Free access
TECHNICAL PAPERS
Aug 31, 2010

Laboratory Investigation of Performance and Impacts of Snow and Ice Control Chemicals for Winter Road Service

Publication: Journal of Cold Regions Engineering
Volume 25, Issue 3

Abstract

This work evaluated the performance attributes and impacts of several alternative deicers along with traditional chloride-based deicers. Four Strategic Highway Research Program tests were conducted to evaluate the ice melting, ice penetration, and ice undercutting capabilities of the deicers of interest, and also their impact to the freeze-thaw resistance of portland cement concrete. Three additional novel methods were utilized to assess the friction coefficient of deiced concrete surfaces, thermal properties of the deicers, and corrosion effects of deicers to metals. The laboratory data shed light on the complexity and challenges in evaluating various deicers. To facilitate scientifically sound decision-making, the writers propose a systematic approach to integrate the information available regarding various aspects of deicers, and to incorporate agency priorities, which is expected to aid agencies in selecting or formulating their snow and ice control chemicals.

Introduction and Research Objectives

In the last two decades, the growing use of chemicals for snow and ice control raised concerns about their effects on motor vehicles, transportation infrastructure, and the environment (Rendahl and Hedlund 1992; Shi et al. 2009d, b; Staples et al. 2004). Motorists and trucking associations have become concerned about the corrosive impact of snow and ice control chemicals (also known as deicers) on their vehicles, because vehicular corrosion has been documented. Menzies (1992) estimated the corrosion of motor vehicles attributable to road salts to cost $2.8 to $5.6 billion per year, including the added manufacturing expense for vehicle protection and preventive maintenance and the cost of owner preventive maintenance and cosmetic-corrosion-related vehicle depreciation. A more recent study (Johnson 2002) estimated the deicer corrosion to cost an average of $32 per vehicle per year. In addition, deicers may cause corrosion damage to the transportation infrastructure such as bridge decks and steel bridges (Virmani 2002). A 1991 study (Transportation Research Board) estimated the cost of installing corrosion protection measures in new bridges and repairing old bridges in the Snowbelt states to fall between $250 and $650 million per year. Indirect costs were estimated to be much greater than the cost of corrosion maintenance, repair and rehabilitation (Yunovich et al. 2002). Vataliano (1992) estimated that the use of road salt to pose a cost of at least $615 per ton on the infrastructure and additional costs associated with vehicular corrosion, aesthetic damage and environmental damage.
Chloride-based salts are the most common chemicals used to serve as freezing-point depressants for winter road service applications (Blackburn et al. 2004). Sodium chloride (NaCl) is the most widely used chemical owing to its abundance and low cost. Magnesium chloride (MgCl2) brines are often used instead of NaCl because laboratory tests have demonstrated that they exhibit better ice melting performance at cold temperatures (Ketcham et al. 1996). According to field studies, calcium chloride (CaCl2) is more effective than NaCl owing to its ability to attract moisture and stay on the roads (Warrington 1998). However, some agencies choose not to use CaCl2 because it does not dry and can cause roads to become slippery (Perchanok et al. 1991). Often, commercially available, corrosion-inhibited versions of these chemicals are used to reduce damaging impacts on motor vehicles and infrastructure.
In recent years, highway agencies have begun to voice a preference for acetate-based deicers such as potassium acetate (KAc), calcium magnesium acetate (CMA), or calcium-magnesium-potassium acetate (CMAK), because they tend to decompose quickly and do not contain chloride (Fay et al. 2008). Acetate-based deicers tend to be much more expensive yet more effective, relative to chloride-based deicers. In addition, agro-based chemicals have begun to offer new options for snow and ice control applications since the late 1990s. They often come from the fermentation and processing of beet juice, molasses, corn, and other agricultural products (Cheng and Guthrie 1998; Better Roads 2001; Albright 2003) and are used either alone or as corrosion inhibitors for other winter maintenance chemicals (Nixon and Williams 2001). Agro-based additives increase cost but may last longer and can be more powerful at melting snow at cold temperatures (Fischel 2001).
According to a survey the writers conducted in 2007 (Fay et al. 2008), most maintenance agencies still depend on chlorides and abrasives for snow and ice control. Sand and other abrasives improve the traction of vehicle tires on snow-and ice-covered pavement by providing a temporary friction layer. NaCl (solid) was most frequently used, followed by abrasives, then MgCl2, agro-based, CaCl2, and finally other deicers. A recent study by Levelton Consultants (2007) indicates that deicing products used in the field are selected based on two main criteria: cost and performance under various road weather conditions; whereas generally there are more factors that need to be considered. Although the hidden costs of road salts to the infrastructure and surrounding environment can be substantial (Shi 2005), such costs are often ignored in formulating highway winter maintenance strategies. Some products for snow and ice control may cost less in regard to materials, labor, and equipment, but cost more in the long run as a result of their corrosion and environmental impacts. A considerable amount of research is needed to fill the knowledge gaps and establish a scientifically robust, defensible decision-making process for winter maintenance.
This work aims to evaluate various deicers in terms of their performance (i.e., ice melted, ice penetration depth, rate of ice undercutting, thermal properties, friction coefficient of deiced concrete surface) and their negative impacts on metals and concrete; to identify promising laboratory tests for evaluating deicers; and to demonstrate how such results can be utilized in a decision-making framework. The research methods and findings are presented as follows.

Experimental

Deicers Tested

For this research, the deicers and reagent-grade chemicals used in various laboratory tests are listed in Table 1. The reagent-grade NaCl (99%, solid) was purchased from ScienceLab.com (Houston). The noninhibited NaCl-based deicer (IceSlicer) and salt-sand mixture (with 10–25% NaCl) were both provided by the Colorado Department of Transportation (CDOT) out of its solid stockpile. IceSlicer consists of naturally occurring complex chlorides (mostly NaCl) and more than 40 trace minerals. The two MgCl2-based deicers were provided by the CDOT out of its liquid tank and by the vendor, respectively, with active ingredient concentration of 27–29%. The three agro-based deicers were provided by CDOT out of its liquid tank and by the vendors, respectively, containing high concentrations of either MgCl2 or CaCl2. The NaAc/NaFm deicer was a blend made from NAAC (solid, 97% anhydrous NaAc, from Cryotech, Fort Madison, IA) and Peak SF (98% granulated NaFm, from the Blackfoot Company, Toledo, OH) at 5050 weight ratio. The liquid deicer CF7 was obtained from Cryotech with 50% KAc. The solid CMA deicer was obtained from Cryotech with 96% of hydrated calcium magnesium acetate. The test solution of CMA, CF7, NAAC, and Peak SF deicers contained 18.4, 0.2, 6.6, and 11.4 mM chloride, respectively, based on chloride sensor measurements. The reagent-grade potassium formate, KFm (99%, solid), was purchased from Alfa Aesar (Ward Hill, MA).
Table 1. List of Deicers Used in Testing
DeicerReagent-grade (r)/commercial product (c)FormMain freezing point depressantActive ingredient concentration (wt%)Tests used
NaClrSolidNaCl99%SHRP 205.1, 3, 5, 8; tribometer; DSC
Salt/sandac10–25%Electrochemical test
NaCl-basedac100%SHRP 205.1, 3, 5, 8; tribometer; DSC; electrochemical test
MgCl2-based 1acliquidMgCl227–29%SHRP 205.2, 4, 6, 8; tribometer; DSC; electrochemica; test
MgCl2-based 2aSHRP 205.2, 4, 6, 8; tribometer; DSC
Agro-based 1acliquidMgCl226%SHRP 205.2, 4, 6; tribometer; DSC; electrochemical test
Agro-based 2CaCl232.4%DSC
Agro-based 3CaCl2NA
NaAc-basedcsolidNa2(C2H3O2)97%SHRP 205.1, 3, 5; tribometer; DSC; electrochemical test
NaFm-basedNa(CHOO)98%SHRP 205.1,3,5,8; tribometer; DSC
NaAc/NaFmc/cNa2(C2H3O2)/Na(CHOO)97%/98%SHRP 205.5,8; tribometer; DSC
KAc-basedcliquidK2(C2H3O2)50%SHRP 205.4, 6, 8; tribometer; DSC; electrochemical test
CMA-basedcsolidCaMg(C2H3O2)296%SHRP 205.8; DSC
KFm-basedrsolidK(CHOO)99%SHRP 205.8
a
Provided by CDOT out of stockpiles or liquid tanks.

Ice Melting Capacity of Deicers

Laboratory measurements of ice melting capacity of various deicers were conducted following the SHRP H205.1 and H205.2 test methods (Chappelow et al. 1992). The SHRP H205.1 test measures the ice melting capacity of solid deicer pellets spread randomly across an ice surface of uniform thickness. The results of the test provide a measurement of the ice melting capacity of the deicer relative to the generated brine, or melted ice. The test utilized 25 mL of deionized water to form a sheet of ice of uniform thickness in a 3.5 cm (radius) Petrie dish. Once frozen, ice extrusions from the ice surface were melted. The sample was then refrozen for 24 h to equilibrate. After equilibration at the desired temperature, 1 g of solid was broadcast over the ice specimen. At 10, 20, 30, 45, and 60 min after application of deicer, the generated brines were removed from the specimen dish and weighed. The generated brine was then reintroduced to the same specimen dish. The process of removal of brine, weight, and reintroduction was completed within 1 min for each sample. The process was completed at 0, -5, -18°C (32, 23, -0.4°F). Special consideration was taken to use separate weighing dishes for each deicer to avoid cross-contamination. For liquid deicing solutions (SHRP H205.2), similar procedures were followed, except that 0.9 g of liquid deicer was pipetted onto the ice surface.

Ice Penetration of Deicers

Laboratory measurements of ice penetration of various deicers were conducted following the SHRP H205.3 and H205.4 test methods. The SHRP H205.3 test measures the penetration behavior of solid deicer pellets on a cylinder of ice. The results of this test provide a measure of the penetration rate of the selected deicer over time.
The test requires a custom-built Plexiglas apparatus with holes drilled to form cavities (Fig. 1). A 0.50 in (1.3 cm)-thick Plexiglas with dimension 2×8 in (5×20cm) had 10 cavities drilled vertically along the margin using a 5/32 in (0.4 cm) drill bit to a depth of 3.5 cm with 1.7 cm spacing between cavities. The cavities were then enlarged at the top with a countersink bit to create a surface cone. These cavities were filled with deionized water and then frozen for 12–24 h. The solid deicers were placed in the same freezer to ensure equilibration to the test temperature. Once completely frozen, any extrusions of ice were melted. The top of the ice surface was then wiped with an absorbent tissue to remove all liquid water and the apparatus was replaced in the freezer for an additional 1–2 h to re-equilibrate. A pellet of solid deicer was weighed and placed on the surface of the ice specimen along with a tracer dye solution (McCormick, red) and was placed in the freezer. The penetration depth of each pellet was measured using a ruler at 10, 20, 30, 45, and 60 min. The process was carried out at 0, -12, and -20°C (32, 10.4, -4°F) in triplicate.
Fig. 1. SHRP ice penetration test apparatus
For liquid deicing solutions (SHRP H205.4), similar procedures were followed, except that 25 mLl of liquid deicer were dyed with 2 drops of (McCormick, red) food coloring. The dyed liquid deicer solution was mixed and allowed to equilibrate to the given test temperature during the freezing of the ice specimen. Three drops of dyed liquid deicer solution were pipetted onto the ice specimen. Often the penetration depth of dyed solutions was difficult to visually assess. When this occurred, a small metal probe was inserted in the cavity until it contacted the ice interface to determine penetration depth. The probe was maintained at the test temperature and did not contribute to any penetration or melting.

Ice Undercutting of Deicers

Laboratory measurements of ice undercutting of various deicers were conducted following the SHRP H205.5 and H205.6 test methods. The SHRP H205.5 test measures the ice undercutting ability of a solid chemical deicer. The results of the test provide an area where the ice-concrete bond has been broken by the deicer as a function of time.
Portland cement concrete (PCC) blocks were made 6×9×2 in (15.2×22.8×5.1cm) and allowed to cure in the mold for 48 h in a wet chamber. The concrete blocks were then demolded and placed in a saturated lime solution for 28 days to cure. Once removed from the lime solution, the substrate was washed and allowed to dry. A 1 in (2.54 cm) wide rubber strip was affixed to the exterior of the concrete block, forming a wall around the top side of the substrate. The interface between the substrate and the rubber wall was sealed using silicon sealant.
Subsequently, 150 mL of deionized water were allowed to freeze on top of a concrete block for 12–24 h. The solid deicers were placed in the freezer to equilibration to the test temperature. Two drops of (McCormick, red) food coloring were deposited at three equally spaced locations across the ice surface and were allowed to freeze for 3–4 h. Two weighed deicer pellets were placed at the center of each frozen red dye stain. Photographs were taken at 10, 20, 30, 45, and 60 min. The process was carried out at 0, -6, -10, and -16°C (32, 21.2, 14, 3.2°F) in triplicate.
The digital pictures taken of the undercut area were imported into Adobe Photoshop CS2. The magic wand, polygon selection, lasso selection, and pixel count tools were used to select the extent of undercut area. A pixel count was obtained of the given selection and an area of undercut substrate was calculated for each photo.
For liquid deicing solutions (SHRP H205.6), similar procedures were followed with the exception that, when the ice specimen was completely frozen, a 0.5 in metal rod was heated and pressed vertically at three evenly spaced locations across the specimen. The resulting liquid was wiped clean, leaving a shallow cylinder free of ice terminating at the concrete surface. Three drops of dyed deicer liquid solution were pipetted into the center of the three shallow holes.

Freeze-Thaw Testing of PCC in the Presence of Deicers

Laboratory measurements of changes to PCC through freeze-thaw testing in the presence of deicers were conducted following the SHRP H205.8 test method entitled “Test Method for Rapid Evaluation of Effects of Deicing Chemicals on Concrete” (1992) with minor modifications. The SHRP H205.8 test evaluates the effects of chemical deicing formulations and freeze-thaw cycling on the structural integrity of small test specimens of non-air-entrained concrete. The method quantitatively evaluates degradation of the specimen through weight loss measurements. This test method is not intended to be used in determining the durability of aggregates or other ingredients of the concrete.
The concrete mix design was specified following the ASTM C 672-91 standard (2003). An ASTM specification C150-07 (2007) Type I/II low-alkali portland cement (Ash Grove Cement Company, Clancy, MT) was used in this study. The chemical composition and physical properties of the cement are available from the writers’ previous paper (Shi et al. 2009e). The fine aggregates used were clean, natural silica sand sifted to allow a maximum aggregate size of 1.18 mm before proportioning and admixing. The coarse aggregates used were crushed limestone with a maximum size of 3/8 in (0.95 cm). The concrete mix design had a water-to-cement ratio (w/c) of 0.51, a coarse-aggregate-to-cement ratio of 2.36, and a fine-aggregate-to-cement ratio of 1.75. The coarse aggregates and fine aggregates were prepared to saturated-surface-dry (SSD) condition before mixing, featuring a moisture content of 0.55% and 3.6%, respectively. The fine and coarse aggregates were added to the 3-cubic-feet (86-L) mixer and mixed until a homogeneous mixture was obtained. Then the cement was added and mixed again until a homogeneous mixture was obtained. Next, water was added from a graduated cylinder and mixed until the concrete is homogeneous and of the desired consistency. The batch was remixed periodically during the casting of the test specimens and the mix container was covered to prevent evaporation. The fresh concrete featured a slump of nearly zero [by the ASTM C143 method (2005)] and air content of 3% after compaction [by the ASTM C173 method (2010b)]. Concrete specimens were made in 11/2 in diameter ×17/8 in height (3.8×4.8cm) polyvinyl chloride piping with a volume of 54cm3. At least four specimens were made for each deicer solution or control solution. To facilitate demolding, one saw cut was made through each specimen mold along the 17/8×(4.8cm) length axis. The cut was then covered with duct tape. In the first 24 h of molding, the concrete specimens were placed on a rigid surface under ambient temperature and at a relative humidity of approximately 50% and covered to present excessive evaporation of water. Next, the specimens were demolded and cured in a moist cure room with relative humidity of 98% for 27 days (which deviates from the curing regime specified by the SHRP method). The 28-day compressive strength of test cylinders [by the ASTM C873 method (2010a)] was 6,619 psi (45.6 MPa), well above the recommended 4,000 psi (27.6 MPa).
Once fully cured the concrete specimens were allowed to dry overnight and weighed. For each deicer, four concrete specimens were placed on a cellulose sponge inside a dish containing 310 mL of deicer solution and then covered with plastic wrap to avoid water evaporation and to slightly compress each test specimen into the sponge. The deicer solutions were made at a 3% by volume for liquids and 3% by weight solution for solids using deionized water. This 3-to-100 ratio was chosen to simulate the field dilution of liquid or solid deicers once applied onto the pavement. One end of each test specimen was in full contact with the sponge. The test specimens (along with the deicer sponge and dish) were placed in the freezer for 16 to 18 h at -17.8±2.7°C (-0.04°F), and every diluted deicer tested froze at this temperature once it was placed in the freezer for some time. Subsequently, the specimens (along with the deicer sponge and dish) were placed in the laboratory environment at 23±1.7°C (73.4 °F) and with a relative humidity ranging from 45 to 55% for 6 to 8 h, at which temperature every diluted deicer tested thawed once it was taken out of the freezer for some time. This cycle was repeated 10 times. The average heating rate and cooling rate was observed to be 0.4°C/min and 1.2°C/min, respectively. After complete thawing following the tenth cycle, test specimens were carefully removed from the dish, individually rinsed under running tap water, and hand-crumbled to remove any material loosened during the freeze-thaw cycling. The largest intact part of each test specimen was then placed in open air to dry for 24 h at 23±1.7°C (73.4 °F) and a relative humidity ranging of 45–55%. After drying, test specimens were weighed and the final weights recorded.

Friction Coefficient of Deiced Concrete Surface

Concrete samples were made (3 cm diameter ×1cm thick) with a w/c of 0.46 and aggregate and sand sizes no greater than 1/6cm. Samples were cured in the moist cure room for 28 days at approximately 100% humidity. The compressive strength of the concrete at seven days was tested to be 4,929 psi (34.0 MPa), by the ASTM C873 method.
Concrete samples and the deicers were brought to the CSM Instruments Laboratory (Needham, MA) where 1.7 mL of deionized water (approximately 2 mm thick) were allowed to freeze (approximately 30 min) on the concrete surface. The freezer temperature was -16.2±1.6°C (2.8±2.8°F) and the freezer humidity was 50.6±11.6 percent, and the lab temperature was maintained at 23.7°C (74.8°F) and 30% humidity. Deicers were then applied to ice surface and allowed to sit for 5 min in the freezer. Samples were then loaded onto the tribometer and run for 100 laps at 3cm/s with a 10 mm diameter and 5 Newtons of force applied. The pin was cleaned between each sample using water and isopropyl alcohol. The liquid deicers were diluted to 3% by volume and two drops were pipetted onto the surface. Solid deicer were weighted and applied so that by weight the deicer did not exceed 3% concentration if all the ice on the sample were melted. Each deicer was run at least in triplicate.

Thermal Properties of Deicers

Laboratory testing was conducted using a differential scanning calorimetry (DSC) to quantify the thermal properties of deicers. The DSC (TA Instruments Q200) was set to run from 25 to -60°C (77 to -76°F) with cooling/heating rates at 2°C per minute. All samples were run as liquids. Solid samples were made into a saturated liquid solution with deionized water. Samples were run at 1.51 dilution. Ten μL of each sample were pipetted into the aluminum test chamber and sealed, and then weighed. All samples were run in triplicate. Data were used from the warming cycle, which were less prone to the supercooling effect and featured better reproducibility than those from the cooling cycle.

Corrosion to Metals

Corrosion to mild steel [ASTM A36 (2008)] and galvanized guardrail steel (Trinity Highway Products) was measured using a Gamry Instruments Potentiostat with an eight-channel electrochemical multiplexer (ECMB). Deicer solutions were 3% by weight for solid and by volume for liquid samples, made with deionized water. Before testing, the metal samples were cleaned with acetone and deionized water, and then dried. Four samples of each metal type were placed in the deicer solution and the open circuit potential (OCP) was monitored for 48 h.
Electrochemical techniques provide an attractive alternative to the gravimetric method [e.g., Pacific Northwest Snowfighters Association (PNS)/National Association of Corrosion Engineers (NACE) test] in terms of allowing for rapid determination of corrosion rates of metals and revealing information pertinent to the corrosion mechanism and kinetics. As such, at 48 h of immersion in the deicer solution, the weak polarization curve of each metal sample was taken to rapidly measure the corrosivity of the deicers. Weak polarization is an experimental technique that measures the current-potential relationship of a metal in the electrolyte when an external potential signal (perturbation) is applied within ±120mV range of its corrosion potential (Ecorr) at a given sweeping rate. Such a current-potential plot, known as a potentiodynamic polarization curve, provides information on the corrosion mechanisms occurring at the metal-electrolyte interface and the instantaneous corrosion rate of the metal, based on which the corrosivity of various deicers can be evaluated.

Results and Discussion

Ice Melting Capacity of Deicers

The ice melting capacity of the selected deicers varied as a function of product type and test temperature. At 0°C, the differences in the 60-min ice melting capacity were relatively small for all liquid and solid deicer products (Fig. 2). At -5°C, the solid deicer products (one NaCl-based, one NaFm-based, and one NaAc-based) performed better than the liquid deicer products (one MgCl2-based and one agro-based). However, at -18°C, liquid deicers outperformed the solid deicers, and the solid deicers based on NaAc or NaFm failed to melt any ice. Overall, the reagent-grade salt [NaCl (r, s)] outperformed all the other products by performing well at all three temperatures. Nixon et al. (2007) reported that at 0°F (-18°C), MgCl2 slightly outperformed NaCl, and at 23°F (-5°C) NaCl performed the best, followed by MgCl2 and the agro-based one, which coincides with this study’s results.
Fig. 2. Results from the SHRP ice melting capacity test, 60 min after application of deicers
Although results from the SHRP ice melting capacity test method provide performance data that can be easily understood and used by field practitioners, reproducibility issues have been reported and the method has been modified in many studies to generate more consistent results (Goyal et al. 1989; Chappelow et al. 1993; Nixon et al. 2005). The rate of dissolution of solid deicers may have been a factor affecting reproducibility, which is dependent on the particle size and the amount of brine needed for reasonably accurate measurements. This study reduced the surface area of the ice based on recommendations by Chappelow et al. (1993) to limit the errors resulting from absorption, but this greatly limited the amount of brine generated, especially at colder temperatures.

Ice Penetration of Deicers

The ice penetration capability of the selected deicers also varied as a function of product type and test temperature. Overall, the liquid deicers outperformed the solids at all temperatures (Fig. 3). At 0°C, the two liquid MgCl2-based deicers, the KAc-based deicer, and the agro-based deicer all penetrated to the bottom within 30 min of the 60-min test. Penetration performance for all deicers gradually diminished as the temperature got colder.
Fig. 3. Results from the SHRP ice penetration test, 60 min after application of deicers
As a group, the solid deicers performed similarly well at 0°C, with the exception of the solid NaAc-based deicer, which failed to penetrate into ice at the three test temperatures. The solid NaFm-based deicer failed to penetrate into ice once the temperature dropped to -12°C. The solid NaCl-based deicer showed penetration at 0°C, but not at -12°C. At -18°C, the solid NaCl-based deicer pellet became lodged, but the generated brine became the agent behind the penetration. Nixon et al. (2007) reported much lower ice penetration rates at 30°F (0°C), less than 6 mm for all deicers tested, compared with 30 mm of ice penetration depth shown in Fig. 3. Yet, the relative deicer performance in these two studies was consistent, where MgCl2 and KAc showed good ice penetration and NaCl showed poor ice penetration.
The writers do not recommend this test method for evaluating solid deicers because the solid pellets often became lodged in the ice column and led to reproducibility issues. This is consistent with the study by Nixon et al. (2007), which indicated that the ice penetration test was not a useful specification test.

Ice Undercutting of Deicers

The ice undercutting abilities of solid deicers were very difficult to measure, as shown from the lack of data in Fig. 4. At temperature ranges between -16°C and -10°C, the solid NaCl-based deicer was the only deicer to reach the substrate; all other solid deicers either demonstrated no notable effect to the ice surface or only partially melted down to the substrate, with some refreezing by 60 min. The solid NaAc-based deicer showed some undercutting at -10°C, but was ultimately refrozen by 60 min. At 0°C, all test deicers generally reached the substrate, however, few generated sufficient brine to begin the undercutting process, i.e., the two liquid MgCl2-based deicers, the KAc-based deicer, and the agro-based deicer. This is consistent with the outstanding ice penetration performance of these four liquid deicers discussed earlier.
Fig. 4. Results from the SHRP ice undercutting test, 60 min after application of deicers
Overall, the liquid deicer products worked better at undercutting and breaking the ice-concrete bond than solids, but no one liquid deicer significantly outperformed others in the group. Generally, the deicers performed better at warmer temperatures, with the exception of the solid NaAc-based deicer showing its peak performance at -6°C. Salt brine (23% liquid, reagent grade), the NaFm-based deicer, and the sodium acetate/formate (NaAc/NaFm, 50% blend of the two)-based deicer were tested, but all showed no ice undercutting at the three test temperatures.
Although results from this test method provide insight and performance data that can be used to guide field applications, the method itself is plagued by reproducibility issues that are difficult to address, especially for solid deicers. In addition, many deicers initially appeared to be undercutting and breaking the ice-concrete bond, but in fact the dye was moving across the surface of the ice.

Freeze-Thaw Resistance of PCC in the Presence of Deicers

The presence of deicers generally degraded the freeze-thaw resistance of PCC. In this test program of various deicers, the controls were deionized water (DI-H2O) and NaCl (reagent grade, solid). Using the PCC samples with air content of greater than 3%, DI-H2O had little effect on the structural integrity of the concrete, whereas NaCl caused approximately 43% weight loss after the SHRP freeze-thaw test. Among the deicers tested, the solid NaCl-based deicer and the liquid KAc-based deicer caused the most weight loss of the concrete sample (approximately 50%), followed by the solid NaFm-based deicer and the solid NaAc/NaFm-based deicer (approximately 24%), whereas the liquid MgCl2-based deicer and the solid CMA-based deicer caused the least weight loss (as shown in Fig. 5).
Fig. 5. Weight loss of PCC specimen following the SHRP H205.8 freeze-thaw test in the presence of various solutions
The significant damage of PCC in the presence of NaCl or NaCl-based deicers can be explained by the role of NaCl in significantly accelerating the freeze-thaw cycles and thus exacerbating the physical and chemical damages. MgCl2 did not seem to have as much impact on the freeze/-thaw resistance of PCC, at least after 10 days of testing. On the other hand, multiple studies have suggested MgCl2-based deicers to be more destructive to PCC than NaCl-based deicers (Cody et al. 1996; Mussato et al. 2004). The difference in the findings can be attributed to the differences in the deicer concentrations (concentrated versus diluted), the PCC samples used (old cored concrete versus new concrete that feature different cement hydration products and concrete microstructure), the test procedures (wet-dry cycles and freeze-thaw cycles), and the test duration. Although previous studies used mostly concentrated deicer solutions, this study tested diluted deicers (assuming a 3-to-100 dilution ratio for all liquid and solid deicers), because the deicer concentrations experienced by the PCC structures and pavements in the field are generally low in the long term. Laboratory findings from either the SHRP H205.8 or other test methods do not necessarily accurately predict the deicer/concrete interactions in the field and research is still needed to establish the correlation between the laboratory data and the field data. Additional work at the writers’ laboratory examined the chemical changes in the cement paste after the SHRP freeze-thaw test and found that MgCl2 did react with the cement hydrates to form fibrous or needlelike crystals characteristic of magnesium chloride hydroxide hydrate (Shi et al. 2009c), which is consistent with the findings by Sutter et al. (2008), Cody et al. (1994, 1996), Deja et al. (1999), and Lee et al. (2000). Given longer exposure time, the MgCl2-based deicer might cause considerable damage to the concrete as well.
The alternative deicers tested all caused significant damage to the PCC, with the only exception being the CMA-based deicer. Because this study did not use aggregates susceptible to alkali silica reactivity (ASR), the exacerbated deterioration of PCC in the presence of KAc, NaFm, and NaAc merits further investigation since it cannot be explained by the mechanism proposed by Rangaraju et al. (2005, 2006).

Friction Coefficient of Deiced Concrete Surface

The friction coefficient on the ice was lower than the friction coefficient on the wet concrete surface (Fig. 6), showing that the ice is more slippery than wet concrete. Partly attributable to the immature nature of this test method, the friction test results had great variance and did not show any significant difference between liquid and solid deicers or between chloride-based deicers and the acetate-or formate-based deicers. Nonetheless, the tribometer test revealed that the agro-based deicer led to the lowest friction coefficients on both the ice and the deiced concrete, whereas the NaCl-based solid deicer had the greatest variance of friction coefficients on the ice. The latter may be attributed to the clay or aggregate material naturally embedded in the NaCl-based solid deicer.
Fig. 6. Box plots of tribometer data with friction coefficient on the y-axis and deicer type on the x-axis; the white boxes show the friction coefficient of deicers on the ice, and the gray boxes show the friction coefficient of the deiced concrete
The current practice for DOTs is to measure the friction coefficient of wet and icy roads with a friction wheel that is attached to the back of a vehicle on a trailer and towed. Al-Qadi et al. (2002) found that the use of friction measurements can improve winter maintenance operations and mobility, but using a device with an extra wheel may not be practical in the field. There have been recent advances in friction measurement devices, such that they can provide real-time data in order to facilitate the determination of deicer application rates based on road surface conditions. Bergström et al. (2003) also reported the use of a portable friction tester to measure friction on cycleways that was able to differentiate various winter road conditions and various pavement materials.
The quantification of friction coefficients of deiced pavement is a new application of the tribometer. Although the data in Fig. 6 demonstrate the feasibility of this technology, the tribometer and test procedures need to be modified to provide laboratory data that are reliable and representative of field experience. The tribometer holds promise for a potential standard test protocol to assess the friction implications of deicer products, if more research is conducted to establish its test results with field friction test data.

Thermal Properties of Deicers

The DSC was used to quantify the thermal properties of deicers. Fig. 7 shows the heating-cycle thermogram for six general types of deicers. The peak at the warmer temperature range for each thermogram defines the characteristic peak, which corresponds well with the effective temperature for the deicer. If the temperature got colder than the characteristic peak, ice crystals would begin to form and lead to icy pavement in the field environment. The DSC thermograms were very reproducible for each deicer at a given dilution rate and heating rate, and thus may serve as a fingerprint tool for quality assurance of deicers. For instance, the NaCl-based deicers were the only deicers consistently featuring two equally strong peaks in the warming cycle.
Fig. 7. DSC thermogram of the heating cycle of six general types of deicers
The two peaks observed in the NaCl-based deicer warming cycle represent two endothermic phase transitions, where the peak on the left represents the separation of the ice from the subcooled NaCl or the pseudoeutectic formation (Table 2) and the peak on the right represents the warmest temperature at which ice crystals begin forming (Huidt and Borch 1991; Koefod 2008). This may also help explain the two peaks observed in the NaFm and NaAc/NaFm-based deicer warming cycles.
Table 2. Effective Temperature and Heat Flow of Deicers (Based on the DSC Thermogram)
  Characteristic temperature peakSecond peak temperatureHeat flow, H(J/g)
DeicerOriginal stateAverage (°C)COVAverage (°F)Average (°C)COVAverage (°F)AverageCOVΔH
MgCl2-based 1Liquid-11.24-0.00411.768   95.05  
MgCl2-based 2Liquid-11.47-0.04311.354   97.53  
Agr-based 1Liquid-11.03-0.04712.146   97.26  
KAc-based 1Liquid-13.91-0.0286.962   90.46  
Agr-based 2Liquid-10.69-0.04112.758   108.87  
Agr-based 3Liquid-3.26-0.06426.132   169.57  
NaCl—basedSolid-4.55-0.01523.81-21.25-0.004-6.25191.770.01262.93
NaCl (reagent)Solid-4.71-0.01923.522-21.02-0.002-5.836179.70.032253.2
NaAc-basedSolid-7.34-0.02318.788   164.6  
NaFm-basedSolid-8.14-0.07617.348-24.56-0.10312.20843.550.042290
NaAc/NaFm-basedSolid-7.73-0.01818.086   205.77  
CMA-basedSolid-4.73-0.01123.486   179.67  
Table 2 shows the characteristic temperature and heat flow of the tested deicers obtained from the DSC thermograms. The table shows that KAc-based deicer had the coldest effective temperature, followed by the MgCl2-based deicer. Table 3 shows the ice melting capacity of the tested deicers from the SHRP test. The table shows that the solid NaCl and solid NaCl-based deicers had the highest ice melting capacity. In general, the SHRP ice melting data had more variance than the DSC data, as indicated by the higher numbers in the coefficients of variance (COV).
Table 3. Ice Melt Capacity for a Limited Number of Deicers at T = 0, –5, –18°C (Based on the SHRP Test)
  Ice Melting Capacity (g ice/g deicer)
  T=0°C (32°F)T=-5°C (23°F)T=-18°C (-0.40°F)
DeicerOriginal stateAverageCOVAverageCOVAverageCOV
MgCl2-based 1Liquid8.681.9%2.958.5%1.8010.0%
Agr-based 1Liquid7.876.3%2.579.8%1.3718.0%
NaCl-basedSolid10.9311.8%5.246.0%0.7327.6%
NaCl (reagent)Solid10.704.3%8.4815.7%1.4317.6%
NaAc-basedSolid9.172.7%5.272.9%0.000.0%
NaFm-basedSolid8.2719.2%4.738.0%0.000.0%
In Table 2, the change in heat flow (ΔH) was calculated by subtracting the total heat of fusion for pure water (334J/g) from the measured heat flow of the characteristic peak. Statistical analysis revealed the following correlation between the DSC data and the SHRP data at 0°C, as shown in Eq. (1):
Ice Melting Capacity=0.66×T+8.58×Log10(ΔH)-4.86(R2=0.91)
(1)
The positive coefficient associated with Log10(ΔH), i.e., 8.58, is consistent with the notion that the more powerful a deicer is, the less external heat it needs to melt the ice (and thus the higher value in ΔH). The high R-square value confirms that there is a strong correlation between the change in heat flow (ΔH) at the characteristic temperature measured with the DSC and the SHRP ice melting capacity at 0°C. The less-than-one R-square can be attributed to experimental error especially in the SHRP test or the different behavior between solid and liquid deicers in the SHRP test. As such, DSC may hold the promise for a reliable standard test protocol to assess the deicer performance under certain road weather conditions. More details related to the development of a DSC-based test protocol for quantifying the performance of liquid deicers can be found in a recent report (Akin and Shi 2010).

Corrosion to Metals

For deicers diluted at 3% by weight or volume (for solid and liquid deicers, respectively), electrochemical testing of their corrosion to mild steel and galvanized steel showed that acetate-based deicers were much less corrosive to mild steel than chloride-based deicers and the agro-based deicer. The corrosion rates [in milli-inches per year (MPY)] of mild steel for NaCl-, MgCl2-, and acetate-based deicers from our electrochemical test (weak polarization) were similar to those reported by Levelton Consultants (2007) who used another electrochemical test method, linear polarization resistance (LPR), to assess the deicer corrosivity to mild steel (ASTM A36).
In general, steel is considered to be passive when its corrosion current density icorr<0.1μA/cm2, and active corrosion occurs when icorr>1.0μA/cm2. As such, it can be concluded that the acetate-based deicers were noncorrosive to mild steel, whereas the chloride-based deicers and the agro-based deicer were very corrosive (Table 4). Nonetheless, the galvanized steel in the acetate-based deicers was found to be corroding at comparably high rates to the other deicers. For both metal types, the specific agro-based deicer product showed high corrosivity, showing little benefits of the agro-based additive and contradicting the claim by a previous study (Kahl 2004).
Data in Table 4 add to the existing knowledge base concerning the corrosive effects of deicers to metals and illustrate that the additives beneficial in reducing the corrosivity of deicers to one metal may not work for a different type of metal. Kennelley (1986) conducted electrochemical and weight loss tests of 14–17 month duration, which indicated that bridge structural metals, including steel, cast iron, aluminum, and galvanized steel corroded considerably less in CMA solutions than in NaCl solution. Callahan (1989) demonstrated that pure CMA could effectively inhibit chloride-induced corrosion of reinforcing steel, whereas CMA as an additive to NaCl did not inhibit the rebar corrosion in concrete. Ushirode et al. (1992) reported that NaAc, urea, and CMA were only marginally effective as corrosion inhibitors for reinforced concrete.
Table 4. Electrochemical Analysis of Deicer Effect to Mild Steel (A36) and Galvanized Steel (Guardrail)
 DeicerCorrosion Rate (MPY)Impedence (kohm.cm2)Ecorr (mV, SCE)Icorr (μA/cm2)
Mild steelMgCl2-based 12.7±1.12.5±0.5-616.0±1.86.0±2.5
Ag-based 14.7±1.92.5±0.5-639.5±6.010.2±4.2
NaCl-based8.1±0.61.7±0.1-745.5±3.017.8±1.4
Kac-based2.5E-03±9.1E-05950.0±50.0-155.3±30.25.5E-03±2.0E-04
NaAc-based7.1E-03±4.1E-03316.7±175.6-204.3±68.66.8E-02±9.3E-02
Salt/sand2.5±0.62.1±0.2-764.3±6.05.4±1.3
NaF-based2.5E-03±2.1E-0487.0±4.6-199.5±12.05.5E-03±6.0E-04
KF-based8.5±11.526.6±48.9-598.0±316.519.8±23.9
Galvanized steelMgCl2-based 11.7±0.21.6±0.3-1037.5±5.03.5±0.6
Ag-based 11.9±0.72.1±0.4-1010.0±8.24.4±0.9
NaCl-based0.9±0.21.5±0.0-1037.5±5.01.9±0.4
Kac-based1.7±0.63.9±1.9-1032.5±5.03.0±0.9
NaAc-based0.9±0.28.9±0.9-1035±19.11.8±0.3
Salt/sand0.8±2.0E-020.7±0.3-1047.5±5.01.6±0.1
NaF-based5.2E-02±5.9E-0217.0±8.5E-02-1003.3±8.30.2±9.1E-02
KF-based1.6±0.91.3±8.5E-02-1060.0±0.04.1±0.6
Note: The salt/sand blend has 10–25% NaCl by weight.

Multicriteria Decision-Making Framework for Deicer Selection

The writers intend to lay the groundwork for a framework that allows an agency to assess the various deicer products on the market or to formulate its own deicer product based on laboratory testing and multicriteria decision-making (MCDM). The notions of alternatives, multiple attributes, conflicts among criteria, decision weights and decision matrix from the deterministic, stochastic, or fuzzy MCDM methods (Triantaphyllou 2000) can be adopted for the selection or formulation of snow and ice control materials. The crux is to strike the right balance in meeting multiple goals of the highway agency, including safety, mobility, environmental stewardship, infrastructure preservation, and economics (Shi 2005).
The following paragraphs describe a deicer composite index that would allow winter maintenance managers to numerically evaluate deicers based on their agency priorities or local needs and constraints. As an example,the CDOT user priorities are applied to the composite index, which would allow the most applicable deicer to be selected based on its relative ranking. Each deicer attribute category or subcategory was assigned a decision weight based on the CDOT users’ input, and more details are available in the project final report by Shi et al. (2009a). As shown in Table 5, the deicer composite index for each deicer product is calculated by multiplying the relevant decision weights with the attribute values indicating where the product’s cost, performance or impacts fall in the specific category or subcategory.
Table 5. Example Decision Matrix Illustrating the Differences Between Three Deicer Products to be Considered by the CDOT
 Deicer attributes for decision-makingAverage decision weight (CDOT)Attribute valueComposite indices
 Noninhibited solid NaClInhibited liquid MgCl2K or Na acetate/formateNoninhibited solid NaClInhibited liquid MgCl2K or Na acetate/formate
Cost-effectivenessLow materials cost per lane mile, also including training, equipment, and material handling7.0098363.056.021.0
Safety/deicer performanceLow effective temperature, ability to use in higher temperatures at lower application rates, high ice melting capacity, and improved pavement friction7.0868442.556.728.3
Corrosion to metalsLow corrosion effect on mild steel, galvanized steel, aluminum, rebar, or dowel bar, and slow penetration into concrete7.9338923.863.571.4
Impacts on pavementOverall low impact on concrete pavement, including resistance to freeze-thaw, ASR, ACR, scaling, strength loss, and expansion8.5648534.268.542.8
 Overall low impact on asphalt pavement, including aggregates (ASR), binder, degradation and disintegration of asphalt pavement, and strength loss7.7677354.354.323.3
Impacts on the environmentOverall low impact on water quality, including total P/N/Cl, total organic carbon (TOC), biological oxygen demand (BOD), chemical oxygen demand (COD), and aquatic toxicity8.3877358.658.625.1
 Overall low impact on plants, including browning/singe, senescence/death, root issues, and native species secession7.8066946.846.870.2
 Overall low impact on soil, including conductivity, heavy metal leaching, microbes, and food web7.8966847.447.463.1
 Overall low impact on wildlife, including attraction, toxicity from ingestion, habitat, and migratory paths8.0636824.248.464.5
 Overall low impact on air quality, including PM 10 and deicer aerosols7.9299771.371.355.4
  Overall deicer composite index46.657.146.5
The decision weights are scaled between 1 and 10, with 1 being least important and 10 being most important. The attribute values are normalized between 1 and 10, which in this work were semiquantitatively estimated based on the overall examination of existing literature, and experimental data and were provided for concept demonstration purpose only. One could also quantitatively calculate these attribute values based on the experimental data alone, assuming that the laboratory test protocols used can effectively capture the relative performance or relative impact of the deicer products. For performance attributes, the attribute values ranged from 1 to 10, with 1 being the worst and 10 being the best. For impact attributes, the attribute values ranged from 1 to 10, with 1 being the most deleterious and 10 being the least.
As shown in Table 5, the deicer composite index was calculated to be 46.6, 57.1, and 46.5 for noninhibited NaCl, inhibited liquid MgCl2, and KAc-or NaAc/NaFm-based deicers, respectively. This illustrates the challenges still faced by the road maintenance agencies, given that none of the deicers evaluated is close to being perfect (which would have a deicer composite index of 100). With the CDOT user priorities, the inhibited liquid MgCl2 deicer products present a better alternative than either the noninhibited NaCl or the KAc-or NaAc/NaFm-based deicers. Nonetheless, such comparisons may no longer hold true when applied to a different agency that has user priorities different from those of CDOT.
This is a preliminary attempt to illustrate how user priorities and experimental data can be integrated into a defensible decision-making process for selecting or formulating snow and ice control materials. There are many caveats in this simplified MCDM framework, such as the use of laboratory data to predict field performance and impacts as well as the lack of an appropriate index to quantify environmental impacts, given that environmental concerns are often site-specific. At this point in time, there is no laboratory test method available for deicing or anti-icing performance and friction coefficient that directly correlates with the performance and friction of deicers in the field. As such, the existing laboratory tests can only provide a baseline to contrast various products under well-controlled conditions, and the findings derived from such tests need to used with caution. Additional research is needed before such a framework can be robust and reliable enough to be adopted by the DOT managers to guide their operations.

Conclusions

Understanding the performance characteristics and negative impacts of deicers is critical to effective and responsible winter maintenance operations. This paper utilized seven tests to quantify the performance of identified deicers and their negative impacts to concrete and metals. The SHRP ice melting, penetration, and undercutting tests together identified four best performing deicers that were all liquids, two of which were MgCl2-based, one agro-based, and one KAc-based. The SHRP freeze-thaw test indicated that the MgCl2-based product and CMA-based deicers had the least impact on PCC. The DSC test showed that the KAc-based deicer had the coldest effective temperature, followed by the MgCl2-based deicer. The tribometer test revealed that the agro-based deicer led to the lowest friction coefficients on both the ice and the deiced concrete, whereas the NaCl-based solid deicer had the greatest variance of friction coefficients on the ice. The acetate-based deicers were found to be noncorrosive to mild steel, but comparably corrosive to galvanized steel as the chloride-based deicers and the agro-based deicer.
All the laboratory test results should be taken with a grain of salt when trying to predict the relative field performance of deicers, because they do not take into account the mixing action and fate/transport of deicers in the field attributable to traffic, UV absorption, gradation and angularity of deicer particles, moisture content and density of snow, pavement type and condition, wind, relative humidity, and possibly other factors. Similarly, the laboratory tests may not reliably predict the various impacts of deicers in the field environment where many site-specific variables are at play. Continued research is needed to develop more reliable laboratory tests that can better simulate the field scenarios. One such research project has recently completed its first phase (Fay et al. 2010).
In addition to identifying two promising test methods (DSC and tribometer), these laboratory data shed light on the complexity and challenges in evaluating various deicers. To facilitate scientifically sound decision-making, a systematic approach is proposed to integrate agency priorities and the available information regarding various aspects of deicers, which is expected to help transportation agencies in selecting or formulating their snow and ice control chemicals.

Acknowledgments

The authors acknowledge the financial support provided by the Colorado Department of Transportation (CDOT) as well as the Research & Innovative Technology Administration (RITA) at the U.S. Department of Transportation for this work. The authors would like to thank the CDOT Research Study Managers Patricia Martinek and Roberto de Dios and the CDOT technical panel. We would also like to extend our sincere appreciation our colleagues at the Western Transportation Institute including Kevin Volkening, Marijean M. Peterson, Doug Cross, Chase Gallaway, Collins Lawlor and Dr. Tuan Anh Nguyen for their assistance with the laboratory testing.

References

Akin, M., and Shi, X. (2010). “Development of standardized test procedures for evaluating deicing chemicals.” Rep., Wisconsin DOT and the Clear Roads Program, Madison, WI.
Albright, M. (2003). “Changes in water quality in an urban stream following the use of organically derived deicing products.” Lake Reservoir Manage., 2(1), 1–10.
Al-Qadi, I. L., Loulizi, A., Flintsch, G. W., Roosevelt, D. S., Decker, R., and Wambold, J. C. (2002). “Feasibility of using friction indicators to improve winter maintenance operations and mobility.” National Cooperative Highway Research Program, NCHRP Web Document 53, 〈http://onlinepubs.trb.org/onlinepubs/nchrp/nchrp_w53.pdf〉 (July 1, 2011).
ASTM. (2003). “Standard test method for scaling resistance of concrete surfaces exposed to deicing chemicals.” ASTM C672/C672M-03, West Conshohocken, PA.
ASTM. (2005). “Standard test method for slump of hydraulic cement concrete.” ASTM C143/C143M-05, West Conshohocken, PA.
ASTM. (2007). “Standard specification for portland cement.” ASTM C150-07, West Conshohocken, PA.
ASTM. (2008). “Standard specification for carbon structural steel.” ASTM A36/A36M-08, West Conshohocken, PA.
ASTM. (2010a). “Standard test method for compressive strength of concrete cylinders cast in place in cylindrical molds.” ASTM C873/C873M-10a, West Conshohocken, PA.
ASTM. (2010b). “Standard test method for air content of freshly mixed concrete by the volumetric method.” ASTM C173-C173M-10b, West Conshohocken, PA.
Bergström, A., Åström, H., and Magnusson, R. (2003). “Friction measurement on cycleways using a portable friction tester.” J. Cold Reg. Eng., 17(1), 37–57.
Better Roads. (2001). “Agricultural byproduct deicers are here to stay.” 〈http://www.betterroads.com/〉 (Dec. 5, 2008).
Blackburn, R. R., Bauer, K. M., Amsler, D. E., Boselly, S. E., and McElroy, A. D. (2004). “Snow and ice control: Guidelines for materials and methods.” NCHRP Rep. 526, Midwest Research Institute, Kansas City, MO.
Callahan, M. R. (1989). “Deicing salt corrosion with and without inhibitors.” Transportation Research Record 1211, Transportation Research Board, Washington, DC, 89–99.
Chappelow, C. C., et al. (1993). “Evaluation procedures for deicing chemicals and improved sodium chloride.” Strategic Highway Research Program, National Research Council, National Academy of Sciences, Washington, DC.
Chappelow, C. C., and Darwin, D. (1992). Handbook of test methods for evaluating chemical deicers, Publication No. SHRP-H-332, National Research Council, Washington, DC.
Chappelow, C. C., McElroy, A. D., Blackburn, R. R., Darwin, D., de Noyelles, F. G., and Locke, C. E. (1992). “Handbook of test methods for evaluating chemical deicers.” Strategic Highway Research Program, National Research Council, National Academy of Sciences, Washington, DC.
Cheng, K. C., and Guthrie, T. F. (1998). “Liquid road de-icing environment impact.” Technical Rep. 498-0670 Prepared for Insurance Corporation of British Columbia, Levelton Engineering, Richmond, BC, Canada.
Cody, R. D., Cody, A. M., Spry, P. G., and Gan, G.-L. (1996). “Experimental deterioration of highway concrete by chloride deicing salts.” Environ. Eng. Geosci., 2(x), 575–588.
Cody, R. D., Spry, P. G., Cody, A. M., and Gan, G.-L. (1994). “The role of magnesium in concrete deterioration.” Rep. HR-355, Iowa Highway Research Board, Ames, IA.
Deja, J., and Loj, G. (1999). “Effects of cations occurring in the chloride solutions on the corrosion resistance of slag cementitious materials.” Proc.: Infrastructure Regeneration and Rehabilitation Improving the Quality of Life through Better Construction—A Vision for the Next Millennium, R. N. Swamy, ed., Sheffield Academy Press, Sheffield, UK.
Fay, L., Akin, M., Wang, S., Shi, X., and Williams, D. (2010). “Developing a test methodology that correlates laboratory testing and field performance in measuring performance characteristics and the friction coefficient of deicing and anti-icing chemicals (Phase 1).” Technical Rep. Prepared for Wisconsin DOT and Clear Roads Program.
Fay, L., Volkening, K., Gallaway, C., and Shi, X. (2008). “Performance and impacts of current deicing and anti-icing products: User perspective versus experimental data.” Proc., 87th Annual Meeting of Transportation Research Board, Transportation Research Board, Washington, DC, 08–1382.
Fischel, M. (2001). “Evaluation of selected deicers based on a review of the literature. “Rep. No. CDOT-DTD-R-2001-15, SeaCrest Group, Louisville, CO.
Goyal, G., Lin, J., and McCarthy, J. L. (1989). “Time, temperature, and relative humidity in deicing of highways using sodium chloride or magnesium chloride with a metal corrosion inhibitor.” Transportation Research Record 1246, Transportation Research Board, Washington, DC, 9–17.
Huidt, A., and Borch, K. (1991). “NaCl-H2O systems at temperatures below 273 K, studied by differential scanning calorimetry.” Thermochim. Acta, 175(x), 53–58.
Johnson, J. T. (2002). “Corrosion costs of motor vehicles.” 〈http://www.corrosioncost.com/pdf/transportation.pdf〉 (Dec. 5, 2009).
Kahl, S. (2004). “Agricultural by-products for anti-icing and de-icing use in Michigan.” SNOW04-009, 6th Int. Symp. on Snow Removal and Ice Control Technology, Transportation Research Board, Washington, DC, 552–555.
Kennelley, K. J. (1986). “Corrosion electrochemistry of bridge structural metals in calcium magnesium acetate.” Dissertation Abstracts International B, 47(4), 328.
Ketcham, S. A., Minsk, L. D., Blackburn, R. R., and Fleege, E. J. (1996). “Manual of practice for an effective anti-icing program: a guide for highway winter maintenance personnel.” Publication No. FHWA-RD-9-202, Army Cold Regions Research and Engineering Laboratory, Hanover, NH.
Koefod, S. (2008). “Eutectic depressants, relationship of eutectic, freezing point, and ice melting capacity in liquid deicers.” Transportation Research Circular, Surface Transportation and Snow Removal and Ice Control Technology, Transportation Research Board, Washington, DC, 73–84.
Lee, H., Cody, R. D., Cody, A. M., and Spry, P. G. (2000). “Effects of various deicing chemicals on pavement concrete deterioration.” Proc., Mid-Continent Transportation Symp., Iowa State Univ., Ames, IA.
Levelton Consultants. (2007). “Guidelines for the selection of snow and ice control materials to mitigate environmental impacts.” NCHRP Rep. 577, Transportation Research Board, National Research Council, Washington, DC.
Menzies, T. R. (1992). “National cost of motor vehicle corrosion from deicing salt.” Proc., CORROSION/91 Symp., Automotive Corrosion and Protection, Robert Baboian, ed., National Association of Corrosion Engineers, Houston.
Mussato, B. T., Gepraegs, O. K., and Farnden, G. (2004). “Relative effects of sodium chloride and magnesium chloride on reinforced concrete: State of the art.” Transportation Research Record 1866, Transportation Research Board, Washington, DC, 59–66.
Nixon, W. A., Kochumman, G., Qiu, L., Qiu, J., and Xiong, J. (2007). “Evaluation of deicing materials and corrosion reducing treatments for deicing salts.” Rep., Project TR 471, Iowa Highway Research Board, Ames, IA.
Nixon, W. A., Qiu, L., Qiu, J., Kochumman, G., and Xiong, J. (2005). “Ice melting performance for ice-control chemicals.” Proc., 84th Annual Meeting of the Transportation Research Board, Transportation Research Board, Washington, DC.
Nixon, W. A., and Williams, A. D. (2001). “A guide for selecting anti-icing chemicals.” IIHR Technical Rep. No. 420, IIHR—Hydroscience & Engineering, Iowa City, IA.
Perchanok, M. S., Manning, D. G., and Armstrong, J. J. (1991). “Highway de-icers: Standards, practices, and research in the Province of Ontario.” Research and Development Branch, Ministry of Transportation, Ontario, Canada.
Rangaraju, P. R., Sompura, K. R., and Olek, J. (2006). “Investigation into potential of alkali-acetate based deicers in causing alkali-silica reaction.” TRB 85th Annual Meeting, Transportation Research Board, Washington, DC.
Rangaraju, P. R., Sompura, K. R., Olek, J., Diamond, S., and Lovell, J. (2005). “Potential for development of alkali-silica reaction in presence of airfield deicing chemicals.” Proc., 8th Int. Conf. on Concrete Pavements, American Concrete Pavement Association, Washington, DC.
Rendahl, N., and Hedlund, S. (1992). “The influence of road deicing salts on motor vehicle corrosion.” Proc., CORROSION/91 Symp.: Automotive Corrosion and Protection, Robert Baboian, ed., NACE, Houston.
Shi, X. (2005). “The use of road salts for highway winter maintenance: An asset management perspective.” Proc., 2005 ITE District 6 Annual Meeting, Institute of Transportation Engineers, Washington, DC, 10–13.
Shi, X., et al. (2009a). “Evaluation of alternative anti-icing and deicer compounds using sodium chloride and magnesium chloride as baseline deicers.” Rep. No. CDOT-2009-01 for Colorado DOT, Denver.
Shi, X., Akin, M., Pan, T., Fay, L., Liu, Y., and Yang, Z. (2009b). “Deicer impacts on pavement materials: Introduction and recent developments.” Open Civ. Eng. J., 3, 16–27.
Shi, X., Fay, L., Peterson, M. M., and Yang, Z. (2009c). “Freeze-thaw damage and chemical change of a portland cement concrete in the presence of diluted deicers.” Mater. Struct., 43(7), 933–946.
Shi, X., Fay, L., Yang, Z., Nguyen, T. A., and Liu, Y. (2009d). “Corrosion of deicers to metals in transportation infrastructure: Introduction and recent developments.” Corrosion Reviews, 27(1-2), 23–52.
Shi, X., Yang, Z., Nguyen, T. A., Suo, Z., Avci, R., and Song, S. (2009e). “An electrochemical and microstructural characterization of steel-mortar admixed with corrosion inhibitors.” Sci. China, Ser. E: Technol. Sci., 52(1), 52–66.
Staples, J. M., Gamradt, L., Stein, O., and Shi, X. (2004). “Recommendations for winter traction materials management on roadways adjacent to bodies of water.” FHWA/MT-04-008/8117-19, Montana DOT, Helena, MT.
Sutter, L., Peterson, K., Julio-Betancourt, G., Hooton, D., Vam Dam, T., and Smith, K. (2008). “The deleterious chemical effects of concentrated deicing solutions on portland cement concrete.” Rep. for South Dakota DOT, Pierre, SD.
Transportation Research Board. (1991). “Highway de-icing: Comparing salt and calcium magnesium acetate.” Special Rep. 235, National Research Council, Washington, DC.
Triantaphyllou, E. (2000). Multi-criteria decision making methods: A comparative study, Springer, New York.
Ushirode, W. M., Hinatsu, J. T., and Foulkes, F. R. (1992). “Voltammetric behavior of iron in cement. Part IV. Effect of acetate and urea additions.” J. Appl. Electrochem., 22(3), 224–229.
Virmani, Y. P. (2002). “Corrosion costs and preventative strategies in the United States.” Publication FHWA-RD-01-156, Federal Highway Administration, Washington, DC.
Vitaliano, D. (1992). “Economic assessment of the social costs of highway salting and the efficiency of substituting a new deicing material.” J. Pol. Anal. Manag., 11(3), 397–418.
Warrington, P. D. (1998). “Roadsalt and winter maintenance for British Columbia municipalities.” Best management practices to protect water quality, Environmental Protection Agency, Washington, DC.
Yunovich, M., Thompson, N. G., Balvanyos, T., and Lave, L. (2002). “Corrosion costs of highway bridges.” 〈http://www.corrosioncost.com/pdf/highway.pdf〉 (Dec. 5, 2009).

Information & Authors

Information

Published In

Go to Journal of Cold Regions Engineering
Journal of Cold Regions Engineering
Volume 25Issue 3September 2011
Pages: 89 - 114

History

Received: Dec 25, 2009
Accepted: Aug 26, 2010
Published online: Aug 31, 2010
Published in print: Sep 1, 2011

Permissions

Request permissions for this article.

Authors

Affiliations

Laura Fay, M.Sc.
Corrosion and Sustainable Infrastructure Laboratory, Western Transportation Institute, Montana State Univ., P.O. Box 174250, Bozeman, MT 59717-4250.
Xianming Shi, Ph.D. [email protected]
P.E.
Civil Engineering Dept., Montana State Univ., 205 Cobleigh Hall, Bozeman, MT 59717-3900 (corresponding author). E-mail: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share