Sie sind auf Seite 1von 28

UPDATE ON ICAR 507: THE SIGNIFICANCE AND APPLICATION OF THE MICRO-DEVAL TEST

G. Daniel Williams MS Candidate in Civil Engineering, The University of Texas at Austin Austin, Texas, USA Kevin Hampel MS Candidate in Civil Engineering, The University of Texas at Austin Austin, Texas, USA John J. Allen Managing Associate Director, International Center for Aggregates Research, The University of Texas at Austin Austin, Texas, USA David W. Fowler Dean T.U. Taylor Professor in Civil Engineering, The University of Texas at Austin Austin, Texas, USA

ABSTRACT The aggregate industry needs a test that better correlates test results to field performance. Micro-Deval has shown potential as a good indicator for field performance. The micro-Deval wet abrasion test for coarse aggregate is studied in this project to determine the ability of the test to predict field performance for various uses and mineralogical backgrounds when used alone or in combination with other aggregate tests. Aggregate properties such as particle shape, surface texture, and mineralogy are studied to determine their effect on the amount of micro-Deval loss. Aggregates were obtained from across the United States and Canada with varying field performance ratings, uses, and mineralogy. Testing is underway and the micro-Deval test is showing promise as an indicator of field performance when used in combination with other aggregate qualifying tests.

Keywords: micro-Deval; abrasion; wet aggregates; coarse aggregate properties; field performance; wet attrition.

Williams, Hampel, Allen, and Fowler

INTRODUCTION From the 1920s to the 1940s, many tests such as the Los Angeles (L.A.) abrasion test (adopted by ASTM in 1939) and the sulfate soundness test (adopted by ASTM in 1931) were created by researchers and adopted by American Association of State Highway and Transportation Officials (AASHTO) and American Society of Testing and Materials (ASTM). These tests had fundamentally sound intentions attempting to model the forces experienced by the aggregates in field conditions, but acceptance limits were determined in relation to the range of test values experienced [1] due to a lack of field performance history and data. Most of these tests gained popularity despite this lack of correlation to field performance and were introduced into aggregate qualification standards. The L.A. abrasion and impact test (AASHTO T 96), for example, is the most widely specified test in North America to determine the impact and abrasion resistance of coarse aggregate [2]. Its development attempted to overcome the short comings of the Deval test which was established in 1878 and adopted by ASTM in 1908. The main short coming of the Deval test consisted of a lack of a correlation with the performance of aggregates in pavements [3]. Research has shown that the L.A. abrasion test is a poor indicator of field performance [4-6]. Senior and Rogers [7] have shown that brittle, crystalline particles tend to shatter under the impact load while fine-grained aggregates such as slates tend to absorb some of the impact and produce lower losses. One major disadvantage of the L.A. abrasion test is its inability to test aggregates in a moist environment. Aggregate in the field is rarely dry and the effects of moisture may significantly alter aggregate mechanical properties [7]. Larson et al. [8] studied the effects of moisture on aggregate tested in the L.A. abrasion apparatus. He found that running 250 revolutions dry followed by 250 revolutions wet produced a better correlation with field performance. He also noted that collecting the entire test specimen from the drum was difficult since the fines tended to adhere to the inside of the drum. These findings gave opening to a new abrasion test designed specifically to include the effects of moisture on the mechanical properties of aggregate.

Williams, Hampel, Allen, and Fowler

The micro-Deval test was developed in France in the 1960s to include the effects of moisture on the mechanical properties of aggregate. Although, the test did not gain popularity in the United States and Canada until the early to mid-1990s, the micro-Deval abrasion test is growing in use across North America. Many state and provincial agencies have begun using the test knowing that the introduction of water affects the behavior of some aggregates. The fact that aggregate in use is rarely dry combined with the relatively short time it takes to get a microDeval abrasion result has encouraged the use of this relatively new test procedure. Some have started using the test for comparison purposes while others have made the test procedure a supplement to qualification standards. Due to little or no correlation between some of the current tests and field performance, some of these agencies have adopted the micro-Deval test despite a lack of confidence in the recommended acceptance limits. In order to gain confidence in the test and set realistic limits for various uses of aggregate (e.g. in concrete pavements, hot mix asphalt, base courses, etc.) the micro-Deval test needs to be correlated with aggregates of known field performance. Early testing performed by Rogers and Senior with the Ontario Ministry of Transportation in the 1980s and 90s has shown a general trend correlating the field performance of aggregates with microDeval loss. While not perfect, the correlation is much better than other tests such as the L.A. abrasion. Correlation between the micro-Deval test results and field performance ratings should be established early. AASHTO has adopted the micro-Deval test procedure (TP 58) and other agencies are looking to adopt the test procedure as well. As more agencies develop acceptance standards, a clear understanding of the effect of aggregate mineralogy, shape, surface texture, and/or use on the micro-Deval abrasion loss is needed. For example, an aggregate used as a base course may not need to be subjected to the same acceptance limits as that same aggregate used in concrete. RESEARCH OBJECTIVE Prior to adoption of the test procedure and its specified allowable limits, careful consideration, testing, and correlation must be performed to reduce the number of aggregates that would be judged incorrectly. Aggregate producers will benefit through satisfactory aggregates
Williams, Hampel, Allen, and Fowler 3

not being labeled unacceptable. Departments of Transportation (DOTs) will gain by avoiding costly pavement repairs due to the degradation of an unsatisfactory aggregate which might have passed a high micro-Deval acceptance limit. It is beneficial to everyone involved in the aggregate industry to establish accurate acceptance limits early based on field performance. This project looks at the correlation of micro-Deval abrasion loss and field performance of aggregates as well as a number of other tests whose results were also correlated with field performance. Each aggregate sample will be subjected to a test suite composed of twelve tests including the micro-Deval, L.A. abrasion, magnesium sulfate soundness, Canadian Freeze-Thaw, AASHTO Freeze-Thaw, Aggregate Crushing Value, Wet Crushing Value, absorption, specific gravity, percent flat and elongated particles, percent crushed particles, and petrographic examination. The results of each of the tests mentioned are being compared with micro-Deval test results to see if a correlation between two tests is able to distinguish better acceptance limits with respect to field performance. AGGREGATE ACQUISITION Initial Survey The first step of the testing process was to determine the extent of knowledge and use of the micro-Deval test. A survey was distributed to the forty-eight contiguous state DOTs, most Canadian provincial transportation ministries, and a few aggregate producers. The survey aimed to obtain information on current aggregate qualifying tests in use, confidence in the results of these tests, knowledge of the micro-Deval test, interest in the micro-Deval test, and knowledge of aggregate sources which might be linked to failures in hot-mixed asphalt, Portland cement concrete, bases, and subbases. 52 of the 55 (95 percent) surveys sent were completed and returned. This high response rate along with the information attained on the form revealed considerable interest in the micro-Deval test. Every responder indicated interest in receiving project updates and findings as well as a copy of the report upon completion. The responses from the survey indicated that most agencies were familiar with micro-Deval, but many were unsure of the role of the micro-Deval test. Many responders, mentioned being familiar with National Center for Asphalt Technology (NCAT) report number 98-4 [9] or NCAT report number 02-09 [10].

Williams, Hampel, Allen, and Fowler

Acquisition Logistics The initial survey response indicated which aggregates could be provided. Along with the identification of sources, field performance ratings were also requested for each source. This gave the ICAR 507 project team an idea of the aggregates it would be receiving and what gaps would need to be filled by future aggregate requests. It was quickly found that providers were more likely to offer aggregates of good and fair field performance. Naturally, most providers did not want to claim their aggregate source as poor so more effort was put into obtaining aggregates of poor field performance. Attempts were also made to ensure there were enough sources for each usage category: hot-mixed asphalt, Portland cement concrete, and base/subbase. Because each use entails different applied forces from transportation, mixing, placement, compaction, and transportation loads, different acceptance criteria may be valid. Collection of aggregates of many usages was easily achieved as many sources were used in two or more categories previously listed. Finally, samples of numerous mineralogical backgrounds were sought. Based on the work performed by Cooley, Jr. et al. [10], it was found that granites, for example, may not need to be subjected to the same acceptance criteria as limestones and sandstones or gravels. Therefore, samples were accumulated from across the United State and Canada so this could be investigated. 31 states and seven provinces have participated in the study to date with communications on-going in six other states. This helps guarantee a broader view and acceptance of the micro-Deval test. The amount of aggregate initially requested was two or three 55 gallon drums of each source. This large amount was needed for two reasons, the project team discussed making specimens for performance testing, and the test suite had not been finalized. The team realized it was more beneficial to use the performance rating found through actual field performance because of redundancy, cost requirements, and time requirements. Henceforth, aggregates were only requested if there was an established field performance for the source. Sieve analysis showed only one fifty-five gallon drum of aggregate was needed if the grading could be limited to the aggregate passing the 25 mm (1 inch) sieve and retained on the 4.75 mm (No. 4) sieve with approximately fifty percent retained on the 12.5 mm (1/2 inch) sieve. When this gradation

Williams, Hampel, Allen, and Fowler

was unable to be provided, calculations were performed to determine the amount of aggregate needed. Field Performance Rating Since one of the goals of this report is to correlate micro-Deval test results with field performance of aggregates, the field performance rating is crucial. Originally, the performance ratings given in the responses to the initial survey were intended to be used as the performance rating of each aggregate source. As the project developed and more published literature was reviewed, it was felt that strictly using the ratings provided in the survey responses could result in a subjective rating system. To define a more objective rating system, two rating systems were found from past research. These were found in published work by Senior and Rogers [7] and Wu et al. [9]. Senior and Rogers [7] developed a rating scale for the field performance evaluation criteria of coarse aggregates used in granular base and asphaltic and Portland cement concrete: Good used for many years with no reported failures, pop-outs, or other signs of

poor durability, Fair used at least once where pop-outs or some reduced service life had resulted,

but pavement or structure life extended for over 10 years, and Poor used once with noticeable disintegration of pavement after one winter,

severely restricting pavement life. The second published rating scale found was that of Wu et al. [9]. Wu et al. looked into characterizing aggregates used in asphaltic concrete only. The scale was as follows: Good used for many years with no significant degradation problem during

construction and no significant pop-outs, raveling, or potholing during service life, Fair used at least once where some degradation occurred during construction

and/or some pop-outs, raveling, and potholing developed, but pavement life extended for over 8 years, and

Williams, Hampel, Allen, and Fowler

Poor used at least once where raveling, pop-outs, or combinations developed

during the first two years, severely restricting pavement [use] ICAR 507 utilized the same three-step rating system (good, fair, and poor) shared by both studies listed above. The good and fair ratings used by the ICAR 507 project were similar to those used by Senior and Rogers since this project dealt with bases, hot-mixed asphalts, and portland cement concretes. The poor rating was similar to the one used by Wu et al. in that two years separated poor from fair instead of the one year used by Senior and Rogers. The final performance rating system accepted for use consisted of the following: Good used for 10 or more years with no reported non-chemical problems, Fair used at least once where minor non-chemically related failures require

minor repairs, but average life extends beyond 10 years, and Poor used at least once where severe degradation or failure occurred within 2

years of service or during construction which severely inhibits and/or prevents the use of the application. In addition to the performance rating scale being developed, a final survey was constructed as well. This survey consisted of a series of questions about each source including which applications the source was used in and any problems experienced due to its use. The goal of the survey was to make the rating system as objective as possible. Following a phone interview, ICAR 507 personnel determined the performance rating based on the information provided. By ICAR 507 personnel determining the field performance rating, all samples could be compared on equal terms. In no case was a sample rated good or poor that was initially rated the other extreme. In some cases though, a sample was initially rated good but was determined to be fair. TESTING METHODOLOGY While choosing the aggregate tests to be included within this study, several factors were taken into consideration. Great importance was placed on tests of widespread current use by state and provincial departments of transportation. Since this research primarily focuses on the microDeval test, attention was given to tests whose results would either correlate with or compliment

Williams, Hampel, Allen, and Fowler

micro-Deval results in comparisons with field performance. Relationship to field performance and the ability of each test to adequately predict field performance was considered. Relationship to basic aggregate properties was also considered. The following is a description of methodology used to determine which tests would be included in this research. Abrasion Tests Los Angeles Abrasion and Impact Test (AASHTO T 96) The L.A. abrasion test was decided on due to its popularity among transportation officials as found through the initial survey. According to Amirkhanian [11], 26 percent of surveyed agencies were unaware where their L.A. abrasion specification loss limits originated. Studies by Minor [4], Rogers et al. [5], and Richard and Scarlett [6] have shown poor correlations between the L.A. abrasion test and field performance. While the L.A. abrasion test can predict the mechanical breakdown of aggregate in stockpiling, transportation, and construction, it does not correlate well with field performance. The L.A. abrasion test [12] calls for an aggregate sample to be placed in a revolving drum along with a set number of steel charges. The drum repeatedly picks up and drops the sample and charges by means of a shelf located inside the drum. While the name of the test implies both abrasion and impact, the L.A. abrasion test correlates well with other impact tests such as the Aggregate Impact value (BS 812: 110) and Aggregate Crushing Value (BS 812: 110) as shown Hudec [13] and Al-Harthi [14]. Micro-Deval Abrasion Test (AASHTO TP 58) The micro-Deval test, developed in France during the 1960s, looks at the effects of moisture on the abrasion resistance of mineral aggregates. The test (AASHTO TP 58 [15]) involves placing 1500 g of soaked, graded aggregate and two liters of water into a five liter jar. Following soaking of the aggregate for a minimum of one hour prior to running the test, 5000 g of steel charges, 9.5 mm (3/8 in) in diameter, are added to the jar in addition to the sample and water. The jar is then placed into the micro-Deval apparatus and rotated at 100 revolutions per minute for two hours. Upon completion of the required number of revolutions, the sample is screened over a 1.18 mm (No. 16) sieve and oven dried to constant mass at 110C (230F).

Williams, Hampel, Allen, and Fowler

Unlike the L.A. abrasion test drum, the micro-Deval test drum does not have a shelf to lift and drop the sample and subject it to impact loads causing fragmentation. Degradation is a product of abrasion between the aggregate particles and steel charges in the presence of water. The micro-Deval test has been shown to correlate well with field performance but its application is still unclear. In 1998 Rogers has suggested that micro-Deval be used as an aggregate qualifying test due to its correlation with field performance [16]. The test could be used to replace the magnesium sulfate test due to the high correlation found by Rogers and Senior [7]. The precision of the micro-Deval test alerts changes in aggregate type at quarries by yielding different losses in the micro-Deval test, which can inform quarry personnel when to perform sulfate soundness testing, saving time and money. Soundness Tests Sulfate Soundness Tests (AASHTO T 104) Soundness tests have been used by transportation agencies and testing laboratories in North America for Many years. Since its birth, many have debated the merit of this test as an indicator of field performance. Although some have found the sulfate test to be an adequate predictor of performance [2, 9, 17-19], some have reported cases where the sulfate tests have lacked the ability to consistently relate to field performance [20-22]. The crystal growth of salts within the pores of aggregates does not subject the aggregate to the same expansive forces as the freezing of water [2]. In addition, several researchers report large variability in the results of soundness testing [23-25]. Nevertheless, the sulfate soundness test is one of the most commonly used qualification test in the United States. Hanna et al. [1] showed that the sulfate soundness test is the soundness test of choice for 31 of the 43 respondents to a national survey. Because of its widespread use, a sulfate soundness test was an obvious choice for this research. Although several reported using magnesium sulfate during phone interviews with DOT representatives, Hanna et al. reported in 2003 that a majority of the states using sulfate soundness testing use sodium sulfate soundness [1]. Despite this, the magnesium sulfate test has several favorable characteristics that warrant its use in this research instead of sodium sulfate. Magnesium sulfate is a much harsher test [26] producing more loss by mass. In addition, magnesium sulfate has been shown to have less variation in solubility in the temperature range of
Williams, Hampel, Allen, and Fowler 9

testing making it a much more dependable and reproducible test [1, 21]. Moreover, sodium sulfate has three different crystalline forms at the temperature of testing making the preparation of solution difficult [21], whereas, magnesium sulfate has only one at that temperature. Researchers have even called for agencies change to the magnesium sulfate test due to the difficulty in preparing and operating the sodium sulfate test [26, 27]. Therefore, the AASTHO T 104 [28] magnesium sulfate soundness test was chosen for use in this research to ensure more reliable results. Canadian Freeze-Thaw Test (CSA A23.2-24A) Although not widespread, some agencies use freezing and thawing tests as a supplement to the sulfate soundness test. Of those that do, the majority test the durability of aggregates by the freezing and thawing of concrete specimens containing those aggregates. Many believe that the tests currently available for the unconfined freezing and thawing of aggregates, such as AASHTO T103, create unrealistically harsh conditions. However, Volger reported that 7 states within the U.S. use some form of unconfined freeze-thaw test on aggregates [29]. As a result, the decision was made to use an unconfined freeze-thaw test in this research. Deciding on a test method proved difficult. Wide variations in the use of AASHTOs T 103 standard are allowed as no cooling rate or absolute minimum temperature is defined [30]. Both of these variables have been shown to affect degradation due to freezing and thawing and the relationship of the results to field performance [31-33]. Moreover, personal communications with state agencies or testing laboratories revealed wide differences in the test method. Some reported freezing aggregate mostly submerged then thawing with a 70 degree forced air draft. Another agency reported vacuum saturating samples and testing them suspended in plastic bags. Still another reported vacuum saturating the aggregate and freezing them in metal pans. All of these methods are roughly a variation of a test that has been shown to be unrealistically harsh. A few other tests have been developed for determining the potential resistance to freezing and thawing. Two of these tests are the Iowa Pore Index Test and the Washington Hydraulic Fracture test. Rogers has shown a good correlation between the Iowa Pore Index test and the durability of aggregates in Ontario, and the final version of the WHFT appears to be satisfactory

Williams, Hampel, Allen, and Fowler

10

[1]. However, through communications with state agencies evidence of widespread use these tests could not be found and were not chosen for this research. The Canadian Standards Association, however, has adopted an unconfined freezing and thawing test of aggregates that has been designed to model actual field conditions and maximize the relationship to field performance [31]. Research at the Ministry of Transpiration of Ontario determined the optimum cooling rate and minimum freezing temperature of unconfined freezethaw tests to maximize loss and relationship to field conditions. These observations are also in accordance with the observations of others. The optimum salt solution strength and the effect of the number of cycles was also determined. As mentioned above, all of these variables have been shown by MTO and other authors to significantly affect freeze-thaw durability [31-33]. The science behind the CSA specification has addressed these issues, whereas AASHTOs T 103 and its variations have not. The CSA standard has also been recommended by the National Cooperative Highway Research Program [1], and the T 103 method has not. Moreover, the CSA test method can be completed with a fraction of the time and difficultly required by the T 103 method. Testing the aggregates for this research with some variation of T 103 would have required the acquisition of equipment the project could not afford, and this would be done for a test that has been shown to be inadequate and is used in various forms by only 7 DOTs. Due to the good correlation with field performance, the ease and quickness of the test, the NCHRP recommendation, and the overwhelming scientific support, the CSA A23.2-24A specification was chosen for this research. However, recognizing that the CSA version of unconfined freeze-thaw testing is not well known among U.S. departments of transportation, a side study of the present micro-Deval research is being conducted to determine if a correlation exists between AASHTO T 103 and the CSA standard. The method used for the Canadian Freeze-Thaw testing follows the CSA A23.2-24A testing standard [34]. After washing, oven drying, and sieving the aggregate, samples are prepared by hand sieving the material according to the following gradation: 3/4 in 1/2 in 1250 grams

Williams, Hampel, Allen, and Fowler

11

1/2 in 3/8 in

1000 grams

3/8 in No. 4 500 grams Each size fraction prepared is then individually placed in autoclavable mason jars. The samples are soaked in the jars for 24 hours in a 3% sodium chloride solution. After soaking, the solution is drained from the samples, and air tight lids are placed on the jars to ensure 100 percent humidity. The samples are cooled to a temperature of -18 C (0 F) for 16 hours overnight. They are removed and allowed to thaw at room temperature for approximately 8 hours. After the fifth cycle the jars are filled with water and rinsed five times. Finally, the samples are oven dried to constant mass at 110 C (230 F) in a convection-type oven and sieved over the original sieve sizes. The percent loss is calculated, and the final loss is determined by the weighted average of the percent loss of the three jars. For the purposes of conducting this test a blast freezer was purchased, and adjustments were made to control the freezing rate according to the optimal freezing rate as determined by MTO [31]. The freezing rate of the freezer was monitored over twelve practice runs to ensure consistent freezing, and fans were placed in the freezer chamber to ensure uniform freezing of all samples. Personnel are present in the afternoons to turn on the freezer for cooling and in the mornings to open the freezer doors for thawing. Although the samples are not removed from the chamber every morning, a high-powered box fan circulates air at room temperature through the chamber. The samples are then rotated as specified before freezing again that afternoon. The remainder of the test is conducted exactly as stated in the specification and above. AASHTO Freeze-Thaw Test (AASHTO T 103) The procedure selected for the AASHTO freeze-thaw test was procedure C of the T 103 specification. The samples will be hand sieved according to the following gradation: 3/4 in to 3/8 in Consisting of: 3/4 in to 1/2 in 330 5 grams 1000 5 grams

Williams, Hampel, Allen, and Fowler

12

1/2 in to 3/8 in 3/8 in to No. 4

670 5 grams 300 5 grams

Each sample will be vacuum saturated by subjecting them to a vacuum with an air pressure not over 3.4 kPa (25.4 mm of mercury) and then introducing de-ionized water to the samples. After saturation they will be placed in Teflon coated baking pans to prevent corrosion with a plastic seal to prevent evaporation. The samples will then undergo twenty-five cycles with 6 hours of freezing to a temperature less than -26C (-15F) followed by 6 hours of thawing to a temperature of 21 to 24C (70 to 75F). Although the procedure calls for thawing in water, this test will be conducted with air thawing. The expense and technical difficulties of acquiring or building a machine to air freeze samples while thawing them in water are beyond the capabilities of this project. In addition, two of three agencies in communication with this project have reported using some sort of air thaw for the unconfined freezing and thawing of aggregates. An inexpensive solution for the unconfined freezing and thawing of aggregate according to the AASHTO T 103 specification was achieved by adapting the blast freezer used for the CSA specification. An industrial timer regulates the cycles of the freezer to provide two 6 hours periods of freezing each day, and other timers regulate a spacer heater that, combined with highpowered fans that ensure uniform heating and cooling, thaw the samples. Through several practice runs and experiments this method has shown the ability to produce consistent and reliable freeze-thaw cycles. Strength Tests Aggregate Crushing Value Test (BS 812:110) Although no AASHTO or ASTM standardized method exists for determining aggregate strength [1], several simple methods have been used by other countries. Variations of the British Aggregate Crushing Value (ACV) have been used for some time in Great Britain, Australia, and New Zealand. This test is thought of favorably and used for qualification purposes in these countries [35], and the ACV was reported by Hanna et all as being a reasonable approach for determining aggregate strength [1]. Therefore, the British Aggregate Crushing Value Standard 812:110 [36] has been selected for use in this study.

Williams, Hampel, Allen, and Fowler

13

The dry method used for the crushing value test follows that which is outlined in the British Standard BS 812:110, and the wet crushing test method used is adapted from that described in the Australian Standard AS 1141.22 (Wet/Dry Strength Variation) [37]. The appropriate equipment was obtained on loan from the Ministry of Transportation of Ontario. The testing equipment consists of thick, hollow steel cylinder which confines the aggregate, a base plate on which the cylinder and aggregate sits, and a steel plunger for applying the load. Also included is a smaller and lighter cylinder to determine the appropriate sample volume. For the ACV test, the oven dry samples are prepared by filling the provided cylinder with aggregate and weighing the sample to the nearest gram. The sample is then poured into the steel cylinder in three lifts, lightly compacting and leveling the aggregate with a metal rod after each lift. The steel plunger is then inserted into the cylinder. A compressive force is added to the aggregate at a rate such that 90,000 lbs (soft conversion of the specified 400 kN) is added uniformly over a period of 10 minutes. The sample is then removed and sieved over the No. 12 sieve, and the percent loss is determined as a ratio of loss over original mass. Wet Aggregate Crushing Value Test (Variation of BS 812:110 and AS 1141.22) A wet version of the Aggregate Crushing Value, which will be further referred to as the Wet Crushing Value (WCV) for this research, is used for aggregate qualification purposes in Australia and in New Zealand. Different crushing strength values can be obtained by testing the aggregate in both oven dry and saturated surface dry conditions. The wet crushing test has been shown to be useful in evaluating an aggregates strength when evaluating both the strength of the aggregate and the fines produced. Also of importance is the variation between the ACV results and WCV results for a given aggregate. Larger variations between the ACV and WCV have been shown to correlate with aggregate performing poorly due to wetting and drying and freezing and thawing. This is a relatively quick and easy test, and therefore an adapted version of the WCV has been adopted for use in this study The test method for the WCV in this research has been adapted from the Australian AS 1141.22 specification and the British Standard BS 812:110 specifications. The WCV is almost identical to the oven dry test except that the sample is crushed in the saturated surface dry condition. Afterwards, the sample is removed and oven dried before sieving. For the purposes of this research and comparing the oven dry and saturated surface dry aggregate strengths and

Williams, Hampel, Allen, and Fowler

14

crushing values, a few additions have been made: the load carried by the aggregate at a deflection of 10% is recorded, and the final deflection of the aggregate at 90,000 lbs is recorded. Other Tests Flat and Elongated Test (AASHTO D 4791) and AIMS Test The flat and elongated test and the Aggregate Imaging System (AIMS) test are included in the test suite to determine if the shape and surface texture of the aggregate affects micro-Deval loss. It is believed that a flat or elongated particle will more than likely have more loss than a round particle in the micro-Deval test, when comparing samples of the same mineralogy, due to the potential of corners chipping off. It has been shown through research that particle shape can significantly affect the field performance of aggregates used in hot-mixed asphalt or railroad ballast [2, 38-40]. According to Hanna et al. [1], most state agencies measure the ratio of particle dimensions rather than measuring the percentage of flat and elongated particles. ICAR 507 is measuring the thickness to width, thickness to length, and width to length ratios to the nearest one half of a ratio. These three ratios are then used to come up with a single number that could be used for comparison purposes. While the flat and elongated test only measures particle shape, the AIMS test was included due to its ability to measure surface texture. The project proposes to send micro-Deval test samples for analysis on the AIMS machine before testing in the micro-Deval apparatus. These samples will be returned, tested according to the micro-Deval specification, and then reanalyzed on the AIMS machine. Correlations will be developed with the micro-Deval test results with the before and after surface textures and particle shape. This procedure looks to determine the effect of particle shape and surface texture on micro-Deval limits. Percent Fractured Particles Test (ASTM D 5821) Similar to the particle shape, the angularity of the aggregate has an effect on attrition. Ekse and Morris [41] have shown angularity to affect the abrasion loss of particles. They found that for a given source, the more angular the aggregate particle is, the higher the loss is. Boucher and Selig [39] also found the same result, noting previously worn particles had much less

Williams, Hampel, Allen, and Fowler

15

attrition than freshly crushed particles. While angularity is not as critical in dense graded hotmixed asphalt mixtures as it is in open graded mixtures [2] its effect on micro-Deval loss is sought. Due to the highly subjective nature of the test, Benson and Ames (507.128) found the both inter-laboratory precision to be poor. For comparison purposes, ICAR 507 plans to have the same person perform all of the percent fractured analysis. Petrographic Examination (ASTM C 295) Knowing the mineralogy of an aggregate can tell a lot about the probable test results through comparison with aggregates of similar mineralogical backgrounds. Studies have been done comparing the petrographic examination to field performance of aggregates with contradicting results. Rhoades and Mielenz [42] found that the quality of natural aggregates can be determined by petrographic analysis. Similarly, Cooper et al. [43] established that a detailed field examination, consisting in large part of a mineralogical determination, was able to predict the overall quality of a potential aggregate source rock with an 86 percent success rate. When this field examination was combined with the micro-Deval test, the success rate jumped to 94 percent. Mielenz (1946) [44] and Boucher and Selig [39] concluded that the petrographic examination aids in the evaluation of other test results and that the information obtained can be used for comparison with unknown aggregates for evaluation purposes. However, both authors say the test results are not sufficient to be used alone in the prediction of performance. Use of the petrographic examination as a supplement to other tests is recommended; therefore, petrographic examination is included in this study. Side Studies AIMS Test vs. Flat and Elongated Test The Aggregate Imaging System (AIMS) test and flat and elongated particle test are being performed on the aggregate samples acquired for the study. As mentioned previously in the Testing Methodology chapter, the AIMS test is being utilized to develop a correlation between the surface texture of aggregate and micro-Deval loss while the flat and elongated particle test is looking to determine a correlation between particle shape and micro-Deval loss.

Williams, Hampel, Allen, and Fowler

16

While the AIMS test shows promise as a surface texture analyzer, its capabilities also include particle shape analysis. The particle shape analysis will be performed with no additional effort while the surface texture is analyzed. According to researchers at Texas A&M University, the AIMS test has a very strong correlation with the flat and elongated particle test. The flat and elongated particle test is being used due to its accepted use in industry but it is a very tedious test. The AIMS test is considerably faster and more objective, but lacks widespread use. If a strong correlation can be found between the results of the two tests, a more efficient means of determining particle shape can be utilized with a side benefit of receiving information on the surface texture of the aggregate as well. Particle Shape Factor Research has shown that particle size and shape can play a significant role in wet attrition tests [41, 45]. Intuition would lead one to believe that all else being equal, the rougher or more angular an aggregate, the more easily the aggregate will be abraded. Knowing this, a method is needed to normalize micro-Deval loss to eliminate the bias introduced by particle shape. Theoretically, an angular, rough aggregate with an exactly identical field performance as a smooth, rounded aggregate should have identical micro-Deval losses. Intuitively, however, this is not the case. In the authors opinion, no good proven method of numerically quantifying aggregate shape for correlation purposes exists. Therefore, as a part of this research two studies are being conducted to determine the effect of aggregate shape and texture on micro-Deval loss and its relation to field performance. Using an adaptation of the AASHTO Flat and Elongated test method, an attempt can be made to numerically quantify aggregate shape [46]. By adopting the standard flat and elongated caliper the thickness vs. width, width vs. length, and thickness vs. length ratios can be determined to the nearest half of a ratio for a specified number of particles of a source. The average ratio for the source can then be determined. This provides numerical information that can be used for correlation purposes. While manipulating these ratios, the authors discovered that a single number can be determined to quantify aggregate shape. The three ratios of a given source can be normalized by the lowest of the three ratios for that source, and then all three normalized ratios can be
Williams, Hampel, Allen, and Fowler 17

multiplied to determine a particle shape factor. Observations made while relating this factor to the shape of the aggregate it described were very promising. Rounded particles have the lowest number, elongated particles have slightly higher numbers, and flat and elongated particles have the highest number. It would be expected that particles would be more susceptible to degradation in this order. One would expect rounded particles to be the least susceptible to micro-Deval degradation and flat and elongated particles to be the most susceptible to degradation. The number could further be manipulated by multiplying the particle shape factor by factors relating to the angularity and roughness to increase or decrease the particle shape factor according to the suspected potential for resulting abrasion loss. Although these values need to be further evaluated experimentally for accuracy, tentative factors have been assigned. Factors of 0.9 and 1.2 can be used for smooth and rough particles respectively, and factors of 0.9 and 1.2 can be used for rounded and angular particles respectively. This factor has shown promise in limited correlations with micro-Deval thus far. It is the hopes of the authors that this factor, or some variation thereof, will at the very least provide a better means of quantifying aggregate shape, or, more desirably, normalize micro-Deval values to eliminate the bias introduced by particle shape and texture. Canadian Freeze-Thaw Test vs. AASHTO Freeze-Thaw Test Recognizing that the CSA version of unconfined freeze-thaw testing is not well known among U.S. departments of transportation, a side study of the present micro-Deval research is being conducted to determine if a correlation exists between AASHTO T 103 and the CSA standard. Attention will also be given to the importance of each test concerning representing field performance. Fifty aggregates will be selected from the aggregates obtained for this research. These aggregates will be selected so that a wide variety of field performances, mineralogical types, and CSA freeze-thaw losses will be represented. A version of Procedure C will be conducted. Aggregate Crushing Value Test vs. Wet Aggregate Crushing Value Test Some feel that the crushing value tests are an indication of potential field performance. Noting that aggregate properties can be different between a wet and a dry aggregate [47], it should be of interest to determine the difference in strength between oven dry and saturated

Williams, Hampel, Allen, and Fowler

18

surface dry aggregates as measured by the Aggregate Crushing Value Test. This has been shown to be very significant in the determination of potential field performance according to a contact at Metso Minerals. The results obtained during this research will show how the strength characteristics of aggregates change. The relationship of each to field performance might yield valuable information, and the difference between the wet and dry values may be important. TESTING OPERATIONS Precision Statements For all but three test methods, a precision requirement similar to the British Aggregate Crushing Value test was adopted (BS 812 110). The ACV requirement states that, provided two test results for a given source fall within 7 % of the mean of the two results, the mean is found to be acceptable. If this is not the case, then two additional tests must be conducted and the mean of the four will then become the result. Provided this requirement is met and the control samples where applicable are also within 7% of the mean, then the results are deemed acceptable. This was concluded to be an acceptable method of ensuring precise results in an efficient manner. This method is not being applied to the Magnesium Sulfate Soundness test, the microDeval test, or the Absorption and Specific Gravity test. For the Absorption test only one sample is tested as this is all that is required per the specification. For the micro-Deval test, three samples of each source are tested and the mean is accepted unless an obvious outlier exists. Control samples from the Brownwood quarry in Texas have been calibrated with Brechin II samples obtained from MTO and are tested every ten tests or every week a sample is tested. The control sample results are monitored to ensure that they remain within the acceptable limits set for in the specification. The micro-Deval testing is being conducted in this manner to match that which is specified in AASHTO TP 58 [15]. The variability of the sulfate tests in general, as shown by previous work [23-25], was too great to apply the method used in the British Standard. Three samples of each source are tested by magnesium sulfate with no more than two of the three being tested in the same run. Two control samples of Brechin II from Ontario are included with each run. If the control samples yield values outside the acceptable limits provided in the AASHTO test specification [28], then a
Williams, Hampel, Allen, and Fowler 19

careful examination is given and additional samples are tested if necessary. One Brechin II control sample and one control sample from the Brownwood quarry in Texas are also used in the Canadian and AASHTO freeze-thaw tests; however, the precision requirements of these tests are subject to those outlined in the British Standard. Over 12 practice freezing and thawing tests have shown that the freezer used for these tests can reliably reproduce freezing and thawing conditions, and the control samples used thus far have been remarkably consistent. Standardizing Gradations The acquired aggregate arrived in a variety of gradations. This created a potential problem in comparing one sample to another. States specify different gradations for different uses and not all states specify the same gradation for the same use. The gradation can affect test results according to Rogers, et al. [48] and Selig and Boucher [49]. This poses the question, how do we meet the needs of all states and provinces when each performs different tests and uses different gradations. Volger has reported that nine of sixteen states responding to a survey standardize gradations before testing [29]. It was decided to create uniform gradations that each sample should meet for comparison purposes. These uniform gradations were mostly specified by testing specifications, but where they were left open according to the specification, a material passing the 1/2 inch sieve and retained on the 3/8 inch sieve was used. This particle size was chosen for two reasons: it is a median range for a typical coarse aggregate used in road construction that should provide a representative test result and the micro-Deval sample most commonly used (grading A) is comprised of fifty percent of 3/8 inch material. Since the ICAR 507 project is determining the role of the micro-Deval test, it was determined 3/8 inch material would be a good representative size for comparison purposes. All sources are tested according to the specified gradations. However, rare instances are occurring where not enough material is available for the all testing. In these cases an attempt is first made to obtain additional material from the same stockpile. If this attempt fails and particles of a larger size of the same sample are available, the aggregate then is crushed to meet the specification. If no larger sizes are available, or if this is not practical, the sample sizes of the soundness, freeze-thaw, and crushing value tests are reduced by no more than 50%. There have

Williams, Hampel, Allen, and Fowler

20

been a few rare instances where not enough material was available and substitutions were required. One case resulted in substituting 1/2 inch material for 5/8 inch material for all tests. Two instances have occurred where an insufficient amount of 3/8 inch material was available and half inch material was substituted. A few more instances occurred where no #4 material was present in any appreciable amounts. In this case the soundness tests and freeze-thaw tests were conducted without this material. In all such cases, the actions taken have been documented and will be reported. CURRENT RESULTS The project is now operating in two laboratories located on the J.J. Pickle Research Campus, part of The University of Texas at Austin. The Construction Materials Research Group just opened a new lab which this project was involved with. A new micro-Deval test machine and a new blast freezer for freeze-thaw testing have been acquired and placed in the laboratory. A convection type oven was also attained and refurbished for use in the new laboratory and a magnesium sulfate soundness test setup was also developed and constructed. A thermostat for the room, independent of the building was installed for better temperature and solubility control of the sulfate soundness test. Arrangements have been made for use of TxDOTs L.A. abrasion test machine. TxDOT has also offered to perform magnesium sulfate soundness testing for comparison purposes. The Ontario Ministry of Transportation (MTO) has also graciously lent the equipment required for Aggregate Crushing Value testing. Both TxDOT and MTO have supplied a wealth of information in setting up and constructing testing equipment. Measures have also been taken to secure an experienced geologist to perform petrographic examinations. The two graduate students that started with the project will be graduating in May, 2005. It was arranged to have another graduate student take over testing and compilation of the final report. The succeeding graduate student began with the project in January to provide a transition time for questions and familiarity with the testing procedures and equipment. Testing of 47 aggregate sources is completed. The results of these sources will be used for two theses scheduled to be completed by June, 2005. Earlier this year it was determined that

Williams, Hampel, Allen, and Fowler

21

testing could be completed by the end of March for 47 of the 117 sources acquired. Most of the remaining sources have some test results but were not far enough along in testing to be completed for the theses planned for June, 2005. ICAR 507 is still expecting a few additional sources to trickle in from agencies that wanted to provide aggregate but were unable to do so earlier for various reasons. The test results thus far look promising. The potential of the micro-Deval test to distinguish the field performance of aggregates used in base, hot-mixed asphalt, and concrete appears to exist. Finding the optimum application of the micro-Deval test is the key. CONCLUSIONS With the pressure to begin constructing longer lasting roads and structures with more marginal aggregates due to depletion of available resources, accurate and reliable testing methods and limits need to be developed to identify appropriate aggregate. Researchers have shown that traditional testing methods are not always suitable alone or even in combination with other tests. However, several agencies have published reports showing micro-Deval to be an outstanding indicator of field performance. However, others have shown results that show microDeval as having poor or mixed correlations with field performance. By investigating the relationship of micro-Deval and other common aggregate tests to field performance, realistic limits can be determined for the qualification of aggregates. The results of this study should provide micro-Deval limits that either alone or combined with other test results will be realistic predictors of field performance.

Williams, Hampel, Allen, and Fowler

22

REFERENCES 1. Amir Hanna, Kevin Folliard, and Kurt Smith, "Aggregate Tests for Portland Cement Concrete Pavements: Review and Recommendations," Research Results Digest, No. 281, NCHRP, 2003. 2. P. Kandhal, and F. Parker, "Aggregate Tests Related to Asphalt Concrete Performance in Pavement," Report No. 405, NCHRP, 1998. 3. Serji Amirkhanian, Douglas Kaczmarek, and James Burati, "Effects of Los Angeles Abrasion Test Values on the Strengths of Laboratory-Prepared Marshall Specimens," Transportation Research Record, No. 1301, 1991, pp. 77-86. 4. C. E. Minor, "Degradation of Mineral Aggregate," Symposium on Road and Paving Materials, ASTM STP 277, Philadelphia, 1959, p. 109. 5. C. A. Rogers, and S. A. Senior, "Predicting Aggregate Performance Using the MicroDeval Abrasion Test," Proceedings: 3rd Annual Symposium for the Center for Aggregates Research, University of Texas, Austin, TX, 1995. 6. Jean A. Richard, and James R. Scarlett, "Airport Engineering ATR-024: A Review and Evaluation of the Micro-Deval Test," Public Works and Government Services of Canada, Report on Project 914222, 1997. 7. S. A. Senior, and C. A. Rogers, "Laboratory Tests for Predicting Coarse Aggregate Performance in Ontario," Transportation Research Record, No. 1301, 1991, pp. 97-106. 8. L. J. Larson, R. P. Mathiowetz, and J. H. Smith, "Modification of the Standard Los Angeles Abrasion Test," Highway Research Record, No. 353, 1971, pp. 15-24. 9. Yiping Wu, Frazier Parker, and Prithvi S. Kandhal, "Aggregate toughness/abrasion resistance and durability/soundness tests related to asphalt concrete performance in

Williams, Hampel, Allen, and Fowler

23

pavements," Transportation Research Record, No. 1638, National Research Council, Washington, DC, USA, 1998, p. 85. 10. L. Allen Cooley Jr, and Robert S. James, "Micro-Deval Testing of Aggregates in the Southeast," Transportation Research Record, No. 1837, National Research Council, 2003, p. 73. 11. Serji N. Amirkhanian, Ervin L. Dukatz, Jr., Douglas Kaczmarek, and Darrell Brownlow, "Effects of Igneous Aggregate Sources with Various Los Angeles Abrasion Test Values on the Strengths of Concrete Mixtures," Cement, Concrete and Aggregates, Vol. 14, No. 2, 1992, p. 86. 12. American Association of State Highway and Transportation Officials, "Standard Test Method for Resistance to Degradation of Small-Size Coarse Aggregate by Abrasion and Impact in the Los Angeles Machine," T 96-02, 2002. 13. Peter P. Hudec, "Aggregate Tests - Their Relationship and Significance," Durability of Building Materials, Vol. 1, No. 3, 1983, p. 275. 14. A. A. Al-Harthi, "A Field Index to Determine The Strength Characteristics of Crushed Aggregate," Bulletin of Engineering Geology and the Environment, Vol. 60, No. 3, 2001, p. 193. 15. American Association of State Highway and Transportation Officials, "Standard Test Method for Resistance of Coarse Aggregate to Degradation by Abrasion in the MicroDeval Apparatus," TP58-00, 1999. 16. J. P. Latham, and C. Rogers, "Canadian experience with the micro-Deval test for aggregates," Geological Society Engineering Geology Special Publication, No. 13, 1998, pp. 139-47. 17. C. A. Vollick, and E. I. Skillman, "Correlation of Sodium Sulfate Soundness of Coarse Aggregate with Durability and Compressive Strength of Air-Entrained Concrete," Proceedings, Vol. 52, American Society for Testing and Materials, 1952, p. 1159.

Williams, Hampel, Allen, and Fowler

24

18.

William Sheftick, "Na2SO4 Soundness Test Evaluation," Cement, Concrete and Aggregates, Vol. 11, No. 1, ASTM, 1989, p. 73.

19.

David K. Miles, "Accelerated Soundness Test for Aggregates," Report No. 500-917, Utah State Department of Highways - Materials and Testing Division, 1972.

20.

Delmar L. Bloem, "Soundness and Deleterious Substances," Significance of Tests and Properties of Concrete and Concrete Making Materials, ASTM STP 169A, No. 169, American Society for Testing Materials, 1966, pp. 497-512.

21.

T. Larson, P. Cady, M. Franzen, and J. Reed, "Special Report No. 80," Highway Research Board, 1964.

22.

Poduru M. Gandhi, and Robert L. Lytton, "Evaluation of Aggregates for Acceptance in Asphalt Paving Mixtures," Proceedings of the Association of Asphalt Paving Technologists, Vol. 53, Scottsdale, AZ, 1984, pp. 525-547.

23.

S. Walker, and C. E. Proudley, "Studies of Sodium and Magnesium Sulfate Soundness Tests," Proceedings, American Society for Testing Materials, Vol. 36, 1936, p. 327.

24.

C.A. Rogers, M.L. Bailey, and B. Price, "Micro-Deval Test for Evaluating the Quality of Fine Aggregate for Concrete and Asphalt," Transportation Research Record, No. 1301, 1991, pp. 68-76.

25.

Leo V. Garrity, and Herbert F. Kriege, "Studies of the Accelerated Soundness Tests," Proceedings, Highway Research Record, Vol. 15, 1935, pp. 237-260.

26.

C. F. Wuerpel, "Modified Procedure for Testing Aggregate Soundness by Use of Magnesium Sulfate," Proceedings, Vol. 39, American Society of Testing Materials, 1939, p. 882.

27.

Ira Paul, "Magnesium Sulphate Accelerated Soundness Test on Concrete on Aggregates," Proceedings of the Twelfth Annual Meeting of the Highway Research Board, Vol. 12, 1932, p. 319.

Williams, Hampel, Allen, and Fowler

25

28.

American Association of State Highway and Transportation Officials, "Standard Test Method for Soundness of Aggregate by Use of Sodium Sulfate or Magnesium Sulfate," T 104-99, 1999.

29.

Ralph Volger, and Gail H. Grove, "Freeze-Thaw Testing of Coarse Aggregate in Concrete: Procedures Used by Michigan Department of Transportation and Other Agencies," Cement, Concrete and Aggregates, Vol. 11, No. 1, American Society for Testing Materials, 1989, pp. 57-66.

30.

American Association of State Highway and Transportation Officials, "Standard Method of Test for Soundness of Aggregates by Freezing and Thawing," T 103-91, 2000.

31.

C. A. Rogers, S. A. Senior, and D. Boothe, "Development of an Unconfined Freeze-Thaw Test for Coarse Aggregates," Technical Report EM-97, Engineering Materials Office, Ontario Ministry of Transportation, 1989.

32.

T. C. Powers, "Basic Considerations Pertaining to Freezing-and-Thawing Tests," Proceedings, Vol. 55, American Society for Testing Materials, 1955, pp. 1121-1155.

33.

Michel Pigeon, Jacques Prevost, and Jean-Marc Simard, "Freeze-Thaw Durability vs. Freezing Rate," Journal of The American Concrete Institute, Vol. 82, No. 5, 1985, p. 684.

34.

Canadian Standards Association, "Method of Test for Resistance of Unconfined Coarse Aggregate to Freezing-and-Thawing," CAN/CSA-A23.2-24A.

35. 36.

Stacy Goldsworthy, Personal Interview, Austin, TX, August 10, 2004. British Standards Institution, "Methods for Determination of Aggregate Crushing Value (ACV)," BS 812: Part 110, 1990.

37.

Standards Association of Australia, "Wet/Dry Strength Variation," Methods for Sampling and Testing Aggregates, Australian Standard, AS 1141.22, 1980.

Williams, Hampel, Allen, and Fowler

26

38.

G. W. Maupin Jr, "Effect of Particle Shape and Surface Texture on the Fatigue Behavior of Asphaltic Concrete," Highway Research Record, No. 313, 1970, p. 62.

39.

Debra L. Boucher, and Ernest T. Selig, "Application of Petrographic Analysis to Ballast Performance Evaluation," Transportation Research Record, No. 1131, 1987, pp. 15-25.

40.

Randy C. Ahlrich, "Influence of Aggregate Properties on Performance of Heavy-duty Hot-mix Asphalt Pavements," Transportation Research Record, No. 1547, National Research Council, Washington, DC, USA, 1996, p. 7.

41.

M. Ekse, and H. C. Morris, "A Test for Production of Plastic Fines in the Process of Degradation of Mineral Aggregates," Symposium on Road and Paving Materials, ASTM STP 277, Philadelpha, 1959, p. 122.

42.

Roger Rhoades, and Richard C. Mielenz, "Petrographic and Mineralogic Characteristics of Aggregates," Mineral Aggregates, ASTM STP 83, American Society for Testing and Materials, 1948, pp. 20-48.

43.

A. J. Cooper, J. S. Villard, and J. D. Ferguson, "Predicting Aggregate Performance of the Iguneous and Metamorphic Rocks in Northern Ontario," International Center for Aggregates Research 11th Annual Symposium, Austin, TX, 2003.

44.

Richard C. Mielenz, "Petrographic Evaluation of Concrete Aggregates," Significance of Tests and Properties of Concrete and Concrete Making Materials, ASTM Special Publication, STP 196C, American Society for Testing and Materials, 1994, p. 341.

45.

Kingsley Goonewardane, "Behavior of Aggregate in the Washington Degradation Test," Journal of Testing and Evaluation, Vol. 5, No. 1, 1977, p. 25.

46. 47.

Chris A. Rogers, Personal Interview, Austin, TX, July 27, 2004. J. P. Latham, A. R. Woodside, and W. D. H. Woodward, "Assessing the Wear Characteristics of Aggregate Exposed at the Road Surface," Geological Society Engineering Geology Special Publication, No. 3, 1998, pp. 149-157.

Williams, Hampel, Allen, and Fowler

27

48.

C. A. Rogers, B. C. Lane, and S. A. Senior, "The Micro-Deval Abrasion Test for Coarse and Fine Aggregate in Asphalt Pavement," International Center for Aggregates Research 11th Annual Symposium, Austin, TX, 2003.

49.

Ernest T. Selig, and Debra L. Boucher, "Abrasion tests for railroad ballast," Geotechnical Testing Journal, Vol. 13, No. 4, 1990, p. 301

Williams, Hampel, Allen, and Fowler

28

Das könnte Ihnen auch gefallen