Está en la página 1de 34

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/254214058

Ice Adhesion —Theory, Measurements and Countermeasures

Article  in  Journal of Adhesion Science and Technology · March 2012


DOI: 10.1163/016942411X574583

CITATIONS READS

50 842

1 author:

Lasse Makkonen
VTT Technical Research Centre of Finland
157 PUBLICATIONS   2,572 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

DICE (Funded by Academy of Finland, grant 268925) View project

All content following this page was uploaded by Lasse Makkonen on 30 July 2018.

The user has requested enhancement of the downloaded file.


Journal of Adhesion Science and Technology 26 (2012) 413–445
brill.nl/jast

Ice Adhesion — Theory,


Measurements and Countermeasures

Lasse Makkonen ∗
VTT Technical Research Centre of Finland, Box 1000, 02044 VTT, Espoo, Finland

Received on 18 February 2011

Abstract
In this paper concepts and models for theoretically estimating ice adhesion are presented. The effects of
temperature, ice salinity and properties of the substrate material on ice adhesion are explained by these
theoretical concepts. Major problems caused by ice adhesion are outlined and the applications of the theory
in combating ice adhesion are discussed. Measurement methods of ice adhesion are described and results of
ice adhesion for various material surfaces are reviewed and interpreted in view of the theory. Prospects for
reducing ice adhesion to solve practical problems are discussed and the methods of anti-icing and de-icing
are summarized.
© Koninklijke Brill NV, Leiden, 2012

Keywords
Ice adhesion, interface mechanics, surface energy

1. Introduction

Adhesion science and technology is usually aimed at improving adhesion. With


ice, however, the problem is of an entirely different nature. Ice adhered to vari-
ous surfaces, such as roads, aircraft wings, ship superstructure (Fig. 1), offshore
structures, TV-towers, overhead power line cables, meteorological instruments and
various machinery components, etc. is a severe problem which causes loss of lives
and huge costs in terms of structural damage and accidents.
Accordingly, the aim of this review is to increase the understanding of ice ad-
hesion with the ultimate intent of combating the problems related to it. Theoretical
aspects of ice adhesion are first discussed and some theoretical concepts are intro-
duced. Ice adhesion experiments are then described at a general level and the factors
that affect the comparison of experimental data are analyzed. Finally, the use of the

* Tel.: +358 20 7224914; Fax: +358 20 722 7007; e-mail: lasse.makkonen@vtt.fi

© Koninklijke Brill NV, Leiden, 2012 DOI:10.1163/016942411X574583


414 L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445

Figure 1. An example of the many practical problems related to ice adhesion: Mechanical ice removal
from ship’s superstructure.

Figure 2. Water drop on a solid surface and the definition of the contact angle θ .

theory and the experimental results in developing countermeasures for problems


related to ice adhesion is discussed and future prospects are highlighted.

2. Theory
2.1. The Relation between Water Contact Angle and Ice Adhesion
Consider a drop of water (w) on a solid (s) with an interface (w, s) and the cor-
responding surface energies γ and a droplet contact angle of θ . The situation is
schematically shown in Fig. 2.
The Young equation for the equilibrium of this situation reads
γw,s + γw cos θ = γs . (1)
Consider next ice (i) that is frozen on the solid (s). In Fig. 2 this would mean that
the drop is frozen into ice. We are now interested in the work that is required to
L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445 415

remove the ice, i.e., break the bond (i, s) and form two new surfaces (i and s) in
the absence of deformations. We call this the thermodynamic work of adhesion Wa
which then is defined as

Wa = γs + γi − γi,s . (2)

Inserting γs from equation (1) into equation (2) shows that

Wa = γi + γw cos θ + (γw,s − γi,s ). (3)

Considering now that the surface energies of water and ice are approximately the
same [1] and assuming that their interfacial energies at the solid interface are also
approximately the same, we obtain

Wa ≈ γw (1 + cos θ ). (4)

According to equation (4) the thermodynamic work of ice adhesion can be closely
approximated by the surface tension of water and the contact angle of water on the
material in question. This is presented graphically in Fig. 3.
Equation (4) and Fig. 3 show that, ideally, we should in ice removal expect a
deterministic dependence between the work of adhesion and the contact angle of
water. Finding a deviation from the curve in Fig. 3 in macro-scale experiments
would imply that work is spent in deformations of the materials or that the ice–
solid contact is somehow complex or incomplete. These possibilities are discussed
in the following sections.

Figure 3. Thermodynamic work of ice adhesion scaled by the surface tension of water as a function
of water contact angle θ .
416 L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445

2.2. The Effect of Material Deformations


The work Ws spent in separating an interface into two surfaces by a perpendicular
force F (x) is
 δx
Ws = F (x) dx, (5)
0
where δx is the distance at which the surfaces are considered to be ‘separated’. Here
both Ws and F (x) are defined as per unit area of the original interface. If the sur-
faces were perfectly smooth and rigid, δx would be of the molecular scale, because
the atomic forces decrease very rapidly with distance. For such ideal surfaces the
mechanical work spent would include a reversible component only and equal the
thermodynamic work of adhesion Wa .
However, real material surfaces are deformed upon adhesional failure. The de-
formations may be either brittle or ductile. Ice itself may fail in both modes, and
it is noteworthy that the elastic modulus of ice strongly depends on its temperature
[2]. The distance δx in ice removal may thus be orders of magnitude larger than in
the ideal situation discussed above. Work may also be done in forming non-planar
failure plains and micro-cracks within the ice.
Accordingly, the work spent in mechanically removing ice from a substrate is
typically much more than the thermodynamic work of adhesion Wa even in the
case of a relatively rigid substrate material. As an example, the fracture energy of
an ice–steel interface has been experimentally determined as 1.1 J/m2 [3] while
Wa from equation (4) gives 0.09 J/m2 for the same interface. This shows that the
irreversible contribution to the work spent in removing ice from surfaces due to
dissipative processes can be much larger than the reversible contribution due to
intermolecular interactions across the interface.
However, since the intermolecular interactions are the cause of the dissipative
processes, we may still expect a strong correlation between Wa and the total work
spent in ice removal [4], similarly to other materials [5]. In any case, the correlation
between Ws and Wa is likely to be less significant when, not only the contact angle,
but also the elastic modulus of the substrate material varies.
The work of ice removal Ws , discussed above, is not the usual way to consider the
adhesion strength of ice, however. Rather, one is interested in the adhesion strength
which is defined as the maximum value Fmax of F (x) required to mechanically
separate ice from the substrate. As discussed above and, e.g., in [5] the connection
between Ws and Fmax is not straightforward. In fact, it is not evident why the ma-
terial deformations during ice removal should affect the adhesion strength of ice at
all.
This can be explained by an indirect effect that arises due to the brittle-ductile
nature of ice. When the substrate material is highly elastic and/or the temperature of
ice is close to its melting point, the ice interface behaves as that of a ductile material
with little tendency to fracture even at high strain rates. On the other hand, on a rigid
substrate and/or cold ice, mechanical removal of ice occurs in a brittle manner. Thus
L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445 417

the failure mechanism at the interface may vary with the flexibility of the interface.
On rigid surfaces, this may theoretically cause a temperature dependence of Fmax ,
so that at low ice temperature, a lower ice adhesion strength would occur. This
is indeed observed at very low temperatures, as discussed later in this paper. At
temperatures close to 0°C there appear to be other mechanisms that override the
effect of deformations. These mechanisms affect the ice–substrate interface already
prior to an attempt to remove the ice and are discussed in the following sections.

2.3. The Effect of Interface Morphology and Crystal Structure of Ice

Obviously, the nominal adhesional force per unit area between ice and a substrate
is weaker when the true interface area is less than the apparent one. This is the case
when the contact is incomplete, e.g., due to minute bubbles of air at the interface.
Also, micropores of the substrate that are not filled with water due to hydrophobicity
of the substrate surface reduce the effective contact area. On the other hand, the true
contact area is larger than the apparent one when the interface includes perturba-
tions, i.e., is rough on the microscopic scale. For example, if the interface includes
micropores that are filled with water, then the true interface area is large. In such a
case, sometimes called ‘locking’ a purely adhesional failure may become impossi-
ble, and the fracture occurs partly cohesionally within the ice material. Therefore,
the adhesion strength of ice on many materials is higher on a microscopically rough
surface than on a smooth surface [6, 7].
The true interface area of the contact of water with a textured surface can be
estimated theoretically in detail. Apparently the fine details that affect the contact
angle hysteresis, such as the sharpness and spacing of the roughness elements, are
important [8, 9]. The theory can be used as guide for estimating ice adhesion when
bulk water is frozen on textured surfaces, for example those that have a ‘lotus’ or
fractal structure in order to make them super-hydrophobic [8–13]. However, these
ideas are perhaps over-optimistic because they are based on the mechanical equi-
librium between capillary forces only. In many cases of ice accretion in nature the
ice is formed from water droplets that impact the surface at a very high speed.
This involves high droplet inertia and often a high stress due to wind, so that water
droplets may penetrate and freeze into the porous structure of the material, par-
ticularly when they are smaller in diameter than the roughness element spacing.
Small droplets that are highly supercooled also freeze so quickly that they do not
re-bounce from a superhydrophobic surface.
Consequently, the adhesion strength of bulk-formed ice on textured surfaces may
be different from that of ice formed by droplet accretion and the sizes of the surface
roughness elements and the impinging droplets may affect the latter. This conclu-
sion is supported by the fact that Teflon has been found to be quite poor in reducing
adhesion of ice produced by droplet accretion [14, 15], while the adhesion strength
of bulk-formed ice on Teflon has been shown to be very low in many other studies
(Section 3).
418 L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445

Figure 4. Crystal structure of ice photographed through a polarized plate on two substrate surfaces,
concrete (above) and Teflon (below).

The size and orientation of crystals in bulk ice at an interface depend on the ma-
terial on which the ice is frozen, see Fig. 4. When ice forms due to droplet accretion
the growth conditions have an effect as well [16–19]. Generally, hydrophilic and
rough surfaces cause smaller and more randomly oriented ice crystals to form. It
has been proposed that the crystal structure at the interface would affect ice ad-
hesion. However, this is difficult to demonstrate experimentally, because materials
show different ice adhesion strengths for other reasons. Theoretically, it is not clear
why such an effect should be significant, since the surface energies of the crystal
faces of ice are not much different from each other [20].
L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445 419

2.4. The Effect of Temperature


2.4.1. The Role of Pre-melting of Ice
One would, of course, expect that a liquid film at the solid–ice interface reduces the
adhesion strength. Anticipating that a very thin layer of water may not mechanically
behave as a liquid, one would also expect that the thicker the liquid film, the lower
the adhesion.
In 1859 Faraday [21] showed by experiments on adhesion of small ice pieces
to each other that there exists a liquid layer on ice at typical natural temperatures.
After a long controversy, and more recent direct measurements, it has been agreed
that this is indeed so. Generally, see, e.g., [22, 23], this has been explained by min-
imizing the total surface free energy when a liquid layer forms, i.e., by assuming
that γi > γw + γw,i . In terms of Fig. 2 and equation (1) this requires assuming that
the contact angle of water on ice is zero. It was shown, however, by Makkonen [1]
that the contact angle of water on ice under relevant experimental conditions is al-
most 40°. This finding outlined that the commonly adopted theory on the surface of
ice based on minimizing the surface free energy is without a basis [1, 2]. A more
appropriate explanation for the liquid layer on ice [1, 24] and its consequences are
briefly explained in the following.
Consider first a surface just exposed by separation assuming that its interatomic
distances are still the same as in the bulk material. Then there is a net inward force
due to the imbalance of atomic forces across the interface. This force per unit area
is a pressure which may be interpreted to cause the reconstruction of the surface
into its final equilibrium state. It is reasonable to assume that the structure and in-
teratomic distances of this reconstructed layer correspond to those formed when
applying the same external pressure on a bulk material. For ice, contrary to most
other materials, an excess pressure causes reduction in the equilibrium melting tem-
perature. This causes the fact that a liquid water film exists on an ice surface even
well below 0°C.
Applying the Lennard–Jones potential of the interatomic forces and the Clau-
sius–Clapeyron equation Makkonen [1] estimated the change in the surface equi-
librium melting temperature δT . This result can be simplified into the following
form:
δT = 290|γi,w − γi,s |, (6)
where δT is in °C. Here γi,s is the interface energy of the ice–solid interface and γi,w
that of the ice–water interface, both in J/m2 . The value of γi,w = 29 × 10−3 J/m2
can be used here [25].
As an example of the results of equation (6), the surface pressure is sufficient to
keep the outermost surface layer of an ice–vapour interface (γi = 75 × 10−3 J/m2 ,
[1]) in a liquid state at temperatures above −13°C. However, the use of equation
(6) for ice in contact with solid materials is not straightforward. First, the interface
energy γi,s is unknown. It can be estimated based on γs and γi but only theoreti-
cally. Using, e.g., the Berthelot rule suggests that, qualitatively, γi,s is high for the
420 L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445

materials for which γs is high. Second, it is not clear if this theory applies to the
cases where the surface energy of the solid is higher than that of ice, and if both the
dispersion and polar components need to be taken into account.
Hence, the predictive value of equation (6) is, perhaps, limited to the low energy
surfaces, such as those polymers that may have γi,s close to the value of the ice–
water interface, i.e., 29 × 10−3 J/m2 . At such interfaces, pre-melting should occur at
temperatures close to 0°C only, whereas for the polymers that have an even smaller
surface energy should involve a liquid film in a wider temperature range. This may
introduce another mechanism, in addition to the one presented in Section 2.1, which
produces a low ice adhesion at low energy surfaces.
Based on the same theory, the thickness of the liquid water layer on ice can be
estimated [1]. The theory suggests that the water film thickness at an interface of ice
decreases from infinity at 0°C to zero at 0◦ C − δT , being proportional to δT −1/3 .
Such relationship has been found also experimentally [26, 27]. More studies would
be necessary to evaluate such a nanoscale film thickness at interfaces and its relation
with ice adhesion. In any case, qualitatively, the theory predicts an increase in the
adhesion strength of ice with decreasing temperature down to the temperature at
which the liquid film disappears.
2.4.2. Interface Cracking Due to Thermal Expansion
It is noteworthy that an ice/substrate interface always forms at the freezing temper-
ature of water, i.e., 0°C for pure water. Thus, at the time of freezing there exists
no direct effect of ambient temperature on the properties of the interface. However,
since the initial ice/substrate interface is at the freezing temperature, any such inter-
face at a sub-freezing temperature must have been cooled to that temperature in a
solid state. Thermal contraction of both the ice and the substrate is always involved
in such a cooling. Obviously, this may initiate failures of the joint at the interface
and cracking of ice when the thermal expansion coefficients of the ice and the sub-
strate material are different from each other. In most cases they are quite different
as the linear thermal expansion coefficient of ice is 50 × 10−6 /°C [20], whereas that
of many other materials it is much smaller. For steel, as an example the value is
11 × 10−6 /°C. This means that the stress that arises at the interface upon simulta-
neous cooling of the ice and the substrate is mainly caused by contraction of the
ice itself. The ice and the substrate material may also cool at different rates due to
their different heat capacity and coefficient of thermal conduction, causing an addi-
tional stress at the interface. The significance of the stress depends on the cooling
rate because ice creeps. When the substrate material remains elastic down to the
temperature in question, this effect is absent.
That failure of an ice–substrate interface actually takes place by cooling only
can be demonstrated by numerical analysis. Figure 5 shows a finite element method
analysis of the shear stress caused by thermal contraction in a rapid temperature
change of 10°C in the U.S. Army Cold Regions Research and Engineering Lab-
oratory (CRREL) ice adhesion test arrangement (see Fig. 7). This analysis and
the corresponding ones for the other stress components indicate that the thermally
L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445 421

Figure 5. Finite element method analysis of the shear stress (in Pa) caused by thermal contraction in
a rapid temperature change of 10°C in the CRREL ice adhesion test arrangement (see Fig. 7).

induced stresses at the ice–substrate interface already at −10°C are higher than
the adhesional and cohesional strength of ice. Thus, in this test arrangement the
ice–substrate interface has been partly broken and/or the ice itself is damaged by
cracking already prior to the adhesion test. In experiments on ice adhesion this ef-
fect can be alleviated by letting the samples cool very slowly so that the creep of
ice partly relaxes the thermal stresses that arise.
One would expect that thermally induced failures of the ice and of the joint at the
interface will reduce the measured value of the adhesion strength. In other words,
the adhesion strength of ice should theoretically decrease with decreasing tempera-
ture due to the effect of differential thermal expansion, except for elastic materials.
This is indeed observed (see Figs 11 and 12) but only at rather low temperatures.
Apparently, the disappearance of the liquid-like layer with decreasing temperature
(Section 2.4.1) provides a more important mechanism than the thermal expansion
at temperatures close to the freezing point of water. Moreover, as long as the liquid-
422 L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445

like layer exists the interface may be ductile so that it adjusts to the thermal stresses
without fracturing.
2.5. The Effect of Water Salinity
Impurities and chemicals affect the adhesion of all materials. This general subject
will not be discussed here. However, in the case of ice adhesion, salt in sea water
requires attention due to its major role in many problems related to accretion of
ice at sea [28]. Salinity at the interface always is a factor when the ice is formed
out of sea water. The physical processes that are involved in saline ice adhesion are
discussed in the following and an attempt is made to, at least qualitatively, explain
some of the observed phenomena.
Saline ice consists of salt-free ice and the so-called brine pockets that contain
saline water. The salt concentration and relative volume of these brine pockets de-
pend on the bulk salinity and temperature of the ice [20]. The reduction in the
adhesion strength of ice caused by salt is apparently due to reduced effective solid–
solid contact area at the ice–substrate interface. It is reasonable to assume that the
adhesion strength depends linearly on the effective contact area. The effective con-
tact area can be related to the brine volume assuming some geometrical form of the
brine pockets.
Oksanen [29] used these ideas to model saline ice adhesion by adopting the ge-
ometry of brine pockets as vertically oriented cylinders. He assumed in his model
that the ice–substrate interface structure is equal to that of a vertical cut of the ice.
However, this model predicted much higher values of the adhesion strength than
observed since the portion of the brine at a plane cut area is small. It, thus, became
obvious that a liquid layer of salt solution exists at the ice–substrate interface. In-
deed, a liquid layer has been observed in adhesion tests at high salinities [30] and
its existence even in the case of very low concentrations of potassium chloride has
been confirmed [31]. The formation of this saline water layer is discussed below.
Ice freezes at its equilibrium freezing temperature, so that it must cool in the
form of ice to the temperature of the environment. During cooling, water freezes
on the walls of the brine pockets [20]. This raises the salt concentration of the
brine pockets allowing the phase equilibrium between the ice and the brine to be
maintained.
The ice that forms within the brine pockets occupies 9% greater volume than
the liquid brine. There is no evidence that this would result in significant micro-
fracturing of ice. Therefore, upon cooling, some brine is evidently forced out of the
brine pockets and must eventually be expelled out of the ice sample. As a result
of this brine expulsion, a layer of high salt concentration forms on the ice surface.
When ice is adhered to a structure and cooled, a concentrated salt layer forms at the
ice/substrate interface too, and reduces the adhesion strength.
Details of the brine expulsion process during cooling of ice are poorly known,
but the existing data indicate that the brine moves along the grain boundaries [32].
It is, therefore, likely that brine expulsion from an ice sample is not always three-
L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445 423

dimensional as the orientation of grain boundaries depends on the crystal structure


of the ice. It is also possible that, in contrast to long-term brine drainage of sea
ice, the preferred direction of brine expulsion from an ice sheet is upwards, since
the upward brine permeability of sea ice is higher than the downward permeability
[33]. Whatever the factors that affect the preferred direction of brine expulsion are,
it is apparent that gravity plays no significant role in the process. This is because
the forces related to expansion of the water in the brine pockets are much higher
than those related to gravity.
Following [34], let us consider a simple case of a cube of ice adhered to a plate,
in order to estimate the amount of salt expelled from the ice to the ice/structure
interface. We will assume that the ice cube consists of polycrystalline isotropic
material with random crystal orientation and that the plate adhered to one wall of
the cube does not affect the brine expulsion process. With these assumptions the
amount of salt, ms , expelled to the surface of an ice sample during cooling from
temperature T1 to temperature T2 is

ms (T2 ) = Mi (Si (Ti ) − Si (T2 )), (7)

where Si is the salinity and Mi the mass of the ice sample. Equation (7) can be
presented as
 
Si (T2 )
ms (T2 ) = Mi Si (T1 ) − Si (T1 ) . (8)
Si (T1 )
Cox and Weeks [35] derived an equation for the salinity ratio Si (T2 )/Si (T1 ). When
using Zubov’s [36] relationship between brine salinity and brine density (ρb in
g/cm3 ),
ρb = 1 + 0.8Sb , (9)

the equation by Cox and Weeks [35] reads


   
Si (T2 ) Sb (T2 ) (1−1/ρpi ) 1 + 0.8Sb (T2 ) 0.8
= exp (Sb (T1 ) − Sb (T2 )) . (10)
Si (T1 ) Sb (T1 ) 1 + 0.8Sb (T1 ) ρpi

Here ρpi is the density of pure ice in g/cm3 .


The salinity of brine, Sb , is defined as the ratio of the mass of salt, ms , to the
mass of brine, mb , so that
ms (T2 )
mb (T2 ) = . (11)
Sb (T2 )
When this amount of brine mb (T2 ) forms a liquid layer on the ice surface, its thick-
ness h will be
1 mb
h= , (12)
ρb Ai
424 L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445

where Ai is the surface area. For a cube of ice with density of ρi and wall length of
L the surface area, Ai , is 6L2 and the mass, Mi , is ρi L3 . It, therefore, follows from
equations (10), (11) and (12) that for a cube of ice with a density of 0.9 g/cm3
 
0.9L3 Si (T1 ) Si (T2 )
h(T2 ) = 1− . (13)
6L2 (1 + 0.8Sb (T2 ))Sb (T2 ) Si (T1 )
Inserting equation (10) into equation (13) and using the value of 0.917 g/cm3 for
ρpi gives an equation for the thickness of the liquid film on the ice surface at T2 :
0.9LSi (T1 )
h(T2 ) =
6(1 + 0.8Sb (T2 )Sb (T2 ))
  
Sb (T2 ) −0.0905 1 + 0.85Sb (T2 )
× 1− (14)
Sb (T1 ) 1 + 0.85Sb (T1 )

 
× exp 0.872(Sb (T1 ) − Sb (T2 )) .

The salinity Sb of brine in ice is an explicit function of temperature which can be


calculated [37] by
Sb (T ) = −3.9921 − 22.700T − 1.0015T 2 − 0.019956T 3 . (15)
When the wall length L and salinity Si (T1 ) of the original ice are known, equations
(14) and (15) give the thickness, h, of the liquid film expelled when the ice is cooled
from T1 to T2 . When we consider h as it affects ice adhesion at a temperature T2 , we
are interested in total brine expulsion that has taken place before the temperature T2
has been reached. Therefore, the temperature T1 that we need to know to calculate
Sb (T1 ) is the temperature from which cooling and brine expulsion have started, i.e.,
the temperature at the ice/water interface during freezing, Tf .
The temperature Tf is basically unknown. This is because salt rejection from
ice during freezing increases the salinity at the ice/water interface. The salinity and
temperature of the interface layer depend on the growth rate of ice and effectiveness
of mixing in the water [19, 38]. These factors are generally unknown. However, the
problem can be solved by determining Sb (Tf ) from the definition of the interfacial
distribution coefficient k ∗ :
Si
k∗ = . (16)
Sb
The experimental data [35, 38, 39] show that at typical growth rates of natural sea
ice k ∗ is approximately a constant at k ∗ = 0.26 regardless of the growth conditions
and water salinity. It has been pointed out by Makkonen [19] that the value of
k ∗ = 0.26 is a reasonable approximation also in the case of icing due to sea spray
droplets. A later analytical solution shows a value of 0.30 in close agreement [40].
Consequently Sb (Tf ) can be approximated by
S1 (T1 )
Sb (Tf ) = (17)
0.26
L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445 425

and the final equation for the thickness of the liquid film on the ice surface at a
temperature T is
0.15
h(T ) = LSi
(1 + 0.8Sb (T ))Sb (T )
  
0.26Sb (T ) −0.0905 1 + 0.8Sb (T )
× 1− (18)
Si 1 + 3.077Si

× exp(3.354Si − 0.872Sb (T )) .

Here Si is the initial salinity of the ice cube and Sb (T ) is solved from equation
(15). An iterative solution for h(T ) as a function of the ultimate salinity could also
be obtained but this would not be justified because equation (17) is only an ap-
proximation. For practical purposes Si can be replaced by the measured ice salinity
in equation (18), as the difference between Si (T1 ) and Si (T2 ) is always small and
equation (18) is insensitive to these small differences.
Examples of the brine layer thickness h calculated at various temperatures and
ice salinities are given in Fig. 6 for an ice cube with L = 10 cm. According to Fig. 6,
the brine layer thickness starts to grow at a certain temperature (e.g., −4.2°C for

Figure 6. Theoretical brine layer thickness on the surface of a three-dimensionally cooled isotropic
10−3 m3 ice cube as a function of its eventual temperature. The curves are for different ice salinities.
426 L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445

Si = 20h). This is because there is no ice formation at temperatures higher than


this. Equation (17) determines the brine salinity, Sb (Tf ), and, together with equation
(15), gives an equilibrium freezing temperature Tf , which is an explicit function
of Si .
At temperatures lower than the equilibrium freezing temperature the brine layer
thickness grows rapidly with decreasing temperature, as liquid water within the ice
matrix freezes and brine is expelled from the ice cube. On the other hand, when
temperatures drops, the brine layer formed on the surface partly freezes, and this
reduces its thickness. Therefore, h reaches a maximum at a certain temperature
(e.g., −6°C for Si = 10h in Fig. 6) and then starts to decrease. It is interesting to
note that according to the results in Fig. 6 the maximum brine layer thickness is
almost independent of ice salinity.
As discussed earlier in this paper, the adhesion strength is probably in some way
related to the calculated liquid layer thickness, which in the case of saline ice would
be the brine layer thickness h. What do the results of equation (18) in Fig. 6 mean in
terms of ice adhesion? First one may note that the values of h calculated by equation
(18) are typically of the order of 102 µm, whereas the liquid-like layer thickness of
fresh water ice is estimated to be of the order of 10−2 µm only [41]. This shows
theoretically that the adhesion strength of saline ice is much smaller than that of
fresh water ice down to the nucleation temperature of the salt.
What actually happens to the ice–substrate bond when brine is expelled to the
interface is unknown. It is also uncertain whether a continuous liquid film forms
at the interface. Instead, a layer of mixed ice and brine may exist at the interface,
allowing some ice/surface contacts to remain. Brine absorption by the structure sur-
face because of its pores and roughness elements may also affect the phenomenon,
although one should note that the calculated values of h are much higher than the
height of the roughness elements on typical plastic, metal and painted surfaces. Fi-
nally, it is possible that some brine is removed from the surface by liquid flow along
the ice/substrate interface.
The equal theoretical maximum values of h at different salinities in Fig. 6 indi-
cate that, if there is weak adhesion at some ice salinity, an equally weak adhesion
will be found for any ice salinity at some temperature. This finding could be useful
in mechanical de-icing, for example. When, say, a ship’s superstructure has been
covered by saline spray ice and the temperature is low, one may minimize the effort
of ice removal by waiting for the ice to cool to the optimum temperature predicted
by Fig. 6.
The results in Fig. 6 suggest that, at a given temperature, ice adhesion should
depend on the salinity. This has been observed in experimental studies at low salin-
ity [30, 42–46] but not in the range of Sw = 5–60h [30, 46], i.e., the reduction of
the adhesion of saline ice from the value of fresh water ice occurs already at a very
small water salinity. This suggests that ice adhesion may be insensitive to variations
in the brine layer thickness when h is large. However, there may be a critical h, at
which the continuous liquid film disappears and the adhesion strength considerably
L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445 427

increases. If such a critical value exists, it probably depends on the roughness of


the substrate surface. When h is larger than such a critical value, one should expect
saline ice adhesion to be almost independent of temperature.
The theory presented above provides a framework for further studies on the ad-
hesion of saline ice. The major problem which should be studied further is the basic
relationship between the liquid layer thickness and the ice adhesion strength. Be-
fore this relationship is determined we can only say that the thicker the brine layer,
the lower the ice adhesion strength.

3. Adhesion Testing and Experimental Relationships


The theory of ice adhesion outlined in Section 2 above underlines that there are
many complex factors involved. Thus, predicting ice adhesion for the design of
countermeasures is also complex. Moreover, all aspects of the theory have not, so
far, been quantified sufficiently to allow purely theoretical estimates of ice adhesion.
Consequently, laboratory and field experiments are still essential in developing and
testing countermeasures for ice adhesion [13, 47–53].
The quantitative values of the measured ice adhesion strength vary from almost
zero to about 1 MPa for the different coatings and test conditions. It is noteworthy
that the highest values in this range are approximately the same as the shear strength
of ice. Indeed, in such tests, traces of ice remaining at the substrate surface after the
test are observed. This shows that the separation occurred at least partly within the
ice material as a cohesional failure, rather than at the ice–substrate interface. Tests
in the cohesion mode are, of course, useless in evaluating the true adhesion strength
of ice with a material as they only measure the strength of the ice material.
There are considerable difficulties in making comparative adhesion strength
measurements even in laboratory conditions because so many factors affect the re-
sults. In addition to such factors as thermal expansion of ice, the variables which
affect adhesion tests on all materials must be considered. These include the load-
ing arrangement, i.e., the stress distribution at the interface during a test, and the
loading rate.
Moreover, in the case of ice, the environmental conditions during and after ice
accretion as well as the time of contact affect the results. The rate at which the
samples are cooled to the test temperature is important in affecting the rigidity of the
interface due to thermal expansion, as discussed in Section 2.4.2. Furthermore, there
is a stochastic effect caused by the brittle nature of ice fracture at low temperatures.
The ultimate failure may be related to randomly spaced dislocations at which the
initial fracturing process at the interface may start. This may also cause a sample
size effect in ice adhesion tests. In case of saline ice, the size effect is unavoidable,
as shown in Section 2.5.
To obtain comparative results it is, therefore, important to remove as many vari-
ables from the test procedure as possible. Several test apparatuses for ice adhesion
tests have been developed with this in mind in various research institutes [30, 47,
428 L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445

54–71]. The principles of different methods have been reviewed by Sayward [72].
Many techniques exist and the size of the ice samples varies from nano-size [66,
67] to microscopic [68] to those faced in large scale applications. Test methods that
are specific to some geometry, such as the leading edge of an airfoil, have also been
developed [73].
Nano-scale experiments on ice adhesion are appealing in that many of the vari-
ables that affect macro-scale ice failure are absent. However, the true contact area
is very difficult to determine in such experiments, and they do not necessarily relate
to the practical situations of ice removal. Therefore, macro-scale ice adhesion mea-
surement methods will be discussed in the following. As an example, the U.S. Army
Cold Regions Science and Engineering Laboratory (CRREL) test arrangement is
shown in Fig. 7. As another example, the ice adhesion test method presently used
at VTT Technical Research Centre of Finland is described in more detail below.

Figure 7. CRREL cylindrical ice adhesion test arrangement (www.crrel.usace.army.mil). The joint
between the coated metal pile and ice is broken by torque.
L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445 429

Figure 8. Ice adhesion test apparatus of VTT and coating samples with ice cylinders frozen on them.
Removed ice pieces are also shown.

In the VTT test the test specimens are 100 mm × 100 mm × 10 mm aluminium
plates on which the coatings have been applied. Ice cylinders with a diameter of
30 mm are frozen on the coating at −10°C by placing a small container filled with
water on the test plate. During freezing the samples are thermally insulated on the
side on which the ice sample is frozen on the plate. Consequently, the latent heat
of freezing is mainly removed by conduction through the aluminium plate. The
purpose of this is to make the freezing front advance from the coating interface
towards the bulk water. This assists in obtaining ice with little or no air bubbles
close to the coating interface.
The ice cylinders are tested by shearing them from the plates after at least 24 h
of storage in a cold room kept at the test temperature. The test device and some
samples are shown in Fig. 8. In the tests, failure at the interface is obtained by a
shear force applied by a belt moving at a constant nominal rate of 3.2 × 10−4 m/s,
which is also measured by a laser device. The resulting force is measured by a
load cell and the force–time curve is recorded. Examples of force–time curves are
shown in Fig. 9. Due to stretching of the belt, used in delivering the force to the ice
cylinder, the increase of the force is somewhat nonlinear with time. The adhesion
strength is determined as the peak shear force divided by the interface area. For
each coating several ice samples are tested and the adhesion strength is calculated
as the mean of the results.
430 L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445

Figure 9. Example of four measurements on the same coating sample in VTT experiments. The ad-
hesion strength in terms of force is based on the mean of the four peak values. Av. force = 441 N,
pressure = 0.49 MPa.

One may evaluate the work done in an ice adhesion test of Fig. 9 by integrating
the force–time curve with respect to the displacement, as the latter is known from
the constant deflection rate applied. Such a calculation gives about 40 J/m2 for the
total work, which can be compared to typical values of about 1 J/m2 for the fracture
energy and about 0.1 J/m2 for the thermodynamic work of adhesion (Section 2.2).
Clearly, the major part of the work spent goes for the irreversible deformations in
the measurement system. This outlines that ice adhesion measurements of this type
provide the adhesion strength only and are inapplicable to determining a relevant
value for the work of adhesion.
A fundamental aspect in measuring ice adhesion is the stress distribution at the
ice–substrate interface upon loading. Ideally, the shear stress should be evenly dis-
tributed on the interface. This situation can be approached by measuring tensile
strength instead of shear strength [57, 61], by a centrifuge adhesion test [63] and by
a test method utilizing laser-pulse induced spallation in which a compressive stress
pulse travels through a substrate disk which has an ice layer grown on its front
surface [69]. However, one needs to consider not only the easiness of the theoreti-
cal interpretation of the ice adhesion measurements, but also their applicability: In
most applications ice removal involves an uneven distribution of stress at the inter-
face. In some applications ice removal in practice is done from curved surfaces and,
accordingly, such laboratory experiments have also been done [73].
In any case, it is essential to understand how the stress in distributed in different
test arrangements. The stress distribution in the basic shear ice adhesion test by
VTT, as calculated by the finite-element method, is shown in Fig. 10.
L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445 431

Figure 10. Shear stress (in Pa) in the VTT test (Fig. 8) as calculated by FEM analysis. The substrate is
at the bottom of the ice sample and the pull is from right to left. The indentation is the ice deformation
by the belt (dimensions exaggerated by a factor of 103 ).

In the VTT test the relative standard deviation of the measured ice adhesion
strength is on average 12% showing that the method provides rather repeatable
results on ice adhesion measurements. A large set of ice adhesion tests on various
materials at different temperatures were made at VTT in the 1980’s using a similar
test procedure but on a larger scale [30, 49]. In those tests the relative standard
deviation was more than in the new method suggesting that it is better to use small
samples in the testing.
Some results of the previous tests at VTT are shown in Figs 11 and 12. The
results in Fig. 11 demonstrate the theoretical predictions (Section 2.4) that the adhe-
sion strength has a maximum at some intermediate temperature. ‘Inerta 400’ which
was manufactured at the time by Teknos Oy is an exception, because it was the
only coating which remained elastic at all test temperatures, thus preventing the
brittle fracturing of the ice due to thermal contraction when cooling down to low
temperatures (see Section 2.4.2).
The results in Fig. 11 and in many other studies referenced in this paper confirm
the theoretical prediction that the materials with a large water contact angle and the
usually related low surface energy tend to have low adhesion strength with ice. In
particular, equation (4) has been verified in that it describes not only the adhesion
energy, but also the adhesion strength on polymers [71]. Experiments on nanoscale
[68] have also confirmed this prediction. There are some experimental results in
which this relationship has been less clear [70], but this is probably due to other
complicating factors causing scatter in the measurement results.
It is noteworthy that polymer surfaces have a high linear thermal expansion co-
efficient: For example the value for poly(vinylchloride) is about the same as for ice.
Following the ideas proposed in Section 2.4.2, this should produce only a small
temperature dependence at low temperatures. This is indeed observed experimen-
432 L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445

Figure 11. Effect of temperature on the ice adhesion strength in shear. Inerta 400 manufactured by
Teknos Oy was a coating that remained elastic at cold temperatures. Vellox 140 was an anti-icing
coating supplied by Clifford W. Estes Co. (USA).

tally (Fig. 11) suggesting that the ice adhesion strength of polymers is affected not
only by the surface energy but also the thermal expansion coefficient.
Figure 12 shows the temperature dependence of the ice adhesion strength for
one substrate material in more detail and also in the case of ice frozen from saline
water. It can be seen that adhesion strength in the case of saline water does not
have a maximum at an intermediate temperature but increases indefinitely. This is
in accordance with the theoretical predictions discussed in Section 2. First, the salt
causes a liquid film that makes the adhesion strength fall to a small fraction of its
value for pure water.
Second, this film makes the interface ductile enough, so that thermal expansion
when going to low temperatures does not cause a reduction in the adhesion strength.
That the adhesion strength of saline ice exceeds that of pure ice at very low tem-
L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445 433

Figure 12. Adhesion strength of ice as a function of temperature for ice formed from fresh water
and water with a 2h NaCl concentration. The substrate material Inerta 160 is an epoxy-based paint
supplied by Teknos Oy.

peratures is because salt eventually nucleates, so that the liquid film completely
disappears at the lowest temperatures.
Adhesion strength of saline ice is typically only 10–20 kPa, except at extremely
low salt concentrations and low temperatures as shown in Fig. 12. However, some-
times an order of magnitude higher values have been reported [45, 74]. The reason
for this discrepancy is probably related to the formation and disappearance of the
liquid layer at the interface. Different methods of freezing the samples, different test
arrangements and varying storage times may all affect the thickness of the liquid
film. Preferred direction of brine expulsion may also depend on crystal orientation
and brine drainage may remove the liquid film. Furthermore, brine drainage may
affect the adhesion strength more on horizontal than on vertical surfaces [44].
The adhesion strength drops to only a fraction of its value for fresh water ice
already at ice salinities below 1h. Also, it has been shown [54, 66, 75] that the
adhesion strength suddenly drops to a very small value when the temperature rises
above the eutectic point of the salt solution, and that this happens also with very low
salt concentrations. This is consistent with Fig. 12 in that the adhesion strength of
saline ice was observed to be considerably higher at −30 to −50°C than at higher
temperatures (the eutectic temperature of NaCl is −23°C).
It was found by Makkonen and Lehmus [30] that the ratio of saline ice adhesion
to fresh water ice adhesion was much higher on a rough polymer surface and on
434 L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445

Figure 13. Field testing of ice adhesion on a navigation lock wall. Samples are made by a bore and
then sheared off by a belt seen in the right hand upper corner.

a concrete surface when compared with other surfaces. It is noteworthy that these
surfaces were much more porous than the other tested surfaces. This suggests that
the liquid film is absorbed by a porous surface, thereby considerably increasing the
adhesion strength. Therefore, roughness of the surface plays a much more important
role in adhesion of saline ice than of fresh water ice.
Measurements of ice adhesion strength have been made not only in the labo-
ratory, but also in the field for naturally formed ice. This has been necessary for
investigating the feasibility of coatings and de-icing systems in practical applica-
tions. Measurements with different coating materials have been made, for example,
onboard ships and offshore structures [76–78] and on navigation lock walls [79, 80]
(Fig. 13).
L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445 435

4. Combatting Ice Adhesion


4.1. Anti-icing Coatings
When developing synthetic coatings having low adhesion to ice, researchers have
looked for relationships between adhesion strength and other material properties
that are easy to measure. In accordance with the theoretical predictions discussed in
Section 2, it has been observed experimentally that low adhesion to ice is generally
connected with low permeability and absorption capacity and high hydrophobicity,
i.e., a high contact angle with water [58–60, 81–84].
Materials that meet these requirements, such as synthetic polymer surfaces, have
proved to be effective in reducing ice adhesion. The adhesion strength with polymer
coatings is an order of magnitude less than on uncoated surfaces. In full-scale tests
on a ship the best coatings were Teflon-4, organosilicone epoxy ‘G’, and a vinyl
polymer sheet with perfluorinated film [43]. Laboratory tests with ice accreted in
a wind tunnel [13] showed that the durability and effectiveness of silicone rubber
made it a promising surface coating even in the case of ice accreted by droplets. Re-
ports of the effect of a poly-bisphenol-A-polycarbonate block copolymer have also
been encouraging [39, 48]. Crouch and Hartley [51] found that a mixture of water-
soluble vinyl pyrrolidone resin and poly(ethylene glycol) in a room temperature
curing silicone gave lowest adhesion in impact tests. A new silicone-based coating
R-2180 supplied by NuSil Technology (USA) has a significantly lower adhesion
strength with ice than Teflon [85, 86].
There are many commercial coatings especially designed for low ice adhesion
available today, see, e.g., www.crrel.usace.army.mil and [63, 85, 87–89]. The most
effective commercial materials are grease-type coatings, which, of course, have lim-
itations in applications due to their low durability. In regard to their usefulness, a
relevant question is: How low should the ice adhesion strength of an effective coat-
ing be?
The best hydrophobic rigid materials existing today provide ice adhesion
strength of the order of 100 kPa [49, 63, 85]. Let us consider a vertical coated
plate and an ice layer with a thickness L and surface area A frozen on it. On such
an ice cover gravity will impose a force gρAL. This divided by the surface area
gives the critical adhesion strength Pc for the ice cover to be released by gravity
alone as
Pc = gρL. (19)
Supposing in equation (19) an ice density ρ of 0.9 g/cm3 and thickness L of ice,
say, 5 cm, gives the critical adhesion strength Pc of 0.44 kPa, which is more than
two orders of magnitude smaller than that provided by typical polymer surfaces.
In other words, only an ice layer that is more than 10 m thick will be released
spontaneously due to gravity for such a surface. In many applications other forces,
such as centrifugal force, winds shear and vibrations contribute to the forces at
the ice–substrate interface. Nevertheless, it is clear from this simple calculation
that none of the hydrophobic solid materials existing today provides an adhesion
436 L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445

strength low enough to solve the practical problems of ice adhesion. Furthermore,
the effectiveness of superhydrophobic nano-rough surfaces is hampered by impact
of small cloud droplets and formation of frost within the texture [90–92].
Therefore, new ideas for anti-icing coating are required and have indeed been
recently introduced as commercial products. As it is difficult to remove ice when
it has frozen on a substrate, an obvious solution would seem to be preventing the
freezing of water on a substrate surface. An obvious approach is to make the sur-
face so hydrophobic that water drops will not remain on the surface for a sufficient
time for nucleation to occur, see [9]. The other approaches are more fundamental
in that they attempt to reduce the adhesion strength of ice by surfaces that prevent
nucleation or freezing even in long-term contact. It is well known that the struc-
ture of water is modified in contact with a surface and that lowering of the freezing
temperature of water occurs at some surfaces, for example to the benefit of some
biological species [93–95]. However, theoretically, the matter is less than clear [96,
97] and the development of such new surfaces has been done mainly by experimen-
tal searching.
One idea worth developing is based on extremely low ice adhesion (less than
10 kPa) measured for an organopolysiloxane resin mixed with alkali metal com-
pound manufactured in Japan in the 1980’ [32, 65, 98]. This coating material was
manufactured by Kansai Paint Inc., but is no more on the production line. Its me-
chanical durability was rather low but it appeared not to be self-sacrificing. This
hydrophobic organopolysiloxane resin contained a small amount of polar ingredi-
ents. Lithium ions act as hydrogen bond breakers in the material. It was assumed
in [65] that the four oxygen atoms from water molecules and two oxygen atoms
present in a carboxylic residue associate together and form an octahedron coor-
dination structure. DSC analysis showed that this water does not have a definite
freezing point [65]. It is assumed by the developers of this coating that “the struc-
tural and energetic differences thereof and synergetic effect from silicone matrix
result in preventing ice adhesion”.
A second idea pursued recently at VTT and elsewhere [95, 99] is to synthesize
natural anti-freeze proteins, apply them in coatings, and rely on their ability to re-
duce the freezing point and thus affect ice adhesion. Preliminary experiments on
prototype coatings have, however, shown that once freezing occurs on these ma-
terials, the adhesion strength is high. It appears, therefore, unlikely that anti-freeze
proteins will provide solutions to practical problem of ice adhesion generally. How-
ever, they may be useful in systems that have a very small size and an undisturbed
environment. They may also have an indirect effect in reducing icing by delaying
the freezing of water.
Yet another approach is the ePaint coating which reduces the adhesion strength
of ice using several processes [86]. This coating is hydrophobic but also includes
phase-change material that is thermally activated. As the coating cools below 0°C
the epoxy-like material contracts, and the embedded solid-phase change material
expands, causing little net change in the surface area of the coating. However, as ice
L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445 437

accretes, liberated latent heat from the ice warms the coating surface. This causes
the phase-change material to warm and to expand. The simultaneous contraction of
the epoxy-like material and expansion of the phase-change material causes shear
stress within the coating and failure of the ice–substrate adhesion bond (see Sec-
tion 2.4.2).

4.2. Mechanical Methods

As discussed above, the use of the present low-adhesion coatings is not very effec-
tive when used alone. However, combined with mechanical and thermal methods,
coatings and paints can make deicing easier. For example, the time needed for de-
icing a surface by heating is reduced considerably when the surface is coated with
a suitable material [100, 101]. Also, the ease of manual and automatic mechanical
ice removal is very sensitive to the strength of adhesion of the ice to the material.
As an example, polymer coatings have proved to be useful in routine service when
deicing, e.g., navigation lock walls [79].
Manual deicing has historically been the only method in combating ice, for ex-
ample at sea. The ice is removed by crew members using mallets, axes, baseball
bats, etc. (see Fig. 1). This method is unsatisfactory because conditions on a slippery
deck during severe icing are hazardous, and de-icing actions are almost impossible
when most needed. The use of motorized cutters for removing ice on a ship [102]
is usually possible only after an icing storm. It is seldom possible at any time to
manually remove ice from the upper parts of the structures most critical to ship’s
stability or from other types of tall structures on ground.
Because of these problems in manual ice removal, automatic and semiautomatic
deicing methods have been developed. The most effective of these devices is a pneu-
matic deicer, a series of tubes standing alone or built into a rubber mat called a boot.
When the boot is inflated with air it expands and breaks the ice adhered to the sur-
face. The principle is in use for protecting aircraft wings from icing [103, 104].
Pneumatic deicers have proven to be effective also on small cylindrical objects and
large flat surfaces, such as radar dishes and ship superstructure [105–107]. Disad-
vantages of this method are the cost and likelihood of damage to the deicers if used
in working areas.
More recent developments are the electro-expulsive (EESS) and electro-impul-
sive (EEIS) separation systems tested in, e.g., shipboard applications and aircraft
[81, 98, 99]. These methods utilize high repulsive forces to impart expulsive move-
ments to the flexible outer layer. The forces are generated by overlapping conduc-
tive ribbons that receive a very high instantaneous current pulse. A lightweight
retrofitable EESS system consists of a 0.5 mm thick polyurethane blanket with
embedded flexible conducting copper ribbons which are paired and separated by
a dielectric. The instantaneous power pulse is very high, but the pulse width is only
about 20 µs. Thus the energy consumption of the EESS is quite low. Test results and
shipboard applications of the EESS system are described by Embry et al. [109].
438 L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445

Economical mechanical deicing can be achieved by using a flexible coating that


moves due to wind drag [78, 110, 111]. The method is not quite satisfactory be-
cause accreting ice makes the coating partly inflexible. To avoid this Alexeiev [112]
covered small-scale models of meteorological masts with flexible coatings, fixing
the coatings with the aid of guy ropes. Vibration of the mast and the guy ropes
then keeps the coating moving sufficiently for ice removal. Also, a large number
of small, conical plastic shelters that move in the wind have been tested with some
success in small-scale tests.
Flexible surface-coating materials whose deicing effect is not based on low ad-
hesion strength but on the ease of ice removal on impact have also been tested on
ships, see, e.g., [102]. The surfaces tested were plastic foam mats, either alone or
covered with a sheet of neoprene rubber. This type of coating is not very successful,
although it makes the ice somewhat easier to remove manually. However, ice has
spontaneously detached from a rubber mat on the outer side of a ship’s bulwark.
High costs, relatively poor durability and problems in attaching the rubber mats
firmly are the main reasons for the limited use of the flexible coatings.
Mechanical deicer based on metal plates and wires moving by means of elec-
tromagnetic induction, activated by discharge from a condenser through a solenoid
situated near the surface of the plates, has been used in protecting meteorological
instruments [110–112] and various other structures [100]. Robotics has also been
developed and demonstrated, e.g., in mechanical de-icing of overhead power line
cables [101, 113]. A piezoelectric de-icing system has also been developed [86].
4.3. Thermal Methods
The most obvious method for preventing icing is heating. However, it is far from
being the most practical because of the large amount of latent heat required to melt
ice or to prevent its formation. For example, the power required for anti-icing, when
the icing rate is 30 kg/(m2 h), is about 2 kW/m2 . For example, anti-icing a cup
anemometer in the field externally typically requires 300–700 W [114]. De-icing
by heating typically requires about 300 kJ/m2 of energy [112, 114].
Large energy requirements of thermal methods have restricted their use mostly
to small objects, although infrared heaters have been used even in the case of an
aircraft [81]. Internal heating is much more effective than external heating unless
the ice layer is transparent and thin [111]. For this reason ultrasonic heating films
have been developed [115, 116]. Another significant problem with heating as a de-
icing or anti-icing method is that the melted water will freeze at some other location
or form icicles. Thus, water drainage must be designed carefully if this method is
to be applied.
One way of heating structures is to use a thermosiphon. This method can be
applied to structures such as masts and handrails, the structures themselves acting
as heat pipes. It is also possible to cover flat surfaces with loops of heat pipes.
These techniques have been applied when constructing a practical thermostatically
controlled thermal deicing system on a ship [117], which can use different heat
L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445 439

sources. The heat consumption of the system has been found to be about 1 kW/m2
in field tests under typical icing conditions. The heat pipe method has also been
used in deicing navigation buoys [118].
Hot water is rather effective in short-term ice prevention and is sometimes used
as a deicing method on ships [119]. A disadvantage of this method, in addition to
high heat consumption, is that water for protecting the upper parts of the structures
falls and flows along the surfaces and may increase icing at lower levels of the
vessel. Thermal and mechanical methods are combined in the so-called seawater
lance, which consists of a high-pressure jet of seawater capable of removing ice by
melting and dynamic pressure. This method has been used on navy vessels [120].
Current-conducting coatings are used for heating small surface areas, such as
automobile and aircraft windows [86]. This method works for larger surfaces too,
when combined with coatings that reduce ice adhesion. Such an approach reduces
the power consumption to a level realistic for some practical applications, such
as wind turbine blades. Detailed analysis of the heat consumption and optimized
design of a wind turbine anti-icing system has been made by Makkonen et al. [116].
The high energy consumption of thermal de-icing is not only due to melting a
sufficiently thick ice layer but due to high loss of heat as conduction to ice and the
substrate while the system is on. This is especially so when heating is applied from
the air and not from the substrate [116]. It is easy to show [117] that the conduction
loss is smaller the quicker the process, i.e., the higher the power used in the de-icing.
Based on this, methods have been developed in which a very high power is applied
directly at the interface by a heated film [121, 122]. This approach is applicable
to many problems, see http://engineering.dartmouth.edu/thayer/research/ice-engg.
html.
In some applications, even a very small amount of rough ice may cause serious
problems. These are related to aircraft wings, helicopter rotors and wind turbines
where aerodynamic penalties due to accreted ice are unacceptable. In these ap-
plications de-icing may not be enough to prevent problems; instead, anti-icing is
required. Such anti-icing systems are presently based on heating, e.g., [116].
4.4. Chemical and Electrical Methods
The application of chemicals on an icing surface to reduce ice adhesion has been
tested, mostly with limited success [102]. Moreover, these chemicals deteriorate
easily by weathering and by the accreting ice [123].
Another kind of chemical measure for preventing ice adhesion is to apply freez-
ing point depressants on surfaces. The major problems of this method are optimiz-
ing the amount of the chemicals and distributing them uniformly on the surface.
When these problems can be overcome, it is possible to reduce icing considerably
by using organic anti-icing fluids, e.g., ethylene glycol [103, 124] or urea [125].
Salts such as calcium nitrate [126] have also been tested and, of course, various
anti-freeze liquids are widely used in removing ice from road surfaces and aircraft.
A detailed review on anti-icing chemicals is available in [86].
440 L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445

Chemical methods can be appropriate for protecting small objects (windshields,


bridge windows, automatic meteorological instruments, etc.) and objects that re-
quire ice removal for only a short time (aircraft before take-off, for example).
Chemicals are difficult to apply on large structures for long-term ice prevention
since their rapid deterioration makes the method uncertain and expensive. Other
disadvantages include making horizontal surfaces slippery and contaminating the
environment. A coating that ejects the encapsulated anti-icing chemical onto the
surface would be one possible solution [127].
Phan and Laforte [128] showed that the adhesion strength of ice accreted on
wires from cloud droplets depends on the applied DC negative electric field. The
effect of applying an electric field when freezing bulk water has also been studied
[129], resulting in the conclusion that the observed small effect is due to the effects
of electrolysis and corrosion on metal surfaces. In some applications electrical phe-
nomena might be of benefit in anti-icing [130, 131], but no practical tests have been
made. An electrostatic model has been proposed for the basic atomic attraction in
ice adhesion [130] and this theory may provide ingredients for further development
[132, 133].

5. Summary
The review and theoretical analysis presented above points out the complexity of
ice adhesion. Many mechanisms related to ice adhesion are still poorly understood
and most practical problems caused by ice adhesion are far from being solved, see,
e.g., [134, 135].
The adhesion strength of ice is related not only to the chemical composition,
surface morphology, stiffness and thermal expansion coefficient of the substrate
material, but critically depends on the temperature and the test arrangement as well.
Further studies on these effects are necessary in order to obtain comprehensive
understanding of ice adhesion.
When salt is involved, as in marine icing and road maintenance, the adhesion
strength of ice depends on even more factors. When the liquid film that forms at the
ice–substrate interface is retained, then the adhesion strength of saline ice is very
small at any salinity. Therefore, the concepts presented in this paper regarding the
formation of the liquid film due to brine expulsion should be tested and developed
further. For the same reason, the mechanisms of brine drainage and movement of
brine along the ice/structure interface should be studied.
Generally, the higher the water contact angle of the material the lower the ice
adhesion strength. However, it seems that the routes to find practical solutions for
many problems caused by ice adhesion using hydrophobic surfaces have been fully
explored and will not result in further significant development. There are notable
exceptions from this general rule suggesting that it may be more fruitful to investi-
gate the materials that would prevent freezing on a surface, or at least make the true
contact area smaller. The organosilicone material with lithium ions and antifreeze
L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445 441

proteins synthesized from insects may provide such solutions. Further fundamental
research on the freezing and properties of water at interfaces is required for devel-
opment in this field.
The work done in mechanically removing ice is typically orders of magnitude
larger than the fracture energy and the thermodynamic work of ice adhesion. This
suggests that there may be room for developing better mechanical de-icing methods
by applying shockwaves or other very high energy pulses with a short duration.
New technologies for manufacturing heated films for thermal anti-icing and de-
icing provide solutions in many applications. That these films now allow heating
at a very high power makes the costs of combating ice more reasonable. However,
energy consuming ice combating methods are expensive to use unless their opera-
tion is optimized by detecting the ice on the surfaces. Therefore, attention should
be paid also to developing reliable ice detection methods.

Acknowledgements
Thank are due to Eila Lehmus, Pieti Marjavaara, Matti Halonen and Erkki Järvi-
nen for assistance in the VTT ice adhesion projects and Kari Kolari for providing
FEM simulations. This work was funded by the Academy of Finland and by the
Nordic Council Top-level Research Initiative project TopNano, as well as the Tekes-
Rescoat project.

References
1. L. Makkonen, J. Phys. Chem. B 101, 6196–6200 (1997).
2. V. F. Petrenko and R. W. Whitworth, Physics of Ice. Oxford University Press, Oxford, U.K.
(1999).
3. Y. Wei, R. M. Adamson and J. P. Dempsey, J. Mater. Sci. 31, 943–947 (1996).
4. K. R. Jiang and L. S. Penn, J. Adhesion 32, 217–226 (1990).
5. K. L. Mittal, Polym. Eng. Sci. 17, 467–473 (1977).
6. M. F. Hassan, H. P. Lee and S. P. Lim, Measurement Sci. Technol. 21, 075701 (2010).
7. M. Zou, S. Beckford, R. Wei, C. Ellis, G. Hatton and M. A. Miller, Appl. Surface Sci. 257,
3786–3792 (2011).
8. S. A. Kulinich and M. Farzaneh, Appl. Surface Sci. 255, 8153–8157 (2009).
9. L. Mishchenko, B. Hatton, V. Bahadur, A. Taylor, T. Krupenkin and J. Alzenberg, ACSNano 4,
7699–7707 (2011).
10. F. Wang, C. Li, Y. Lv, F. Lv and Y. Du, Cold Regions Sci. Technol. 62, 29–33 (2010).
11. J. Bico, U. Thiele and D. Quere, Colloids Surfaces 206, 41–46 (2002).
12. S. A. Kulinich and M. Farzaneh, Langmuir 25, 8854–8856 (2009).
13. D. K. Sarkar and M. Farzaneh, J. Adhesion Sci. Technol. 23, 1215–1237 (2009).
14. J. R. Stallabrass, Canad. Aeronaut. Space J. 9, 199–204 (1963).
15. C.-L. Phan, P. McComber and A. Mansiaux, Trans. Can. Soc. Mech. Eng. 44, 204–208 (1978).
16. O. B. Naselle, L. Levi and F. Prodi, J. Glaciol. 33, 120–122 (1987).
17. J.-L. Laforte, C. L. Phan, B. Felin and R. Martin, Special Report 83-17, pp. 83–92, U.S. Army
Cold Regions Research and Engineering Laboratory, CRREL (1983).
442 L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445

18. M. Lyyra, M. Jäntti and J. Launiainen, in: Proceedings of the International Offshore and Naviga-
tion Conference, Polartech 86, Technical Research Centre of Finland, Helsinki, Finland, Vol. 1,
pp. 484–496 (1986).
19. L. Makkonen, Cold Regions Sci. Technol. 14, 163–171 (1987).
20. P. V. Hobbs, Ice Physics. Clarendon Press, Oxford, UK (1974).
21. M. Faraday, Phil. Mag. Ser. 4 17, 162–169 (1859).
22. N. H. Fletcher, Phil. Mag. Ser. 8 7, 255–269 (1962).
23. J. G. Dash, Science 246, 1591–1592 (1989).
24. N. Fukuta, J. Physique 48, 503–509 (1987).
25. S. C. Hardy, Phil. Mag. Ser. 8 27, 471–484 (1977).
26. Y. Furukawa, M. Yamamoto and T. Kuroda, J. Cryst. Growth 82, 665–674 (1987).
27. H. Dosch, A. Lied and J. H. Bilgram, Surface Sci. 327, 145–164 (1995).
28. L. Makkonen, Special Report 89-5, pp. 277–309, U.S. Army Cold Regions Research & Engi-
neering Laboratory, CRREL (1989).
29. P. Oksanen, Publications 10, 36 pp., Technical Research Centre of Finland, VTT (1983).
30. L. Makkonen and E. Lehmus, Report No. 12, 53 pp., Finnish–Soviet Committee on Utilization
of Oil and Gas Resources in Frozen Sea Areas (1988).
31. K. Yano and D. Kuroiwa, in: Proceedings of the International Symposium on Snow and Ice
Control Research, U.S. Army Cold Regions Research & Engineering Laboratory, Hanover, NH,
USA, Paper 185, pp. 30–34 (1978).
32. M. Wakatsuchi and T. Saito, Annals Glaciol. 6, 200–202 (1985).
33. N. Ono and T. Kasai, Annals Glaciol. 6, 298–299 (1985).
34. L. Makkonen and E. Lehmus, in: Proceedings of the 9th International Conference on Port and
Ocean Engineering Under Arctic Conditions (POAC), University of Alaska, Fairbanks, Alaska,
USA, Vol. I, pp. 45–55 (1987).
35. G. F. N. Cox and W. F. Weeks, Report 354, 85 pp., U.S. Army Cold Regions Research and
Engineering Laboratory, CRREL (1975).
36. N. N. Zubov, Arctic Ice. U.S. Navy Electronics Laboratory, San Diego, CA, USA (1945).
37. V. L. Tsurikov, Oceanologia 5, 463–472 (1965).
38. W. F. Weeks and G. Lofgren, in: Physics of Snow and Ice, Vol. 1, pp. 579–597. Institute of Low
Temperature Science, Japan (1967).
39. A. S. Myerson and D. J. Kirwan, Ind. Eng. Chem., Fundamentals 16, 414–429 (1977).
40. L. Makkonen, Appl. Phys. Lett. 96, 091910 (2010).
41. H. H. G. Jellinek, J. Colloid Interface Sci. 25, 192–205 (1967).
42. A. V. Panyushkin, Z. I. Shvayshteyn and N. A. Sergatcheva, Arctic Antarctic Res. Instit. Trudy
298, 59–70 (1972).
43. A. G. Tkachev and V. P. Malyshev, Kholodilnaya Tekhnika 8, 15–18 (1976).
44. D. M. Berenger, R. Y. Edwards Jr. and J. P. Nadreau, Research Project Report 85-1, Arctic
Petroleum Operators’ Association (APOA) (1985).
45. W. M. Sackinger and P. A. Sackinger, in: Proceedings of the International Offshore and Naviga-
tion Conference, Polartech 86, Technical Research Centre of Finland, Helsinki, Finland, Vol. 1,
pp. 512–527 (1986).
46. V. V. Panov, Arctic Antarctic Res. Instit. Trudy 334, 1–263 (1976).
47. A. V. Panyushkin, Z. I. Shvayshteyn, N. A. Sergatcheva and V. S. Podokshik, Arctic Antarctic
Res. Instit. Trudy 298, 71–77 (1972).
48. H. H. G. Jellinek and I. Chodak, Special Report 83-17, pp. 97–102, U.S. Army CRREL (1982).
L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445 443

49. L. Makkonen, in: Proceedings of the 12th Annual Adhesion Society Meeting, The Adhesion
Society, Hilton Head Island, SC, USA, Paper 234, 4 pp. (1989).
50. L. Makkonen, Atmospheric Icing of Sea Structures. Monograph 84-2, U.S. Army CRREL
(1984).
51. V. K. Croutch and R. A. Hartley, J. Coatings Technol. 64, 41–53 (December 1992).
52. C. Laforte, J.-L. Laforte and J.-C. Carriere, in: Proceedings of the 10th International Workshop
on Atmospheric Icing of Structures (IWAIS), Brno, Czech Republic (2002).
53. S. Frankenstein and A. M. Tuthill, J. Cold Regions Eng. 16, 83–96 (2002).
54. A. V. Panyushkin, Yu. D. Sinitchkin and N. A. Sergatcheva, Arctic Antarctic Res. Instit. Trudy
317, 99–102 (1975).
55. E. H. Andrews and N. A. Lockington, J. Mater. Sci. 19, 73–81 (1975).
56. V. A. Igoshin, E. L. Tyunina and I. N. Cherskii, Soviet J. Friction Wear 6, 78–83 (1985).
57. M. Javan-Mashmool, C. Volat and M. Farzaneh, Hydrol. Processes 20, 645–655 (2006).
58. M. L. Chu, R. J. Scavuzzo and W. V. Olsen, in: Proceedings 3rd International Workshop on
Atmospheric Icing of Structures (IWAIS), BC Hydro, Vancouver, Canada, pp. 273–280 (1986).
59. L.-O. Anderson, C.-G. Golander and S. Persson, J. Adhesion Sci. Technol. 8, 117–132 (1994).
60. D. N. Anderson and A. D. Reich, Tests of the performance of coatings for low ice adhesion.
Technical Memorandum 107399, 14 pp., NASA (1997).
61. B. Somlo and V. Gupta, Mech. Mater. 33, 471–480 (2001).
62. N. Sonwalkar, S. S. Sunder and S. K. Sharma, Appl. Spectroscopy 47, 1585–1593 (1993).
63. C. Laforte and A. Beisswenger, in: Proceedings of the 11th International Workshop on Atmo-
spheric Icing of Structures (IWAIS), Montreal, Canada, Paper IW-53 (2005).
64. W. D. Bascom, R. L. Cottington and C. R. Singleterry, J. Adhesion 1, 246–263 (1969).
65. H. Murase and K. Nanishi, Annals Glaciol. 6, 146–149 (1985).
66. S. Alansatan and M. Papadakis, in: Proceedings of the General, Corporate and Regional Aviation
Meeting and Exposition, SAE Technical Paper Series, No. 991584 (1999).
67. K. Matsumoto and Y. Daikoku, Int. J. Refrigeration 32, 444–453 (2009).
68. X. Fan, P. Ten, C. Clarke, A. Bramley and Z. Zhang, Powder Technol. 131, 105–110 (2003).
69. P. Archer and V. Gupta, J. Mech. Phys. Solids 46, 1745–1771 (1998).
70. H. Saeki, T. Ono and A. Ozaki, in: Proceedings of the IAHR Symposium on Ice, International
Association for Hydraulic Research, Quebec City, Canada, Vol. 2, pp. 641–649 (1981).
71. A. J. Meuler, J. D. Smith, K. K. Varanasi, J. M. Mabry, G. H. McKinley and R. E. Cohen, Appl.
Mater. Interfaces 2, 3100–3110 (2011).
72. J. M. Sayward, Special Report 79-11, U.S. Cold Regions Research & Engineering Laboratory,
CRREL (1979).
73. A. G. Kraj and E. L. Bibeau, Renewable Energy 35, 741–746 (2010).
74. N. S. Stehle, in: Proceedings of the IAHR Symposium on Ice, Reykjavik, Iceland, Paper 5.3
(1970).
75. L. E. Raraty and D. Tabor, Proc. Roy. Soc. A 245, 184–210 (1958).
76. A. V. Panyushkin, B. V. Rozenzweig, Yu. B. Petrov, L.Ye. Gurvitch and N. A. Sergatcheva, Arctic
Antarctic Res. Instit. Trudy 298, 83–90 (1972).
77. H. Saeki, T. Ono, T. Takeuchi, S. Kanie and N. Nakazawa, in: Proceedings of the Fifth Offshore
Mechanics and Arctic Engineering Conference (OMAE), Hokkaido University, Tokyo, Japan,
Vol. IV, pp. 534–540 (1986).
78. T. Ozeki and R. Yamamoto, in: Proceedings of the 18th IAHR International Symposium on Ice,
International Association for Hydraulic Research, Sapporo, Japan, pp. 153–160 (2006).
444 L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445

79. G. Frankenstein, J. Wuebben, H. Jellinek and R. Yokota, in: Proceedings of Symposium on Inland
Waters for Navigation, Flood Control and Water Diversions, pp. 1487–1496, American Society
of Civil Engineers, Fort Collins, CO, USA (1976).
80. L. Makkonen, O. Erikoinen and E. Lehmus, in: Proceedings of the International Offshore and
Navigation Conference, Polartech’86, Technical Research Centre of Finland, Helsinki, Finland
(VTT Symposium, No. 71), Vol. I, pp. 725–734 (1986).
81. H. Saito, K.-I. Takai and G. Yamauchi, Mater. Sci. Res. Int. 3, 185–189 (1997).
82. V. F. Petrenko and S. Peng, Can. J. Phys. 81, 387–393 (2003).
83. M. Landy and A. Freiberger, J. Colloid Interface Sci. 25, 231–244 (1967).
84. L.-O. Anderson, Thesis 1993:23L, LuleåUniversity of Technology, ISSN 0280-8242 (1993).
85. S. L. Sivas, B. Ruegler, R. Thomaier and K. Hoover, in: Proceedings SAMPE Fall Technical
Conference, Cincinnati, Ohio, USA (2007).
86. C. C. Ryerson, Report ERDC/CRREL TR-09-4, 342 pp., U.S. Army Cold Regions Research &
Engineering Laboratory, CRREL (2009).
87. S. Kimura, T. Sato and K. Kosugi, in: Proceeding of the Boreas VI Meeting, Pyhätunturi, Finland
(2003).
88. A. Dolan, H. Dodiuk, C. Laforte and S. Kenig, J. Adhesion Sci. Technol. 23, 1907–1915 (2009).
89. R. Karmouch, S. Coude, G. Abel and G. G. Ross, Electrical Power & Energy Conference
(EPEC). IEEE, Montreal, Canada (2009).
90. K. K. Varanasi, T. Deng, J. D. Smith, M. Shu and N. Bhate, Appl. Phys. Lett. 97, 234102 (2010).
91. S. A. Kulinich and M. Farzaneh, Cold Regions Sci. Technol. 65, 60–64 (2011).
92. S. A. Kulinich, S. Farhadi, K. Nose and X. W. Du, Langmuir 27, 25–29 (2011).
93. N. Du, X. Y. Liu and L. C. Hew, J. Biol. Chem. 278, 36000–36004 (2003).
94. E. Kristiansen and K. E. Zachariassen, Cryobiology 51, 262–280 (2005).
95. K. Siegmann, A. Kaufmann and M. Hirayama, Erneuerbare Energien 2007, 38–41 (2007).
96. P. M. Wiggins, Microbiol. Rev. 54, 432–449 (1990).
97. V. M. Gun’ko, V. V. Turov, V. M. Bogatyrev, V. I. Zarko, R. Leboda, E. V. Goncharuk, A. A.
Novza, A. V. Turov and A. A. Chuiko, Adv. Colloid Interface Sci. 118, 125–172 (2005).
98. H. Murase, K. Nanishi, H. Koruge, T. Fujibayashi, K. Tamura and N. Haruta, J. Appl. Polym. Sci.
54, 2051–2062 (1994).
99. N. Rehfeld, in: Proceedings of the European Coatings Congress, Nürenberg, Germany (2009).
100. H. G. Hanamoto, J. J. Gagnon and B. Pratt, Special Report 80-18, U.S. Army Cold Regions
Research & Engineering Laboratory, CRREL (1972).
101. C. Volat, M. Farzaneh and A. Leblond, in: Proceedings of the 11th International Workshop on
Atmospheric Icing of Structures, IWAIS, Montreal, Canada, 11 pp. (2005).
102. L. I. Churakov, V. K. Savinukh and R. N. Bobrov, Inst. Inzhenerov Vodnogo Transporta, Novosi-
birsk, Trudy 94, 3–8 (1976).
103. G. J. Hartraft, Report FAA-RD-72-78, Federal Aviation Admin. (1972).
104. Anon, Safety Advisor, Weather No. 2. AOPA Air Safety Foundation (2004).
105. J. R. Stallabrass, Mechanical Engineering Report MD-51, National Research Council, Ottawa,
Canada (1970).
106. S. F. Ackley, K. Itagaki and M. Frank, J. Glaciol. 19, 467–478 (1977).
107. T. Tabata, Low Temperature Sci., Series A 21, 173–221, Defense Research Board, Ottawa,
Canada, Translation T95J (1963).
108. G. W. Zumwalt and R. A. Friedberg, in: Proceedings of the 24th AIAA Aerospace Sciences Meet-
ing, Reno, NV, USA, AIAA Paper 86-0545 (1986).
L. Makkonen / J. Adhesion Sci. Technol. 26 (2012) 413–445 445

109. G. D. Embry, R. W. Erskine, L. A. Haslim, T. T. Lockyer and P. T. McDonough, Naval Engineers


J. 102, 55–66 (1990).
110. I. Strangeways, in: Proceedings of the Third International Workshop on Atmospheric Icing of
Structures (IWAIS), Vancouver, Canada, pp. 293–299 (1986).
111. R. Ross and G. W. Zumwalt, in: Proceedings of the Third International Workshop on Atmospheric
Icing of Structures (IWAIS), BC Hydro, Vancouver, Canada, pp. 310–309 (1986).
112. J. K. Alexeiev, WMO Technical Note 135, pp. 7–16, World Meteorological Organization (1974).
113. M. A. Allaire and J. L. Laforte, US Patent 6,518,497 (2003).
114. L. Makkonen, P. Lehtonen and L. Helle, J. Atmos. Oceanic Technol. 18, 1457–1469 (2001).
115. J. L. Palacios, E. C. Smith and J. L. Rose, in: Proceedings of the American Helicopter Society,
64th Annual Forum, Montreal, Canada (2008).
116. L. Makkonen, T. Laakso, M. Marjaniemi and K. J. Finstad, Wind Eng. 25, 1–21 (2001).
117. B. Song and R. Viskanta, J. Thermophys. 4, 311–317 (1990).
118. T. Okihara, M. Kanamori, A. Kamimura, N. Hamada, S. Matsuda, J. Buturlia and G. Miskolczy,
in: Proceedings of the 15th AIAA Thermophysics Conference, USA (1980).
119. B. S. Larkin and S. Dubuc, Report ESA SP112, 1, pp. 529–535, European Space Agency (1976).
120. V. V. Panov, Polar Geography 2, 166–186 (1978).
121. V. F. Petrenko, in: Proceedings of the International Conference on Physics and Chemistry of Ice,
Bremenhaven, Germany (2006).
122. G. Botura, D. Sweet and D. Flosdorf, in: Proceedings of the 43rd AIAA Aerospace Sciences
Meeting and Exhibit, Reno, Nevada, USA (2005).
123. S. F. Ackley, K. Itagaki and M. Frank, Internal Report 351, AD-777947, U.S. Army Cold Regions
Research & Engineering Laboratory, CRREL (1973).
124. H. Gerger, WMO Technical Note 135, pp. 1–6, World Meteorological Organization (1974).
125. C. C. Bates, in: Proceedings of the 23rd International Congress of Navigation, Ottawa, Canada
(1973).
126. Y. P. Semenova, Draft Translation 411, pp. 98–107, ADA003215, U.S. Army Cold Regions Re-
search & Engineering Laboratory, CRREL (1974).
127. J. Ayres, W. H. Simendinger and C. M. Balik, J. Coatings Technol. Res. 4, 473–481 (2007).
128. C. L. Phan and J. L. Laforte, Cold Regions Sci. Technol. 4, 15–25 (1981).
129. P. Paasivuori, A. Etvushenko, E. Järvinen and L. Makkonen, in: Proceedings of the PO-
LARTECH’94 Conference, Lulea University of Technology, Luleå, Sweden, pp. 232–242 (1994).
130. I. A. Ryzhkin and V. F. Petrenko, J. Phys. Chem. B 101, 6267–6270 (1997).
131. V. F. Petrenko, J. Appl. Phys. 86, 5450–5456 (1999).
132. S. T. Wang, US Patent 6,402,093 (2002).
133. V. F. Petrenko, WO Patent WO/1998/057851 (1998).
134. E. M. Petrie, Metal Finishing 107 (2), 56–59 (2009).
135. N. Dalili, A. Edrisy and R. Carriveau, Renewable Sustainable Energy Rev. 13, 428–438 (2009).

View publication stats

También podría gustarte