93-110.
1823
1824
1825
1826
Talbot, C.J., Aftabi, P., 2004, Geology and models of salt extrusion at Qum Kuh, central Iran.
Journal of the Geological Society of London 161, 321-334.
Thorsteinsson, T, Waddington, E.D., 2002. Folding in strongly anisotropic layers near ice-sheet
centers. Annals Glaciology 35, 480-486.
1827
Tison J. L., Hubbard, B., 2000. Ice crystallographic evolution at a temperate glacier: Glacier de
1828
Tsanfleuron, Switzerland. Geological Society London Special Publications 176, 23-38.
1829
1830
1831
1832
1833
1834
1835
Turcotte, D. L., Schubert, G., 1982. Geodynamics: applications of continuum physics to geological
problems. John Wiley & Sons. New York. 450 pp.
Unwin, B., Wingham, D., 1997. Topography and dynamics of Austfonna, Nordaustlandet, Svalbard,
from SAR interferometry. Annals Glaciology 24, 403-408.
Urai J,L., Spiers, C. J., Zwart, H. J., Lister, G.S., 1986. Weakening of rock salt by water during long
term creep. Nature 324, 554-557.
van de Wal, R.S.W, Boot, W. Van de Broeke, M.R., Smeets, C.J.P.P., Reijmer, C.H., Donker,
1836
J.J.A., Oerlmans, J. 2008. Large and rapid melt-induced velocity changes in the ablation zone
1837
of the Greenland ice sheet. Science 321, 111-113.
1838
1839
1840
1841
Ville, L., 1856. Notice geologique sur les salinesde Sahrez et les gites de sel gemme de Rang el
Melah et d’Ain Hadjera (Algerie). Annales des Mines 15, 351-410.
Volozh, Y., Talbot, C., Ismail-Zadeh, A., 2003. Salt structures and hydrocarbons in the Pricaspian
basin. AAPG. Bulletin 87, 313-334.
63
1842
1843
1844
1845
1846
1847
1848
1849
1850
1851
1852
1853
Vreeland, R. H., Rosenzweig, W.D., Powers, D.W., 2000. Isolation of a 250 million-year-old
halotolerant bacterium from a primary salt crystal. Nature 207, 897-900.
Waddington, E.D., Bolzan, J.F., Alley, R.B., 2001. Potential for stratigraphic folding near ice-sheet
centers. J. Glaciol., 47, 639-648.
Walker, J.C.F., Waddington. E.D., 1988. Early discoverers XXXV: Descent of glaciers: some early
speculations of glacier flow and ice physics. J. Glaciol., 34, 342-348.
Wang, Y., Thorsteinssson, T., Kipfstuhl, J., Miller, H., Dahl-Jensse, D., Shoji, H., 2002. A vertical
girdle fabric in the NorthGRIP deep ice core, North Greenland. Annals Glaciology 35, 515-520.
Warren, J K. 1997, Evaporites, brines and base metals: Fluids, flow and 'the evaporite that was'',
Australian Journal of Earth Sciences 44, 149 -183.
Warren, J., 1999. Evaporites, their evolution and economics. Blackwell Science, Oxford, UK. 438
pp.
1854
Washburn, A.L. 1979. Geocryology. Edward Arnold, London, 406 pp.
1855
Weinberg, R.F., 1993. The upward transport of inclusions in Netwonian and power-law salt diapirs.
1856
1857
Tectonophysics 228, 141-150.
Weinberger, R., Begin, Z.B., Waldman, N., Gardosh, M., Baer, G., Frumkin, A., Wdowinski, S.,
1858
2006. Quaternary rise of the Sedom diapir, Dead Sea basin, In Enzel, Y., Agnon, A., Stein, M.,
1859
(Eds.) New frontiers in Dead Sea paleoenvironmental research: Geological Society of America
1860
Special Paper 401, 33-51.
1861
Wenkert, D.D., 1979. The flow of salt glaciers, Geophysical Research Letters 6, 523-526.
1862
Wilson, C.J.L. Zhang, Y., 1996. Development of microstructure in the high-temperature deformation
1863
of ice. Annals Glaciology 23, 293-302.
1864
Wilson, C.J.L., Marmo, B., 2000, www.virtualexplorer.com.au/ VEjournal/Volume2.
1865
Wingham, D.J., Siegert, M J., Shepherd, A., Muir, A.S., 2006. Rapid discharge connects Antarctic
1866
1867
subglacial lakes. Nature 440, 1033-1036.
Worrall, D.M., Snelson, S., 1989. Evolution of the northern Gulf of Mexico, with emphasis on
1868
Cenozoic growth faulting and the role of salt. In: Bally, A., Palmer, A., (Eds.), The geology of
1869
North America-an overview. Geological Society of America, Boulder, Colorado, pp. 97–138 (v.
1870
A).
64
1871
1872
1873
1874
Worster, M.G., Wettlaufer J.S., 1999. In Shyy, W. (Ed) Fluid Dynamics at Interfaces, Cambridge
University Press, Cambridge, p. 339.
Wu, S., Bally, A.W., Cramez, C, 1990. Allochthonous salt, structure and stratigraphy of the northeastern Gulf of Mexico, Part 11; Structure. Marine and Petroleum Geology 7, 334-370.
1875
Yatsu, E., 1988. The nature of weathering. Sozosha,Tokyo,624 pp.
1876
Yuen, D.A., Saari, M.R., Schubert, G., 1986. Explosive growth of the heating instabilities in the
1877
1878
down-slope creep of ice sheets. J. Glaciol., 32, 314-320.
Zirngast. M., 1996. The development of the Gorleben salt dome (northwest Germany) based on
1879
qualitative analysis of peripheral sinks. Geological Society of London Special Publication 100,
1880
203-226.
1881
Zwally, H.J., et al., 2006. Eos Trans. AGU 87 Fall meet suppl. Abstract C14B-00.
1882
1883
Talbot and Pohjola Figure captions
1884
1885
Fig. 1. Photographs illustrating similarities between bodies of ice (In) and salt (Sn): IA: Outlet
1886
glaciers draining from Vatnajökull ice cap, southern Iceland (Photos of ice by VP; of salt by CT
1887
unless otherwise indicated). SA: Perspective satellite view looking east at 4 numbered
1888
extrusions of Hormoz salt at different stages of development in the Zagros Mountains, SE of the
1889
city of Lar, Iran. The dark tones are due to clays in soils of insoluble Hormoz materials left on
1890
the surface after salt dissolution. 1: Vigorous young salt dome rises to 2.2 km above sea level
1891
in the western end of a short Zagros anticline. 2: A namakier without visible salt on its surface
1892
flows 7 km south from diapir breaching Kuh-e-Gach anticline. 3: Another namakier flows north
1893
from a more subdued salt dome in Kuh-e-Burkh anticline. 4: Only gravels of Hormoz materials
1894
remain on the plain north of the remains of a salt dome in an amphitheatre of Cretaceous-
1895
Eocene limestones at the west end of Kuh-e-Gach. Image by
1896
NASA/GSF/METI/ERSDAC/JAROS and the U.S/Japan ASTER Science team and is available
1897
at: http://asterweb.jpl.nasa.gov/content/02_gallery/images/salt-iran-view
1898
IB: Vatnajökull ice cap, southern Iceland, from the south. SB: Salt dome and southern namakier of
1899
Kuh-e-namak (Dashti), Iran from south (28°17'S, 51° 43'E). IC: Franz Josef glacier, New
1900
Zealand from www. SC: Looking up at NE corner of salt dome at Kuh-e-namak (Dashti), Iran.
65
1901
ID: Ice pinnacles of ice towering over terminal face of Franz Josef glacier, New Zealand from
1902
picasaweb.google.com/ SD: Looking NW over northern namakier of mid-reaches of Kuh-e-
1903
namak (Dashti), Iran
1904
1905
Fig. 2. More photographs comparing ice (In) and salt (Sn) on different scales. IA: The advancing
1906
front of Franz Josef Glacier, New Zealand. Debris-rich basal ice facies seen below upwarped
1907
foliation in layered ice. SA: Dirty folded and jointed salt on NE corner of northern namakier of
1908
Kuh-e-namak (Dashti), Iran. IB: Ice pinnacle on Franz-Josef glacier, New Zealand, from
1909
www.virtualtourist.com. SB: salt pinnacle of Ku-e-namak (Dashti) (photo by M.P.A. Jackson).
1910
IC: Icicles and regelated debris bands at the base of Nordenskiöldbreen, Svalbard. SC:
1911
Stalactites and stalagmites of secondary halite fronting river cave well above base of northern
1912
namakier of Kuh-e-namak (Dashti). ID: Looking up-glacier at open upright folding of
1913
stratification with axial plane foliation, Austre Lovénbreen, Svalbard (from Hambrey and
1914
Lawson, 2000, fig 8).SD: Recumbent similar-type fold in mylonitised distal salt of the
1915
axisymmetric “fried egg” salt fountain Zagros Mountains (at 28º00’N, 54º55’E). IE: 'Similar' fold
1916
in fine-grained and coarse bubbly ice foliation with vertical axial plane, Vadret del Forno,
1917
Switzerland (from Hambrey and Lawson 2000, fig 6).SE: Similar folds with axial plane foliation
1918
in Tertiary salt of Qum Kuh, central Iran (see Talbot and Aftabi, 2004).
1919
1920
Fig. 3. Comparison of ice and salt cycles. A: Ice usually flows from where and soon after it forms.
1921
B: Salt sequences have to be buried and squeezed back to the surface for namakiers to
1922
become analogues of glaciers.
1923
1924
Fig. 4. Categories of salt diapirs. A: The crest of a “passive” diapir stays near the depositional
1925
surface as it is downbuilt by other rocks accumulating on the surrounding source layer. B:
1926
Differential loading drives reactive diapirs to pierce overburdens weakened and thinned by faulting,
1927
here as is common in Iran, at a releasing bend along a strike-slip fault. If reactive diapirs reach the
1928
surface they may rise further by active growth through pre-growth overburden and/or downbuilding
1929
by syn-growth overburden. C: Rare active diapirs upbuild through pre-growth ductile then syn-
1930
growth overburden.
66
1931
1932
Fig. 5. Types of salt structures (after Jackson and Talbot, 1994) with extrusive salt analogues of
bodies of crystalline ice shaded.
1933
1934
Fig. 6. Topographic profiles of extrusions of salt in the Zagros Mountains of Iran (from fig. 3 in
1935
Talbot, 1998) superposed to illustrate: A: how they grow with time and B, when no longer
1936
supplied by salt from depth, collapse to the theoretical profile of a viscous droplet (dashed) and
1937
then degrade, mainly by dissolution.
1938
1939
Fig. 7. Natural scale centred profiles of extrusions of salt of different ages in Iran and their internal
1940
structures. A. N-S profile of Hormoz salt (~500 Ma) that rose up an opening pulled-apart along
1941
a NS transfer fault in the Zagros Mountains to extrude Kuh-e-Jahani, probably the largest
1942
current salt fountain in Iran, at 28°37’N, 52°25’E (from Talbot et al., 2000). Star indicates flow
1943
separation point beneath summit dome. B. N-S profile of Tertiary salt extruding at Qum Kuh at
1944
34°48’N, 50°41’E on the central plateau of Iran (fr om Talbot and Aftabi, 2004). Notice the cap
1945
soil of insoluble components on the surface (black) thickening downstream as the salt dissolves
1946
and parts of the outer carapace of brittle dilated salt (grey in inserts) in divides between rivers
1947
that erode down toward confined salt. Folds in colour bands within the salt are also indicated.
1948
C: Folds near the southern (feather-edged) retreating terminus of Kuh-e-namak (Dashti). Salt
1949
relief shown is about 45 m. D: Tank-track fold ~100 m short of terminus in northern namakier
1950
off Kuh-e-namak (Dashti). Figure for scale right foreground.
1951
1952
Fig. 8. Photographs of grain shape fabrics in glacier ice and namakiers of salts of different ages in
1953
Iran. Pictures of Ice were taken between crossed polars. IA: Thin slice of coarse grain ice
1954
specimen taken from the front of Engabreen, Norway. SA: a halite “pegmatite” in Kuh-e-
1955
Bachoon, Zagros Mountains (29º00’N, 52º15’E) from Talbot, 1998. IB: Thin slice of
1956
heterogeneous and recrystallised fine grained ice extracted from a tunnel blasted into the base
1957
of Engabreen, Norway. SB: Part of an automated map by electron backscatter diffraction of
1958
polished and chemically etched sample of bimodal salt from the Eyvanekey plateau, ~50 km
1959
east of Tehran. Different orientations of crystal lattice appear in different shades (from
1960
Schleder and Urai 2006 fig. 7A.) IC: Thin slice of homogenous medium grained ice from a core
67
1961
drilled into Storglaciären, Sweden. SC: Bi-modal grain shape fabric in Hormoz salt on southern
1962
namakier Kuh-e-namak (Dashti). ID: Basal ice facies from the ice tunnel beneath Engabreen,
1963
Norway, showing pinched ice layers interbedded with debris rich strata The vertical side is ca
1964
70 cm high SD: Aligned ellipsoidal megacrysts of halite with length/width ratios of 1.5:1 in a
1965
halite gneiss; pipe is 15 cm long, Kuh-e-namak (Dashti). IE Basal ice stratigraphy from the ice
1966
tunnel beneath Engabreen, Norway, showing vertical compression of ice close to larger stones
1967
frozen into the ice (dark objects). The vertical side is ca 70 cm high. SE: Salt mylonites in which
1968
halite grains can reach length/width ratios 4:1 along ductile shear zones typically only dm thick,
1969
here thicker in Tertiary salt in the Eyvanekey plateau, 50 km east of Tehran, Iran.
1970
1971
Fig. 9. Vertical profiles of viscous fluids with passive marker layers gravity spreading over
1972
horizontal substrates with different bottom boundary conditions. A1-3: Three stages in spreading
1973
of a block of silicone putty over a horizontal high-friction substrate (fig. 9-10, Ramberg, 1981).
1974
Distal stratification curls downward distally in tank track fold signaling high-frictional base. A4-6:
1975
Three stages in similar blocks of silicone putty gravity spreading over mercury (figs. 9-7 & 9-12,
1976
rearranged from Ramberg 1981). Stratification curls upward distally over free-slip base. B-F:
1977
Thin vertical slices of experimental extrusions of transparent viscous silicone polymer (from fig.
1978
12, Talbot and Aftabi, 2004). Flow folds in passive markers in the upper limbs of tank-track fold
1979
(numbered) record the number of pauses in their extrusion history. B: Fundamental recumbent
1980
tank-track fold developed with a simple upper limb records steady extrusion. C-E Increasing
1981
number of flow folds in the upper limb of tank-track fold record number of pauses in loading
1982
history. Scale bars in mm. F: Tank track fold undulating over a no-slip irregular slope
1983
(experiment by Rosemary Talbot).
1984
1985
Fig. 10. Passive stratification within salt develops trains of internal similar-type flow folds with an
1986
axial planar grain shape fabric in association with every significant obstruction in their channel floor
1987
they flow beyond (from Talbot, 1981). A: Different regimes of structures shown in B are developed
1988
to different degrees down the namakier. B: Instead of eroding obstructions, the salt infills troughs in
1989
its bed and flows beyond the obstruction by detaching along mylonite zones ≤1 dm thick. C: Only
1990
some of the distal salt flows beyond some of the last obstructions.
68
1991
Fig.11. Stress relief fractures newly exposed beneath an irregular top of the salt in
1992
Drakhsharnamak quarry, NW Eyvanekey Plateau, 60 km east of Tehran (photo taken January
1993
2002). Most of these fractures, a few decimetres apart, are horizontal and slightly oblique to the
1994
salt layering (that dips gently left); many are infilled by relatively insoluble residual soils washed
1995
in from above. The lowest fracture here is estimated to be about 6 m beneath the top of salt
1996
and is unusually long. In the upper slopes of most salt extrusions, the planar fracture
1997
anisotropy may parallel and emphasise a strong gneissose foliation and dies out downward,
1998
passing into sound salt at depths between 4 and 10 m.
1999
2000
Fig. 12. Working model for slow normal flow and short rapid surges in namakiers supplied from a
2001
country-rock vent at an essentially steady rate. A. Most of the time gravity spreads dry salt at
2002
cm/dm/year and develops a tank track fold as extruded salt is stored at high levels by strong
2003
elastic carapace. B. Salt surges at dm/day for a few days over near basal mylonite zones after
2004
3 or 4 rain storms per decade have weakened outer carapace.
2005
Figure 1
Click here to download high resolution image
Fig1-color
Click here to download high resolution image
Figure 2
Click here to download high resolution image
Fig2-color
Click here to download high resolution image
Figure 3
Click here to download high resolution image
Figure 4
Click here to download high resolution image
Figure 5
Click here to download high resolution image
Figure 6
Click here to download high resolution image
Figure 7
Click here to download high resolution image
Fig7-colour
Click here to download high resolution image
Figure 8
Click here to download high resolution image
Fig-8 colour
Click here to download high resolution image
Figure 9
Click here to download high resolution image
Figure 9-colour
Click here to download high resolution image
Figure 10
Click here to download high resolution image
Figure 11
Click here to download high resolution image
Figure 12
Click here to download high resolution image
Table 1
Click here to download Table: Table 1.doc
Table 1: comparison of properties of ice and salt
Chemistry
Name
Crystallography
Typical crystallisation pressure
Moh’s hardness
Density at 1 bar
Ice
H20
Water ice
Hexagonal/trigonal
Up to 253 MPa at -22°C
(Shumskii, 1964, p.180)
1.5 at 0°C to 6 at –78.5°C
3
916.8 Kg m-
Melting point
Solubility in water
~ 0°C
Brittle shear strength
6 kg cm at 3°C (anisotropic)
Typical flow stresses
0.001% of Young’s modulus
Typical viscosity en masse
10
Pa “warm”
14
10 Pa s “cold”
(Shumskii, 1964)
-1
1-10 m a when dry
4
-1
10-10 m ar with basal slip
-6 -1
17-29x10 K (Shumskii1964)
-6
2 -1
1.33x10 m s (Drewry 1986)
mm to dm
1-5 (usually near 3)
Pressure solution creep
dislocation creep, dynamic
recrystalisation, basal slip
Typical flow rates in nature
Linear thermal expansivity
Thermal diffusivity
Typical grain size
Flow law exponent
Flow mechanisms
infinite
-2
10-11
Salt
NaCl
halite
Face centred cubic
66-379 MPa depending on brine concentration
(Yatsu 1988 p. 86).
3
3
Pure halite 2170 Kg m- ,
3
Rock salt often taken as 2200 Kg m~ 700°C (~973 K)
At 25°C, 1 cm of rain can potentially dissolve
0.1667 cm of salt (Lockner, 1995). In
experiment, water with 300 gr/L of NaCl
2
dissolved 0.2 gr/s/ m (Alkattan et al., 1997).
5 MPa at 100ºC (Kawamoto & Shimamoto,
1997)
~7 MPat ~250ºC
(Kawamoto & Shimamoto, 1997)
13-16
Pa s (damp)
10
18
10 Pa s (dry) depending on grain size &
temperature (Warren 1999 p. 154)
-1
-1
1-10 m a (0.5 m dayr when wet)
-5
-1
4.2x10 K (Gevantman, 1981)
-6
2 -1
3.6x10 m s (Durham et al., 1981)
5-20 mm (Warren 1999 p. 159)
1 (damp) 3—8 (dry)
Pressure solution creep, dislocation creep,
dynamic recrystalisation.
Table 2
Click here to download Table: Table 2.doc
Table 2. Comparison of formation and deformation
Material
Formation
lithification
downslope flow
Loss mchanisms
Ice
salt
Snow largely sedimented from above(random c
axes) but also by freezing of infiltrated melt water
resulting in more oriented fabrics
Compaction and crystal growth due to vapour
transport, and strain induced recrystalisation
Compaction to shear in decades
Ablation, sublimation, calving
Crystallisation from brines evaporating
to dryness or epitaxial growth in
permanent evaporating brines.
By cementation
Power law (n=3) penetrative creep, basal slip,
dynamic recrystalisation
Subdued on accumulation zone, ≥10 m in serac
zones
Active strain markers Stratigraphy: grain size, bubble & impurity content,
Foliation, crevasses, crevasse traces & infills, grain
boundaries, bubble elongation
Stratigraphy, foliation (difficult to see)
Passive strain markers
Gravity spreading
mechanisms
Surface relief
6
Burial, diapir, extrusion over 10 a
Erosion (mainly dissolution) over whole
area of extrusion, local river erosion
Power law (n=3) penetrative creep,
basal slip, dynamic recrystalisation
≤5 m on summit dome, increasing to
≥20 m in lower reaches
Non-salt layers, veins
Bedding, obvious grain shape and
orientation foliation,
Folds boudins, stripes, thrusts, ogives, lags
Easy to see folds, boudins, shears,
structures
Difficult to see
thrusts, lags.
Supra-, en-, sub-, lateral, terminal
En- and supra
moraines
Upright, surface waves, time transients
Reclined, recumbent
Dynamic folds
Flow parallel to boundaries so convergent
Where flow is convergent, Tank-Track Fold Easy
Kinematic folds
bed topography related, subtle in low relief
to see on surface relief
Significant by basal tools where basal friction
None, no basal tools instead salt infills
Bedrock erosion
hollows and flows beyond detaching
in-salt mylonites
Impurities and water content, fossil in
Flow regime control Temperature, water content, strong/weak beds
intrusions, mainly rain in extrusions
Clean salt attached no slip.
subglacial impurities common. Cold ice frozen to bed, warm
Types of basal
ice subject to basal slip & regelation
contact
Entrainment, plucking, quarrying and ploughing common
Very minor equivalents where salt flows
Basal erosion
over soil substrates