Next Article in Journal
Asynchronously H Tracking Control and Optimization for Switched Flight Vehicles with Time-Varying Delay
Next Article in Special Issue
Experimental Investigation on Ice–Aluminum Interface Adhesion Strength under Heating Conditions
Previous Article in Journal
Real-Time Ground Aeroservoelastic Test for Slender Vehicles Based on Condensed Aerodynamic Force Loading
Previous Article in Special Issue
Three-Dimensional Trajectory and Impingement Simulation of Ice Crystals Considering State Changes on the Rotor Blade of an Axial Fan
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Round-Robin Study for Ice Adhesion Tests

1
Department Paint Technology, Fraunhofer Institute for Manufacturing Technology and Advanced Materials IFAM, 28359 Bremen, Germany
2
Laboratoire International des Matériaux Antigivre/Anti-Icing Materials International Laboratory, Department of Applied Sciences, Université du Québec à Chicoutimi, Saguenay, QC G7H2B1, Canada
3
Department of Aerospace and Mechanical Engineering, University of Notre Dame, Notre Dame, IN 46628, USA
4
Power Generation Technologies and Materials Department, RSE, Ricerca sul Sistema Energetico, 29122 Piacenza, Italy
5
Materials Science and Environmental Engineering, Faculty of Engineering and Natural Sciences, Tampere University, 33720 Tampere, Finland
6
INTA-Instituto Nacional de Técnica Aeroespacial, Área de Materiales Metálicos, Ctra. Ajalvir Km 4, 28850 Madrid, Spain
7
Department of Structural Engineering, Norwegian University of Science and Technology (NTNU), 7491 Trondheim, Norway
8
School of Aerospace, Transport and Manufacturing, Centre for Propulsion and Thermal Power Engineering, Cranfield University, Cranfield MK43 0AL, UK
9
Department of Mechanical and Industrial Engineering, University of Toronto, Toronto, ON M5S 3G8, Canada
10
US Army Engineer Research & Development Center, Cold Regions Research & Engineering Laboratory, Hanover, NH 03755, USA
11
RISE Research Institutes of Sweden, Department Materials and Surface Design, SE-114 86 Stockholm, Sweden
12
Faculty of Engineering, University of Nottingham, University Park, Nottingham NG7 2RD, UK
*
Author to whom correspondence should be addressed.
Aerospace 2024, 11(2), 106; https://doi.org/10.3390/aerospace11020106
Submission received: 18 December 2023 / Revised: 12 January 2024 / Accepted: 14 January 2024 / Published: 24 January 2024
(This article belongs to the Special Issue Deicing and Anti-Icing of Aircraft (Volume III))

Abstract

:
Ice adhesion tests are widely used to assess the performance of potential icephobic surfaces and coatings. A great variety of test designs have been developed and used over the past decades due to the lack of formal standards for these types of tests. In many cases, the aim of the research was not only to determine ice adhesion values, but also to understand the key surface properties correlated to low ice adhesion surfaces. Data from different measurement techniques had low correspondence between the results: Values varied by orders of magnitude and showed different relative relationships to one another. This study sought to provide a broad comparison of ice adhesion testing approaches by conducting different ice adhesion tests with identical test surfaces. A total of 15 test facilities participated in this round-robin study, and the results of 13 partners are summarized in this paper. For the test series, ice types (impact and static) as well as test parameters were harmonized to minimize the deviations between the test setups. Our findings are presented in this paper, and the ice- and test-specific results are discussed. This study can improve our understanding of test results and support the standardization process for ice adhesion strength measurements.

1. Introduction

Icephobic surfaces are of great interest for various technical applications, including aviation, energy, and automotive sectors [1,2,3,4,5,6,7,8,9,10,11,12]. The aim of using such materials is to prevent or delay ice accretions on technical surfaces and/or reduce the adhesion of ice to the extent that it can be removed easily by external measures (e.g., gravity, vibrations, and heating) without costly energy input. The developers of such icephobic materials are facing a lack of standardized test methods for performance evaluation and technology transition and deployment. Various test methods have been developed over the past decades [13,14,15,16,17,18,19,20], but efforts to compare the test results with each other are limited, resulting in uncertainties of general rules for the identification of icephobic surfaces [21,22,23,24,25,26,27]. This study is aimed at the delivery of test data derived from identical coating and tape materials to exclude any uncertainties in the comparability of test surfaces. This includes the selection of robust (mechanical and chemical wise) as well as long-term stable surfaces for the round-robin study.
The use of lab-based tests is generally accompanied by great uncertainty regarding the significance of test results in terms of their final technical application. Some researchers have suggested different values for low ice adhesion coatings, including Hejazi et al. (2013), who suggested a threshold of 100 kPa [28], and Dou et al. (2014), who suggested a value of τice ≤ 27 kPa for ice detachment by a strong breeze [29]. The data contrast with the test data summarized by Work and Lian (2018), with averaged values for aluminum (a non-icephobic material) ranging from 27 to 122 kPa for centrifuge tests [22]. Other researchers have introduced the so-called adhesion reduction factor (ARF) by dividing the adhesion strength of a predefined benchmark material by that of the test surface [13]. This accounts for the nonuniform stress distributions during ice adhesion tests and allows for comparison against a state-of-the-art material. In this case, the data quality of the benchmark material needs to be at a certain level (standard deviations preferably below 10%) to avoid any misinterpretations. Nonetheless, it is uncertain whether it will be possible to define general values that are valid for all types of ice adhesion tests and ice adhesion thresholds in the future. This is not only due to the different test designs but also because of the variable and complex (de-)icing scenarios of the target applications. However, this study aims to deliver ice adhesion strength data for defined surfaces, allowing the comparison of test designs and the development of steps to lead future research topics.
The different types of ice adhesion measurements vary in fundamental principles regarding how the ice is formed and removed. Additionally, variances in temperature and test duration result in further deviations in ice adhesion values [30]. Conducting different ice adhesion measurements under harmonized test conditions and on identical test surfaces will allow the comparison of these results and improve the understanding of differences in test designs. Table 1 summarizes the contributors in this paper.
This paper summarizes the results of ice adhesion measurements in the various test facilities on identical test surfaces. The test surfaces were proven to possess appropriate stability over the period of testing in a parallel study [31]. The results are grouped with regard to the test types and ice types, compared to each other, and the observed similarities and differences are discussed.

2. Materials and Methods

Neither standard test methods nor general parameter sets were available for the ice adhesion tests. To conduct round-robin tests, basic test types were identified, and test parameter sets were defined to achieve maximum comparability of the test results. The test surfaces and pretreatment procedures (handling and cleaning) were identical in order to exclude any further deviations from the test types.

2.1. Test Sample Preparation

The sizes and geometries of the test samples differed due to the various test designs in this round-robin test. Prior to coating application, all test samples were sanded (3M 320-grit sandpaper) and cleaned with isopropanol. The preparation of the samples for this round-robin test was performed in a single process with identical material batches to avoid deviations due to changes in material composition, handling, or environmental conditions. We used four different test surfaces that represented a general range of wettability and roughness and that were proven to be robust in a previous study [31]. The chemical (for cleaning purposes) and mechanical (for repeated icing/de-icing cycles) stability assured a high comparability of the test surfaces for the different test facilities.
For the coating type “Primer”, an epoxy primer (Aerodur 37045 Barrier Primer White with Hardener S66/22R; Akzo Nobel, Sassenheim, The Netherlands) was used in a mixing ratio of 2:1 by volume. Material preparation and application were carried out according to supplier specifications using a spray gun (SATA Jet 90 with Ø 1.3 mm, pressure of 1.6 bar, and distance of 40 cm) under standard conditions (temperature of 21 °C and relative humidity of 40%). After the coating application, samples were stored at room temperature in a clean environment for 12 h prior to thermal curing at 60 °C for 60 min. The resulting dry film thickness of the primer was 40 ± 10 µm (according to DIN EN ISO 2808:2019 with byko-test 8500 P Fe/NFe, Byk Gardner, Geretsried, Germany) [32]. Figure 1 shows the test samples after primer application.
The coating type “Standox” is a clear coat that is used for repairs in the automotive industry. It was applied on top of the primer coating (described above) to ensure good adhesion properties. Standocryl VOC-Premium Clear K9540 with Hardener VOC 10–20 was purchased (Standox GmbH, Wuppertal, Germany), and material preparation was carried out according to supplier specifications in a mixing ratio of 3:1 (by volume). Application and curing parameters were identical to those of the primer coating. The same application method was used for the coating type “PUR C25”, which is a noncommercial 2-component formulation based on silanized polyisocyanate-curing acrylic resin as described in [33]. The total film thickness (incl. primer) for the samples Standox and PUR C25 was 70 ± 15 µm [32].
The last material in this study was “PTFE Tape” (extruded polytetrafluoroethylene PTFE film tape 5490; 3M Deutschland GmbH, Neuss, Germany) which was applied on primed test samples according to supplier specifications. The film thickness of the tape was 90 µm.
Samples for 15 partners were prepared as described and delivered by Fraunhofer IFAM. Additionally, CRREL prepared test samples as follows: The materials of interest (Primer, Standox, PTFE tape) were applied to aluminum substrates for testing. Before the coating application, all aluminum substrates were polished with lapping films through p4000-grit papers, resulting in an average roughness of approximately 0.362 μm, measured using a non-contact optical profilometer (Model ST400, Nanovea, Irvine, CA, USA) equipped with a confocal chromic sensor (ISO 25178). Following the polishing process, the substrates were rinsed with high-purity water, dried with ethanol, and stored in airtight bags until used for coating or testing. At the time of use, substrates were removed from the sealed bags, soaked in sulfuric acid (pH 1.5) for 5 min, rinsed with high-purity water, and then dried by wiping with acetone or isopropanol. The substrates were then tested for ice adhesion or had coatings applied within 30 min to limit the formation of aluminum oxide on the surfaces. All coatings were applied and cured according to manufacturer specifications.

2.2. Surface Characterizations

For the assessments of the ice adhesion, we selected 4 different materials that represented a reasonable range of wettability and roughness properties and that were proven to be robust in repeated tests, as reported in [31]. Surface characterization was conducted prior to the shipment of the test samples to the partners for ice adhesion measurements. Wettability tests were performed with the Drop Shape Analyzer DSA 100S (Krüss GmbH, Hamburg, Germany), according to relevant specifications (DIN EN ISO 19403-2) [34]. Surface free energy (SFE) was determined by measuring the dynamic contact angle of 3 liquids—water, diiodomethane, and ethylene glycol (droplet application of 0.2 µL/s and a total volume of 6.0 µL)—and calculated according to the method of Owens, Wendt, Rabel, and Kaelble (OWRK). The water contact angle (WCA) was extracted from this measurement. The water sliding angle (WSA) was determined with a water droplet volume of 20 µL and a tilting speed of 60°/min. The sliding angle was defined as the angle at which the advancing and receding angles of the water droplet moved at least 1 mm from the starting point [35]. Contact angle hysteresis (CAH) was determined at this sliding angle or at the maximum tilting angle of 90° (in the case where the water droplet did not run off) by calculating the difference between advancing and receding angles. The tilting method was chosen because it delivers consistent results. Roughness data Ra (arithmetic average value of the roughness profile) and Rz (maximum height of the profile) were determined using a Perthometer M2 (Mahr GmbH, Göttingen, Germany). Surface parameters are expressed in Table 2 as the means of 6 measurements from 3 test samples. Additional random samples for each participating partner were controlled against these data to demonstrate the comparability of delivered test samples of this series.
At CRREL, wettability experiments were performed with a Model 590 contact angle goniometer (Ramé-Hart, Succasunna, NJ, USA) with an automated liquid dispensing system, tilting base, and camera using the Dropimage Advanced software package. Six measurements were performed for each material surface. For sliding (WSA) and static (WCA) contact angle measurements, 20 µL of MilliQ® was dispensed onto the material surface. Static angle measurements were determined from the initial contact angle measurement prior to tilting at t0. Tilting of the base for WSA measurements was performed at 60° per min (1° per second) with automated measurements performed every 0.5 s. The sliding angle was determined at the point when the leading edge of the drop slid out of the camera’s frame (approximately 1 mm of movement from the initial position), and the volumetric measurement for the drop determined using the Dropimage Advanced software fell below 20 µL. Contact angle hysteresis was calculated using the trailing and leading contact angle from 1 measurement prior to the drop moving out of the frame and volumetric measurement reduction.
Roughness was determined by obtaining 1 × 1 mm scans using the optical profilometer (described above) in a 5 µm step size in both x and y directions. The scanned surfaces were analyzed using MountainsMap 7.4. The surface form was removed using a 2nd-order polynomial fit. Using the removed surface form, 6 horizontal profile lines (in the direction of scanning) were extracted at equidistant lengths. Roughness and waviness from the extracted profile were separated using a Gaussian filter with a cut-off of 250 µm, and Ra and Rz values were reported. Results are summarized in Table 3.
Results for test surfaces, prepared by IFAM, indicate surface free energies (SFE) from 15.1 mN/m to 38.5 mN/m (Table 2). Water contact angles (WCA) were determined from 83° to 110°. This property range was expected to be sufficient for the round-robin study. A further increase in wettability (higher SFE and lower WCA) bears the risk of cohesive ice failure instead of quantifiable ice adhesion results. For surfaces with lower wettability (lower SFE and higher WCA, including superhydrophobic surfaces), no materials were identified that fulfilled the requirements for this round-robin study in terms of robustness and long-term stability.
The comparison of the resulting surface properties from CRREL and IFAM preparations showed significant deviations in roughness for the Primer and Standox coating materials. This may be caused by differences in the used substrates, material batches, and application techniques. For the CRREL preparation, the Primer surface showed a significantly lower roughness, but the Standox material showed an increased roughness. For the PTFE tape, the roughness data are in a comparable range, indicating no effects based on the different roughness measurement techniques but showed differences between the coating types for CRREL and IFAM preparations. The trends for contact angle measurements fit well despite the different methods. The only significant difference was observed for the WCA of PTFE tape (CRREL 95° and IFAM 110°), which may be the result of static and dynamic contact angle measurements using different volumes and fitting methods. These findings will be considered during the result assessments for the ice adhesion data and emphasize the need to prepare test samples for comparison tests in a single facility.

2.3. Ice Adhesion Test Methods

The test methods in this study cover a wide range of designs and are grouped into direct mechanical tests (push or pull), centrifuge tests, and mode I tests, as described below.

2.3.1. Direct Mechanical Tests (RSE, INTA, NTNU, RISE, ND, and CRREL)

The described tests used static ice for the assessment of ice adhesion; graphical schemes and images are summarized in Table 4.
RSE conducted direct mechanical tests by using a homemade apparatus for shear test in pull mode equipped with an electromechanical testing system, INSTRON 4507 [36]. The specimens with a cylindrical shape were frozen in an aluminum alloy mold at −19 °C overnight, and then the mold was fixed into the machine and the sample was extracted from the ice at a speed of 0.3 mm/s. The force F needed to pull the sample off the mold was recorded.
INTA’s double-lap shear (pull) method is a modification of a test described by Ferrick et al. (2006) [37] in which optimizations have been adapted: The dented edges of the mold were replaced by wedge-shaped molds, which helped to decrease the cohesive ice fracture tests that were probably caused by the stress generated by the upward-slanting roughness elements. Regarding the method, the ice was prepared in-mold using adhesive tape to retain the deionized water and the coupon; the ice was frozen at −8 °C overnight inside an ultra-low temperature freezer (Arctiko ULTF series). One hour before the test, the adhesive tape was removed, and any remaining ice accreted over the sample edges was carefully but quickly removed using a blade. The molds were placed in the freezer again for 1 additional hour. The test blocks prepared using this method were then fixed to an Instrom 5882 Universal Machine and placed inside a climate chamber (refrigerated with liquid nitrogen). The samples were left for 5 additional minutes before beginning the test in order to stabilize the temperature. The displacement speed was set to 0.3 mm/s, and the test was initialized until the samples were completely out of the mold. The Fmax value of the loading curve was used to calculate the ice adhesion strength.
At RISE, an ice shear test (pull mode) was performed by using a plastic cuvette attached to the surface with an inverted lab jack and filled with 1 mL of ultraclean water (Milli-Q, Type 1) through a hole in the cuvette [38]. The assembly was then placed in a freezer (−8 °C) for 180 min. The ice adhesion strength was measured with a modified slip/peel tester (IMASS SP-2000) equipped with a force sensor and a Peltier cooling plate. The equipment was kept in a climate room at 23 °C and 50% RH. Immediately prior to the measurement, the sample was transferred from the freezer to the Peltier plate, which was maintained at −8 °C.
NTNU measured ice adhesion strength by vertical shear test rig (push mode) using an Instron 5944 Universal Machine equipped with a home-built cooling chamber and testing system. A polypropylene tube mold with a 1 mm-thick wall and a 28 mm inner diameter was placed onto the coatings acting as an ice mold; then, the pressure of a 200 g metal cylinder was applied to prevent water leakage. Subsequently, 5 mL of deionized water was syringed into the mold, and the mold was transferred into a freezer at −8 °C for 180 min to ensure complete freezing. Before the test, the samples were transferred from the freezer to the cooling chamber and stabilized at −8 °C for 30 min. During ice adhesion tests, a force probe with a 5 mm diameter propelled the tube-encased ice columns at a velocity of 0.3 mm s−1, and the probe was located close (less than 1 mm) to the tested coating surface to minimize the torque on the ice cylinder. The loading curve was recorded, and the peak value of the shear force (Fmax) was used for the calculation of the ice adhesion strength.
ND employed a horizontal push-type device to measure the ice adhesion. On the test sample plate, the ice was created inside an aluminum ring with a 1 inch inner diameter. The ring with the ice was then pushed horizontally using a rod that was attached to the load cell. The output voltage reading of the load cell was converted into the force to determine the ice adhesion. During the measurement, the ambient temperature was kept at −8 °C, and the pushing speed was controlled at 0.3 mm/s.
For ice adhesion tests conducted by CRREL, freshwater columnar ice was grown on the substrates [39]. This method did not use molds to facilitate surface freezing, but used a growth from the melt procedure instead. Ice growth was conducted at −8 °C. Under these conditions, approximately 1.5 h was required to grow the 1 cm-thick laminate of ice on the material surfaces. The ice adhesion peel test (IAPT) developed at CRREL [40] was carried out in tensile or shear delamination modes. In the tensile mode, the ice is lifted away from the substrate; in the shear mode, the ice is pushed off the substrate along its surface. The testing geometry was fitted inside a universal load frame with machined baseplates and custom load heads. Load and displacement were recorded as a function of time during the test.
For the calculation of the ice adhesion strength (τ0), the following equation was used by all contributors,
τ 0 = F m a x A ,
where Fmax is the maximum force and A is the contact area at the ice substrate interface. Quantitative results indicate that the ice was delaminated by purely adhesive mechanisms with no residual ice remaining on the material surfaces after testing. The ice laminate was removed as a single piece without cohesive failure. A summary of methods for ice adhesion measurements using direct mechanical tests is provided in Table 4.

2.3.2. Centrifuge Tests (AMIL, IFAM, NU, TAU, and P-A)

The contributors AMIL, IFAM, and NU conducted centrifuge tests using static ice. Additionally, AMIL, IFAM, TAU, and P-A used impact ice, accreted in ice wind tunnels, for the assessments. Generally, the centrifuge test used centripetal forces to apply shear stress to the ice and remove it from the test surface. Separation is detected when the ice hits the centrifuge wall and is correlated to the rotational speed of the centrifuge rotor. This speed (angular velocity ω in rad/s) was used to calculate the shear strength of ice to the substrate according to the following equation,
τ = F A = m i c e   ω 2   r   A ,
where mice is the mass of ice (kg), r is the radius of the rotating beam at the mid-length ice position (m), and A is the surface area of the adherent interface (m2) [13]. The calculated values express the adhesive strength of the ice. This method is specified in ISO/TS 19392-6:2023 [41].
AMIL conducted centrifuge tests with static and impact ice. The tests were performed using 32 mm-wide × 6.4 mm-thick aluminum 6061-T6 flat bars cut to a 340 mm length; the surface materials for this round-robin study were tested in a cold chamber, a closed-loop icing wind tunnel, and a centrifuge to conduct the tests under controlled environmental conditions. Figure 2 summarizes the test equipment used at AMIL [19].
IFAM conducted centrifuge tests with static and impact ice. Test samples (EN AW 5083, dimensions of 220 mm × 30 mm × 4 mm) with the surface materials of this round-robin study were tested in an ice lab that includes a closed-loop ice wind tunnel and a centrifuge to conduct the tests under controlled environmental conditions. Figure 3 summarizes the test equipment used at IFAM.
NU measured the ice adhesion strength using the centrifugal method with static ice. The coated specimens possessed dimensions of 50 mm × 20 mm × 1 mm, and the test was conducted in an environmental chamber (ALPHA 1550-40H) with controlled temperature (e.g., −8 °C). Glaze ice was formed on the coating surface with a silicone mold. The mold was kept on top of the ice block during the test, and its weight was also counted for the calculation. Figure 4 illustrates the schematic diagram of the formation of the glaze ice and the testing configuration at NU.
Ice accretion and centrifugal ice adhesion tests (CATs) were performed at the Ice Laboratory at TAU. Ice was accreted with an icing wind tunnel (IWiT) which is located in a climate-controlled cold room. Figure 5 shows icing test facilities, and more information can be found in [43]. Ice was accreted on flat samples in an area of 30 mm × 30 mm. Typically, in TAU’s icing tests, a wind speed of 25 m/s, a temperature of −10 °C in the IWiT, and an acceleration speed of 300 rpm/s were used in the CATs. Similar to previous studies, mixed glaze-type ice was used [44], but in this present study, the parameters in the iWiT were selected as wind speed of 15 m/s and temperature of −8 °C. In the CATs, 200 rpm/s was used as an acceleration speed for the test samples.

2.3.3. Mode I Tests (CU, ConU)

Mode I tests were conducted by 2 participants, Cranfield University (CU), UK, and Concordia University (ConU), Canada, in their respective icing wind tunnels [45,46] using impact ice. Both test devices are similar in principle and were adapted from the Andrews and Lockington blister test [47,48]. The device used in this test consisted of a hollow cylinder of 30 mm (CU) or 40 mm (ConU) in diameter with an inner hole of 4 mm in diameter. In the CU device, the cylinder was made of aluminum 2024-T3 with a front face coated with the material under investigation. Alternatively, in the ConU setup, the cylinder was used as a sample holder where the coated substrates could be secured on the surface using a cap, as shown in Figure 6c.
The inner hole of the cylinder was covered by a thin PTFE disc of 6 mm in diameter and 50 µm in thickness (CU) or a thin rubber elastomer flushed to the surface of the substrate (ConU) prior to testing. This acted as a defect to initiate a crack at the ice/substrate interface as well as to cover the hole to avoid unwanted ice accretion (Figure 6b).
The devices were positioned in the test section of the icing wind tunnel at CU and ConU such that the coated surfaces were perpendicular to the airflow (Figure 6a,d). Ice was accreted on the front surface, and when a sufficient thickness was obtained to ensure plain stress condition, gas was allowed through the hole with gradually increasing pressure until ice detached from the substrate. The thickness of ice prior to removal was 15–20 mm for both CU and ConU. The spray was left on during the entire mechanical test. The type of fracture (adhesive, cohesive, or mixed) as well as the critical pressure needed to remove the ice were monitored and used to calculate the fracture energy and the tensile strength of the ice. Separation was detected when there was a drop in the pressure rise of the applied force.
The fracture energy (FE) required to open the crack can be calculated from the critical pressure Pc, the thickness of ice, and the size of the flaw [48,49]. The fracture toughness and the tensile strength can be obtained from the fracture energy using the average grain size as a typical defect size (Equation (3)),
σ T = F E × E i 1 ν i 2 × 1 π × a g ,
where Ei and νi are the Young’s modulus and the Poisson ratio of ice, respectively, and ag is the average grain size of the ice [48].

2.4. Test Parameter Definition

For ice adhesion measurements, test conditions and resulting ice types are of high relevance for data interpretation. In this study, ice formation types were divided into “static ice” and “impact ice”. Static ice refers to ice that is formed from liquid water, poured in a mold or alike, and allowed to freeze directly onto the test surface. Impact ice refers to ice accretions, formed in an ice wind tunnel with impacting water droplets onto the test surfaces, leading to the formation of an ice layer for subsequent ice adhesion testing. Icing conditions as well as parameters for the ice removal during the tests were discussed between the partners and coordinated to obtain the greatest conformity possible.
Regardless of the method used, all test surfaces were cleaned prior to testing by using isopropanol and soft tissue. This was defined in a pre-phase of the round-robin study along with further basic parameters: The temperature for all tests was set to −8 °C, and deionized water was used unless otherwise stated in Table 5 and Table 6. This harmonization was conducted to improve the comparability of test results.

3. Results

The results of this study are presented in subsections, following the structure of the previous Section 2.3. The results section includes graphs with means and standard deviations for each test method and the tested materials. Raw data are included in the Appendix A.

3.1. Direct Mechanical Tests Using Static Ice (RSE, INTA, NTNU, RISE, ND, and CRREL)

For the mechanical ice adhesion tests, seven methods were used by six different laboratories. Figure 7 summarizes the results by indicating mean and standard deviations of the measurement data, based on the raw data, included in Appendix ATable A1.
The results for mechanical ice adhesion tests show the highest data range for the Primer and Standox coating materials: from cohesive failure (RISE) to 26 kPa (CRREL, tension) and from 700 kPa (RSE) to 19 kPa (CRREL, tension) for the Primer and Standox materials, respectively. These coating materials were expected to have no low ice adhesion properties, resulting in higher ice adhesion strengths compared to the PUR C25 and PTFE tape materials. This is shown by the pull- and push-based shear tests in this study. The standard deviations were comparably high for the INTA results. The result discrimination was less distinct for the NTNU results. However, all test results derived from the materials prepared and delivered by Fraunhofer IFAM confirmed the basic expectations in the ice adhesion ranking.
For CRREL, it was necessary to prepare test samples in parallel, resulting in different surface properties compared to samples of IFAM preparations (see Table 3). This adds uncertainties to the result interpretation. However, CRREL used the same PTFE tape material (3M tape 5490) as IFAM for the sample preparations. This allowed for the best comparability in this study, and results for the mechanical tests indicated the following result ranking: 323 kPa (INTA) > 189 kPa (NTNU) > 158 kPa (ND) > 102 kPa (CRREL) > 71 kPa (RISE). For the CRREL tension test, the data were the lowest (32 kPa) in this study. In this configuration, the ice was lifted directly off the surface with minimal sliding or shear components along the interface. Unlike shear delamination modes, tensile delamination involves minimal interfacial sliding friction between the ice and substrate materials during delamination. As a result, different forces govern delamination in tension vs. shear modes, and different relative rankings can be expected.
A detailed assessment of test parameter dependencies provided additional findings:
  • The standard deviations showed that there is good agreement between RSE and INTA results. For both tests, ice formations were conducted overnight. For the other round-robin tests, icing times between 35 min and 180 min were used (see Table 5). Additionally, the test samples were completely embedded in the ice compound for the RSE and INTA tests instead of only one flat homogeneous surface that was covered by ice (see Table 4). These differences might have led to increased ice adhesion results, especially for PUR C25 and PTFE tape, compared with the remaining shear test results.
  • The deviation in the ice formation temperature for RSE (−19 °C instead of the harmonized −8 °C for the rest of the test program, Table 5) did not seem to affect the results significantly.
No correlations could be observed amongst the mechanical tests between ice adhesion results and area of ice coverage, displacement type (push/pull), and displacement speed.

3.2. Centrifuge Tests (NU, IFAM, AMIL, TAU, and P-A)

Centrifuge tests were conducted by five contributors using different types of ice. The results are summarized in Figure 8. The figure includes results for static and impact ice formations, the latter accreted in ice wind tunnels with wind speeds as indicated. Mean and standard deviations are shown; these are derived from raw data displayed in Appendix ATable A2.
Results for static ice formations were delivered by three partners. The PUR C25 material showed the lowest ice adhesion strength data for the materials in testing, with 116 kPa (AMIL), 70 kPa (NU), and 41 kPa (IFAM). For the Primer and Standox materials, no consistent material ranking trend could be observed.
For impact ice, accreted at a wind speed of 15 m/s, no consistent material ranking was observed for the two delivering partners AMIL and TAU. This may be linked to deviations in ice conditioning times, with 17 h for the TAU facility and 10 min for the AMIL facility. For partners IFAM and P-A (impact ice, 40 m/s), the following material ranking was identified: Primer ≥ Standox ≥ PTFE tape > PUR C25.
A direct comparison of different ice types can be conducted for test results from AMIL and IFAM. AMIL tested ice adhesion using static ice as well as impact ice accreted at 15 m/s. For the PUR C25 and Standox materials, no significant differences in ice adhesion strength data amongst the ice types were observed. For the Primer material, the ice adhesion strength for static ice was significantly lower (242 kPa) compared to that of impact ice (446 kPa). It can be postulated that the high surface roughness of the coating causes a mechanical interlocking of ice due to the impinging of water droplets under freezing conditions. A similar finding has been reported for an aluminum surface [50].
IFAM observed cohesive ice failure for Primer coating (qualitative result), regardless of the ice type. For the PTFE tape and PUR C25 materials, no significant differences between the ice types were observed. A quantifiable deviation was observed for the Standox coating, with 118 kPa for impact ice (40 m/s) and 185 kPa for static ice. This is not in accordance with the findings for the Primer coating in AMIL tests. Standox and Primer surfaces showed different surface roughness. The effects of the difference in surface roughness on the results of different ice types remain unclear and underline the need for systematic assessments considering various ice types in the material evaluations.
For further parameter assessment, tests were grouped into the ice types “static” and “impact”. For tests using static ice, the iced areas and ice masses increased amongst the test designs as follows: NU (1.38 cm2; 1.3 g) < IFAM (9 cm2; 3 g) < AMIL (11.2 cm2; 7 g). No potential correlations were observed in the measurement data for PUR C25, but for the Standox coating, increasing ice adhesion strength was observed for higher ice masses.
Additionally, the shear stress evolution during the centrifuge tests differed significantly in this study and resulted in the same ranking as that for the ice masses: NU << IFAM < AMIL (see Figure 9). The lower the centrifuge acceleration speed, the higher the resolution for the low ice adhesion region (e.g., < 100 kPa) and the higher the cumulative stresses over the test duration. The observed low ice adhesion strength for Standox coating in NU tests (109 kPa)—the lowest acceleration speed in this study—could be linked to an increased test duration/cumulative stress. An increased acceleration speed (shorter test times) may then lead to increased ice adhesion strength data for tests at IFAM (185 kPa) and AMIL (397 kPa). However, for the PUR C25 and PTFE tape materials with expected low ice adhesion, this effect was not observed.
In a parallel IFAM study with increasing acceleration speeds of 100 rpm/s, 200 rpm/s, and 300 rpm/s and a fixed ice mass of 3 g, a slight increase in the ice adhesion strength was observed for the highest acceleration speed in the test (Standox: 185 kPa, 185 kPa, and 217 kPa) [31]. However, the observed difference was not significant considering the standard deviations, and factual correlations remain unclear due to the observed multiparameter dependency.
For impact ice adhesion tests, the four test designs in this study used comparable ice areas (9 cm2 to 11.2 cm2), but the ice masses differed significantly with IFAM~3 g << AMIL~7 g < TAU~8 g < P-A~9 g (see Table 6). These results were accompanied by increasing ice thickness and different ice shapes for the tests. However, correlations with ice adhesion test data were not observed for any of the tested materials. This also applies to the parameter shear stress evolution, which was lowest for IFAM (200 rpm/s), followed by P-A, TAU, and AMIL in a narrow range. The high complexity of parameter setups prevents the clear identification of dependencies, which remains an open topic in this study.

3.3. Mode I Tests (CU, ConU)

Mode I tests were conducted by two partners in this study. Impact ice was used, and test parameters were harmonized between the facilities (see Table 6). The results, including means and standard deviations, are summarized in Figure 10 for the four test surfaces; these results are based on the raw data included in Appendix ATable A3.
The ranking for the test surfaces differed between the partners. The most significant difference was found for the Primer coating. The best agreement could be observed for the PTFE tape material.
The remaining parameter difference between the facilities after the harmonization process was the iced area; this measured 7.07 cm2 for CU and 12.6 cm2 for ConU (assuming that the ice masses are comparable). The calculated ice volumes were ~11 cm3 for CU and ~12 cm3 for ConU. In this study, the impact of the test parameter on the ice adhesion test results was unclear.

4. Discussion

This study aimed to compare ice adhesion test results using identical test surfaces but different test designs at 13 partner facilities. In the definition phase, the homogenization of test parameters was conducted, and test designs were grouped according to the ice types and ice removal techniques used. Table 7 summarizes the results and provides a material ranking based on the absolute values for the averaged ice adhesion strength data.
The Primer (high roughness and hydrophilic) and Standox (low roughness and hydrophilic) materials were expected to show higher ice adhesion test results compared to the PUR C25 (low roughness and hydrophobic) and the PTFE tape (high roughness and hydrophobic). The results from each partner facility were mainly in accordance with these expectations. However, the absolute values among the facilities differ significantly and do not allow for a general definition of a specific value for low ice adhesion surfaces.
For the mechanical tests, it was observed that icing times and/or geometries of the ice/test surface interface have significant effects on ice adhesion strength. The longer icing times (overnight) as well as the complete immersion of the test samples during the icing process in INTA and RSE tests led to higher shear forces, especially for PUR C25 and PTFE tape. For the remaining mechanical tests, reasonable comparability in ice adhesion strength data was identified between pull-based (RISE) and push-based (NTNU, ND, and CRREL) tests.
For the centrifuge tests in this study, the direct comparison of static and impact ice in tests at AMIL and IFAM did not show differences for the potential low ice adhesion surfaces. The comparison of all centrifuge test designs showed the most obvious differences for the ice-covered area, the ice mass and shape, and the evolution of the shear stress. These parameters may have contrary effects on the measurement data, preventing the identification of correlations between test parameters and ice adhesion strength results. Further systematic studies would improve our understanding of the most decisive test parameter.
The result comparison for mechanical shear tests and centrifuge tests with comparable ice preparation times (from 10 min for impact ice to 200 min for static ice) shows, for the PUR C25, a comparably narrow range of mean data: from a minimum of 41 kPa (IFAM centrifuge using static ice) to a maximum of 141 kPa (NTNU push-based test using static ice). For PTFE tape, the range of measurement data increased from 71 kPa (RISE pull-based test using static ice) to 243 kPa (P-A centrifuge test using impact ice). For the materials for which a high ice adhesion strength was expected (Standox and Primer), the result ranges increased further, also indicating a potential risk for false positive (low) ice adhesion evaluations. However, the available results provided no indications about how to mitigate these risks regarding a specific test design. For surface evaluations, it is thus recommended to perform different test methods, rather than focusing on one test, to reduce uncertainties.
The Mode I tests in this study showed significantly higher ice adhesion strength values. The test designs differed significantly from the mechanical and centrifuge tests. Failure types (tension vs. shear) and other differences were discussed. During the ice adhesion measurements for the Mode I tests, the respective ice wind tunnel, including the water spray system, was active. This is of high relevance for actual technical applications and may have led to completely different data, but our understanding of its relevance to the assessment of low ice adhesion surfaces needs to be improved.
The use of absolute ice adhesion values is discussed in the literature because of the uneven force distribution for most of the test designs, e.g., in [5,42]. Amongst others, this resulted in the introduction of the adhesion reduction factor (ARF) to set the values in relation to a benchmark material [13]. In this study, for each test method, the Standox material (as unmodified PUR material) was defined as the test-specific reference material. Regarding the ice adhesion value, the percentage of reduction was calculated for PUR C25 (Figure 11) and PTFE tape (Figure 12).
In this study, the percental reduction for PUR C25 ranged from 0% for ConU to −82% for RISE. No basic trend could be observed for ice types or test designs. The same trend was observed for the PTFE tape results, for which the resulting span was even larger compared to the respective Standox reference material: from +72% for the CRREL tension test to −80% for the RISE push test. CRREL prepared test samples in parallel to the IFAM preparations. Despite the highest possible diligence in the processes, differences in surface properties occurred, which affected the ice adhesion test results, thus highlighting the necessity of single-source surface preparations for comparison tests.
In conclusion, this study highlights the extreme difficulty of comparing ice adhesion measurements not only among different methodologies, but also within the same measurement techniques. A harmonization of parameters, especially from the method of ice formation on the test surface, and an optimization of measurement parameters to lower the standard deviation should be performed in order to produce data that are more comparable among laboratories. Based on the results of this study, it is unlikely that a unified test standard for ice adhesion measurements can be developed. Future topics should address the identification of the most relevant test parameters and work on correction factors in the dependence of ice adhesion measurement techniques, preferably supported by data from relevant technical applications for icephobic materials. In addition, test conditions should be selected to be as similar to application conditions as possible.

Author Contributions

Conceptualization, N.R. and V.S.; Data curation, N.R.; Investigation, N.R., J.-D.B., M.Y., M.B., H.K., J.M., J.H., M.-L.P., A.D., E.A.-S., M.J. and X.H.; Methodology, N.R.; Project administration, N.R.; Resources, N.R. and E.A.-S.; Supervision, V.S.; Validation, J.-D.B., H.S., M.B., H.K., J.M., J.H., M.-L.P., A.D., E.A.-S., M.J., X.H. and V.S.; Visualization, N.R.; Writing—original draft, N.R., J.-D.B., M.Y., H.K., J.M., J.H., M.-L.P., A.D., E.A.-S., M.J. and X.H.; Writing—review and editing, H.S., M.B. and V.S. All authors have read and agreed to the published version of the manuscript.

Funding

The RSE activity has been financed by the Research Fund for the Italian Electrical System under the Three-Year Research Plan 2022–2024 (DM MITE n. 337, 15.09.2022), in compliance with the Decree of 16 April 2018. Further contributors to this research received no external funding.

Data Availability Statement

The data presented in this study are available in Appendix A of this manuscript.

Acknowledgments

The authors would like to thank their organizations and institutions for providing the test infrastructure and supporting this activity. We thank the Fraunhofer IFAM for the financial support to prepare and deliver the test specimen to the contributing partners.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A

Table A1. Raw data [kPa] from ice adhesion tests—direct mechanical tests.
Table A1. Raw data [kPa] from ice adhesion tests—direct mechanical tests.
Contributor RSEINTARISENTNUNotreDameCRRELCRREL
Test TypePull-Out/Static
Temperature (Ice Formation −19 °C)
Pull/StaticPull/StaticPush/StaticPush/StaticPush/StaticTension/Static
Material#Run 1#Run 1MeanStdev#Run 1Run 2Run 3MeanStdev#Run 1MeanStdev#Run 1MeanStdev#Run 1Run 2Run 3MeanStdev#MeanStdev#MeanStdev
Primer158611577 1621579426 1coh 203 207225252
256112490 2668500574 2coh 204
351013624 3705495584 3coh 179
455314402 4484479537 4coh
548915532
643616606
759417677
850418no data
959319600
105622043854471 55483 coh 19514 22823 1339 262
Standox2160126814 8148447589 8301 308 288292310
22no data27663 9163437647 9343 232
2368828667 10300568768 #369 282
2469329678 11395647553 #390
257513074770062 472195 35139 27439 29712 1177 191
PUR C253115836401 12258442442 #86 171 727983
3227037100 13268332453 #42 114
3329138290 14211305321 #71 139
3419439301 15142321416 #52
3531740415274100 32699 6319 14128 786 no datano datano datano data
PTFE-Tapeno data 5337326384 581 154 155148171
6311232353 6109 170
no datano data7258321384323517237144 24218947 15812 1028 321
Q-test analysis with aberrant 99% coh = cohesive failure
Table A2. Raw data [kPa] from ice adhesion tests—centrifuge test.
Table A2. Raw data [kPa] from ice adhesion tests—centrifuge test.
Contributor Nottingham UniversityIFAMAMILAMILTAUIFAMP-A
Test TypeCentrifuge/StaticCentrifuge/StaticCentrifuge/StaticCentrifuge/15 m/sCentrifuge/15 m/sCentrifuge/40 m/sCentrifuge/40 m/s
A = Aluminum; C = Composite
Material#Run 1Run 2MeanStdev#Run 1Run 2Run 3Run 4Run 5Run 6Run 7MeanStdev#Run 1Run 2Run 3MeanStdev#Run 1Run 2Run 3MeanStdev#Run 1MeanStdev#Run 1Run 2Run 3Run 4Run 5Run 6Run 7Run 8MeanStdev#Run 1Run 2Run 3MeanStdev
Primer1397378 Z1cohcohcoh AMIL1236234274 AMIL1411499525 176 Z1cohcohcoh A3358no data290A:
2469410 Z2cohcohcoh AMIL2225284196 AMIL2434386421 256 Z2cohcohcoh A428739625631758
3397372 Z3cohcohcoh 3110 Z3cohcohcoh C2411299237
4416345 449 C3366438no dataC:
5410416 C1029048042536885
6384497
741042940938 coh 24233 44654 7327 coh A+C:34977
Standox13106148 Z9218 AMIL3303397551 AMIL3298327363 8117 Z11111157 C676205no data
1470106 Z10148 AMIL4472220440 AMIL4343341360 985 Z1318111664118 C7197335475
15119109 Z11204165 11109 A164144 C11194475447
1696133 Z13165203 A3100128
17102106 A1189
10921A2191 18524 397119 33924 10417 11837 301154
PUR C25185558 Z14282959 AMIL5132131107 AMIL5122104122 1282 Z145054583243 A1114129161A:
197863 Z1529 AMIL613111082 AMIL6107109115 1361 Z15518248 A2369412811043
207363 Z17243272453937 14no data Z183958 C11177974
217078 Z1861172955 1567 C4727464C:
227387 B25240 C55485757717
7010 4115 11620 1138 7011 5214 A+C:9033
PTFE-Tape8164148 Z5160123 no data no data 538 Z513189 C8no data278258
9168181 Z6143147124129158144111 632 Z6133122 C9no data282281
10119189 Z7105105129 791 Z8941101411039580121128 C12104281215
11144207 Z8150122118
1216014816325 13118 no data no data 5432 11220 24366
Ice shed may have been caused by impact of ice from adjacent sample
Table A3. Raw data [kPa] from ice adhesion tests—Mode I tests.
Table A3. Raw data [kPa] from ice adhesion tests—Mode I tests.
Contributor Concordia UniversityCranfield Univ.
Test TypeMode I/Impact 40 m/sMode I/Impact 40 m/s
Material#Run 1Run 2Run 3MeanStdev#Run 1Run 2Run 3Run 4Run 5MeanStdev
Primer1103713581124 1coh2740377035003880
211429611427 2coh35304630cohcoh
31236106410341154157 3675615
PTFE-Tape4170015581635 419902080224025002320
5187417651422 52320
61524167821531701216 2242184
Standox7233628292193 734303120413027803530
8268225002524 850003170367035603250
92831252425752555210 3564622
PUR C2510207127942427 1039602280231026302380
11221127392689 1136302880264031003740
122818319221372564375 2955625

References

  1. Antonini, C.; Innocenti, M.; Horn, T.; Marengo, M.; Amirfazli, A. Understanding the Effect of Superhydrophobic Coatings on Energy Reduction in Anti-Icing Systems. Cold Reg. Sci. Technol. 2011, 67, 58–67. [Google Scholar] [CrossRef]
  2. Kulinich, S.A.; Farhadi, S.; Nose, K.; Du, X.W. Superhydrophobic Surfaces: Are They Really Ice-Repellent? Langmuir 2011, 27, 25–29. [Google Scholar] [CrossRef] [PubMed]
  3. Kreder, M.J.; Alvarenga, J.; Kim, P.; Aizenberg, J. Design of Anti-Icing Surfaces: Smooth, Textured or Slippery? Nat. Rev. Mater. 2016, 1, 15003. [Google Scholar] [CrossRef]
  4. Golovin, K.; Kobaku, S.P.R.; Lee, D.H.; Di Loreto, E.T.; Mabry, J.M.; Tuteja, A. Designing Durable Icephobic Sur-faces. Sci. Adv. 2016, 2, e1501496. [Google Scholar] [CrossRef] [PubMed]
  5. Wohl, C.J.; Berry, D.H. Contamination Mitigating Polymeric Coatings for Extreme Environments; Wohl, C.J., Berry, D.H., Eds.; Part II: Ice Contamination-Mitigating Coatings; Advances in Polymer Science 284; Springer Nature: Cham, Switzerland, 2019; pp. 52–214. [Google Scholar]
  6. Asadollahi, S.; Farzaneh, M.; Stafford, L. On the icephobic behavior of organosilicon-based surface structures de-veloped through atmospheric pressure plasma deposition in nitrogen plasma. Coatings 2019, 9, 679. [Google Scholar] [CrossRef]
  7. Huang, X.; Tepylo, N.; Pommier-Budinger, V.; Budinger, M.; Bonaccurso, E.; Villedieu, P.; Bennani, L. A survey of icephobic coatings and their potential use in a hybrid coating/active ice protection system for aerospace applications. Prog. Aerosp. Sci. 2019, 105, 74–97. [Google Scholar] [CrossRef]
  8. Liu, G.; Yuan, Y.; Liao, R.; Wang, L.; Gao, X. Fabrication of a porous slippery icephobic surface and effect of lubri-cant viscosity on anti-icing properties and durability. Coatings 2020, 10, 896. [Google Scholar] [CrossRef]
  9. Esmeryan, K.D. From ExtremelyWater-Repellent Coatings to Passive Icing Protection—Principles, Limitations and Innovative Application Aspects. Coatings 2020, 10, 66. [Google Scholar] [CrossRef]
  10. Milles, S.; Vercillo, V.; Alamri, S.; Aguilar-Morales, A.I.; Kunze, T.; Bonaccurso, E.; Lasagni, A.F. Icephobic per-formance of multi-scale laser-textured aluminum surfaces for aeronautic applications. Nanomaterials 2021, 11, 135. [Google Scholar] [CrossRef]
  11. Parent, O.; Ilinca, A. Anti-icing and De-icing Techniques for Wind Turbines: Creitical Review. Cold Reg. Sci. Technol. 2021, 65, 88–96. [Google Scholar] [CrossRef]
  12. Mora, J.; García, P.; Carreño, F.; González, M.; Gutiérrez, M.; Montes, L.; Agüero, A. Setting a comprehen-sive strategy to face the runback icing phenomena. Surf. Coat. Technol. 2023, 465, 129585. [Google Scholar] [CrossRef]
  13. Laforte, C.; Beisswenger, A. Icephobic Material Centrifuge Adhesion Test. In Proceedings of the IWAIS 2005, Montréal, QC, Canada, 12–16 June 2005. [Google Scholar]
  14. Arianpoura, F.; Farzaneh, M.; Kulinich, S.A. Hydrophobic and ice-retarding properties of doped silicone rubber coatings. Appl. Surf. Sci. 2013, 265, 546–552. [Google Scholar] [CrossRef]
  15. Soltis, J.; Palacios, J.; Eden, T.; Wolfe, D. Evaluation of Ice-Adhesion Strength on Erosion-Resistant Materials. AI-AA J. 2015, 53, 1825–1835. [Google Scholar] [CrossRef]
  16. Janjua, Z.A.; Turnbull, B.; Choy, K.-L.; Pandis, C.; Liu, J.; Hou, X.; Choia, K.-S. Performance and Durability Tests of Smart Icephobic Coatings to Reduce Ice Adhesion. Appl. Surf. Sci. 2017, 407, 555–564. [Google Scholar] [CrossRef]
  17. Orchard, D.; Clark, C.; Chevrette, G. Reducing Aviation Icing Risk: Ice Adhesion Measurement in the NRC’s Altitude Icing Wind Tunnel. In Proceedings of the SAE AeroTech Conference, FortWorth, TX, USA, 26–28 September 2017. [Google Scholar]
  18. Rønneberg, S.; Laforte, C.; Volat, C.; He, J.; Zhang, Z. The effect of ice type on ice adhesion. AIP Adv. 2019, 9, 055304. [Google Scholar] [CrossRef]
  19. Brassard, J.D.; Laforte, C.; Guerin, F.; Blackburn, C. Icephobicity: Definition and Measurement Regarding Atmos-pheric Icing. In Contamination Mitigating Polymeric Coatings for Extreme Environments; Wohl, C.J., Berry, D.H., Eds.; Springer International Publishing: Cham, Switzerland, 2019; pp. 123–144. [Google Scholar] [CrossRef]
  20. Tetteh, E.; Loth, E. Reducing Static and Impact Ice Adhesion with a Self-Lubricating Icephobic Coating (SLIC). Coatings 2020, 10, 262. [Google Scholar] [CrossRef]
  21. Schulz, M.; Sinapius, M. Evaluation of Different Ice Adhesion Tests for Mechanical Deicing Systems; SAE Technical Paper 2015-01-2135; SAE International in United States: Warrendale, PA, USA, 2015. [Google Scholar] [CrossRef]
  22. Work, A.; Lian, Y. A Critical Review of the Measurement of Ice Adhesion to Solid Substrates. Prog. Aerosp. Sci. 2018, 98, 1–26. [Google Scholar] [CrossRef]
  23. Rønneberg, S.; He, J.; Zhang, Z. The need for standards in low ice adhesion surface research: A critical review. J. Adhes. Sci. Technol. 2020, 34, 319–347. [Google Scholar] [CrossRef]
  24. Emelyanenko, K.A.; Emelyanenko, A.M.; Boinovich, L.B. Water and ice adhesion to solid surfaces: Common and specific, the impact of temperature and surface wettability. Coatings 2020, 10, 648. [Google Scholar] [CrossRef]
  25. Nazifi, S.; Firuznia, R.; Huang, Z.; Jahanbakhsh, A.; Ghasemi, H. Predicitve model of ice adhesion on non-elastomeric materials. J. Colloid Interface Sci. 2023, 648, 481–487. [Google Scholar] [CrossRef] [PubMed]
  26. Stendardo, L.; Gastaldo, G.; Budinger, M.; Pommier-Budinger, V.; Tagliaro, I.; Ibánez-Ibánez, P.F.; Antonini, C. Reframing ice adhesion mechanisms on a solid surface. Appl. Surf. Sci. 2023, 641, 158462. [Google Scholar] [CrossRef]
  27. Nistal, A.; Sierra-Martín, B.; Fernández-Barbero, A. On the Durability of Icephobic Coatings: A Review. Materials 2024, 17, 235. [Google Scholar] [CrossRef] [PubMed]
  28. Hejazi, V.; Sobolev, K.; Nosonovsky, M. From superhydrophobicity to icephobicity: Forces and interaction analysis. Sci. Rep. 2013, 3, 2194. [Google Scholar] [CrossRef] [PubMed]
  29. Dou, R.; Chen, J.; Zhang, Y.; Wang, X.; Cui, D.; Song, Y.; Jiang, L.; Wang, J. Anti-icing coating with an aqueous lubri-cating layer. ACS Appl. Mater. Interfaces 2014, 6, 6998–7003. [Google Scholar] [CrossRef]
  30. Makkonen, L. Ice Adhesion—Theory, Measurements and Countermeasures. J. Adhes. Sci. Technol. 2012, 26, 413–445. [Google Scholar] [CrossRef]
  31. Rehfeld, N.; Speckmann, B.; Stenzel, V. Parameter Study for the Ice Adhesion Centrifuge Test. Appl. Sci. 2022, 12, 1583. [Google Scholar] [CrossRef]
  32. DIN EN ISO 2808:2019; Beschichtungsstoffe—Bestimmung der Schichtdicke. Beuth Verlag GmbH: Berlin, Germany, 2019.
  33. Rehfeld, N.; Speckmann, B.; Schreiner, C.; Stenzel, V. Assessment of Icephobic Coatings—How Can We Monitor Performance Durability? Coatings 2021, 11, 614. [Google Scholar] [CrossRef]
  34. DIN EN ISO 19403-2; Beschichtungsstoffe—Benetzbarkeit—Teil 2: Bestimmung der freien Oberflächenenergie fester Ober-flächen durch Messung des Kontaktwinkels. Beuth Verlag GmbH: Berlin, Germany, 2020.
  35. DIN EN ISO 19403-7; Beschichtungsstoffe—Benetzbarkeit—Teil 7: Messung des Kontaktwinkels bei Neigetisch-Experimenten (Abrollwinkel). Beuth Verlag GmbH: Berlin, Germany, 2020.
  36. Balordi, M.; Cammi, A.; Santucci de Magistris, G.; Chemelli, C. Role of micrometric roughness on anti-ice proper-ties and durability of hierarchical super-hydrophobic aluminum surfaces. Surf. Coat. Technol. 2019, 374, 549–556. [Google Scholar] [CrossRef]
  37. Ferrick, M.G.; Mulherin, N.D.; Haehnel, R.B.; Coutermarsh, B.A.; Durell, G.D.; Tantillo, T.J.; Welser, E.S.; Cano, R.J.; Smith, R.J.; Martinez, E.C. Double Lap Shear Testing of Coating Modified Ice Adhesion to Liquid Oxygen Food Line Bracket, Space Shuttle External Tank; No. ERDC/CRREL-TR-06-11; Engineering Research and Development Center Hanover nh Cold Regions Research and Engineering Lab: Hanover, NH, USA, 2006. [Google Scholar]
  38. Chernyy, S.; Järn, M.; Shimizu, K.; Swerin, A.; Pedersen, S.U.; Daasbjerg, K.; Makkonen, L.; Claesson, P.; Irutha-yaraj, J. Superhydrophilic Polyelectrolyte Brush Layers with Imparted Anti-Icing Properties: Effect of Counter ions. ACS Appl. Mater. Interfaces 2014, 6, 6487–6496. [Google Scholar] [CrossRef]
  39. Asenath-Smith, E.; Hoch, G.R.; Erb, C.T. Adhesion of freshwater columnar ice to material surfaces by crystallization from the melt. J. Cryst. Growth 2020, 535, 125563. [Google Scholar] [CrossRef]
  40. Lovell, A.R.; Hoch, G.R.; Donnelly, C.J.; Hodge, J.M.; Haehnel, R.B.; Asenath-Smith, E. Shear and Tensile Delamination of Ice From Surface: The Ice Ahdesion Peel Test (IAPT). ERDC/CRREL Tech. Note, TN-21-1; U.S. Army Engineer Research and Development Center, Cold Regions Research and Engineering Laboratory, USA. 2021. Available online: https://apps.dtic.mil/sti/trecms/pdf/AD1147237.pdf (accessed on 13 January 2024).
  41. ISO/TS 19392-6:2023 (E); Paints and Varnishes—Coating Systems for Wind-Turbine Rotor Blades—Part 6: Determination and Evaluation of Ice Adhesion Using Centrifuge. ISO: Geneva, Switzerland, 2023.
  42. Memon, H.; De Focatiis, D.S.A.; Choi, K.-S.; Hou, X. Durability enhancement of low ice adhesion polymeric coat-ings. Prog. Org. Coat. 2021, 151, 106033. [Google Scholar] [CrossRef]
  43. Koivuluoto, H.; Hartikainen, E.; Niemelä-Anttonen, H. Thermally Sprayed Coatings: Novel Surface Engineering Strategy towards Icephobic Solutions. Materials 2020, 13, 1434. [Google Scholar] [CrossRef] [PubMed]
  44. Niemelä-Anttonen, H.; Kiilakoski, J.; Vuoristo, P.; Koivuluoto, H. Icephobic Performance of Different Surface Designs and Materials. In Proceedings of the International Workshop on Atmospheric Icing of Structures, IWAIS2019, Reykjavik, Iceland, 23–28 June 2019; p. 5. [Google Scholar]
  45. Hammond, D.; Luxford, G.; Ivey, P. The Cranfield University Icing Tunnel. In Proceedings of the 41st Aerospace Sciences Meeting and Exhibit, Reno, NV, USA, 6–9 January 2003; AIAA 2003-901. Available online: https://arc.aiaa.org/doi/pdf/10.2514/6.2003-901 (accessed on 13 January 2024).
  46. Grasso, M.J. Development of a Mode I Test Rig for Quantitative Measurements of Ice Adhesion Using Tensile Stress. Master’s Thesis, Concordia University, Montreal, QC, Canada, 2019. [Google Scholar]
  47. Andrews, E.H.; Stevenson, A. Fracture energy of epoxy resin under plane strain conditions. J. Mater. Sci. 1978, 13, 1680–1688. [Google Scholar] [CrossRef]
  48. Andrews, E.H.; Lockington, N.A. The cohesive and adhesive strength of ice. J. Mater. Sci. 1983, 18, 1455–1465. [Google Scholar] [CrossRef]
  49. Pervier, M.L.A.; Hammond, D. Measurement of the fracture energy in mode I of atmospheric ice accreted on different materials using a blister test. Eng. Fract. Mech. 2014, 1, 223–232. [Google Scholar] [CrossRef]
  50. Rønneberg, S.; Zhuo, Y.; Laforte, C.; He, J.; Zhang, Z. Interlaboratory Study of Ice Adhesion Using Different Tech-niques. Coatings 2019, 9, 678. [Google Scholar] [CrossRef]
Figure 1. Test samples for the round-robin study after primer application.
Figure 1. Test samples for the round-robin study after primer application.
Aerospace 11 00106 g001
Figure 2. General view of the test equipment at AMIL. (A,B) are static and impact ice substrates, respectively. (C) shows a general view of the AMIL closed-loop ice wind tunnel. (D) shows a general view of the CAT vat. (E) shows the view inside the centrifuge.
Figure 2. General view of the test equipment at AMIL. (A,B) are static and impact ice substrates, respectively. (C) shows a general view of the AMIL closed-loop ice wind tunnel. (D) shows a general view of the CAT vat. (E) shows the view inside the centrifuge.
Aerospace 11 00106 g002
Figure 3. View inside the Fraunhofer IFAM ice lab with closed-loop ice wind tunnel. (a) Test sample preparation for ice accretion using a silicone mold (static ice, (b) left) or a specimen holder for ice wind tunnel insertion (impact ice, (b) right). (c) View inside the centrifuge with test sample with ice formation (upper) and test sample after fixation (lower).
Figure 3. View inside the Fraunhofer IFAM ice lab with closed-loop ice wind tunnel. (a) Test sample preparation for ice accretion using a silicone mold (static ice, (b) left) or a specimen holder for ice wind tunnel insertion (impact ice, (b) right). (c) View inside the centrifuge with test sample with ice formation (upper) and test sample after fixation (lower).
Aerospace 11 00106 g003
Figure 4. (a) Glaze ice formation on a sample surface and (b) the testing configuration using the centrifuge method at NU (adapted from [42]).
Figure 4. (a) Glaze ice formation on a sample surface and (b) the testing configuration using the centrifuge method at NU (adapted from [42]).
Aerospace 11 00106 g004
Figure 5. Icing wind tunnel (IWiT), centrifugal ice adhesion tester (CAT), and examples of mixed glaze ice accreted on test samples at Tampere University (TAU), Finland.
Figure 5. Icing wind tunnel (IWiT), centrifugal ice adhesion tester (CAT), and examples of mixed glaze ice accreted on test samples at Tampere University (TAU), Finland.
Aerospace 11 00106 g005
Figure 6. Mode I test rig installed in Cranfield icing wind tunnel (a), schematic drawing of the device used by CU (b), Mode I sample holder used by ConU (c), and ice accretion process on the substrate (d).
Figure 6. Mode I test rig installed in Cranfield icing wind tunnel (a), schematic drawing of the device used by CU (b), Mode I sample holder used by ConU (c), and ice accretion process on the substrate (d).
Aerospace 11 00106 g006
Figure 7. Results for mechanical ice adhesion tests using static ice. Bar exceeding the ice adhesion axis refers to the qualitative result: high ice adhesion strength, resulting in cohesive ice failure.
Figure 7. Results for mechanical ice adhesion tests using static ice. Bar exceeding the ice adhesion axis refers to the qualitative result: high ice adhesion strength, resulting in cohesive ice failure.
Aerospace 11 00106 g007
Figure 8. Results for centrifuge ice adhesion tests. Bars exceeding the ice adhesion axis refer to the qualitative result: high ice adhesion strength, resulting in cohesive ice failure.
Figure 8. Results for centrifuge ice adhesion tests. Bars exceeding the ice adhesion axis refer to the qualitative result: high ice adhesion strength, resulting in cohesive ice failure.
Aerospace 11 00106 g008
Figure 9. Shear stress evolution for centrifuge tests in this study. Graphs (IFAM—100 rpm/s, IFAM—300 rpm/s (in grey)) were adapted from [31].
Figure 9. Shear stress evolution for centrifuge tests in this study. Graphs (IFAM—100 rpm/s, IFAM—300 rpm/s (in grey)) were adapted from [31].
Aerospace 11 00106 g009
Figure 10. Results for Mode I ice adhesion tests.
Figure 10. Results for Mode I ice adhesion tests.
Aerospace 11 00106 g010
Figure 11. Ice adhesion reduction (%) for each test in this study, based on the test-specific results for Standox (reference) and PUR C25.
Figure 11. Ice adhesion reduction (%) for each test in this study, based on the test-specific results for Standox (reference) and PUR C25.
Aerospace 11 00106 g011
Figure 12. Ice adhesion reduction (%) for each test in this study, based on the test-specific results for Standox (reference) and PTFE tape.
Figure 12. Ice adhesion reduction (%) for each test in this study, based on the test-specific results for Standox (reference) and PTFE tape.
Aerospace 11 00106 g012
Table 1. List of contributors to the ice adhesion round-robin tests.
Table 1. List of contributors to the ice adhesion round-robin tests.
InstitutionUsed
Abbreviation
CountryIce TypeTest Type
Power Generation Technologies and Materials DepartmentRSEItalyStaticShear: pull
Instituto Nacional de Técnica AeroespacialINTASpainStaticShear: pull
RISE Research Institutes of SwedenRISESwedenStaticShear: pull
Norwegian University of Science and TechnologyNTNUNorwayStaticShear: push
University of Notre DameNDUSAStaticShear: push
Cold Regions Research & Engineering LaboratoryCRRELUSAStaticShear push and
tensile peel
Université du Québec à Chicoutimi AMILCanadaStatic + impactCentrifuge
Fraunhofer IFAMIFAMGermanyStatic + impactCentrifuge
University of NottinghamNUUKStaticCentrifuge
Tampere UniversityTAUFinlandImpactCentrifuge
Partner-AP-A---ImpactCentrifuge
Concordia UniversityConUCanadaImpactMode I
Cranfield UniversityCUUKImpactMode I
Table 2. Surface properties of the 4 selected materials (IFAM preparations).
Table 2. Surface properties of the 4 selected materials (IFAM preparations).
PrimerStandoxPUR C25PTFE Tape
SFE (mN/m)38.5 (±0.3)36.0 (±1.0)18.0 (±0.8)15.1 (±0.4)
WCA (°)83 (±1)86 (±1)100 (±1)110 (±2)
WSA (°)>9067 (±4.7)41 (±2.8)29 (±3.4)
CAH (°)40 (±2.3)36 (±3.1)26 (±1.8)22 (±2.1)
Ra (µm)1.5 (±0.07)0.07 (±0.007)0.05 (±0.010)0.11 (±0.015)
Rz (µm)8.3 (±0.26)0.38 (±0.035)0.29 (±0.038)0.68 (±0.088)
Description:Hydrophilic
roughness: high
Hydrophilic
roughness: low
Hydrophobic
roughness: low
Hydrophobic
roughness: moderate
Table 3. Surface properties of the 3 selected materials (CRREL preparations).
Table 3. Surface properties of the 3 selected materials (CRREL preparations).
PrimerStandoxPTFE Tape
WCA (°)87 (±4.7)88 (±4.1)95 (±1.5)
WSA (°)88 (±4.6)57 (±7.5)35 (±2.5)
CAH (°)30 (±16.0)35 (±5.4)20 (±1.9)
Ra (µm)0.49 (±0.02)0.67 (±0.21)0.09 (±<0.01)
Rz (µm)2.36 (±0.07)6.92 (±1.38)0.53 (±0.03)
Comparison with IFAM preparation:Wettability: comparable,
roughness: lower
Wettability: comparable,
roughness: higher
Wettability: lower,
roughness: comparable
Table 4. List of direct ice adhesion test methods used by the round-robin contributors.
Table 4. List of direct ice adhesion test methods used by the round-robin contributors.
RSE pull test device
Aerospace 11 00106 i001Aerospace 11 00106 i002
INTA double-lap shear test
Aerospace 11 00106 i003Aerospace 11 00106 i004
RISE modified slip/peel method
Aerospace 11 00106 i005Aerospace 11 00106 i006
NTNU vertical shear test
Aerospace 11 00106 i007Aerospace 11 00106 i008
Notre Dame (ND) push test device
Aerospace 11 00106 i009Aerospace 11 00106 i010
CRREL ice adhesion peel test–
shear test method
Aerospace 11 00106 i011Aerospace 11 00106 i012
CRREL ice adhesion peel test–
tensile test method
Aerospace 11 00106 i013Aerospace 11 00106 i014
Table 5. Summary of test parameters for tests using static ice.
Table 5. Summary of test parameters for tests using static ice.
Test
Facility
Test NameSample
Geometry
Iced Area
[cm2]
Freezing from…Ice
Characterization
Temp.
[°C}
Icing Time
[min]
De-Ionized WaterHandling of
Iced Samples
Conditioning Time after Sample Handling:
X min
Test Temp. [°C]Test Specific Information
PULLRSEshear: pullcylindrical17.02Aluminum moldsclear ice: mass 50 g−8overnightY5−8Ice formation −19 °C; displacement speed: 0.3 mm/s
INTADouble
lap shear
flat17.5bulk water, sealed mold sidesclear ice, mass 3.5 governightY5displacement speed: 0.3 mm/s
RISEshear: pullflat1bulk waterplastic cuvetteclear ice mass: 0.9 g/sample180Y2–5displacement speed: 0.3 mm/s
PUSHNTNUice shear testflat6.15bulk water, poly-propylene molds sealed by siliconeclear ice−8120Y5−8displacement speed: 0.3 mm/s
NDshear: pushflat5.07water in steel ring (Ø 1 inch)clear ice180
(distilled)
Y5displacement speed: 0.3 mm/s
CRRELpeel test
(shear)
flat12Mold free crystallization from the meltclear ice, columnar90N5displacement speed: 0.01 mm/s
TENSIONCRRELpeel test
(tension)
flat12Mold free crystallization from the meltclear ice, columnar−890N5−8displacement speed: 0.01 mm/s
CENTRIFUGEAMILcentrifugeflat11.2bulk water
silicone moulds
clear ice,
about 7 g
−835Y20−8Radius 17 cm;
acceleration 300 rpm/s
IFAMcentrifugeflat9bulk water
silicone moulds
clear ice,
mass 3 g
90Y15Radius 11 cm; accelaration 200 rpm/s
NUcentrifugeflat1.38bulk water
silicone moulds
glaze ice,
mass 1.31 g
180Y5Radius 16.75 cm; acceleration 30 rpm/s
Table 6. Summary of test parameters for tests using impact ice.
Table 6. Summary of test parameters for tests using impact ice.
Test FacilityTest NameSample
Geometry
Iced Area
[cm2]
Freezing from…Ice
Characterization
Temp. [°C]Velocity [m/s]LWC [g/m3]MVD [µm]De-Ionized WaterHandling of
Iced samples
Conditioning Time after Sample Handling:
X min
Test
Temp.
[°C]
Test Specific Information
CENTRIFUGEAMILcentrifugeflat11.2supercooled droplets8 mm (±2 mm) thick−8150.827Y10−8Radius 17 cm; acceleration 300 rpm/s
TAUcentrifugeflat9supercooled droplets~9.5 mm thick; ice mass~8 g150.820Y17 hRadius 17 cm;
acceleration 200 rpm/s
IFAMcentrifugeflat9supercooled dorplets~4 mm thick; ice mass~3 g401.320Y15Radius 11 cm;
acceleration 200 rpm/s
P-Acentrifugenot specified10supercooled dropletsice mass 9–10 g400.520
(distilled)
Nnot specifiedRadius 18.5 cm, acceleration 150 m/s
MODE IConUMode Icylinder end12.6ice wind tunnel10 mm thick−8400.520N“spray on” during testing−8pressure rise
10 bar/s
CUMode Icylinder end7.0715 mm thick400.520N
Table 7. Summary of ice adhesion test results in this study, including a material ranking for increasing means of ice adhesion strength, with cells colored in white < light grey < dark grey < black.
Table 7. Summary of ice adhesion test results in this study, including a material ranking for increasing means of ice adhesion strength, with cells colored in white < light grey < dark grey < black.
TYPETest
Facility
Test NameSample
Geometry
Ice TypeIce Adhesion [kPa]
Mean (stdev)
Difference from General Test Program
PrimerStandoxPUR C25PTFE Tape
PULLRSEShear: pullCylindricalStatic544
(71)
700
(62)
274
(100)
No dataIce formation at
−19 °C overnight
INTADouble-lap shearFlatStatic554
(83)
472
(195)
326
(99)
323
(51)
Overnight
PUSHNTNUIce shear testFlatStatic195
(14)
274
(39)
141
(28)
189
(47)
RISEShear: pushFlatStaticcohesive351
(39)
63
(19)
71
(44)
NDShear: pushFlatStatic228
(23)
297
(12)
78
(6)
158
(12)
CRRELPeel test
(shear)
FlatStatic133
(9)
117
(7)
No data102
(8)
Sample preparation at CRREL
TENSIONCRRELPeel test (tension)FlatStatic26
(2)
19
(1)
No data32
(1)
CENTRIFUGEAMILCentrifugeFlatStatic242
(33)
397
(119)
116
(20)
No data
IFAMCentrifugeFlatStaticcohesive185
(24)
41
(15)
131
(18)
NUCentrifugeFlatStatic409
(38)
109
(21)
70
(10)
163
(25)
CENTRIFUGEAMILCentrifugeFlatImpact
15 m/s
446
(54)
339
(24)
113
(8)
No data
TAUCentrifugeFlatImpact
15 m/s
73
(27)
104
(17)
98
(56)
54
(32)
IFAMCentrifugeFlatImpact
40 m/s
cohesive118
(37)
52
(14)
112
(20)
Partner P-ACentrifugeCylindricalImpact
40 m/s
349
(77)
301
(154)
90
(33)
243
(66)
MODE IConUMode ICylinder endImpact
40 m/s
1154
(157)
2555
(210)
2564
(375)
1701
(216)
CUMode ICylinder endImpact
40 m/s
3675
(615)
3563
(622)
2955
(625)
2242
(184)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Rehfeld, N.; Brassard, J.-D.; Yamazaki, M.; Sakaue, H.; Balordi, M.; Koivuluoto, H.; Mora, J.; He, J.; Pervier, M.-L.; Dolatabadi, A.; et al. Round-Robin Study for Ice Adhesion Tests. Aerospace 2024, 11, 106. https://doi.org/10.3390/aerospace11020106

AMA Style

Rehfeld N, Brassard J-D, Yamazaki M, Sakaue H, Balordi M, Koivuluoto H, Mora J, He J, Pervier M-L, Dolatabadi A, et al. Round-Robin Study for Ice Adhesion Tests. Aerospace. 2024; 11(2):106. https://doi.org/10.3390/aerospace11020106

Chicago/Turabian Style

Rehfeld, Nadine, Jean-Denis Brassard, Masafumi Yamazaki, Hirotaka Sakaue, Marcella Balordi, Heli Koivuluoto, Julio Mora, Jianying He, Marie-Laure Pervier, Ali Dolatabadi, and et al. 2024. "Round-Robin Study for Ice Adhesion Tests" Aerospace 11, no. 2: 106. https://doi.org/10.3390/aerospace11020106

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop