Vous êtes sur la page 1sur 23

Cold Regions Science and Technology 60 (2010) 107–129

Contents lists available at ScienceDirect

Cold Regions Science and Technology


j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / c o l d r e g i o n s

Review

A review of the engineering properties of sea ice


G.W. Timco a,⁎, W.F. Weeks b
a
Canadian Hydraulics Centre, National Research Council of Canada, Ottawa, ON, K1A 0R6, Canada
b
6533 SW 34th Avenue, Portland, OR 97239-1077, USA

a r t i c l e i n f o a b s t r a c t

Article history: Operations in ice-covered waters require good engineering to ensure the safety of personnel and the
Received 9 July 2009 environment. In polar regions, the presence of sea ice is the main factor hindering the operations. It affects
Accepted 6 October 2009 shipping, and oil and gas exploration and development. This paper looks at the stage of knowledge and
applications of the engineering properties of sea ice. The physical properties (microstructure, thickness,
Keywords:
salinity, porosity, and density) and the mechanical properties (tensile, flexural, shear, uni-axial compression
Sea ice
Thickness
and multi-axial compression strength, borehole strength, failure envelope, creep, elastic and strain modulus,
Mechanical properties Poisson's ratio, fracture toughness and friction) are explored. The paper outlines these properties for both
Strength first-year sea ice and Old Ice (i.e. second-year and multi-year sea ice). Although some properties are
Engineering reasonably well understood (microstructure, salinity, flexural strength, compressive strength, and elastic
First-year sea ice modulus), others are not. Knowledge of Old Ice is particularly limited.
Old Ice Crown Copyright © 2009 Published by Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
2. Growth and microstructure of sea ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
2.1. First-year sea ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
2.2. Old Ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
3. Ice thickness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
3.1. First-year sea ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
3.2. Old Ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
4. Ice salinity and porosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
4.1. First-year sea ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
4.2. Old Ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
5. Ice density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
5.1. First-year sea ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
5.2. Old Ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
6. Tensile strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
6.1. First-year sea ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
6.2. Old Ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
7. Flexural strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
7.1. First-year sea ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
7.2. Old Ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
8. Shear strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
8.1. First-year sea ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
8.2. Old Ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
9. Compressive strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
9.1. First-year sea ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
9.2. Old Ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

⁎ Corresponding author.
E-mail addresses: garry.timco@nrc.gc.ca (G.W. Timco), w-f-weeks@comcast.net (W.F. Weeks).

0165-232X/$ – see front matter. Crown Copyright © 2009 Published by Elsevier B.V. All rights reserved.
doi:10.1016/j.coldregions.2009.10.003
108 G.W. Timco, W.F. Weeks / Cold Regions Science and Technology 60 (2010) 107–129

10. Multi-axial loading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118


10.1. First-year sea ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
10.2. Old Ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
11. Borehole ice strength measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
11.1. First-year sea ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
11.2. Old Ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
12. Sea ice creep . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
13. Elastic and strain modulus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
13.1. First-year sea ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
13.2. Old Ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
14. Poisson's ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
14.1. First-year sea ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
14.2. Old Ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
15. Fracture toughness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
15.1. First-year sea ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
15.2. Old Ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
16. Friction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
17. Final comments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126

1. Introduction Chukchi Seas. Such considerations become particularly important if


the area of interest is in the Antarctic where ice types can be very
Activity is clearly increasing in the polar marine areas of the world. different and studies of ice properties are still fairly limited. Inferring
This includes increases in shipping and in offshore construction in properties of sea ice in other regions of the world must take proper
support of both commercial and tourist operations. Of particular account all of the factors that affect the properties.
current interest is the expansion in offshore oil and gas exploration
and production activities in diverse geographic regions where ice- 2. Growth and microstructure of sea ice
covered waters are of common occurrence. Safe and economical
activities in such regions, typically polar, require very careful and 2.1. First-year sea ice
insightful engineering in that hostile and unusual operating condi-
tions coupled with a fragile environment present significant new Sea ice is a complex material that is composed of solid ice, brine,
challenges and dictate extra diligence. gas, and depending upon the temperature, various types of solid salts.
The presence of sea ice is clearly the main factor contributing to the Environmentally controlled variations in the mechanisms of sea ice
complexity of operations in these regions. What do we know about growth can result in several different grain structures, depending upon
this material? Fortunately, there has been a considerable amount of the prevailing conditions. The most common grain structures include
research performed to measure sea ice properties, especially during granular, columnar, and discontinuous columnar. Although granular
the activity boom in the Canadian and American Beaufort Sea in the ice can form in several different ways, a common method is through
1970s, 1980s and early 1990s. Several review articles have been the churning action that occurs when waves pass through mats of
written to summarize these findings (e.g. Weeks and Assur, 1967, frazil crystals that can occur on the sea surface during the initial stages
1968; Schwarz and Weeks, 1977; Mellor, 1983, 1986; Weeks and of sea ice formation or in open leads in pack ice. Granular ice is usually
Ackley, 1982, 1986)1. Since these review articles were written, activity isotropic (ice properties do not depend on the direction of measure-
on understanding the properties of sea ice has been ongoing. The ment). Also when brine and salt inclusions occur in granular ice, they
present paper attempts to consolidate this knowledge and commu- usually occur between the ice crystals instead of within the crystals.
nicate it in a coherent fashion. Since sea ice properties are highly Columnar ice characteristically forms either at the sea surface when
variable depending upon the material's environmental history, this is conditions are relatively calm or beneath an initial granular frazil layer
challenging, particularly considering the fact that the literature is once vertical motions have subsided. Such ice is comprised of
widely scattered. Although the present paper focuses primarily on columnar crystals that may extend through the complete thickness
mechanical properties, a number of environmental factors affecting of the ice sheet and are aligned parallel to the direction of heat flow. In
property values are discussed in passing. Both first-year and Old Ice most situations this means that the columns are elongated vertically.
are considered. In columnar crystals the brine and salt inclusions are located in a series
There are several things that one should keep in mind when of vertically oriented planes within the ice crystals. The spacings
applying the information contained in this paper. It will be seen that between these planes are slightly less than 1 mm and vary inversely
the great majority of the results discussed come from sites in the with growth rate. In that ice is hexagonal, the planar or basal direction
Arctic and, more specifically, from sites in the North American Arctic. in each crystal where the inclusions are located is normal to the
In applying these results elsewhere, it is important to consider site- direction of the optic or c-axis. As the presence of such inclusions
specific environmental conditions. What is the salinity of the results in a decrease in ice to ice contacts, such planes are planes of
seawater, how protected is the site (are you dealing with pack ice weakness. Furthermore columnar ice comes in two varieties. In one,
or fast ice), are Old Ice types expected to be present, etc.? For example, the ice properties in the horizontal plane are independent of direction.
because of the presence of very large rivers draining into the marginal This means that although the c-axes of the crystals are horizontal, they
seas to the north of the Russian mainland, ice there could be quite are oriented randomly within that plane, resulting in overall
different from ice typical of the coastal regions of the Beaufort or directional independence in the horizontal plane. In the other,
although all the crystals also have their c-axes oriented in the
1
Also of interest is the excellent review article by Gold (1977) on the engineering horizontal plane, they are no longer random, but all point in the
properties of freshwater ice. same direction. In such ice, properties measured parallel to and
G.W. Timco, W.F. Weeks / Cold Regions Science and Technology 60 (2010) 107–129 109

perpendicular to the c-axis alignment direction can be quite different. boundaries within a lattice of essentially pure ice (Light et al., 2003;
Such strong c-axis alignments are frequently observed in the lower Wettlaufer, 1998). It is also important to remember that, even though
parts of thick fast ice sheets and are believed to reflect the mean the upper surface of an ice sheet may be very cold during the winter
current direction beneath the ice (Weeks and Gow 1980; Stander and (close to the ambient air temperature), the lower surface is always at
Michel 1989). the freezing point of seawater (usually −1.8 °C). The temperature
Studies of grain structure are typically carried out via the profile across the sheet is frequently close to linear during the growth
preparation of thin sections, which are then examined using phase. During spring when the air warms-up, the temperature
petrographic techniques similar to those applied to thin sections of gradient in the ice exhibits a “C-shape” which flattens as the air
rock. To do this, an ice sample is first mounted onto a glass slide. temperature increases (Johnston and Timco, 2002). As there are a
Then the thickness of the ice is reduced to a millimetre or less via the number of different salts in the ice, the relationships specifying the
use of a microtome. The resulting thin section is examined using a temperatures at which different salts crystallize are complex (Assur,
polarizing microscope. Needless to say, the implementation of such 1958; Marion et al., 1999). Consideration of all these factors makes
procedures is time consuming. As good descriptions of the details of understanding and characterizing sea ice challenging.
the grain structure variations can be found in the literature (Weeks
and Gow, 1980; Weeks and Ackley, 1982, 1986; Gow et al., 1987a,b; 2.2. Old Ice
Weeks 1998, in press) the reader is referred to these papers for
additional information. Here the important fact is that sea ice can Ice that has survived one summer melt season is termed second-
exhibit granular structures which are isotropic as well as columnar year ice. It is a form of Old Ice. [Note that in the following the
structures that show a variety of anisotropic alignments. These dif- capitalization Old Ice will indicate ice that has survived one or more
ferences are, in turn, reflected in variations in the mechanical summer melt seasons.]. Second-year ice normally does not exceed a
properties of the ice. thickness of 2.5 m and is a two-layered system in which the top layer
Ice has a very selective lattice which incorporates very few has survived one melt season and is underlain by a layer of first-year
impurities (e.g. Fluorine can replace Oxygen under certain growth ice. Property measurements on second-year ice have been sparse
scenarios). When seawater solidifies, the resulting ice sheet tries to (Bjerkelund et al., 1985; Johnston, 1998; Eicken et al., 2002) and little
reject the salts present in seawater. However, it does not do this is known about this ice type. The problem is exacerbated because
completely and some salt is trapped in brine pockets within the often no clear distinction is made between second-year and multi-
growing ice sheet (see Fig. 1). The amount of brine trapped is affected year ice, which leads to second-year ice being incorrectly classified as
by a variety of factors including the salinity of the seawater and the multi-year ice (see Johnston and Timco, 2008). In fact, it is often very
growth rate of the ice. First-year sea ice typically has an average difficult to confidently make these distinctions even with on-ice field
salinity in the range of 4 to 6 parts per thousand (‰) salt. This is measurements. For example, Eicken et al. (2002) reported measuring
significantly lower than the salinity of seawater, which is typically 32 salinities of highly desalinated, thin second-year ice during the SHEBA
to 35‰. The brine, gas and solid salts are usually trapped at sub-grain field program.

Fig. 1. Thin section of first-year ice showing the ice platelets and the brine pockets along the grain boundaries (photo courtesy of M. Johnston).
110 G.W. Timco, W.F. Weeks / Cold Regions Science and Technology 60 (2010) 107–129

Old Ice that has survived two or more summer melt seasons is tions on the reason why the floes were identified as Old Ice. The floe
termed multi-year ice. In contrast to first-year ice, multi-year ice usually observations were made by highly experienced Captains and Ice
has a very low salinity. As such, there is little porosity in the ice and it is Service Specialists from the Canadian Ice Service. This study clearly
considerably stronger than first-year sea ice. Further, as there is little illustrates the fact that Old Ice comes in many disguises and it is
salt, the change in the volume of brine in the ice with increasing frequently difficult to know if such ice should be classified as second-
temperature is small. There is, however, a general decrease in strength year or multi-year unless detailed observations are available. In some
with rising temperature, mostly resulting from internal melting at the cases there was even uncertainty in identifying first-year ice. The old
grain boundaries and the temperature dependence of the strength of the adage that multi-year ice can be identified by its characteristic blue
pure ice matrix. In many ways, the properties of multi-year ice are closer colour was certainly disproven in this research program.
to those of freshwater ice and glacial ice than they are to first-year sea An appreciation for the total range of structural variation possible
ice. As with second-year ice, the oldest layer of undeformed multi-year in naturally occurring sea ice can be obtained by examining the
ice is on top while the bottom layer is first-year sea ice. schematic drawings presented by Cherepanov (1974) [see also Fig. 15
In contrast to first-year ice whose growth and thickness is primarily in Weeks (1998)].
controlled by the overlying meteorological conditions, multi-year ice
thicknesses frequently are the result of both meteorological and 3. Ice thickness
mechanical conditions. This difference affects the microstructure of
the ice. Cox et al. (1984a) and Richter-Menge and Cox (1984) made thin The thickness of the ice is one of the most important engineering
sections of a large number of samples of multi-year ice. The sections properties. For example, the way that ice fails is a direct function of the
show that the grain structure of multi-year ice is quite varied and can ice thickness. Ice loads on offshore structures increase significantly
consist of a mix of granular ice, mixed granular and columnar ice, and with increasing ice thickness. The bearing capacity of an ice cover
pulverized brecciated ice (large 10 mm to 50 mm angular fragments depends largely upon its thickness. The speed at which ships can
surrounded by a fine-grain matrix). This chaotic microstructure is a move through ice-covered waters is directly related to the ice
result of mechanical deformation (ice ridging and rubbling) and thermal thickness. The height and size of pressure ridges and ice pile-up
growth. The columnar ice was observed with a range of inclinations features is directly influenced by the ice thickness.
relative to the surface of the ice which is an obvious indication of Observations of sea ice covers in any region of the world show
deformation. However, there were substantial regions where the regions of level ice and also ice that is rafted or ridged with a resultant
columnar grains remained aligned with the vertical direction of ice. thickness greater than that of level ice (see e.g., Timco and Burden,
There is a very large variation in the grain size, even over small distances. 1997). The relative portion of each ice type (i.e. level ice, ridged or
Fig. 2 shows a thin section of a sample of multi-year ice. Note the absence rafted) varies depending upon the geographic location and the time of
of salt in the sample and the large variation in the grain structure. year. The topic of ridging, rafting and other forms of ice thickening is
Overall, in contrast to first-year sea ice that sometimes can be highly vast and will not be treated in this paper. The following discussion is
anisotropic, multi-year ice is frequently isotropic. However, unde- focused on the thickness of level sea ice.
formed portions of multi-year ice floes can be highly anisotropic.
Recently Johnston and Timco (2008) compiled a “Guide to Old Ice”. 3.1. First-year sea ice
This guide provides a considerable amount of information on recent
measurements on multi-year ice including a collection of photographs The thickness of first-year ice is directly controlled by the ambient
of multi-year and second-year ice floes along with detailed descrip- air temperature, the freezing time (i.e. length of the cold season),
snow type and thickness, wind speed, ocean heat flux and surface
radiation balance. Arctic ice is always thicker than ice in more
temperate climates largely due to the first two factors mentioned
above.
The expected thickness of a first-year ice cover can be estimated by
the freezing degree method (see e.g. Ashton, 1986, pp 234). If a steady
state condition is assumed and the heat transfer between the water
and ice is negligible, the growth rate is determined by the energy
balance at the ice–water interface as:

ϕi dt = ρi Ldhi ð1Þ

where ϕi is the heat flux from ice to the air, ρi is the ice density, L is
the latent heat of fusion of ice, t is the time, and hi is the ice thickness.
If the assumption is further made that the top ice surface temperature
is the same as the air temperature, then

ϕi = ki ðTb −Ta Þ = hi ð2Þ

where Tb and Ta are the temperatures at the bottom and top of the ice
sheet (i.e. the assumption implies that Ta is also the air temperature)
and ki is the thermal conductivity of ice. Integration of Eq. (1) using
Eq. (2) and letting hi = 0 at t = 0, gives:

0:5 0:5
h = ð2ki =ρi LÞ ½ðTb −Ta Þt ð3Þ

Fig. 2. Vertical thin section of multi-year ice. Note the relative absence of the salt This is commonly referred to as the Stefan equation. Since the air
pockets and the large variation in grain structure (photo courtesy of M. Johnston). temperature will vary with time, the application of this equation makes
G.W. Timco, W.F. Weeks / Cold Regions Science and Technology 60 (2010) 107–129 111

use of the sum of the number of freezing degree days [i.e. Σ (Tb −Ta) t] geographic areas delineated by Masterson et al. (2000) for each zone
for the region of interest. for the Northern Hemisphere. Masterson et al. (2000) have provided
Direct application of this equation to any region will always over quantitative predictions of local ice pressure as a function of contact
predict the ice thickness due to the assumptions that have been made area and global ice pressure as a function of ice thickness for each of
in deriving it. The equation does not take into account the effect of the the three zones. Their calculations indicate decreasing pressures with
snow cover which will insulate the top surface thereby making it decreasing severity of ice conditions. Masterson and Frederking
warmer than the air temperature. Wind speed is also not considered (2006) have subsequently applied the freezing degree day concept
and this will affect the heat transfer rate at the ice surface. The ocean to estimates of global loads on offshore structures as part of the
heat flux is also not considered and this will affect the heat transfer at Canadian Standards Association (CSA) Code S471 General Require-
the growth interface. For these reasons, the equation is usually ments, Design Criteria, the Environment, and Loads. This concept has
applied with an “α” factor to empirically account for these effects. also been integrated into the International Standards Organization
When appropriate values for the various parameters are inserted, this (ISO) Arctic Structures Code.
equation simplifies to:

0:5
h = 0:035α½ΣðTb −Ta Þt ð4Þ 3.2. Old Ice

where h is in meters, T is in °C and t is in days. The “α” value is always Multi-year ice can be extremely thick. As previously noted, the
less than one. For example, the authors have used this equation and thickness is usually a combination of thermal growth and mechanical
compared the calculated to the measured ice thickness for the consolidation through pressure-ridging processes. There is consider-
Canadian Beaufort Sea and found that the best fit occurs with α of able debate regarding the thickness of Old Ice in the Arctic. Kwok and
approximately 0.75. This factor will vary for other geographic regions. Rothrock (2009) report a drastic decrease in thickness of ice in the
The equation only deals with the thermodynamics of ice growth and Arctic Ocean. They used submarine and ICESat records and reported
not the ice thickening from mechanical deformation. The limit of its that the average ice thickness decreased between 1980 and 2008 from
applicability for thermodynamic growth is about 2 m. 3.64 m to 1.89 m. Their measurement region does not include coastal
Level first-year sea ice in the Canadian and American Beaufort Sea sites. Eicken et al. (1995) reported measured average thickness of
typically reaches a thickness of about 2 m at the end of the winter (see 2.86 m for Old Ice in the Eurasian sector of the Arctic Ocean. Johnston
e.g. Melling et al., 2005). Other regions such as Sakhalin, Caspian, et al. (2009) have compiled nearly 5000 direct measurements from 31
Bohai Bay, etc. typically have thinner sea ice. studies, spanning 51 years in the Canadian Arctic. Fig. 4 shows the
Masterson et al. (2000) have defined three zones of different levels exceedance probabilities for relatively level floes and deformed
of severity of ice conditions and climate in offshore areas based pressure ridges (after Johnston et al., 2009). The figure shows that
primarily on freezing degree days. They have defined Zone I for the multi-year floes and pressure ridges represent two very distinct
Arctic (annual freezing degree days 3000 to 4000 °C-days), Zone II populations. On average, the mean thickness of a relatively level
(annual freezing degree days approximately 2000 °C-days), and Zone multi-year feature is 5.6 m (±2.2 m). The mean thickness of a
III (annual freezing degree days 1200 °C-days or less). Fig. 3 shows the pressure ridge is, on average, 9.9 m (±4.7 m). The most formidable

Fig. 3. Map showing the geographic areas for each of the three ice Zones for the Northern Hemisphere (after Masterson et al., 2000).
112 G.W. Timco, W.F. Weeks / Cold Regions Science and Technology 60 (2010) 107–129

In a sense by specifying the brine volume2, one integrates the influence


of both ice temperature (Ti) and salinity (Si) in that these two terms
specify its values at temperatures between −0.5 °C≥Ti ≥ −22.9 °C
(Eq. 2, Frankenstein and Garner, 1967)
 
49:185
vb = Si + 0:532 ð6Þ
j Ti j

An alternative method is also available (Cox and Weeks, 1982,


1983) that provides an estimate of the volume of gas present.
Unfortunately this procedure requires high quality density measure-
ments that are difficult to obtain in the field (Timco and Frederking,
1996). Brine volume is usually quoted in terms of the volume in parts
per thousand, similar to the salinity. Alternatively, it can be expressed
as a volume fraction. (For example, a brine volume of 20‰ is
equivalent to a brine volume fraction of 0.020). Note that Eq. (6)
cannot be used to predict salt or brine content in decaying sea ice.
There is no reliable method known to do this.
Although the knowledge of the brine volume in sea ice is useful on
Fig. 4. Exceedance probabilities for relatively level floes and deformed pressure ridges
its own, it is also useful to know the amount of gas present in the ice.
for Old Ice (after Johnston et al., 2009).
In certain instances (especially when brine drainage has occurred) the
air volume can be significant. This is particularly the case in the above
floe sampled in the studies cited here had a mean thickness of 11.3 m sea level portion of Old Ice floes. For this reason it is usually better to
(from Johnston, 2008). The most massive pressure ridge had a mean express the porosity of the ice as the total porosity vT.
thickness of 24.7 m (from Kovacs, 1975).
vT = vb + va ð7Þ
4. Ice salinity and porosity
where vb is the relative brine volume and va is the relative air volume.
Equations for calculating the total porosity can be found in Cox and
4.1. First-year sea ice
Weeks (1983).
Sea ice decays rapidly during the spring and early summer as the
Ice salinity (Si) is usually expressed as the fraction by weight of the
result of increases in air temperature and incoming solar radiation.
salts contained in a unit mass. It is quoted as a ratio of grams per
Associated with this decay are decreases in ice strength and ice
kilogram of seawater, that is, in parts per thousand (‰, or ppt, or psu —
thickness. Once the air temperature warms to approximately −8 °C, the
practical salinity units). It is usually measured by taking a core of sea
great majority of the solid salts present in the ice at low temperatures
ice and then (quickly) cutting the core into discrete pieces, placing
have dissolved back into the liquid phase. If the ambient air temperature
them in individual air-tight containers, and letting the ice melt. Once
rises further, the brine pockets rapidly increase in size, interconnect and
melted, a salinity meter is used to measure the electrical conductivity
form “brine drainage channels”. The channels increase the permeability
of the meltwater, and based on its conductivity and temperature, a
of the ice (Eicken et al., 2002) which results in a desalination of the ice
salinity value is obtained. This approach provides an average value for
(Lake and Lewis, 1970; Cole and Shapiro, 1998). Such higher
the particular vertical section of the ice. Quickness of operation is
temperatures also frequently result in the formation of surface melt
essential since the brine begins to drain from the ice as soon as the core
ponds during the late spring and summer. All of these factors combine to
is removed from the ice cover.
produce ice that is both thinner and significantly weaker.
In sea ice, there is usually some salinity variation with depth in the
ice sheet (Cox and Weeks, 1988). This depth dependence of the salinity
4.2. Old Ice
changes throughout the winter as the salt within the ice migrates
downward through the ice. There can be, however, marked salinity
Kovacs (1996a) has also compiled salinity values on multi-year ice
variations even within a small sample. In many cases, therefore, the
and developed a relationship between the bulk salinity and the floe
average value of a salinity profile is used as a first approximation of
thickness (hFi) as:
salinity for an ice sheet.
Cox and Weeks (1974) examined the average salinity (Si) of 2
growing sea ice and found that there was a consistent variation with Si = 1:85 + ½80217:9 = hFi  ð8Þ
the ice thickness (hi). Kovacs (1996a) expanded this data base and
developed a general expression that relates the bulk salinity of the ice where the floe thickness is in cm with a range up to 900 cm (9 m). There
to its ice thickness: is considerable scatter in the measured data about the fitted curve
resulting in an r2 value of 0.22 [Here r is the correlation coefficient and r2
Si = 4:606 + ½91:603 = hi  ð5Þ provides an estimate of the percentage of the scatter in the data (22%)
that is explained by the regression curve]. This value should be
where the ice thickness is in cm, up to a maximum value of 200 cm contrasted with the r2 value for first-year ice (Eq. (5)) that was 0.73.
(2 m). This approach assumes that there is no salinity variation with Measured average salinity values for Old Ice with a thickness in the
depth through the ice sheet, which is a reasonable first approximation range of 2 to 4 m were 1.5 to about 5‰. For thicker ice floes, there was
for sea ice. Note that this equation does not apply for ice that is less scatter with mean salinities decreasing to values on the order of 1.5
warming in the spring (i.e. decaying ice), or for second-year or multi- to 2‰. Recent measurements reported in Johnston and Timco (2008)
year ice. In these cases, a majority of the salt has drained from the ice 2
It should be noted that the literature uses the term “brine volume” with units of
and salinities are much lower — typically 0.5 to 4‰. parts per thousand. This value is a simple ratio of the volume of liquid brine to the total
Historically sea ice property variations have been analyzed in terms volume of the ice sample. As such, it does not have the units of volume. It is a ratio and
of the brine volume νb (i.e. the amount of liquid) present within the ice. not a true volume.
G.W. Timco, W.F. Weeks / Cold Regions Science and Technology 60 (2010) 107–129 113

reinforce earlier results by Cox and Weeks (1974) that the salinity of
multi-year hummocks ranges from 0 to 4.2‰. Ice in the sails of
hummocks has a very low salinity (less than 1‰) that changes little
during the summer months; likely because there is little salt left to drain.
Salinities below the sails were higher (although still low) and much
more variable. There were also higher salinities “spikes” in the profiles
(up to about 4‰). These may result from seawater infiltrating into voids
in hummocked ice and freezing in an essentially closed system.

5. Ice density

5.1. First-year sea ice

Knowledge of the density of sea ice is important in many


applications. These applications fall into one of two categories. In Fig. 5. Plot of the density versus temperature for four different salinities for gas-free sea
one category, in which the ice moves out of the sea water, the density ice. These density values represent the upper limit of the density for any given salinity
and size of the ice blocks determine the weight of the ice. This and temperature. The addition of gas (i.e. air) inclusions in the ice will decrease the
situation has applications, for example, as ice rides up the face of a density (after Timco and Frederking, 1996).

conical structure, the weight of the ice exerts a load on it. Similarly,
heights of natural ice pile-ups are controlled by the ice density, ice to be relatively insensitive to temperature changes with a base value of
thickness and driving force of the ice. In this category, small variations approximately 0.92 Mg m− 3. The exception here is a pronounced rise
in the density value do not change the load or height estimates by a that occurs at near-melting temperatures. If gas is present in a sample,
great degree. In the second category, when the ice is displaced down the density will be lower than these values.
into the sea water, there is a buoyancy force which is proportional to Unless precise density values for specific ice samples are required,
the difference in density between the ice and the sea water. Here, 0.920 Mg m− 3 should serve as a reasonable estimate for first-year sea
small differences in density can make a large difference in the ice.
buoyancy force. This situation has application on the loads that a ridge
keel can exert on a structure, and on the potential for ice interacting 5.2. Old Ice
with the propeller of a vessel moving through an ice cover.
There are basically four different techniques that have been used A number of measurements of the density of multi-year ice are
to measure the density of sea ice. These include mass/volume, available (Cherepanov, 1966; Kovacs and Mellor, 1971; Hibler et al.,
displacement (submersion), specific gravity, and freeboard and ice 1972; Ackley et al., 1974; Langleben and Pounder, 1963; Richter-
thickness techniques. Details can be found in Timco and Frederking Menge and Cox, 1984; Sinha, 1984; Jeffries et al., 1988). Fortunately,
(1996). Density measurements on first-year can be found in a number for ice below the waterline there does not appear to be a large
of reports (Malmgren, 1927; Butkovich, 1956; Weeks and Lee, 1958; difference between the in situ density of first-year and multi-year ice.
Langleben, 1959; Pounder and Little, 1959; Pounder and Stalinski, However, samples above the waterline show a wider range (0.72 to
1960; Brown, 1963; Langleben and Pounder, 1963; Kohnen, 1972; 0.91 Mg m− 3) and generally lower values.
Nakawo, 1983; Timco and Frederking, 1983a,b, 1986; Sinha, 1984, Hibler et al. (1972) and Ackley et al. (1974) made detailed ice density
1986a; Urabe and Inoue, 1986). In most field measurements, density profile measurements on multi-year ice floes at the 1971–1972 AIDJEX
was measured as part of a larger field operation, and almost always in stations. Measurements of the average densities representative of the
support of some mechanical property testing program. complete ice sheet were between 0.910 and 0.915 Mgm− 3. They also
Reported values vary over a wide range from 0.72 Mgm− 3 to found that the higher the freeboard, the lower the average density as
0.94 Mg m− 3, with an average of approximately 0.91 Mgm− 3. The given by the empirical relation:
reasons for the spread are partly real and partly a function of the
measurement technique used. Accurate measurements which represent ρ = −0:194f + 0:974 ð9Þ
the in situ density of first-year sea ice range from 0.84 to 0.91 Mg m− 3
for the ice above the waterline, and 0.90 to 0.94 Mg m− 3 for the ice where ρ is the bulk density in Mg m− 3 and f is the freeboard in meters.
below the waterline. A number of these test programs have been carried
out in the early spring when the ambient air temperature is relatively 6. Tensile strength
high. Samples are typically small with dimensions on the orders of tens-
of-centimetres, and with a mass up to 2 to 3 kg. There is a distinct The tensile strength is a fundamental property of sea ice. It defines
difference between tests that were performed on samples where brine the maximum tensile stress that the ice can sustain before failure.
drainage was limited as compared to samples where brine drainage was Observations of large areas of sea ice often show a large number of
significant. Samples in which brine drainage was limited gave open leads. These leads result from a tensile failure of the ice. The
consistently higher density values that better represented the in situ tensile strength is important for predicting both large-scale ice
density of sea ice. There is also a distinct difference between densities movements and local ice forecasting. Also, the tensile strength is a
measured on ice above and below the waterline. For the samples with key parameter in defining the failure envelope of sea ice (to be
appreciable brine drainage from the upper part of the ice sheet, values discussed later in this paper). It also represents a key failure mode
for first-year ice ranged from 0.84 to 0.91 Mg m− 3. Below the waterline, when ice interacts with an offshore structure.
density values are much more consistent varying over a smaller range There have not been a large number of tests performed to measure
(0.90 Mg m− 3 to 0.94 Mgm− 3). the true tensile strength of sea ice. This mainly reflects the fact that they
Timco and Frederking (1996) have used the Cox and Weeks (1983) are difficult and time consuming due to the required precise nature of
equations to calculate the density of gas-free sea ice versus temperature the test set-up. For a reliable test, the sample needs to be perfectly
for four different ice salinities (0, 2, 5, and 10‰). The results are shown aligned. To achieve this, the sample needs to be carefully machined. In
in Fig. 5 and can be considered as placing an upper-bound on density addition, the machining should be carried out at fairly low temperature
values at a particular temperature and salinity. In general density proves (<−24 °C) to minimize brine drainage. Secondly, some procedure needs
114 G.W. Timco, W.F. Weeks / Cold Regions Science and Technology 60 (2010) 107–129

to be devised to maintain contact between the sample and the platens of


the test machine. Otherwise the tensile failure will occur between the ice
and the platen instead of within the ice sample. As might be expected,
this is a particular problem at near-melting temperatures when large
brine volumes occur. Fortunately, procedures for mounting end-caps on
tensile samples are now available (Cole et al., 1985; Lee, 1986; Sinha,
1989a; Richter-Menge and Jones, 1993; Richter-Menge et al., 1993),
although they are time consuming. Even so, pure tensile tests are
essentially a laboratory procedure.

6.1. First-year sea ice

There have been a small number of tests of the tensile strength of


first-year sea ice (Peyton, 1963, 1966; Dykins, 1967; Kuehn et al., 1990;
Richter-Menge and Jones, 1993). Fig. 6 shows the results as a function
of temperature for loading in the horizontal plane and strain rates
varying from 10− 3 to 10− 5 s− 1. Note that, as would be anticipated,
there is a general decrease in the tensile strength with increasing
temperature. Richter-Menge and Jones (1993) have also shown that
there is a drop off in strength with increasing porosity (brine + gas)
but it is not linear. Fig. 7 shows a plot of the tensile strength (σt) as a
function of total porosity (vT) for all horizontally-loaded samples. If the
(small) effect of strain rate is neglected, a fit to the data yields
Fig. 7. Plot of the tensile strength of first-year sea ice as a function of total porosity for
horizontally-loaded sea ice.
σt = 4:278vT−0:6455 MPa ð10Þ

6.2. Old Ice


with an r2 = 0.72. Thus, the tensile values are a strong function of the
temperature (and brine volume), but are little influenced by the There have been only a small number of measurements of the tensile
loading rate. Values of the tensile strength loaded in the horizontal strength of Old Ice. Cox and Richter-Menge (1985) measured the tensile
direction of a sea ice sheet range between 0.2 MPa and 0.8 MPa. strength of ice collected from multi-year pressure ridges. They found a
The tests by Peyton (1966) and Dykins (1967, 1968) included large variation in the tensile strength that correlated with variations in
loading with different orientations relative to the growth direction of the structure of their samples. They noted that in brecciated ice, failure
the ice. For columnar ice, both authors reported that the tensile commonly occurred in the finer grained portions of the samples when
strength was about three times higher when loading took place with the columnar portions of the sample were oriented in the hard-fail
tensile stresses exerted parallel to the growth direction of the ice (i.e. direction (i.e., columns parallel to the tensile axis). Also when there was
with vertically-loaded specimens). This is not a surprising result. a large range in the grain sizes present in a sample, there did not appear
Values of the tensile strength in this case ranged up to 2 MPa with to be a significant correlation between grain size and tensile strength.
vertical loading of the sample. Sammonds et al. (1998) have measured the uni-axial tensile
strength of undeformed MY sea ice which was collected from a ∼7 m
thick floe located near Graham Island in the Canadian Archipelago.
Tensile strength tests were performed at a temperature of − 10 °C on
samples with salinities less than 2‰. Seven tests were performed over
a range of strain rates (10− 7 to 10− 2 s− 1) with two loading directions.
Tensile strength values ranged from 0.47 to 1.02 MPa for horizontal
loading and 1.39 to 1.44 MPa for vertical loading. These tests were
performed as part of a very comprehensive test program to outline the
failure envelope and fracture behaviour of sea ice.
Sinha (1989a) reported on a test technique for measuring the
tensile strength of multi-year ice using a closed loop test apparatus.
He reported on the results of two tests. Tensile strengths were on the
order of 0.5 MPa with no strong rate dependence for tests conducted
at 10− 6 and 10− 4 s− 1. Of course, the small number of tests precludes
examining any definitive trends.
Additional tests on the tensile strength of multi-year ice would be
extremely useful. However, they require laboratory control for proper
results. Measured values of the tensile strength of Old Ice give a range
from 0.5 MPa to 1.5 MPa.

7. Flexural strength

The flexural strength of ice is not a basic material property. The test
for flexural strength creates non-uniform stress fields in the ice and
assumptions are required about the material behaviour in order to
Fig. 6. Plot of the tensile strength of first-year sea ice as a function of temperature for interpret the test results. Thus, the flexural strength is generally regarded
horizontally-loaded sea ice. as an index test. Because of its importance and use in ice engineering
G.W. Timco, W.F. Weeks / Cold Regions Science and Technology 60 (2010) 107–129 115

problems, however, a large number of investigators have measured this et al., 1988) and sea ice (Butkovich, 1956, 1959; Weeks and Anderson,
property. In measuring the flexural strength, two different approaches 1958; Butkovich, 1959; Brown, 1963; Tabata, 1967; Dykins, 1968,
have been used: cantilever beam tests, and simple beam tests. For the in situ 1971; Airaksinen, 1974; Tabata et al., 1975; Vaudrey, 1977; Frederking
cantilever test, the ice is cut to form three sides of a beam with the fourth and Hausler, 1978; Saeki et al., 1981; Shapiro and Weeks, 1995; Timco
side uncut and connected to the floating ice sheet. An increasing vertical and Frederking, 1983b; Williams et al., 1992).
load is applied to the free end of the beam until it breaks at the root of the The strength of ice may depend on a large number of parameters
beam. This test has the advantage of being relatively easy to perform on a including the temperature, the loading direction on the ice, the ice
large beam, and of maintaining the temperature gradient in the ice sheet. grain structure, the grain size, the test type (cantilever or simple
Usually the test results are analyzed in terms of simple elastic beam beam), the loading rate, beam size and, for sea ice, the ice salinity and
theory. For the simple beam test, the beam is completely cut free of the ice brine volume. This large number of parameters makes developing
sheet and loaded at three (or four) equidistant points such that the centre strong correlations between the flexural strength of sea ice and the
load is parallel to, but opposed to, the load at the ends of the beam. More above sample characteristics challenging.
often this test is performed in a laboratory on smaller samples of ice cut Several researchers have attempted to relate the strength of sea ice
from the ice cover. With either test arrangement, the equations that are to the brine volume or total porosity of the ice. There is a good reason for
used to calculate flexural strength assume that the ice in the beam is both this. It is generally assumed that as the total porosity in the ice increases,
homogeneous and perfectly elastic. If this were to be the case, then the the strength should decrease since there is less “solid ice” that has to be
flexural strength is approximately equal to the tensile strength. Although broken. Timco and O'Brien (1994) have done the most comprehensive
this may be true for some materials, it is certainly not the case for sea ice. analysis of this problem. They compiled a database of about 2500
The IAHR Committee on Ice Problems has published some guidelines for reported measurements on the flexural strength of freshwater ice (1556
determining the flexural strength of ice (Schwarz et al., 1981). beams) and sea ice (939 beams). Timco and O'Brien (1994) showed that
Considering these limitations, it is realistic to ask why bother with the data for first-year sea ice could be described by:
these tests and how relevant are they as an engineering property for ice.
pffiffiffiffiffi
The flexural strength is significant for several reasons. First and most σf = 1:76 expð−5:88⁎ vb Þ ð11Þ
importantly, many real sea ice failures occur in flexure (pressure ridge
formation, icebreaking vessels failing ice, bending of ice on conical-shaped where σf is the flexural strength of the ice (in MPa) and the brine
structures or collars (see e.g. Brown and Määttänen, 2009). Therefore the volume (vb) is expressed as a brine volume fraction (Fig. 8). The data
test is a reasonable approximation for reality in many applications. Also for the equation was compiled from a large number of investigators,
the test can be carried out in the field using portable equipment in that and from a variety of geographic locations, in both polar and
failures typically occur under comparatively small loads. From a research temperate climates. Therefore it should be quite representative of
viewpoint, the test is very flexible allowing the investigator to study the the flexural strength of sea ice in most regions. The brine volume used
effects on strength of changes in sample orientation in the horizontal to represent the ice beam for any test was taken to be the average
plane as well as variations in the vertical profile. It also allows one to vary, brine volume, determined from the average temperature and salinity
within limits, the volume of ice subjected to stress. In addition, the nature of the beam. Thus, to calculate the flexural strength, it is only
of the samples makes it easy to examine correlations between the necessary to know the average temperature and salinity of the ice.
observed failure surface and visible flaws in the ice. It must be emphasized that this relationship is valid only for cold,
growing ice. In the spring, as the ice begins to decay, the internal brine
7.1. First-year sea ice channel network within the ice begins to interconnect and brine (i.e.
salt) can drain from the ice. Such warm ice has an open internal
There have been a large number of researchers who have measured structure and hence high porosity even though the salt concentration
the flexural strength of both freshwater ice (Frankenstein, 1959, 1961; is low. Calculating the flexural strength using Eq. (11) would suggest
Lavrov, 1969; Timco and Frederking, 1982a; Gow and Ueda, 1984, Gow that flexural strengths can be quite high. One problem is that the

Fig. 8. Flexural strength versus the square root of the brine volume for first-year sea ice (after Timco and O'Brien, 1994).
116 G.W. Timco, W.F. Weeks / Cold Regions Science and Technology 60 (2010) 107–129

flexural strength more accurately depends upon the total porosity strength values. Old Ice has a low salinity compared to first-year ice
(not just the brine porosity). However, very few scientists have and thus strength values would be higher than first-year ice compared
measured the ice density during the flexural strength tests so it is at the same temperature. The strength value would not be expected to
currently not possible to calculate the total porosity (brine plus air) be as high as those for freshwater ice since there is some salt present.
for their tests. For warm ice, where water appears on the surface, Therefore estimates for the flexural strength of Old Ice would be on
Eq. (11) would not be appropriate. In such cases, brine drainage the order or 0.8 to 1.1 MPa in the winter and 0.4 to 0.6 MPa in the
within the ice may be appreciable and flexural strength values in the summer.
range of 100 to 150 kPa are probably more appropriate. Although there are no reported small-scale measurements of the
There are a few other important features of the flexural strength of flexural strength of Old Ice, it should be noted during the 1960s and
sea ice that should be noted: 1970s, Panarctic Oil Ltd. placed very high loads (drill rigs) on
(thickened) Old Ice in the Archipelago in Canada's Arctic. Special
• Loading rate: Very few investigators report the loading rate for their
criteria were developed for calculating the strength and creep of this
tests. Typically, the time-of-loading for tests of this type are on the
type of ice. Unfortunately details on this are not publically available
order of 0.5 s to 30 s. With this loading time it is generally considered
and remain in the grey literature.
that there is not a strong functional dependence of the loading rate on
the flexural strength of ice. Some tests (e.g. Tabata et al., 1967, 1975;
8. Shear strength
Määttänen, 1975) have been performed with a much higher loading
rate and these show an increase in strength with loading rate. How-
Ice covers interacting with structures or existing ice features are
ever results of Enquist (1972) and Määttänen (1975) suggest that if
often subjected to biaxial stress conditions involving tensile and
corrections are made to account for the inertial forces associated with
compressive stresses or a shear stress. Thus, the shear strength of ice is
the displacement of water during such tests, the increases disappear
an important mechanical property. Shear testing, however, presents
and the flexural strength becomes essentially independent of loading
difficulties both in performing experiments and interpreting results.
rate. This latter result is appealing inasmuch as tensile strength has
This is mainly due to problems associated with generating a stress
been found to be essentially independent of loading rate.
condition that corresponds to the one assumed when analysing the
• Beam size: There has been considerable discussion in the literature
test results. Normally it is assumed that a uniform shear stress is
on the influence of beam size on the strength of ice. Parsons et al.
generated on a plane of failure, but in many (if not most) cases,
(1992) have shown that there is not a large size effect on the flexural
indeterminate normal stresses are also generated. Torsion, direct shear
strength of sea ice. An examination of the plots of large beams
and punching are the primary methods of generating shear stresses.
versus small beams in Timco and O'Brien (1994) support this result.
• Test type: Timco and O'Brien (1994) did a comprehensive compar-
8.1. First-year sea ice
ison of the results of flexural strength measured using either the
simple beam or the cantilever beam approach. They found that there
There have been relatively few measurements of the shear
was a significant difference for freshwater ice. This has been
strength of sea ice. Butkovich (1956) conducted tests with a double
attributed to stress concentrations at the root of a cantilever beam
shear device in which a cylindrical specimen (76 mm in diameter,
that results in a too low strength (see e.g. Svec et al., 1985). This
300 mm long) was fixed at both ends to a cylindrical support. The
does not have a large effect for the more ductile sea ice.
central section was loaded at right angles to the long axis of the
• Loading direction: There are two issues with regard to loading
specimen. Shear occurred on two circular surfaces 76 mm in diameter
direction; viz, crystal orientation effects and top/bottom tension
and the total area of these two surfaces was used to calculate the shear
tests (i.e. full thickness tests). First, Kayo et al. (1983) show a
strength. The specimens were made from cores of first-year sea ice
difference in strength between small samples loaded with 3-point
(salinity 6‰) in which the long axis was parallel to both the growth
arrangement and beams cut vertically and horizontally from the ice
direction and to the long axis of the columnar grains. Average shear
sheet, with vertically specimens having strength typically 2 to 2.5
strength was 1600 kPa in the temperature range of − 5 to −7 °C and
higher than the horizontal specimens. For horizontal specimens,
2300 kPa in the range of −10 to −13 °C. Pounder and Little (1959)
both in frazil ice and in columnar ice that has a random c-axis
carried out single direct shear tests on sea ice of various temperatures
orientation in the horizontal plane, the orientation of the sample in
and grain structures. They obtained strengths in the range of 20 to
the horizontal direction has no effect on the observed strength.
1000 kPa for a wide range of conditions. The values ranged from 650
However, when a strong c-axis alignment exists, there are
to 850 kPa for summer ice in the Arctic. They did not provide
appreciable differences between horizontal beams cut in the
consistent information with regard to sample temperatures or
hard-fail direction [the long axis of the beam cut perpendicular
salinities. Paige and Lee (1967) and Dykins (1971) conducted single
to the c-axis] and beams cut in the easy-fail direction [the long axis of
shear tests using a standard 76 mm diameter core as a specimen. The
the beam cut parallel to the c-axis]. Full thickness tests made during
specimen, of length equal to the diameter, was loaded over a
cold periods on freshwater ice that places the cold tops of the beams
semicircular area at each end. The loading was oriented so that a
in tension and the warm bottoms in compression give higher values
failure plane was generated along the length of the specimen. A relief
than comparable tests that place the tops in compression and the
hole was provided in the central section to reduce confinement
bottoms in tension (Gow et al., 1987c). Tests by Kayo et al. (1983)
effects. Paige and Lee (1967) carried out tests on specimens made
show that there is little evidence for a difference in strength between
from cores of natural sea ice. Because the long axes of these cores were
these loading directions for sea ice.
in the growth direction, the shear failure plane of the specimen was
In summary, the flexural strength of sea ice ranges from about 1 MPa parallel to the long axes of the columnar grains. The strengths were in
and deceases with brine volume using a functional relationship outlined the range of 500 to 1200 kPa with a significant dependence on brine
in Eq. (11). For warm sea ice, typical values are on the order of 100 to volume. Dykins (1971) used laboratory-grown columnar-grained
150 kPa. The strength is not strongly affected by the loading rate. saline ice and obtained strengths in the range of 100 to 250 kPa.
All of the measurements reported above have a problem of uncertain
7.2. Old Ice normal stresses on the failure plane. As a result, Frederking and Timco
(1984a, 1986) utilized an asymmetric 4-point bending system to
The authors could find no reported measurements of the flexural produce shear failures in sea ice specimens. For granular sea ice, they
strength of Old Ice. However it is possible to infer some representative reported an average shear strength of 550 kPa ± 120 kPa for an average
G.W. Timco, W.F. Weeks / Cold Regions Science and Technology 60 (2010) 107–129 117

test temperature of −13 °C± 2 °C and salinity of 4.2‰ ± 0.5. These


conditions correspond to a brine volume of 18‰. Frederking and Timco
(1986) measured the shear strength of columnar sea ice and reported
values in the range between 550 kPa and 900 kPa. They found that when
temperatures were higher, shear strengths were lower with total
porosity providing a better correlation to strength changes than brine
volume alone. In general, horizontal shear strength values were higher
than values obtained when the shear failure was oriented vertically (i.e.
when the failure plane was oriented parallel to the crystal elongation).
Frederking and Timco (1986) combined their results with those of Paige
and Lee (1967) and found that for the vertical shear strength (σs) of
columnar ice could be represented by
 rffiffiffiffiffiffiffiffiffi
vT
σs = 1500 1− ð12Þ
390

where vT is the total porosity in parts per thousand and the shear
strength is given in kPa.
Overall, there is considerable scatter in the measured shear strengths
of sea ice. Moreover, many of the test results were generated using test
techniques which impose unrealistic (and unknown) normal stresses
on the failure plane. Shear strength values from the more reliable tests
ranged from 400 to 700 kPa for granular ice and 550 kPa to 900 kPa for
columnar sea ice. Much more research is required to better understand
the shear strength of sea ice. It is certainly one of the least understood Fig. 9. Photograph showing a typical test arrangement for a uni-axial compression test
on sea ice.
properties of sea ice.
It should be noted that in engineering practice, the shear strength
is not usually explicitly used. Since ice tends to fracture rather than to
flow in a crack-free, volume-conserving manner, the shear strength is Urabe and Inoue, 1986; Wang, 1979; Wang and Poplin, 1986; Moslet,
actually governed by the tensile strength of the ice. Since most ice 2007) have measured the compressive strength of small samples of
engineering issues occur at higher loading rates, the compressive sea ice. It has been found that a number of factors influence the
strength is much higher than the tensile strength. Thus, ice loaded measured strength value. These factors can either be intrinsic
with a shear condition would fail in tension rather than in shear. (temperature, salinity, density, ice type, crystal size and orientation)
or test condition (rate of loading, confinement conditions, loading
8.2. Old Ice direction, sample size, stiffness of the test machine, and sample
preparation techniques).
There are no reported measurements of the shear strength of Old Timco and Frederking (1990, 1991) developed a model to calculate
sea ice. Limited test results on freshwater ice by Frederking et al. the strength of sea ice sheets. They compiled 283 compressive
(1988) using the asymmetric 4-point approach gave average values in strength test results on first-year sea ice. The test information
the range of 1100 kPa for columnar freshwater ice at − 10 °C. Direct included the temperature, ice salinity, bulk ice density, grain
measurements of the shear strength of Old Ice would be useful. structure, loading direction, number of tests, test results and the
investigators. They derived the equations for the uni-axial compres-
9. Compressive strength sive strength of first-year sea ice for several different grain structures.
For horizontally-loaded columnar ice, the uni-axial strength is
The compressive strength is another of the fundamental properties rffiffiffiffiffiffiffiffiffi
of sea ice. Observations of both large and small-scale sea ice failures
show that ice often fails in compression. These can occur during the
σc = 37ð ε̇Þ
0:22
½ 1−
vT
270  : ð13Þ
formation of large compression pressure ridges (which also include
out-of-plane failures), or by crushing against an offshore structure. For vertically-loaded columnar ice, the uni-axial strength is
Milling of ice pieces by a ship's propellers are another type of
rffiffiffiffiffiffiffiffiffi
compressive failure of ice. Because of its importance, this property has
been extensively studied for sea ice.
σc = 160ð ε̇Þ
0:22
½
1−
vT
200
; ð14Þ
Fig. 9 shows a photograph of a typical test arrangement. Cores of
sea ice are cut into cylinders which are typically 20 to 25 cm high. The and for granular ice, the uni-axial strength is
samples are placed in a test apparatus that can generate high rffiffiffiffiffiffiffiffiffi
(typically 50 kN) loads. A load cell is included in-line to measure
the applied load. Often, compliant platens are used to minimize edge
σc = 49ð ε̇Þ
0:22
½ 1−
vT
280 : ð15Þ
effects. Extensometers are often attached to the sample to measure
the deformation of the ice during the test. From this information and where vT is in parts per thousand (ppt). Eq. (13) to Eq. (15) relate the
the measured applied load, the strain rate and Effective Modulus of uni-axial compressive strength of sea ice explicitly to grain type,
the sample can be obtained. loading direction, loading strain rate and total porosity, and implicitly
to ice salinity, temperature and density. They give a good represen-
9.1. First-year sea ice tation of the compressive strength of sea ice sheets. The applicable
range of strain rates for these equations is from 10− 7 s− 1 to about
Many investigators (Frederking and Timco, 1980, 1983, 1984b; 2 × 10− 4 s− 1. At higher strain rates, premature (brittle) failure of the
Sinha, 1981, 1983a,b, 1984; Timco and Frederking, 1983a, 1984, 1986; ice can occur in some, but not all instances. This strain rate begins the
118 G.W. Timco, W.F. Weeks / Cold Regions Science and Technology 60 (2010) 107–129

onset of the so-called “ductile-to-brittle” transition. In this higher


strain rate range, compressive strength tests performed in the field
will mostly fail in a brittle manner with the ice sample catastroph-
ically breaking apart into a number of separate pieces. This could be a
real effect or an artefact of the restricted experimental techniques for
field measurements. High quality compressive strength tests where
the ice is loaded at high rates are virtually impossible to perform in the
field. Jones (1997) for example reports from high quality laboratory
tests that the strength of sea ice does not decrease for higher strain
rates. His data suggest that the uni-axial strength continues to
increase to strain rates up to 101 s− 1. At this loading rate, measured
strength values were on the order of 8 to 12 MPa.
Recently Moslet (2007) used a specially designed field-portable
testing system to study Svalbard fast ice. There were two particular
foci in this multi-year study (∼ 540 samples): to minimize the time
between the removal of the sample from the ice sheet and testing and Fig. 10. Plot of the compressive strength of the ice as a function of temperature. The
to provide additional information on the strength of warm, decaying corresponding flexural strength of the ice is indicated by the dashed line.
sea ice (0 to − 11 °C). Although the observed trends were generally
similar to the results of earlier work, there were differences in that
Moslet reported higher strength values at higher porosities than
estimated by the Timco and Frederking (1990, 1991) equations. He
also suggested that strength and behaviour may not be unique where compressive strengths (σc) ranged from 7 to 15 MPa. Measure-
functions of state and as a result cannot be predicted by ice properties ments showed a stress rate (σ̇) dependence of the form:
alone. If this proves to be the case, the environmental history of each
0:23
sample may also be of importance. σc = 29:5 σ̇ c ð17Þ
The large number of measurements of the compressive strength of
sea ice provides some design data pertinent to the problem of ice where the stress rate is in MPa s− 1. These strengths are comparable to
forces on offshore structures. In such calculations, the desired good quality freshwater ice and are higher than compressive
information is typically the compressive strength of thick first-year strengths measured on first-year sea ice that, for the same range of
ice. As tests are both difficult and expensive, only a very limited stress rates, varied from 0.5 to 4 MPa (Timco and Frederking, 1986).
number of full thickness strength tests have been completed (Lee The uni-axial compressive strength of multi-year has been
et al., 1986; Chen and Lee, 1986). However, based on these tests, measured by Timco and Frederking (1982b), Sinha (1984), and Cox
mathematical expressions have been developed that give estimates of et al. (1984a,b, 1985a,b). These measurements showed that although
full thickness compressive strengths based on small-scale tests. Timco the strength of multi-year ice is similar to the strength of first-year ice
and Frederking (1990, 1991) developed a model to predict the full when the ice is very cold (i.e. −20 °C), multi-year ice can be
thickness strength by dividing the ice sheet into nine separate layers considerably stronger than first-year ice when the ice is warmer. This
and then calculating the compressive strength of each layer based on is a reflection of the fact that, as the temperature of multi-year ice
the temperature, salinity, density and grain structure. Fig. 10 shows a increases, there are only small differences in its brine volume. As
plot of the compressive strength of a 1 m thick sheet of sea ice as a stated previously, multi-year ice can contain of a great mix of grain
function of air temperature. In this case, it was assumed that the types within a sample. Weeks (1985) showed that when the variation
surface temperature of the ice was the same as the air temperature, in ice strength within and between different ridges was examined, it
the ice salinity was 5‰ and the density was 0.9 Mg m3. The figure was found that in all cases the main factor contributing to the
shows that the compressive strength is a strong function of the observed variance was the differences within cores as opposed to
loading strain rate (ε̇) and less dependent upon the temperature until between ridges. Richter-Menge and Cox (1985) showed that when
close to the melting point. Compressive strength values ranged from the effect of internal sea ice structure was examined, highest
0.4 MPa to about 5 MPa. strengths were observed when the loading occurred parallel to the
Kovacs (1996b) used a different approach and suggested the crystal elongation (i.e. perpendicular to the crystal c-axis direction).
strength was represented by When the elongation direction of the crystals was oriented parallel to
the plane of maximum shear (30° to 60° to the loading direction),
ð1 = nÞ m
σc = B2 ε̇ ϕB ð16Þ there was an appreciable decrease in strength. A similar decrease in
strength occurred when the samples contained a mix of granular and
where the values of the parameter B2 is 2.7 × 103, n = 3 and m =−1, and columnar ice.
ε̇ and ϕB are the strain rate (range from 10− 6 s− 1 to about 10− 3 s− 1)
and bulk porosity (in ‰). The compressive strength σc is given in MPa. 10. Multi-axial loading
Although using this approach clearly could be convenient, it would be
prudent to have local checks in the setting under study. Several researchers have measured the strength of ice under
In summary, the compressive strength is a strong function of the complex stress and confinement conditions in both the laboratory and
loading rate up to a rate of approximately 10− 4 s− 1. Typical values field. The purpose of these tests is to determine the complete failure
range from 0.5 MPa to over 5 MPa. Eq. (13) to Eq. (15) can be used to envelope for ice. Here the term failure envelope refers to a relation
estimate the value if the grain structure, salinity and temperature of the that describes when the ice yields for any combination of compressive
ice are known. or tensile stress states. True tri-axial tests (σ1 > σ2 > σ3) are rare and
require specialized equipment (Hausler, 1981; Gratz and Schulson,
9.2. Old Ice 1994). Even biaxial tests (σ1 > σ2 = σ3), frequently referred to as
conventional tri-axial tests, can require involved procedures (Cox and
The uni-axial compressive strength of undeformed second-year ice Richter-Menge, 1985). Several authors have used these experimental
at Mould Bay in the Canadian Arctic was reported in Sinha (1985), results to try to define the failure envelope for a range of ice types.
G.W. Timco, W.F. Weeks / Cold Regions Science and Technology 60 (2010) 107–129 119

10.1. First-year sea ice

Multi-axial tests have been performed on sea ice in the field (Timco
and Frederking, 1983a, 1986) and on saline ice in the laboratory by
several authors (Hausler, 1981; Cox and Richter-Menge, 1985; Gratz
and Schulson, 1994, 1997; Smith and Schulson, 1994; Schulson and
Nickolayev, 1995).
The field tests by Timco and Frederking (1983a, 1986) used a
technique developed by Frederking (1977) on freshwater ice as shown
in Fig. 11. With this approach, the ice is loaded in one direction while
confined in a second direction, thereby allowing deformation in only
one direction. Fig. 12 shows a photograph of the test arrangement. By
measuring both the applied and the confining loads as well as knowing
the orientation of the confined ice, it is then possible to determine the
failure stresses for a variety of stress states.
Frederking (1977) found that when columnar freshwater ice is
constrained from moving in the plane of the ice sheet, as shown in the
Type A loading condition, the strength was increased by a factor of four
at a strain rate of 10− 7 s− 1 (to about 2.4 MPa) and by a factor of two at
10− 4 s− 1 (to about 20 MPa). When the crystals were confined in a
direction normal to the axis of elongation as shown in the Type B loading
condition, little or no change was observed. When similar tests were
carried out on granular sea ice (Timco and Frederking, 1983a), the
confined compressive results were about 20% higher than for the
unconfined specimens. In contrast, when columnar specimens were
tested, there was up to a four-fold increase in the uni-axial strength
when the loading conditions were of Type A. No significant change was
Fig. 12. Photograph showing the test arrangement for a field confined compression test.
observed when the conditions were of Type B. Timco and Frederking
(1986) later expanded their series of tests obtaining results for all five
test configurations shown in Fig. 11. The results of their tests for a failure envelope for both columnar and granular sea ice and freshwater
temperature of −10 °C and a strain rate of 2 × 10− 4 s− 1 are summa- ice (Timco and Frederking, 1984). The comparison shows that the failure
rized in Fig. 13. The failure envelope is shown to be highly anisotropic for envelope is significantly larger for freshwater ice. It should be noted that
columnar sea ice and relatively isotropic for granular and discontinuous the relative difference between the two ice types will change as a
columnar sea ice. function of temperature in that with decreasing temperature, as there is
Timco and Frederking (1984) compared the failure envelope of sea less brine in the sea ice, its strength will approach that of freshwater ice.
ice and freshwater ice. The comparison was made for ice loaded in the Laboratory tests on saline ice have further elucidated the failure
horizontal direction of the original ice sheet, at a temperature of −12 °C envelope and behaviour of this ice. Both the failure stress and the failure
and strain rate of 2 × 10− 4 s− 1. This strain rate and temperature mode are sensitive to confinement with the failure mode changing from
produces ductile failure of the ice. Fig. 14 shows a comparison of the axial splitting to shear faulting in the loading plane (Smith and Schulson,

Fig. 11. Confinement conditions for testing the failure envelope of ice (after Frederking, 1977).
120 G.W. Timco, W.F. Weeks / Cold Regions Science and Technology 60 (2010) 107–129

Fig. 13. Failure envelope of sea ice for granular and discontinuous columnar sea ice, and also columnar sea ice (after Timco and Frederking, 1986).

1994). The transition from brittle to ductile has been found to occur Under moderate confinement, failure occurred by Coulombic shear
when the strain rate falls below a critical level with the transition strain faulting where the fault plane is typically inclined by 28–30° to the
rate first increasing and then decreasing with increasing across column direction of shortening (see Schulson, 2002). Finally, when the ice was
confining stress (Schulson and Nickolayev, 1995). Further work on this loaded under high confinement (R > 0.2), failure occurred in a raft-like
subject by Schulson et al. (2006) has combined measurements of the manner accompanied by splitting across the columns. Comparable
brittle compressive strength of first-year S2 sea ice (S2 = c-axes failure modes have been observed in earlier studies of saline ice
randomly oriented within the horizontal plane) with the earlier (Schulson and Nickolayev, 1995). In contrast to Fig. 14 which shows
measurements of the tensile strength of similar sea ice by Richter- the failure envelopes in the ductile range, Fig. 15 shows that the
Menge and Jones (1993) thus allowing the complete brittle failure failure envelopes appear to be similar in both freshwater ice and sea
envelope to be constructed. The envelope was found to be symmetric ice in the brittle range. This suggests that the processes controlling
about the loading path R = (σ2 /σ1) = 1 as the result of the isotropic fracture are also similar. Derradji-Aouat (2003) has developed multi-
character of the material within the horizontal plane. However, the surface failure criterion for saline ice in the brittle range based on a
envelope is asymmetric with respect to the compressive/compressive large number of published experimental data.
and the tensile/tensile quadrants as the result of the relative weakness
of sea ice under tension. 10.2. Old Ice
Fig. 15 shows a comparison of the brittle failure envelopes at
−10 °C for first-year sea ice with that of freshwater ice having a There have only been a few studies of the multi-axial strength of
similar structure (from Iliescu and Schulson, 2004). Three different Old Ice.
failure modes were observed. When the ice was unconfined (R = 0), Sinha (1985) measured the confined compressive strength of
axial splitting occurred along the loading direction with the splits second-year ice at Mould Bay in the Canadian Arctic. He used the Type
running parallel to the long axis of the columnar grains. There A conditions (Fig. 11) and found that the confinement increased the
appeared to be no preference for fracture along the grain boundaries. strength considerably but decreased the rate sensitivity. He found that
the strength of this second-year ice was comparable to the confined

Fig. 14. Comparison of the failure envelope for first-year sea ice and freshwater ice for Fig. 15. Failure envelope of the brittle strength of saline ice (after Iliescu and Schulson,
both granular and columnar sea ice (after Timco and Frederking, 1984). 2004).
G.W. Timco, W.F. Weeks / Cold Regions Science and Technology 60 (2010) 107–129 121

strength of laboratory-grown columnar freshwater ice tested in a 11.1. First-year sea ice
laboratory environment. Loading rates were set below 10− 6 s− 1 so
the ice failed in the ductile range. Nevertheless, the high strength There have been a number of measurements of the borehole
values for second-year ice agree with the latter observations of Iliescu strength of first-year sea ice (Blanchet et al., 1997; Masterson, 1996;
and Schulson (2004) showing the similarity in failure envelope of sea Masterson et al., 1997; Johnston et al., 2000, 2001, 2002; 2003a,b). In
ice with that of freshwater ice. general, the borehole strength of mid-winter, cold first-year ice is on
Sammonds et al. (1998) included a number of tri-axial compres- the order of 25 MPa to 30 MPa (Blanchet et al., 1997). In a series of
sion tests on multi-year sea ice. Four main types of behaviour were papers, Johnston et al. show that the strength of first-year sea ice
observed depending upon the confinement conditions: decreases rapidly during the spring season in the Arctic. This is
discussed in more detail below.
• Under uni-axial compression at high strain rates (10− 3 to 10− 2 s− 1) Timco and Johnston (2002) have analyzed the borehole jack data for
brittle fracture was observed and was characterized by multiple first-year ice and compared them to calculated values of the flexural
axial splitting. strength. They then analyzed these strengths as they changed during the
• If even a small confining pressure was present, splitting was winter and spring seasons in the Canadian Arctic. Fig. 17 shows the
inhibited with fracture occurring via the formation of a narrow calculated flexural strengths and the measured borehole strengths, both
shear fault inclined at 45 ± 3°. This failure process is different than normalized to their mid-winter values (flexure = 0.71 MPa, borehole=
the Coulombic shear faulting discussed above (see Schulson, 2002). 26 MPa). The trends are in excellent agreement. Both strengths show
• As the confining pressure increased, plastic deformation was that in mid-May, the ice had about 60–70% of its mid-winter strength. By
accompanied by substantial cracking activity. early June, the ice had about 50% of its mid-winter strength. After the
• At the highest confining pressures, cracking was completely first week of June, only the borehole jack measurements provided
inhibited and the deformation becomes completely plastic. information about the degradation in ice strength. Measurements
showed that by the end of June, the ice had only 15 to 20% of its mid-
winter strength. The ice strength was stable during the month of July,
11. Borehole ice strength measurements when only 10% of the mid-winter ice strength remained.

The borehole jack test has been used extensively to characterize 11.2. Old Ice
the in situ strength of ice. For these studies, a typically 15 to 20 cm
wide core is taken from the ice sheet and the borehole jack is lowered Johnston et al. (2003c) discuss results from all publicly available
down into the resulting hole in the ice cover. The jack consists of a and privately funded studies into the borehole strength of Old Ice
high-strength stainless-steel hydraulic cylinder with a single or (Fenco, 1977; Geotech, 1983, 1984; Sinha, 1986b, 1991; Johnston
double-acting piston that pushes on the sidewalls of an augured et al., 2003b). Fig. 18 shows the borehole strengths from all the field
hole in the ice (Sinha, 1986b; Kivisild, 1992; Masterson and Graham, studies plotted as a function of temperature. The figure shows that the
1992; Masterson, 1996, Masterson, in press). Pressure is applied to the strength of multi-year ice is relatively independent of temperature
loading plate(s) hydraulically by means of the piston located inside from approximately − 20 °C to −5 °C. Within that temperature range,
the jack body and activated by a pump at the surface of the ice. Fig. 16 most borehole strengths are between 20 and 35 MPa. In comparison,
shows a photograph of a double-acting borehole jack placed inside a once the temperature of the multi-year ice reaches about −5 °C, data
core hole in sea ice. Oil pressure and plate displacement are measured show that the ice strength may decrease. Measurements showed
and used to determine the failure stress. This type of in situ testing is some temperate multi-year ice (above a temperature of −5 °C) had
widely used for soils and rock strength determination. However, in very little strength (3.6 MPa) while other warm multi-year ice still
contrast to the uni-axial strength tests, the stress conditions and had appreciable strength (up to 26 MPa). The strength data show that
stress-state for this test are complex (see Masterson, 1992). Thus, this in the summer, multi-year ice could be weak or very strong.
is a form of a multi-loading of the ice. However the unknown factors of Fig. 19 shows a comparison of the measured borehole strength of
ice grain structure, loading contact area, etc. make interpretation of first-year, second-year and multi-year ice (after Johnston and Timco,
the results problematic. Nevertheless, the borehole jack test has been 2008) as a function of date in the Canadian Arctic. The figure shows
used extensively to study the in situ strength of freshwater ice and sea the depth-averaged strength (i.e. strength averaged over all tests in
ice. Its main advantages are that the ice being tested is not removed
from the ice sheet, it is comparatively easy to obtain strength profiles
under in situ conditions even on thick ice, and multiple tests can be
performed quickly.

Fig. 17. Comparison of the normalized measured borehole strength and calculated
Fig. 16. Photograph showing a borehole jack in a core hole in first-year sea ice (photo flexural strength for first-year sea ice in the Canadian Arctic (after Timco and Johnston,
courtesy of R. Lanthier). 2002).
122 G.W. Timco, W.F. Weeks / Cold Regions Science and Technology 60 (2010) 107–129

the same borehole) as a function of Julian Day for first-year ice,


second-year ice and multi-year ice. There are some clear trends. First,
the strength of all three ice types decrease in the spring months but at
different rates. Second, the decrease in strength of first-year ice is
more rapid than that of Old Ice. Third, although not clearly shown
here, the mid-winter strength of all three ice types are approximately
the same with a typical strength of 30 MPa. The same mid-winter
strength for all three ice types is a reflection of the fact that most of the
salt in first-year sea ice is in its solid state when the ice is cold. As such,
there is little brine porosity in the ice, regardless of its overall salinity
in the different ice types, and their strengths are comparable. The fact
that the strength of first-year, second-year and multi-year ice showed
differing degrees of decay (in terms of strength) as the summer
progressed is a direct result of the different salinities (hence brine
volumes) of the three ice types, and the possibility that different decay
processes occur in the different ice types.
This latter finding is somewhat surprising when viewed in terms of Fig. 19. Comparison of first-year, second-year and multi-year borehole strength during
the failure envelope generated by controlled field and laboratory tests. the decay season (after Johnston et al., 2003b).

As discussed above, there was little difference observed between the


failure envelope for sea ice and freshwater ice in the brittle range large tanker “MV Arctic” (B. Gorman, personal communication). It is
(Iliescu and Schulson, 2004). Similarly Sinha (1985) showed that the not clear, however, whether this was due to lower ice strength or a
strength of second-year ice and high quality freshwater columnar ice relatively thin multi-year ice floe.
were the same when loading and confinement were in the plane of • There is no visual method for judging if multi-year ice has decayed.
the ice cover. Thus, there is an apparent disconnection between the • The large range of strength values for multi-year ice at elevated tem-
borehole jack tests and the controlled stress results with regard to the peratures close to the melting point indicates that temperature cannot
decay process in ice. It should be noted however that Fig. 18 clearly be used as a method for defining low strength for multi-year ice.
shows little temperature dependence at temperatures colder than
about −5 °C. Warmer than this temperature, there is a rapid decrease 12. Sea ice creep
in strength in roughly half of the samples. This suggests that
additional tests on the failure envelope of Old Ice in this higher The long-term deformation or creep of sea ice sheets is important
temperature range would be beneficial. in several engineering applications. For example, the load exerted on
Little is known about the actual decay process in multi-year ice. As an offshore structure by an ice sheet which is undergoing thermal
discussed above, there have been only a limited number of expansion is controlled by the creep properties of the ice. This relates
measurements on multi-year ice throughout the winter and spring to the in-plane properties of the ice. Usually ice failures here appear to
seasons. The information available is not conclusive for making any be ductile in nature. On the other hand, the load bearing capacity of an
concrete statements about the decay process. There are a few things to ice sheet is controlled by the properties related to the flexural creep.
note in this regard for multi-year ice: Applications include landing aircraft on ice covers, floating ice drill
pads, ice roads, and storage of materials on ice covers. In these cases, it
• The limited number of borehole jack tests at higher temperatures is important that the ice remain in the non-brittle state. Brittle failure
does not show strong temperature dependence, indicating that is often catastrophic in these situations.
multi-year ice does not decay in the same manner as first-year ice. Creep is usually measured by placing a material under constant
This is understandable based on the phase relationships of the salts stress or constant load and monitoring its deformation as a function of
and the low salinity of multi-year ice. time. In spite of its importance, there have been relatively few direct
• Multi-year ice can decay. Fig. 18 shows a decrease in strength very creep measurements on sea ice. There are several reasons for this. For
close to the melting point of the ice. Further, large multi-year ice one, it is very difficult to perform long-term tests on natural sea ice
floes have sometimes broken apart very easily when hit with the without the occurrence of significant brine drainage during the testing
period. At best this would mean that although the total void volume
might remain essentially constant during a test, the ratio of the
volume of air to the volume of brine in the sample would be
continuously increasing. Another aspect here relates to the numerous
factors that control creep and the significant time that it would take to
try to elucidate each factor.
Sinha (1978, 1979, 1989b) discussed the rheology of freshwater
ice and he describes four deformation mechanisms which are
important for polycrystalline ice:
• elastic deformation from the atomic bonds changing length;
• delayed elasticity from sliding at the grain boundaries;
• viscous deformation from the movement of dislocations within
grains; and
• deformation from the microcracks in the ice.
The total strain in the ice is usually considered as the resultant sum
of these separate strain components; i.e.

Fig. 18. Borehole jack strength as a function of ice temperature, Old Ice (after Johnston T e d v c
et al., 2003c). εij = εij + εij + εij + εij ð18Þ
G.W. Timco, W.F. Weeks / Cold Regions Science and Technology 60 (2010) 107–129 123

where the superscripts refer to the total strain (T), the elastic strain microstructural changes once it deforms and a number of interde-
(e), the delayed elastic strain (d), the viscous strain (v) and the strain pendent micromechanisms operate simultaneously. These processes
due to cracking (c). Because ice is an anisotropic material, the strain and their interdependence make understanding creep behaviour of
tensor (i j) is used. The first term, which represents the elastic sea ice difficult.
behaviour, and the second term, which represents the delayed elastic It should be mentioned that measurements of creep on multi-year ice
response, are of particular importance for situations where relatively where made during the on-ice drilling by Panarctic Oils Ltd. in the
small strains are important. The concept of delayed elasticity is Canadian archipelago over a period of 12 years. These data are very
interesting since this strain is not permanent. Although it is fully valuable but, as previously mentioned, they are not in the public domain.
recoverable after removal of the load, its recovery is not instanta- Also, Schulson and Duval have recently published a book that deals with
neous. This delayed elastic creep is sometimes referred to as primary the science of sea ice creep and fracture (Schulson and Duval, 2009).
creep or recoverable creep. The third term in Eq. (18) represents the
viscous creep (or secondary creep) of the ice. It is permanent, non- 13. Elastic and strain modulus
recoverable deformation. It becomes apparent when the delayed
elastic strain rate approaches zero. When this occurs, the viscous The ratio of the stress (σ) to the strain (ε) during elastic behaviour
strain rate begins to dominate with a rate dependence of the form ε̇v is known as the Elastic Modulus (E) of the material (i.e. σ = E · ε). For
(t) = Bσn where B is parameter dependent primarily on temperature elastic behaviour this is often termed “Young's Modulus”. There is a
and ice type. The value of the exponent n is normally taken to be 3 great deal of confusion about this parameter with sea ice (and also
(Glen's law). After a long time under compressive loading, the strain freshwater ice). The purely elastic modulus for sea ice relates to the
rate increases and tertiary creep begins. When that happens, elastic displacements within the ice lattice (i.e. the first term in
microcracks begin to form at grain boundaries. These microcracks, Eq. (18)). This property can be determined by either measuring the
in sufficient numbers, begin to coalesce and cause accelerating propagation of elastic waves in the ice sheet, or by measuring the
deformation rates. This is the fourth term in this equation. ultrasonic velocity in a small ice sample. Any mechanical measure-
Sinha (1978, 1979, 1989b) has presented equations for freshwater ment of the Modulus (either by small beam tests or in situ cantilever
ice and the interested reader is referred to these papers for additional beam) is not truly elastic since the second term of Eq. (18) always
details. Since these equations were developed for freshwater ice, it is comes into play. In this latter case, the term Elastic (or Young's)
certain that other factors such as salinity, brine volume, sea ice flaws, Modulus is not correct. Several authors correctly note this and they
etc, would play a key role in the rheology of sea ice. Nevertheless, the express measurements of the stress/strain ratio as either the Effective
key points illustrated above are important for sea ice engineering. The Modulus or the Strain Modulus. As might be expected, the Elastic
lack of knowledge of the rheological behaviour of sea ice restricts Modulus is always higher than the Strain Modulus.
applying basic material properties to key engineering applications.
As discussed above, there are two main areas of interest here. The 13.1. First-year sea ice
first relates to ice loads on offshore structures due to long-term creep
loading. Sanderson (1988) has developed two simple theories for Weeks and Assur (1967, 1968) have reviewed the in situ seismic
predicting the ice creep buckling force. One approach estimates the determinations of the Elastic Modulus. Measured values vary from 1.7
force from the buckling wavelength where as the other approach uses to 5.7 GPa when measured by flexural waves and from 1.7 to 9.1 GPa
the time for the buckling event. Unfortunately it is not possible at when determined by body-wave velocities. This difference is
present to use the basic mechanical properties of sea ice to estimate reasonable in that the flexural wave velocity is controlled by the
loads. Instead, empirical data on loads exerted by creep failure against overall properties of an ice sheet while the body-wave velocity is
a structure must be used (Timco and Johnston, 2004). controlled by the high velocity channel in the commonly colder, less
The other engineering application where sea ice rheology is saline and stronger upper section of the ice. There is a pronounced
important relates to the bearing capacity of ice covers. Once again, increase in the value of E with decreasing temperature (see
bearing capacity determinations cannot be done from first principles Peschanskii, 1960) and decreasing brine volume (Anderson, 1958).
based on sea ice mechanical properties. Instead, empirical data is As a result, it would be expected that there is a significant seasonal
used. Masterson (2009) has written an excellent overview on bearing change in elastic parameters such as the longitudinal plate wave
capacity of ice covers. He notes that there are two basic design steps velocity. This has been clearly been shown by the data collected by
that are completed for a structural analysis of a loaded ice sheet. These Hunkins (1960) who documented systematic velocity variations
are: during the year on Arctic Ocean pack ice ranging from a low of
∼2.3 km s− 1 in August to a high of 3.2 km s− 1 in late February.
• ensure that the ice sheet maximum tensile extreme fibre stress is
Most dynamic measurements of the Elastic Modulus are deter-
less than the allowable flexural stress for ice; and
mined from small samples that have been removed from the ice sheet.
• ensure that the ice sheet's short-term and long-term deflections are
Results of a typical series of tests performed by Langleben and
less than the available freeboard.
Pounder (1963) are shown in Fig. 20. Note that the E values at very
Models for these predictions assume that the ice structure acts as low brine volumes are characteristically in the range of 9 to 10 GPa.
an elastic, homogeneous, isotropic plate on an elastic foundation. These values are similar to those for freshwater ice at high loading
Clearly that is not the case. Again, empirical data (e.g. Gold, 1971) are rates (Gold, 1977). Note also that E appears to decrease as a linear
used to provide full-scale experience on bearing capacity of the ice function of brine volume (vb) with a relationship:
cover.
In that the rheological properties of sea ice are very complex and E = 10−0:0351vb GPa ð19Þ
poorly understood, additional research in this area would greatly
increase our knowledge. However, the overall difficulties of The results shown in the figure are averages based on over 300
performing accurate creep tests with a sample of sea ice that drains measurements. Slesarenko and Frolov (1974) measured the Elastic
during the test would cast uncertainty on the results. Moreover, Sinha Modulus of saline ice and show results with similar trends to that
(1990) has argued that the complete history and state of the ice must shown in Fig. 20 but about 4 to 5% lower. Interested readers can
be taken into account when dealing with the rheological behaviour. consult the paper by Gammon et al. (1983) for a discussion on the
That is, the complete form of Eq. (18) must be used in any creep measurement and results of the elastic constants for sea ice and other
application. This reflects the fact that ice experiences irreversible ice types by Brillouin spectroscopy.
124 G.W. Timco, W.F. Weeks / Cold Regions Science and Technology 60 (2010) 107–129

14.1. First-year sea ice

High frequency measurements are generally done by using either


seismic methods on ice covers, or ultrasonic measurements of ice cores.
Langleben and Pounder (1963) measured the pulse velocity and
resonance of sea ice cores and using elasticity theory, determined values
for Poisson's ratio with a mean value of 0.295 ± 0.009. Weeks and Assur
(1967, 1968) analyzed Lin'kov's (1958) seismic results and proposed the
following equation for the dynamic Poisson's ratio (μD) for sea ice:

μD = 0:333 + 0:06105 expðTi = 5:48Þ ð20Þ

where Ti is the ice temperature in °C. Sinha (1987) has analyzed


published measurements of Poisson's ratio and concluded that
measurement frequencies on the order of a few MHz are required to
obtain the elastic component if the grain size of the sea ice is 0.5 mm.
Lower frequencies (few kHz) can be used if the ice is coarse-grained.
There have also been a few measurements of the “Effective” Poisson's
Fig. 20. Elastic Modulus as a function of brine volume for first-year sea ice (after
Langleben and Pounder, 1963). ratio for sea ice. Wang (1981) performed measurements of the Effective
Poisson's ratio on columnar sea ice and found that lateral deformations
were anisotropic. This means that the ice deformed much more in the
Static measurements of E are more variable and difficult to interpret
direction perpendicular to the crystal columns (horizontal direction)
than are dynamic measurements because, as discussed above, sea ice
than in the direction parallel to the crystal columns (vertical direction).
invariably behaves viscoelastically when subjected to appreciable
He found that for the majority of tests, Poisson's ratio in the horizontal
stresses for finite periods of time. Nevertheless, it is these Effective
direction ranged from 0.8 to 1.2, while in the vertical direction it ranged
Modulus values that are applicable to most engineering problems such
from 0 to 0.2. Saeki et al. (1981) used an instrumented ring on samples
as bearing capacity or ice forces on structures. There have been many
of columnar sea ice tested to measure the uni-axial compressive
reported measurements of the Effective Modulus of sea ice (Dykins,
strength. They reported that the value was a strong function of
1971; Tabata, 1960, 1966, 1967; Tabata and Fujino, 1964; Tabata et al.,
temperature and loading rate. Lower values were reported with higher
1967; Lainey and Tinawi, 1981; Sinha, 1989c; Moslet, 2007). In general,
temperatures. In the temperature range of −6 to −8 °C, the value
the Effective Modulus shows a considerable amount of scatter but there
changed from 0.1 at a stress rate of 0.01 MPas− 1 to a maximum value of
is a tendency for decreasing values as the brine volume increases.
0.48 at a stress rate of approximately 5 MPa s− 1. For higher rates, the
Typical measured values of the Effective Modulus range from 1 to 5 GPa.
value decreased significantly. At higher temperatures (− 2.5 °C),
These values increase with increasing loading rate and begin to
Poisson's ratio varied from 0.05 to a maximum of 0.18 over this same
approach those of the Elastic Modulus at very high loading rates.
stress rate range. Murat and Lainey (1982) measured the longitudinal
and transverse strains on simply supported beams loaded in flexure. The
13.2. Old Ice tests were performed at different temperatures and loading rates on
columnar ice of the S2 variety that had an average salinity of 5‰. It was
Langleben and Pounder (1963) completed 36 measurements on found that Poisson's ratio decreased with increasing stress rate and
Old Ice at Isachsen. In general, the results were about 5% lower than decreasing temperature. They report on values of Poisson's ratio of 0.48
those measured on first-year ice. The summary data points are also at stress rates of 0.01 MPa s− 1, decreasing to values between 0.35 and
shown in Fig. 20. 0.4 for a stress rate of 0.6 MPa s− 1. They proposed the following
Cox et al. (1984a,b) provide Effective Modulus values obtained by equation for the Effective Poisson's ratio (μE):
using samples from multi-year pressure ridges. Their results are
comparable in magnitude to the values reported by Tratteberg et al. −0:29
μE = 0:24ð σ̇ = σ̇ 1 Þ + μD ð21Þ
(1975) for freshwater ice. Again the Effective Modulus was observed
to increase with increasing strain rate and decreasing temperature.
The considerable scatter in the data was attributed to the large where σ̇ is the stress rate and σ̇1 is a unit stress rate (1 kPa s− 1) and
variations in the internal structure of the samples (Richter-Menge and μD is given by Eq. (20). Murat and Lainey (1982) also assumed that the
Cox, 1985). strain rate could be approximated by the stress rate divided by the
Overall, very little is known about the Elastic or Effective Modulus Effective Modulus. Then, the above equation becomes:
of Old Ice. It would be useful to have more measurements on this ice
type. In that this is a simple test to perform, it is surprising that so few −0:29
μE = 0:0024ð ε̇= ε̇1 Þ + μD ð22Þ
values are available.
where ε̇ is the strain rate and ε̇1 is a unit strain rate (1 s− 1). They
14. Poisson's ratio suggest that for short-term bearing capacity calculations (50 kPa s− 1),
an Effective Poisson's ratio value of 0.42 would be more suitable than
Poisson's ratio (μ) is defined as the ratio of the lateral strain to the the dynamic value of 0.33. One reason that Poisson's ratio often
longitudinal strain in a homogeneous material for a uni-axial loading exceeds 0.33 in such applications is that creep deformation is
condition. It is an important engineering property that has received contributing to the measured strain.
surprisingly little attention. Similar to the Modulus discussed above, the The Effective Poisson's ratio for sea ice is still very poorly understood.
effects of the sea ice viscoelasticity come into play with Poisson's ratio. There are a large number of factors that influence its value including the
Thus, unless the value is measured using high frequency techniques loading rate, temperature, grain size, grain structure, loading direction,
where only elastic response is involved, it is more appropriate to call the state of microcracking, etc. Additional systematic research would help to
measured values the Effective Poisson's ratio. clarify the influence of these parameters.
G.W. Timco, W.F. Weeks / Cold Regions Science and Technology 60 (2010) 107–129 125

14.2. Old Ice ice (see e.g. Dempsey, 1989). These values may be more representative
of Old Ice than those measured on first-year ice. Measurements of the
There are no reported measurements of Poisson's ratio for Old Ice. fracture toughness of Old Ice would be very useful.

15. Fracture toughness


16. Friction
The fracture toughness (Kc) is a material property that describes
There has been a number of friction tests performed with sea ice on
the stress required to make a crack of a known size propagate. It is
various substrates. Frederking and Barker (2002a,b) have summarized
directly related to the strain energy release rate (G) by the expression
the results of these tests. They also performed some interesting friction
(for plane strain);
measurements using an apparatus that could apply various levels of
2 2 (high) normal force on the sample. They measured the friction of
K1c = GE = ð1−μ Þ ð23Þ
laboratory-grown sea ice on a number of different substrates. The tests
of saline ice on steel (Frederking and Barker, 2002a) showed that there
where E is the Young's modulus and μ is Poisson's ratio. The “1” in this
was a significant variation in kinetic friction coefficient by about a factor
equation refers to the Mode 1 failure in which the crack surfaces are
of two as speed decreased from 0.1 m/s to a stop. The static friction on
opening in the direction normal to the crack faces (i.e. the “opening”
restarting motion was about five times greater than the kinetic friction
mode). The fracture toughness is usually measured by using 4-point
at 0.1 m/s. The condition of the test surface had a very significant effect
beam geometry. To perform the test, a notch is cut along the centre of
on friction. Kinetic friction at a speed of 0.1 m/s on corroded steel was
the beam approximately one-third through its thickness, and the edge
about 0.14 compared to 0.04 on smooth painted steel. Static friction on
of the notch is sharpened with a razor blade. A load is applied to the
the corroded steel was 0.45 compared to 0.25 on painted steel. Even
beam and the load at failure, or the Crack-Opening-Displacement of
with constant speed there is considerable variation in the friction
the notch can be used to calculate the fracture toughness.
coefficient, with a variation of 25 to 30% regardless of whether the
It is not entirely clear how the fracture toughness should be used in
surface was painted or corroded steel.
ice engineering. When the concept was first applied to address the
Frederking and Barker (2002b) used the apparatus to measure the
issue of ice forces on offshore structures, it was felt that this approach
friction coefficient of saline ice on various materials (concrete, wood,
would help to explain why measured full-scale loads were lower than
steel, and ice). Two general trends were apparent from the tests:
those predicted by simple physics-based models. This was never fully
friction was higher at (i) lower speeds and (ii) on rough materials.
realized and the use of the fracture mechanics concepts fell out of
They also observed that the friction was higher for the first pass of ice
fashion for awhile. However, more recently some large-scale tests
over the test samples, and was lower for subsequent passes. The
performed by Kennedy et al. (1994) and Dempsey et al. (1999) have
average friction coefficient of sea ice on smooth concrete, painted
provided good insight into the tensile failure of ice sheets.
steel and sea ice was about 0.05 for speeds greater than 50 mm/s and
increased to about 0.1 at 10 mm/s. The average friction coefficient of
15.1. First-year sea ice
sea ice on rough concrete and corroded steel was about 0.1 at speeds
greater than 100 mm/s and increased to 0.2 at 10 mm/s. Temperature
The fracture toughness of first-year sea ice has been measured by a
has a weak effect on friction coefficient, with slightly higher values at
large number of investigators especially in the 1980s and early 1990s
−2 °C than at −10 °C. There was also a weak trend of lower friction
(Urabe et al., 1980; Urabe and Yoshitake, 1981a,b; Timco and
coefficient at higher contact pressures.
Frederking, 1983b; Shen and Lin, 1986; Urabe and Inoue, 1986; Tuhkuri,
Barker and Timco (2003) measured the friction of a large block of ice
1987; DeFranco et al., 1991; DeFranco and Dempsey, 1994; Stehn,
sliding on a sand/gravel shore. Four different friction coefficients were
1994). The fracture toughness depends on the loading rate and the ice
measured corresponding to the four modes of movement of the block on
type, with less variation due to temperature and grain size. Typical
the beach: static, bulldozing, transition and sliding. The friction
values of the “opening mode” (i.e. Mode-1) fracture toughness (K1c) for
coefficient decreased as the movement mode changed from static to
small samples were in the range of 115 kPa m0.5. Dempsey (1989) and
sliding. The friction coefficient was found to be a function of the velocity
Mulmule and Dempsey (2000) have found that a proper experimental
of the interaction and generally decreased with increasing velocity.
arrangement is essential to ensure that the value measured represents
Friction values ranged between 0.6 and 0.2 with higher values for the
the true fracture toughness of ice. Many of these earlier studies did not
static friction which reduced through the bulldozing and transition mode
do this, and so their reported K1c values are not material properties for
to the lowest value when the ice was sliding along the sand/gravel shore.
ice. Totman et al. (2007) have developed a method to measure the
correct K1c on small samples in the laboratory but their tests indicated
that it was difficult to achieve stable crack growth. 17. Final comments
Kennedy et al. (1994) and Dempsey et al. (1999) report on large-
scale field programs to measure the fracture toughness of sea ice. This roller coaster ride of trying to understand the present state-of-
Dempsey et al. have shown that for thick first-year sea ice, the fracture knowledge of sea ice properties has been both interesting and
toughness is on the order of 250 kPa m0.5. This value is higher than frustrating. Clearly there is quite good understanding of a number of
those reported from measurements on small samples. This suggests the properties of sea ice, especially first-year sea ice. Table 1 summarizes
that there may be a size effect. However, Dempsey's measurements on the basic state-of-knowledge of the properties and their potential
larger pieces of ice were performed at a lower loading rate than were applications in ice engineering. The state-of-knowledge of each
the measurements on smaller pieces. As a result, creep deformation property has been summarized by:
contributed to the work of fracture in the larger bodies, but not in the
smaller ones, thereby raising the apparent fracture toughness of the • Good — there are many measurements and the results are well
cover. Further studies to define the size effect would be beneficial. correlated with the intrinsic properties of the ice;
• Reasonable — there are reasonable number of measurements but the
15.2. Old Ice results have not been well correlated with the intrinsic properties of
the ice;
There are no reported measurements of the fracture toughness of • Limited — only a few tests are available but the tests appear to be of
Old Ice. There have been a number of tests to measure K1c on freshwater good quality;
126 G.W. Timco, W.F. Weeks / Cold Regions Science and Technology 60 (2010) 107–129

Table 1
Summary of the state-of-knowledge and application of the properties of sea ice.

Knowledge Application

First-year ice Old Ice

Microstructure Good (but locally highly variable) Reasonable Controls strength and failure behaviour
(especially multi-axial loading situations).
Ice thickness Good Limited Extremely important property used in determining
ice forces, ice failure behaviour, bearing capacity,
ship resistance in ice, etc.
Salinity and porosity Good (fairly systematic) Reasonable Controls strength and failure behaviour.
Density Reasonable (consistent trends but Limited Influences strength (especially as the ice decays),
many confusing values in the literature) buoyancy (ice pieces broken by an icebreaker interacting
with a propeller), heights of ice rubble piles, etc.
Tensile strength Limited Limited Important failure process for local and mesoscale failures.
Flexural strength Good Non-existent Important failure process for local and mesoscale failures.
Important for icebreaker design and ride-up and pile-up processes.
Shear strength Poor Non-existent Important failure mode for local failures.
Compressive strength Good Reasonable Important for ice crushing failures on structures,
propeller–ice interactions, etc.
Multi-axial strength Reasonable Poor Important for ice crushing failures on structures
especially with large contact areas.
Borehole strength Reasonable Reasonable Important for situations where ice failure occurs
under confined conditions.
Creep behaviour Poor Non-existent Important for bearing capacity, ice roads, landing strips, etc.
Elastic and strain modulus Good (literature is full of confused terminology) Poor Important for bearing capacity behaviour.
Poisson's ratio Poor Non-existent Used in bearing capacity calculations.
Fracture toughness Reasonable Non-existent Important failure process for local and mesoscale failures.
Friction Poor (many variables to consider) Non-existent Important for ice–ice interactions and ice–structures
interactions, especially on sloping structures and vessels.

• Poor — the tests are either of dubious quality and/or very few in References
number; and
• Non-existent — values of this property may be guessed by analogy Ackley, S.F., Hibler, W.B., Kugrzuk, F., Kovacs, A., Weeks, W.F., 1974. Thickness and
roughness variations of arctic multi-year ice. Oceans '74, IEEE International
BUT there are no reported tests. Conference on Engineering in the Ocean Environment, vol. 1, pp. 109–117. Halifax,
NS, Canada.
As discussed in the paper and evident in the table, many properties Airaksinen, K., 1974. Free beam tests and friction tests at Pond Inlet, NWT. Polarforschung
J44, 71–75.
are poorly understood. This lack of understanding is often attributed Anderson, D.L., 1958. Preliminary results and review of sea ice elasticity and related
to the complex nature of sea ice due to the large number of factors that studies. Transactions of the Engineering Institute of Canada 2, 116–122.
affect its behaviour. However, for many properties there is insufficient Ashton, G.D. (Ed.), 1986. River and Lake Ice Engineering. Water Resources Publication,
Littleton, CO, USA.
information to allow the development of a coherent picture. This is Assur, A., 1958. Composition of sea ice and its tensile strength. Arctic Sea Ice, vol. 598. U.S.
either due to a lack of data, or the fact that not all of the essential Nat. Acad. Sci. Publication, pp. 106–138.
properties were measured during the test program. In most cases, Barker, A., Timco, G.W., 2003. The friction coefficient of a large ice block on a sand/
gravel beach. Proceedings 12th Workshop on the Hydraulics of Ice Covered Rivers,
temperature and salinity are measured and from this, the brine
CRIPE'03, Edmonton, Canada.
volume can be calculated if the ice is not decaying. Often, however, the Bjerkelund, C.A., Lapp, D.J., Ramseier, R.O., Sinha, N.K., 1985. The texture and fabric of
microstructure and density are not measured and this hampers the second-year sea cover at Mould Bay, Prince Patrick Island, NWT, April 1983.
Proceedings 1985 IEEE International Geoscience and Remote Sensing Symposium,
full interpretation of the data.
IGARSS'85, vol. 1, pp. 426–431. Amherst, MA, USA.
Knowledge of Old Ice is extremely poor. This might appear surprising Blanchet, D., Abdelnour, R., Comfort, G., 1997. Mechanical properties of first-year sea ice
since it is the most severe type of ice in the Arctic regions. Measurements at Tarsiut Island. Journal of Cold Regions Engineering 1 (1), 59–83.
on Old Ice are costly, and due to its greater thickness and strength Brown, J.H., 1963. Elasticity and strength of sea ice. In: Kingery, W.K. (Ed.), Ice and
Snow. MIT Press, Cambridge, MA, pp. 79–106.
compared to first-year ice, are more difficult to obtain. Also, it is not as Brown, T.G., Määttänen, M., 2009. Comparison of Kemi-I and Confederation Bridge cone
“readily available” geographically as first-year ice. Nevertheless, this ice ice load measurement results. Cold Regions Science and Technology 55, 3–13.
causes the most damage to vessels (Kubat and Timco, 2003) and Butkovich, T.R., 1956. Strength studies of sea ice. Snow Ice and Permafrost Research
Establishment (SIPRE). U.S. Army Research Report RR20. Wilmette, Ill., USA.
produces the highest ice loads on offshore structures (Jeffries and Butkovich, T.R., 1959. On the mechanical properties of sea ice, Thule, Greenland, 1957.
Wright, 1988; Timco and Johnston, 2004; Frederking and Sudom, 2006). Snow Ice and Permafrost Research Establishment (SIPRE). U.S. Army Research
Clearly, more effort should be spent to better understand Old Ice. Report RR54. Wilmetre, IL., USA.
Chen, A.C.T., Lee, J., 1986. Large-scale ice strength tests at slow strain rates. Proc. Offshore
The authors hope that this review paper will assist the engineering Mechanics and Arctic Engineering (OMAE), vol. 4, pp. 374–378. Tokyo, Japan.
community to make better, more informed decisions regarding oper- Cherepanov, N.V., 1966. Structure of sea ice of great thickness. Problems of Arctic Ice
ating in polar regions. Research 267, 13–18.
Cherepanov, N.V., 1974. Classification of ice of natural water bodies. Oceans '74, IEEE
International Conference on Engineering in the Ocean Environment, vol. 1, pp. 97–101.
Halifax, NS, Canada.
Acknowledgements Cole, D.M., Shapiro, L.H., 1998. Observations of brine drainage networks and
microstructure of first-year sea ice. Journal of Geophysical Research C: Oceans
103 (C10), 21,739–21,750.
The authors would like to thank Bob Frederking for helpful Cole, D.M., Gould, L.D., Burch, W.D., 1985. A system for mounting end caps on ice
comments on the manuscript. Funding for GWT from the Program of specimens. Journal of Glaciology 31 (109), 362–365.
Energy Research and Development (PERD) through the ice–structure Cox, G.F.N., Weeks, W.F., 1974. Salinity variations in sea ice. Journal of Glaciology 13
(67), 109–120.
interaction activity of the Offshore Environmental Factors program is Cox, G.F.N., Weeks, W.F., 1982. Equations for determining the gas and brine volumes in
gratefully acknowledged. sea ice samples. CRREL Report 82-30, Hanover, N.H., USA.
G.W. Timco, W.F. Weeks / Cold Regions Science and Technology 60 (2010) 107–129 127

Cox, G.F.N., Weeks, W.F., 1983. Equations for determining the gas and brine volumes in Gammon, P.H., Kietfe, H., Cloutier, M.J., Denner, W.W., 1983. Elastic constants of
sea ice samples. Journal of Glaciology 29 (102), 306–316. artificial and naturally ice samples by Brillouin spectroscopy. Journal of Glaciology
Cox, G.F.N., Weeks, W.F., 1988. Profile properties of undeformed first-year sea ice. 29 (103), 433–460.
CRREL Report 88-13, Hanover, N.H., USA. Geotech, 1983. Multi-year ice strength testing program for Gulf Canada Resources Inc.
Cox, G.F.N., Richter-Menge, J.A., 1985. Triaxial compression testing of ice. Proceedings of Report submitted to Gulf Canada, Calgary, AB, Canada. (APOA Project 200).
the Conference “Arctic 85”, ASCE, pp. 476–488. San Francisco, CA, USA. Geotech, 1984. Multi-year ice strength test program, phase II. Report 9100 submitted to
Cox, G.F.N., Richter, J.A., Weeks, W.F., Mellor, M., 1984a. A summary of the strength and Gulf Canada, Calgary, AB, Canada.
modulus of ice samples from multiyear pressure ridges. Proceedings 3rd Gold, L.W., 1971. Use of ice covers for transportation. Canadian Geotechnical Journal 8,
International Offshore Mechanics and Arctic Engineering Symposium, ASME, NY, 170–181.
vol. III, pp. 126–133. New Orleans, LA, USA. Gold, L.W., 1977. Engineering properties of freshwater ice. Journal of Glaciology 19 (81),
Cox, G.F.N., Richter, J.A., Weeks, W.F., Mellor, M., Bosworth, H., 1984b. The mechanical 197–212.
properties of multi-year sea ice, Phase I: test results. CRREL Report 84-9. 105 pages, Gow, A.J., Ueda, H.T., 1984. Flexural strengths of freshwater model ice. Proceedings
Hanover, NH, USA. International Association for Hydraulic Research, IAHR Symposium on Ice Problems,
Cox, G.F.N., Richter-Menge, J.A., Weeks, W.F., Bosworth, H.W., Perron, N., Mellor, M., vol. 1, pp. 73–82. Hamburg, Germany.
Durell, G., 1985b. Mechanical properties of multiyear sea ice, Phase II: test results. Gow, A.J., Ackley, S.F., Buck, K.R., Golden, K.M., 1987a. Physical and structural
CRREL Report 85-16. 81 pages, Hanover, NH, USA. characteristics of Weddell Sea pack ice. CRREL Report 87-14. 70 pages, Hanover,
Cox, G.F.N., Richter, J.A., Weeks, W.F., Mellor, M., 1985a. A summary of the strength and NH, USA.
modulus of ice samples from multi-year pressure ridges. Journal of Energy Gow, A.J., Tucker, W.B., Weeks, W.F., 1987b. Physical properties of summer sea ice in the
Resources Technology 107 (3), 93–98. Fram Strait, June–July, 1984. CRREL Report 87-16. 81 pages, Hanover, NH, USA.
DeFranco, S.J., Dempsey, J.P., 1994. Crack propagation and fracture resistance in saline Gow, A.J., Ueda, H.T., Richard, J.A., 1987c. Flexural strength of ice on temperate lakes. U.S.
ice. Journal of Glaciology 40, 451–462. Army CRREL Report 78-9, Hanover, NH, USA.
DeFranco, S.J., Wei, Y., Dempsey, J.P., 1991. Notch acuity effects on fracture toughness of Gow, A.J., Ueda, H.T., Govoni, J.W., Kalafut, J., 1988. Temperature and structure
saline ice. Annals of Glaciology 15, 230–235. dependence of the flexural strength and modulus of freshwater model ice. CRREL
Dempsey, J.P., 1989. The fracture toughness of ice. Proc. IUTAM/IAHR Symposium on Report 88-6, Hanover, NH, USA. CRREL Report 88-6, Hanover, NH, USA.
Ice/Structure Interaction, pp. 109–145. St. John's, Nfld., Canada. Gratz, E.T., Schulson, E.M., 1994. Preliminary observations of brittle compressive failure
Dempsey, J.P., Adamson, R.M., Mulmule, S.V., 1999. Scale effects on the in situ tensile of columnar saline ice under triaxial loading. Annals of Glaciology 19, 33–38.
strength and fracture of ice. Part II: first-year sea ice at Resolute, NWT. Intentional Gratz, E.T., Schulson, E.M., 1997. Brittle failure of columnar saline ice under triaxial
Journal of Fracture 95, 347–366. compression. Journal of Geophysical Research 102 (B3), 5091–5107.
Derradji-Aouat, A., 2003. Multi-surface failure criterion for saline ice in the brittle range. Hausler, F.U., 1981. Multi-axial compressive strength tests on saline ice with brush-type
Cold Regions Science and Technology 36 (1), 47–70. loading platens. IAHR International Symposium on Ice, vol. 2, pp. 526–539. Lulea,
Dykins, J.E., 1967. Tensile properties of sea ice grown in a confined system. In: Oura, H. Sweden.
(Ed.), Proceedings of the International Conference on Low Temperature Science. Hibler, W.D., Ackley, S., Weeks, W.F., Kovacs, A., 1972. Top and bottom roughness of a
Physics of Snow and Ice, vol. 1. Institute of Low Temperature Science, Sapporo, multiyear ice floe. Proceedings of the International Association for Hydraulic
Japan, pp. 523–537. Research IAHR Symposium on Ice, pp. 130–142. Leningrad, U.S.S.R.
Dykins, J.E., 1968. Tensile and flexural properties of saline ice. In: Riehl, N., Bullemer, B., Hunkins, K., 1960. Seismic studies of sea ice. Journal of Geophysical Research 65,
Engelhardt, H. (Eds.), Physics of Ice. Plenum Press, New York, NY, pp. 251–270. 3,459–3,472.
Dykins, J.E., 1971. Ice engineering— material properties for a limited range of Iliescu, D., Schulson, E.M., 2004. The brittle compressive failure of fresh-water columnar
conditions. US Navy Civil Engineering Laboratory, Technical Report R720, Port ice loaded bi-axially. Acta Materialia 52 (20), 5,723–5,735.
Hueneme, CA. Jeffries, M.G., Wright, W.H., 1988. Dynamic response of Molikpaq to ice–structure
Eicken, H., Lensu, M., Leppäranta, M., Tucker III, W.B., Gow, A.J., Salmela, O., 1995. interaction. Proc. OMAE'88, vol. IV, pp. 201–220. Houston, TX, USA.
Thickness, structure, and properties of level summer multi-year ice in the Eurasian Jeffries, M.O., Sackinger, W.M., Frederking, R., Timco, G.W., 1988. Initial mechanical and
sector of the Arctic Ocean. Journal of Geophysical Research 100, 22697–22710. physical–structural property measurements of old sea and brackish ice from the Ward
Eicken, H., Krouse, H.R., Kadko, D., Perovich, D.K., 2002. Tracer studies of pathways and Hunt Ice Shelf, Canada. Proceedings IAHR Symposium on Ice, vol. 1, pp. 177–187.
rates of meltwater transport through Arctic summer sea ice. Journal of Geophysical Sapporo, Japan.
Research 107 (C10), 8046. doi:10.1029/2000JC000583. Johnston, M., 1998. Influence of ice microstructure on microwave scattering properties
Enquist, E., 1972. On the ice resistance encountered by ships operating in the continuous of sea ice. Ph.D. Thesis no. 16884, Universite Laval, Quebec City, Canada, 311 pages.
mode of icebreaking. Swedish Academy of Engineering Sciences in Finland, Report No. Johnston, M., 2008. Characterizing multi-year ice in the high Arctic: evaluating two
24, Helsinki, Finland. ground-based EM sensors. Report by NRC Canadian Hydraulics Centre, CHC-CTR-073.
Fenco, 1977. 1977 Winter field ice survey offshore Labrador. Report submitted to Total 70 pages (Controlled), Ottawa, ON, Canada.
Eastcan Exploration Ltd., Calgary, AB., Canada. Johnston, M., Timco, G.W., 2002. Temperature changes in first year arctic sea ice during
Frankenstein, G.E., 1959. Strength data on lake ice. SIPRE Technical Report 59, Wilmette, the decay process. Proceedings of the 16th IAHR International Symposium on Ice,
IL, USA. vol. 2, pp. 194–202. Dunedin, New Zealand.
Frankenstein, G.E., 1961. Strength data on lake ice. SIPRE Technical Report 80, Wilmette, Johnston, M., Timco, G.W., 2008. Understanding and identifying Old Ice in summer. NRC
IL, USA. Canadian Hydraulics Centre Report CHC-TR-055, Ottawa, Ont., Canada.
Frankenstein, G.E., Garner, R., 1967. Equations for determining the brine volume of sea Johnston, M., Frederking, R., Timco, G.W., 2000. Seasonal decay of first-year sea ice. NRC
ice from −0.5 to −22.9 °C. Journal of Glaciology 6 (48), 943–944. Canadian Hydraulics Centre Report HYD-TR-058, Ottawa, Ont., Canada.
Frederking, R.M.W., 1977. Plane-strain compressive strength of columnar-grained and Johnston, M., Frederking, R., Timco, G.W., 2001. Decay induced changes in the physical and
granular snow-ice. Journal of Glaciology 18, 505–516. mechanical properties of first-year sea ice. Proceedings Port and Ocean Engineering
Frederking, R.M.W., Hausler, F., 1978. The flexural behaviour of ice from in situ under Artic Conditions, POAC'01, vol. 3, pp. 1395–1404. Ottawa, Canada.
cantilever beam tests. Proceedings International Association for Hydraulic Johnston, M., Frederking, R., Timco, G.W., 2002. Properties of decaying first-year sea ice:
Research, IAHR Symposium on Ice Problems, vol. 1, pp. 197–215. Lulea, Sweden. two seasons of field measurements. Proc. 17th Int. Sym. on Okhotsk Sea and Sea Ice,
Frederking, R.M.W., Timco, G.W., 1980. NRC ice property measurements during the pp. 303–311. Hokkaido, Japan.
Canmar Kigoriak trials in the Beaufort Sea, winter 1979–80. DBR Paper No. 947, Johnston, M., Frederking, R., Timco, G.W., 2003a. Properties of decaying first-year sea
DBR/NRC Rept., Ottawa, ON, Canada. ice at five sites in Parry Channel. Proc. 17th Int. Conf. on Port and Ocean Engineering
Frederking, R.M.W., Timco, G.W., 1983. Uni-axial compressive strength and deforma- under Arctic Conditions (POAC), vol. 1, pp. 131–140. Trondheim, Norway.
tion of Beaufort Sea ice. Proceedings International Conference on Port and Ocean Johnston, M., Frederking, R., Timco, G.W., 2003b. Property changes of first-year ice and
Engineering under Arctic Condition, POAC 83, vol. I, pp. 89–98. Helsinki, Finland. old ice during summer melt. NRC Canadian Hydraulics Centre Report CHC-TR-010,
Frederking, R.M.W., Timco, G.W., 1984a. Measurement of shear strength of granular/ TP14098E, Ottawa, Ont., Canada.
discontinuous columnar sea ice. Cold Regions Science and Technology 9, 215–220. Johnston, M., Timco, G.W., Frederking, R., 2003c. In situ borehole strength measure-
Frederking, R.M.W., Timco, G.W., 1984b. Compressive behaviour of Beaufort Sea ice under ments on multi-year sea ice. Proceedings of the 13th International Offshore and
vertical and horizontal loading. Proceedings OMAE Symposium, vol. III, pp. 145–149. Polar Engineering Conference, pp. 445–452. Honolulu, Hawaii, USA.
New Orleans, USA. Johnston, M.E., Masterson, D., Wright, B., 2009. Multi-year ice thickness: knowns and
Frederking, R.M.W., Timco, G.W., 1986. Field measurements of the shear strength of unknowns. Proceedings 20th POAC Conference, Paper POAC09-120, Lulea Univer-
columnar-grained sea ice. Proceedings 8th International Association for Hydraulic sity of Technology, Lulea, Sweden.
Research Symposium on Ice, vol. I, pp. 279–292. Iowa City, U.S.A. Jones, S.J., 1997. High strain-rate compression tests on ice. Journal of Physical Chemistry. B
Frederking, R., Barker, A., 2002a. Friction of sea ice on steel for condition of varying 101, 6099–6101.
speeds. Proceedings of the 12th International Offshore and Polar Engineering Kayo, Y., Kawasaki, T., Minami, T., Tozawa, S., Tanaka, A., Abdelnour, R., 1983. Field study
Conference, pp. 766–771. Kitakyushu, Japan. on mechanical properties of sea ice at east coast of Hokkaido. Proceedings POAC'83,
Frederking, R., Barker, A., 2002b. Friction of sea ice on various construction materials. vol. 1, pp. 109–118. Helsinki, Finland.
Proceedings of the 16th IAHR International Symposium on Ice, vol. 1, pp. 442–449. Kennedy, K.P., Mamer, K.J., Dempsey, J.P., Adamson, R.M., Spencer, P.A., Masterson, D.M.,
Dunedin, New Zealand. 1994. Large scale ice fracture experiments: Phase 2. Proc. 12th IAHR Ice Symposium,
Frederking, R., Sudom, D., 2006. Maximum ice force on the Molikpaq during the April vol. 1, pp. 315–324. Trondheim, Norway.
12, 1986 event. Cold Regions Science and Technology 46 (3), 147–166. Kivisild, H.R., 1992. In situ borehole testing in ice: a historical perspective. Proceedings
Frederking, R.M.W., Svec, O., Timco, G.W., 1988. The shear strength of ice. Proceedings 9th IAHR Symposium on Ice, vol. 2, pp. 841–855. Banff, AB, Canada.
International Association for Hydraulic Research Symposium on Ice, vol. 3, pp. 76–88. Kohnen, H., 1972. Seismic and ultrasonic measurements on the sea ice of Eclipse Sound
Sapporo, Japan. near Pond Inlet, N.W.T. on Northern Baffin Island. Polarforschung 42, 66–74.
128 G.W. Timco, W.F. Weeks / Cold Regions Science and Technology 60 (2010) 107–129

Kovacs, A., 1975. A study of multi-year pressure ridges and shore ice pile-up. APOA Pounder, E.R., Stalinski, P., 1960. General Properties of Arctic Sea Ice. Assoc. Science
Project Report 89. 45 pages, Hanover, NH, USA. Hydrology, Pub. No 54, pp. 25–34.
Kovacs, A., 1996a. Sea Ice Part 1. Bulk salinity versus ice floe thickness. CRREL Report 96-7, Richter-Menge, J.A., Cox, G.F.N., 1984. Structure, salinity and density of multi-year sea
Hanover, NH, USA. ice pressure ridges. Proceedings OMAE'84, vol. 2, pp. 194–198. Dallas, TX, USA.
Kovacs, A., 1996b. Sea Ice Part II. Estimating the full-scale tensile, flexural, and compressive Richter-Menge, J.A., Cox, G.F.N., 1985. A preliminary examination of the effect of
strength of first-year ice. CRREL Report 96-11. 17 pages, Hanover, NH, USA. structure on the compressive strength of ice samples from multi-year pressure
Kovacs, A., Mellor, M., 1971. Sea ice pressure ridges and ice islands. Report by Creare Inc. for ridges. Journal of Energy Resources Technology V107, 99–102.
Arctic Petroleum Operators Association, Technical Note, vol. 122. Hanover, NH, USA. Richter-Menge, J.A., Jones, K.F., 1993. The tensile strength of first-year sea ice. Journal of
Kubat, I., Timco, G.W., 2003. Vessel damage in the Canadian Arctic. Proceedings 17th Glaciology 39 (133), 609–618.
International Conference on Port and Ocean Engineering under Arctic Conditions, Richter-Menge, J.A., Claffey, K.J., Walsh, M.R., 1993. End-capping procedure for cored ice
POAC'03, vol. 1, pp. 203–212. Trondheim, Norway. samples used in tension tests. Journal of Glaciology V39 (133), 698–700.
Kuehn, G.A., Lee, R.W., Nixon, W.A., Schulson, E.M., 1990. The structure and tensile Saeki, H., Ozaki, A., Kubo, Y., 1981. Experimental study on flexural strength and elastic
behavior of first-year sea ice and laboratory grown saline ice. ASME Journal of modulus of sea ice. Proceedings of Port and Ocean Engineering under Arctic
Offshore Mechanics and Arctic Engineering 112 (4), 357–363. Conditions, POAC'81, vol. 1, pp. 536–547. Quebec City, PQ, Canada.
Kwok, R., Rothrock, D.A., 2009. Decline in Arctic sea ice thickness from submarine and Sammonds, P.R., Murrell, S.A.F., Rist, M.A., 1998. Fracture of multi-year sea ice. Journal
ICESat records: 1958–2008. JGR 36, L15501. doi:10.1029/2009GL039035. of Geophysical Research 103 (C10), 21,795–21,815.
Lainey, L., Tinawi, R., 1981. The mechanical properties of sea ice — a compilation of Sanderson, T.J.O., 1988. Ice mechanics: risks to offshore structures. Graham & Tronman,
available data. Canadian Journal of Civil Engineering 11 (4), 884–923. London. pp. 148–151.
Lake, R.A., Lewis, E.L., 1970. Salt rejection by sea ice during growth. Journal of Schulson, E., 2002. Compressive shear faults in ice: plastic vs. Coulombic faults. Acta
Geophysical Research 75 (3), 583–597. Materialia 50 (13), 3415–3424. doi:10.1016/S1359-6454(02)00154-4.
Langleben, M.P., 1959. Some physical properties of sea ice II. Canadian Journal of Physics Schulson, E.M., Nickolayev, O.Y., 1995. Failure of columnar saline ice under biaxial
37, 1438–1454. compression: failure envelopes and the brittle to ductile transition. Journal of
Langleben, M.P., Pounder, E.R., 1963. Elastic parameters of sea ice. In: Kingery, W.D. Geophysical Research 100 (B11), 22,383–22,400.
(Ed.), Ice and Snow. MIT Press, USA, pp. 69–78. Schulson, E.M., Duval, P., 2009. Creep and Fracture of Ice. Cambridge University Press,
Lavrov, V.V., 1969. Deformation and strength of ice. Gidrometeorologicheskoe Izdatel'stvo. Cambridge, UK.
Leningrad. Schulson, E.M., Fortt, A.L, Iliescu, D., Renshaw, C.E., 2006. Failure envelope of first-year
Lee, R.W., 1986. A procedure for testing cored ice under uniaxial tension. Journal of Arctic sea ice: the role of friction in compressive fracture. Journal of Geophysical
Glaciology 32 (112), 540–541. Research 111 (C11S25). doi:10.1029/2005JC003235.
Lee, J., Ralston, T.D., Petrie, D.H., 1986. Full-thickness sea ice strength tests. Proceedings Schwarz, J., Weeks, W.F., 1977. Engineering properties of sea ice. Journal of Glaciology
IAHR Ice Symposium, vol. 1, pp. 293–306. Iowa City, Iowa, USA. 19 (81), 499–531.
Light, B., Maykut, G.A, Grenfell, T.C., 2003. Effects of temperature on the microstructure Schwarz, J., Frederking, R., Gavrillo, V., Petrov, I.G., Hirayama, K.I., Mellor, M., Tryde, P.,
of first-year Arctic sea ice. Journal of Geophysical Research 108 (C2), 3051. Vaudrey, K.D., 1981. Standardized testing methods for measuring mechanical
doi:10.1029/2001 JC000887. properties of sea ice. Cold Regions Science and Technology 4, 245–253.
Lin'kov, E.M., 1958. Study of the elastic properties of an ice cover in the Arctic. Vestnik Shapiro, L.H., Weeks, W.F., 1995. Controls on the flexural strength of small plates and
Leniagradskogo Universiteta 13, 17–22 (in Russian). beams of first-year sea ice. “Ice Mechanics-95”, Joint Mechanics Meeting of the
Määttänen, M., 1975. On the flexural strength of brackish water in in situ tests. American Society for Civil Engineering–Engineering Mechanics Division/American
Proceedings POAC'75, vol.1, pp. 349–359. Fairbanks, AL, USA. Society of Mechanical Engineers–Applied Mechanics Division / Society of
Malmgren, F., 1927. Norwegian north polar expedition with the “Maud”, 1918–1925. Engineering Science, AMD-Vol. 207, pp. ASME, UCLA, USA, pp. 179–188.
Scientific Results 1 (5). Shen, W., Lin, S.Z., 1986. Fracture toughness of Bohai Bay sea ice. Proc. 5th OMAE
Marion, G.M., Farren, R.E., Komrowski, A.J., 1999. Alternative pathways for seawater Symposium, vol. IV, pp. 354–357. Tokyo, Japan.
freezing. Cold Regions Science and Technology 29 (3), 259–266. Sinha, N.K., 1978. Rheology of columnar-grained ice. Experimental Mechanics 18 (12),
Masterson, D.M., 1992. Interpretation of in situ borehole ice strength measurement 464–470.
tests. Proceedings IAHR Symposium on Ice, vol. 2, pp. 802–815. Banff, AB, Canada. Sinha, N.K., 1979. Grain-boundary sliding in polycrystalline materials. Philosophical
Masterson, D.M., 1996. Interpretation of in situ borehole ice strength measurement Magazine. A 40 (6), 825–842.
tests. Canadian Journal of Civil Engineering 23 (1), 165–179. Sinha, N.K., 1981. Rate sensitivity of compressive strength of columnar-grained ice.
Masterson, D.M., 2009. State of the art of ice bearing capacity and ice construction. Cold Experimental Mechanics 21 (6), 209–218.
Regions Science and Technology 58 (3), 99–112. doi:10.1016/j.coldregions.2009.04.002. Sinha, N.K., 1983a. Field test 1 of compressive strength of first year sea ice. Annals of
Masterson, D.M., in press. Ice strength: In situ measurement. In: Eicken, H., Gradinger, R., Glaciology 4, 253–259.
Salganek, M., Shirasawa, K., Perovich, D., Leppäranta, M., editors. Field techniques for Sinha, N.K., 1983b. Field tests on rate sensitivity of vertical strength and deformation of
sea ice research. University of Alaska Press, Fairbanks, AK, p. 181–213. first-year columnar-grained sea ice. Proceedings POAC 83, vol. 1, pp. 231–242.
Masterson, D.M., Graham, W.P., 1992. Development of the original ice borehole jack. Helsinki, Finland.
Proceedings IAHR Symposium on Ice, vol. 2, pp. 748–759. Banff, AB, Canada. Sinha, N.K., 1984. Uniaxial compressive strength of first-year and multi-year sea ice.
Masterson, D.M., Frederking, R., 2006. Experience with the Canadian Standards Association Canadian Journal of Civil Engineering 11, 82–91.
Offshore Structures Code. Proceedings 16th International Offshore and Polar Sinha, N.K., 1985. Confined strength and deformation of second-year columnar-grained
Engineering Conference, ISOPE'06, vol. 1, pp. 14–19. San Francisco, CA, USA. sea ice in Mould Bay. Proceedings OMAE'85, vol. 2, pp. 209–291. Dallas, TX, USA.
Masterson, D.M., Graham, W.P., Jones, S.J., Childs, G.R., 1997. A comparison of uniaxial and Sinha, N.K., 1986a. Young Arctic frazil sea ice: field and laboratory strength tests.
borehole jack tests at Fort Providence ice crossing, 1995. Canadian Geotechnical Journal of Materials Science 21 (5), 1533–1546.
Journal 34, 471–475. Sinha, N.K., 1986b. The borehole jack: is it a useful tool? Proc. of 5th Int. Offshore Mechanics
Masterson, D.M., Frederking, R.M.W., Truskov, P.A., 2000. Ice force and pressure and Arctic Engineering Symposium (OMAE), vol. IV, pp. 328–335. Tokyo, Japan.
determination by zone. Proceedings ICETECH 2000, pp. 383–390. St. Petersburg, Russia. Sinha, N.K., 1987. Effective Poisson's ratio of isotropic ice. Proceedings OMAE'87, vol. IV,
Melling, H., Reidel, D.A., Gedalof, Z., 2005. Trends in thickness and extent of seasonal pp. 189–195. Houston, TX, USA.
pack ice, Canadian Beaufort Sea. Geophysical Research Letters 32 (24), 1–5. Sinha, N.K., 1989c. Closed-loop controlled tensile strength testing method for multi-
Mellor, M., 1983. Mechanical behavior of sea ice. USA CRREL Monograph 83-1, Hanover, year sea ice. Proceedings OMAE, vol. IV. The Hague, The Netherlands, pp. 1–6.
N.H., USA. Sinha, N.K., 1989a. Microcrack-enhanced creep in polycrystalline material at elevated
Mellor, M., 1986. Mechanical behavior of sea ice. In: Untersteiner, N. (Ed.), Geophysics temperatures. Acta Metallurgica 37, 3107–3118.
of Sea Ice, vol. 146. Plenum Press, New York, pp. 165–281. Sinha, N.K., 1989b. Experiments on anisotropic and rate-sensitive strain ratio and
Moslet, P.O., 2007. Field testing of uniaxial compression strength of columnar sea ice. modulus of columnar-grained ice. Journal of Offshore Mechanics and Arctic
Cold Regions Science and Technology 48 (1), 1–14. Engineering 111, 354–360.
Mulmule, S.V., Dempsey, J.P., 2000. LEFM size requirements for the fracture testing of Sinha, N.K., 1990. Is minimum creep rate a fundamental material property? Proc. 9th
sea ice. International Journal of Fracture 102 (1), 85–98. OMAE Conference, vol. IV, pp. 283–288. Houston, TX, USA.
Murat, J.R., Lainey, L.M., 1982. Some experimental observations on the Poisson's ratio of Sinha, N.K., 1991. In Situ multi-year ice strength using NRCC borehole indentor. Proc. of
sea ice. Cold Regions Science and Technology 6, 105–113. 10th Int. Offshore Mechanics and Arctic Engineering Symposium (OMAE), vol. IV,
Nakawo, M., 1983. Measurements on air porosity of sea ice. Annals of Glaciology 4, 204–208. pp. 229–236. Stavanger, Norway.
Paige, R.A., Lee, C.W., 1967. Preliminary studies on sea ice in McMurdo Sound, Slesarenko, Y.E., Frolov, A.D., 1974. Comparison of elasticity and strength characteristics
Antarctica, during Deep Freeze 65. Journal of Glaciology 6 (46), 515–528. of salt-water ice. Proceedings of the IAHR Symposium on Ice, vol. 2, pp. 85–87.
Parsons, B.L., Lai, M., Williams, F.M., Dempsey, J.P., Snellen, J.B., Everard, J., Slade, T., Leningrad, Russia.
Williams, J., 1992. The influence of beam size on the flexural strength of sea ice, Smith, T.R., Schulson, E.M., 1994. Brittle compressive strength of salt-water columnar
freshwater ice and iceberg ice. Philosophical Magazine. A 66 (6), 1017–1036. ice under biaxial loading. Journal of Glaciology 40 (135), 265–276.
Peschanskii, I.S., 1960. Arctic and Antarctic sea ice (R). Problemy Arktiki i Antarktiki, vol. 4, Stander, E., Michel, B., 1989. The effect of fluid flow on the development of preferred
pp. 111–129. St. Petersburg, Russia. orientations in sea ice: laboratory experiments. Cold Regions Science and
Peyton, H.R., 1963. Some mechanical properties of sea ice. In: Kingery, W.D. (Ed.), Ice and Technology 17 (2), 153–161.
Snow-Processes, Properties and Applications. MIT Press, Cambridge, MA, pp. 107–113. Stehn, L., 1994. Fracture toughness and crack-growth of brackish ice using chevron-
Peyton, H.R., 1966. Sea ice strength. University of Alaska Report UAG-182. Geophysical notched specimens. Journal of Glaciology 40, 415–426.
Institute, Fairbanks, AK, USA. 187 pp. Svec, O.J., Thompson, J.C., Frederking, R.M.W., 1985. Stress concentrations in the root of
Pounder, E.R., Little, E.M., 1959. Some physical properties of sea ice. Canadian Journal of an ice cover cantilever: model tests and theory. Cold Regions Science Technology
Physics 37, 443–473. 11, 63–73.
G.W. Timco, W.F. Weeks / Cold Regions Science and Technology 60 (2010) 107–129 129

Tabata, T., 1960. Studies on mechanical properties of sea ice V. Measurement of flexural Tuhkuri, J., 1987. The applicability of LEFM and the fracture toughness of sea ice. Proc.
strength. Low Temperature Science, Series A 19, 187–201. 9th International POAC Conference, vol. I, pp. 21–32. Fairbanks, Alaska, USA.
Tabata, T., 1966. Studies on the mechanical properties of sea ice IX. Measurement of the Urabe, N., Yoshitake, A., 1981a. Strain rate dependent fracture toughness K1c of pure ice and
flexural strength in situ. Low Temperature Science, Series A 24, 259–268. sea ice. Proc 6th IAHR Ice Symposium, vol. II, pp. 551–563. Quebec City, PQ, Canada.
Tabata, T., 1967. The flexural strength of small sea ice beams. Physics of Snow and Ice, Urabe, N., Yoshitake, A., 1981b. Fracture toughness of sea ice — in situ measurement and its
vol. I. Hokkaido University, Japan, pp. 481–497. Part 1. application. Proc 6th POAC Conference, vol. I, pp. 356–365. Quebec City, PQ, Canada.
Tabata, T., Fujino, K., 1964. Studies on the mechanical properties of sea ice VIII. Urabe, N., Inoue, M., 1986. Mechanical properties of Antarctic sea ice. ASME OMAE
Measurement of the flexural strength in situ. Low Temperature Science, Series A 23, Symposium, vol. 4, pp. 303–309. Tokyo, Japan.
157–166. Urabe, N., Iwasaki, T., Yoshitake, A., 1980. Fracture toughness of sea ice. Cold Regions
Tabata, T., Fujino, K., Aota, M., 1967. Studies on the mechanical properties of sea ice: the Science and Technology 3, 29–37.
flexural strength of sea ice in situ. Physics of Snow and Ice, vol. I(1). Hokkaido Vaudrey, K., 1977. Ice engineering—— study of related properties of floating sea ice
University, Japan, pp. 539–550. sheets and summary of elastic and viscoelastic analyses. U.S. Naval Civil
Tabata, T., Suzuki, Y., Aota, M., 1975. Ice study in the Gulf of Bothnia: II. Measurement of Engineering Laboratory, Report TR860, Port Hueneme, CA.
flexural strength. Low Temperature Science, Series A 33, 199–206. Wang, Y.S., 1979. Crystallographic studies and strength tests of field ice in the Alaskan
Timco, G.W., Burden, R.P., 1997. An analysis of the shapes of sea ice ridges. Cold Regions Beaufort Sea. Proc. POAC 79, vol. I, pp. 651–665. Trondheim, Norway.
Science and Technology 25, 65–77. Wang, Y.S., 1981. Uniaxial compression testing of Arctic sea ice. Proceedings of the Sixth
Timco, G.W., Frederking, R.M.W., 1982a. Comparative strength of freshwater ice. Cold Int. Conf. POAC, vol. 1, pp. 346–355. Quebec City, PQ, Canada.
Regions Science and Technology 6, 21–27. Wang, Y.S., Poplin, J.P., 1986. Laboratory compressive tests of sea ice at slow strain rates
Timco, G.W., Frederking, R.M.W., 1982b. Compressive strength of multi-year ridge ice. from a field test program. Proc. OMAE 86, vol. 4, pp. 379–384. Tokyo, Japan.
Proceedings Workshop on Sea Ice Ridging and Pile-up. NRC/DBR Technical Memo Weeks, W.F., 1985. The variation of sea ice strength within and between multiyear
134, Calgary, Al, Canada. pressure ridges in the Beaufort Sea. Journal of Energy Resources Technology 107 (2),
Timco, G.W., Frederking, R.M.W., 1983b. Confined compressive strength of sea ice. Proc. 167–172.
POAC 83, vol. I, pp. 243–253. Helsinki, Finland. Weeks, W.F., 1998. Growth conditions and the structure and properties of sea ice. In:
Timco, G.W., Frederking, R.M.W., 1983a. Flexural strength and fracture toughness of sea Leppäranta, M. (Ed.), Physics of Ice-covered Seas, vol. 1. Helsinki University Printing
ice. Cold Regions Science and Technology 8, 35–41. House, Helsinki, Finland, pp. 25–104.
Timco, G.W., Frederking, R.M.W., 1984. An investigation of the failure envelope of Weeks, W.F., in press. On Sea Ice. University of Alaska Press, Fairbanks AL, USA.
granular/discontinuous-columnar sea ice. Cold Regions Science and Technology 9, Weeks, W.F., Anderson, D., 1958. An experimental study of strength of young sea ice.
17–27. Trans. American Geophysical Union 4 (4), 641–647.
Timco, G.W., Frederking, R.M.W., 1986. Confined compression tests: outlining the failure Weeks, W.F., Lee, S.O., 1958. Observations on the physical properties of sea ice at
envelope of columnar sea ice. Cold Regions Science and Technology 12 (1), 13–28. Hopedale, Labrador. Arctic II (3), 135–155.
Timco, G.W., Frederking, R.M.W., 1990. Compressive strength of sea ice sheets. Cold Weeks, W.F., Assur, A., 1967. The mechanical properties of sea ice. U.S. Army CRREL
Regions Science and Technology 17, 227–240. Monograph II-C3, Hanover, NH, USA.
Timco, G.W., Frederking, R.M.W., 1991. Seasonal compressive strength of Beaufort Sea ice Weeks, W.F., Assur, A., 1968. The mechanical properties of sea ice. Proceedings, Conference
sheets. In: Jones, S., et al. (Ed.), Proceedings IUTAM-IAHR Symposium on Ice/Structure on Ice Pressures Against Structures, Quebec City, Laval University, National Research
Interaction, St. John's, Nfld. Springer Verlag, Berlin, Heidelberg, pp. 267–282. Council of Canada NRC-DBR Report Tech. Memo. No. 92, pp. 25–78. Ottawa, ON, Canada.
Timco, G.W., O'Brien, S., 1994. Flexural strength equation for sea ice. Cold Regions Weeks, W.F., Gow, A.J., 1980. Crystal alignments in the fast ice of Arctic Alaska. Journal
Science and Technology 22, 285–298. of Geophysical Research 85 (C2), 1,137–1,146.
Timco, G.W., Frederking, R.M.W., 1996. A review of sea ice density. Cold Regions Science Weeks, W.F., Ackley, S.F., 1982. The growth, structure and properties of sea ice. USA
and Technology 24, 1–6. CRREL Monograph 82-1, Hanover, N.H., USA.
Timco, G.W., Johnston, M.E., 2002. Sea ice strength during the melt season. Proceedings Weeks, W.F., Ackley, S.F., 1986. The growth, structure, and properties of sea ice. In:
of the 16th IAHR International Symposium on Ice, vol. 2, pp. 187–193. Dunedin, Untersteiner, N. (Ed.), The Geophysics of Sea Ice. NATO ASI Series, vol. 146. Plenum
New Zealand. Press, New York, pp. 9–164.
Timco, G.W., Johnston, M.E., 2004. Ice loads on the caisson structures in the Canadian Wettlaufer, J.S., 1998. Introduction to crystallization phenomenon in natural and
Beaufort Sea. Cold Regions Science and Technology 38, 185–209. artificial sea ice. In: Leppäranta, M. (Ed.), Physics of Ice-covered Seas, vol. 1.
Totman, C.A., Uzorka, O.E., Dempsey, J.P., Cole, D.M., 2007. Sub-size fracture testing of FY Helsinki University Printing House, pp. 105–194. Helsinki, Finland.
sea ice. Proceedings of the 6th International Conference on Fracture Mechanics of Williams, F.M., Everard, J., Butt, S., 1992. Ice and snow measurements in support of the
Concrete and Concrete Structures, vol. 3, pp. 1683–1690. Catania, Italy. operational evaluation of the Nathanial B. Palmer in the Antarctic winter
Tratteberg, A., Gold, L.W., Frederking, R., 1975. The strain rate and temperature environment. NRC/IMD Report TR-1992-14, St. John's, NL, Canada.
dependence of Young's modulus of ice. Proceedings IAHR Symposium on Ice
Problems, pp. 479–486. Hanover, NH, USA.

Vous aimerez peut-être aussi