05.01.2013 Views

Fundamental Properties of Asphalts and Modified Asphalts, III

Fundamental Properties of Asphalts and Modified Asphalts, III

Fundamental Properties of Asphalts and Modified Asphalts, III

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Fundamental</strong> <strong>Properties</strong> <strong>of</strong> <strong>Asphalts</strong> <strong>and</strong><br />

<strong>Modified</strong> <strong>Asphalts</strong>, <strong>III</strong><br />

Quarterly Technical Progress Report<br />

April 1-June 30, 2008<br />

June 2008<br />

Prepared for<br />

Federal Highway Administration<br />

Contract No. DTFH61-07-D-00005<br />

By<br />

Western Research Institute<br />

www.westernresearch.org


TABLE OF CONTENTS<br />

TASK 1. COORDINATION ........................................................................................................1<br />

TASK 2. CONTINUED RESEARCH.........................................................................................3<br />

Subtask 2-1. Moisture Damage.................................................................................................3<br />

Subtask 2-2. Aging....................................................................................................................7<br />

Subtask 2-2.1. Impact <strong>of</strong> Water on Aging ..........................................................................7<br />

Subtask 2-2.2. Support <strong>of</strong> the Mechanistic-Empirical Pavement Design Guide ..............11<br />

Subtask 2-3. Nanotechnology: AFM Analysis <strong>of</strong> Asphalt Thin-Film Microstructure<br />

Phenomenology..................................................................................................................45<br />

Subtask 2-4. Low-Temperature <strong>Properties</strong> .............................................................................55<br />

Subtask 2-5. <strong>Modified</strong> <strong>Asphalts</strong>..............................................................................................57<br />

Subtask 2-6. Validation Site Monitoring ................................................................................71<br />

TASK 3. OTHER RESEARCH ACTIVITIES (IDIQ)............................................................73<br />

TASK 4. INFORMATION DEPLOYMENT ...........................................................................75<br />

Subtask 4-1. Publications, Presentations, Newsletters, Flyers <strong>and</strong> Brochures .......................75<br />

Subtask 4-2. Website Maintenance.........................................................................................77<br />

Subtask 4-3. Research Database—Contractor Team Activities .............................................79<br />

Subtask 4-4. Research in Progress Database ..........................................................................81<br />

Subtask 4-5. Support <strong>of</strong> the Mechanistic-Empirical Pavement Design Guide .......................85<br />

Subtask 4-6. Semi-Annual Meetings ......................................................................................87<br />

i


TASK 1. COORDINATION<br />

The goals <strong>of</strong> this task are to maintain contact with current State-<strong>of</strong>-the-Art <strong>and</strong> State-<strong>of</strong>-the-<br />

Science in fundamental research on asphalts <strong>and</strong> modified asphalts, to keep abreast <strong>of</strong> current<br />

<strong>and</strong> on-going science <strong>and</strong> technology in asphalt research, <strong>and</strong> to learn the needs <strong>of</strong> the asphalt<br />

technical community.<br />

During this quarter, WRI hosted a four-day RHEOBIT asphalt rheology course taught by Dr.<br />

Ge<strong>of</strong>frey Rowe <strong>and</strong> Dr. David Anderson. Course topics ranged from basic asphalt properties to<br />

master curve principles <strong>and</strong> from basic testing techniques to sources <strong>of</strong> measurement error.<br />

H<strong>and</strong>s-on sessions were conducted with spreadsheets <strong>and</strong> RHEA analysis s<strong>of</strong>tware, <strong>and</strong> with<br />

laboratory instrumentation.<br />

During this quarter, WRI welcomed a two-day visit with Dr. Ernest Bastian <strong>and</strong> Dr. Terry<br />

Arnold, both <strong>of</strong> FHWA <strong>and</strong> co-COTRs for the <strong>Fundamental</strong> <strong>Properties</strong> <strong>of</strong> <strong>Asphalts</strong> <strong>and</strong> <strong>Modified</strong><br />

<strong>Asphalts</strong> contract with WRI. Ongoing research was discussed with individual researchers <strong>and</strong><br />

the project status was discussed with Ray Robertson, Fred Turner, <strong>and</strong> Mike Harnsberger, WRI.<br />

Next quarter WRI will host the 2008 Petersen Asphalt Research Conference (July 14-16) <strong>and</strong> the<br />

Pavement Performance Prediction (P3) Symposium (July 16-18). Details can be obtained at<br />

www.westernresearch.org.<br />

1


SUBTASK 2-1. MOISTURE DAMAGE<br />

Statement <strong>of</strong> Problem<br />

TASK 2. CONTINUED RESEARCH<br />

<strong>Asphalts</strong> <strong>and</strong> aggregates demonstrate different performance characteristics, especially in the<br />

presence <strong>of</strong> moisture. Moisture damage mechanisms are poorly understood. From a chemical<br />

st<strong>and</strong>point, the question is: How are moisture damage <strong>of</strong> asphalt concrete <strong>and</strong> the chemistry <strong>of</strong><br />

the asphalt <strong>and</strong> aggregate materials related?<br />

Approach<br />

The work described below represents an attempt to gain insight into moisture damage using the<br />

Hamburg Wheel Test to subject asphalt concrete specimens to stress under a steel wheel in a<br />

water bath. The damage that occurs to the specimen results in the loss <strong>of</strong> material from the<br />

sample into the water bath. This material is both inorganic (aggregate) <strong>and</strong> organic (asphalt)<br />

<strong>and</strong>, when properly sampled, the chemical make-up <strong>of</strong> each can be determined. This chemical<br />

characterization can then be used to establish correlations, or lack there <strong>of</strong>, between pavement<br />

failures due to the presence <strong>of</strong> moisture <strong>and</strong> chemical make-up <strong>of</strong> the asphalt <strong>and</strong> aggregate.<br />

Another approach can be taken using theoretical calculations that represent interactions between<br />

asphalt <strong>and</strong> aggregate to help define what chemical characteristics <strong>of</strong> each material will result in<br />

favorable interactions. In other words, this method theoretically predicts what types <strong>of</strong><br />

molecules in asphalt will interact favorable with certain aggregate molecules. These predictions<br />

eventually require a pro<strong>of</strong>-<strong>of</strong>-concept experiment in the laboratory. Therefore, based on theory,<br />

certain types <strong>of</strong> molecules can be added into or removed from an asphalt to test for changes in<br />

adhesion characteristics that could support or disagree with the predictive calculations. This<br />

requires complex separations <strong>of</strong> asphalt into various fractions that are based on the overall,<br />

chemical make-up <strong>of</strong> the material.<br />

Goal<br />

The goal <strong>of</strong> this research is to definitively determine how the chemistry <strong>of</strong> asphalt <strong>and</strong> aggregate<br />

affect roadway performance in the presence <strong>of</strong> moisture. To do this, a better fundamental<br />

underst<strong>and</strong>ing <strong>of</strong> asphalt chemistry <strong>and</strong> aggregate material properties <strong>and</strong> their relationship to<br />

pavement performance is needed. Underst<strong>and</strong>ing their fundamental properties makes it possible<br />

to decipher <strong>and</strong> better define the phenomena that occur at the asphalt-aggregate interface <strong>and</strong><br />

truly predict the performance <strong>of</strong> the pavement. The ultimate goal for the body <strong>of</strong> this work is to<br />

eventually develop the material science for asphalt pavements.<br />

3


Support <strong>of</strong> FHWA Strategic Goals<br />

The work conducted in this subtask is in support <strong>of</strong> the FHWA strategic goal that pertains to the<br />

optimization <strong>of</strong> pavement performance. There is a broad-based need for a better underst<strong>and</strong>ing<br />

<strong>of</strong> the mechanisms <strong>of</strong> moisture damage <strong>and</strong> advanced methods for the detection <strong>and</strong> mitigation <strong>of</strong><br />

moisture damage. Progress in these areas will lead to improved material selection for new<br />

pavements as well as enhanced capability for maintenance <strong>of</strong> existing roads.<br />

Work Conducted this Quarter<br />

Work this quarter has involved advancing our capabilities on the Hamburg Wheel Tracking<br />

(HWT) Device <strong>and</strong> developing our know-how on the Gyratory Compactor for the experimental<br />

work <strong>of</strong> the “Fines <strong>and</strong> Organics” project. This study is designed to fully characterize the<br />

inorganic <strong>and</strong> organic material generated during HWT testing. The primary aim <strong>of</strong> this project is<br />

to eventually determine whether the material liberated from a failing core sample can be<br />

correlated with the failure <strong>of</strong> a mix design <strong>and</strong>, furthermore, whether these results correlate with<br />

previous moisture damage studies as well as concurrent studies <strong>of</strong> asphalt-aggregate interfacial<br />

phenomena. To date, testing <strong>of</strong> numerous road cores obtained from various sites has occurred.<br />

Because <strong>of</strong> the varying depths <strong>of</strong> the road cores used, shims <strong>of</strong> different thicknesses were<br />

designed <strong>and</strong> built in the shape <strong>of</strong> the test specimens from aluminum. The shims were built in<br />

the shape <strong>of</strong> test specimens because prototype versions <strong>of</strong> the same width as the wheel resulted in<br />

rutting <strong>of</strong> the bottom <strong>of</strong> the sample. Aluminum was the material <strong>of</strong> choice because it was readily<br />

available, does not react with water, is durable <strong>and</strong> non-compressive, is relatively inexpensive in<br />

comparison to other materials, <strong>and</strong> is easy to work with. Our system is now capable <strong>of</strong> rut<br />

testing cores at a wide variety <strong>of</strong> depths.<br />

During testing it has become increasingly apparent that the inorganic <strong>and</strong> organic sample<br />

collections from HWT may be troublesome. First, the material is generated slowly in the water<br />

bath during an extended time span. It was originally thought that it might be possible to collect<br />

the fines as the experiment progressed, however that my not be the case. Additionally, it is not<br />

yet clear by what method samples can be collected in the case <strong>of</strong> a major failure that involves<br />

adherence <strong>of</strong> a large amount <strong>of</strong> material to the wheel <strong>and</strong> yields very large aggregate samples.<br />

Most likely, useful information cannot be obtained from these mixes since they appear to be a<br />

representation <strong>of</strong> the entire mix. At this point it is believed that collecting two samples at the end<br />

<strong>of</strong> the test will be sufficient. Both a coarse <strong>and</strong> a fine sample will then be isolated for further<br />

testing. The results <strong>of</strong> the initial mineral analyses <strong>of</strong> these samples will guide further sample<br />

collection method development. If, for example, both the course <strong>and</strong> fine material are<br />

representative <strong>of</strong> the entire mix design the sample collection method will have to be redeveloped.<br />

This brings the discussion to another possible sampling issue: collecting material that is smaller<br />

than 30 microns. This would prove to be very challenging <strong>and</strong> might require distillation <strong>of</strong> water<br />

away from the fine material rather than filtration to remove the fines from water.<br />

Finally, after several trials, it is evident that there will be enough organic material to extract from<br />

the aggregate for spectroscopic analysis. If the method <strong>of</strong> sample collection is altered, this may<br />

also cease to be the case.<br />

4


In other work, the Superpave Gyratory Compactor is now being brought on line to produce<br />

samples for the fines <strong>and</strong> organics project. A visit to the Wyoming Department <strong>of</strong> Transportation<br />

was taken by Ryan Boysen, Mike Farrar, <strong>and</strong> Eric Kalberer to get familiarized with the process<br />

as it is being utilized in the field. Since then, numerous, additional components for mixing,<br />

compacting, <strong>and</strong> characterizing gyratory compacted samples have been purchased. Current<br />

efforts are being directed at developing simple mix designs to evaluate the fines <strong>and</strong> organics<br />

project.<br />

Another project that is being developed under the moisture damage task is to revisit the use <strong>of</strong><br />

chromatography to separate asphalt into fractions that could prove useful for the study <strong>of</strong><br />

moisture damage. Chemical functional groups have been identified through past work at WRI as<br />

being strong or weak binders to aggregate in the presence <strong>of</strong> water. Current research on<br />

interfacial properties <strong>of</strong> asphalt <strong>and</strong> aggregate are starting to provide detailed information on<br />

what might truly be binding to the rock. The use <strong>of</strong> high performance liquid chromatography<br />

(HPLC) will be a useful tool in isolating pinpointed fractions <strong>of</strong> the asphalt that could be used as<br />

a pro<strong>of</strong>-<strong>of</strong>-concept for the surface studies. This work relates to some <strong>of</strong> the initial size exclusion<br />

chromatography <strong>and</strong> ion exchange chromatography that was developed at WRI during the SHRP<br />

program. Other recent developments at WRI have involved automating the SARA (saturates,<br />

aromatics, resins, asphaltenes) separation. New methods incorporating past <strong>and</strong> current<br />

developments can be used to isolate subfractions <strong>of</strong> particular interest to moisture damage<br />

studies for doping experiments among other things. For example, if theoretical or interfacial<br />

studies attribute good binding characteristics to an acidic fraction <strong>of</strong> the aromatics fraction in<br />

asphalt, it would then be possible to isolate assorted acidic fractions <strong>of</strong> aromatics from asphalt<br />

<strong>and</strong> to eventually perform doping experiments.<br />

Work to be Conducted Next Quarter<br />

• Continue with the optimization <strong>of</strong> the Hamburg Wheel Tracking Device <strong>and</strong> begin full<br />

utilization <strong>of</strong> the Gyratory Compactor to test mixes with varying compositions from the<br />

SHRP material database. Initial extractions <strong>of</strong> organics from the fines <strong>and</strong> subsequent<br />

spectroscopic studies will be initiated. Additionally, mineral analyses <strong>of</strong> samples will<br />

begin to determine whether proper sample collection methods are being utilized.<br />

• An HPLC instrument will be ordered to begin work on isolating asphalt fractions <strong>of</strong><br />

interest for pro<strong>of</strong> <strong>of</strong> concept studies.<br />

Problems <strong>and</strong> Solution to Problems<br />

Some minor details during process development occurred. These centered on being able to<br />

incorporate cores <strong>of</strong> varying depths for testing. It is believed that the core-mold <strong>and</strong> shim have<br />

effectively alleviated these difficulties. The lack <strong>of</strong> manpower has been essentially alleviated by<br />

the hiring <strong>of</strong> Ryan Boysen. A full-time technician may still be needed for the operation <strong>of</strong> both<br />

the Hamburg <strong>and</strong> Gyratory Compactor as the project progresses. No further problems were<br />

encountered this past quarter.<br />

5


SUBTASK 2-2. AGING<br />

SUBTASK 2-2.1. IMPACT OF WATER ON AGING<br />

Statement <strong>of</strong> Problem<br />

The presence <strong>of</strong> moisture has been shown to increase the rate at which aging progresses in<br />

paving asphalts. Since aging ultimately leads to embrittlement <strong>and</strong> pavement failure,<br />

determining <strong>and</strong> modeling the influence <strong>of</strong> moisture is important to the prediction <strong>of</strong> pavement<br />

performance.<br />

Approach<br />

The sensitivity <strong>of</strong> asphalts to environmental factors is being determined using laboratory PAV<br />

aging tests on unmodified asphalts with <strong>and</strong> without moisture in the oven. Analytical tests<br />

applied include spectroscopic (FTIR) <strong>and</strong> rheologic (DSR) <strong>of</strong> the aged materials. Master curve<br />

<strong>and</strong> shift factors are used to quantify changes.<br />

Goal<br />

Develop relationships that predict the long-term aging <strong>of</strong> asphalts in the presence <strong>of</strong> moisture.<br />

Support <strong>of</strong> FHWA Strategic Goals<br />

The ability to reliably reproduce in the laboratory the aging that occurs in-sevice is vital to the<br />

highway community. With the development <strong>of</strong> a correlation between laboratory aging data <strong>and</strong><br />

field pavement performance, an agency may be able to predict the advent <strong>of</strong> distresses in the<br />

field, therefore, improving the cost effectiveness <strong>of</strong> the preventative maintenance <strong>and</strong>/or<br />

rehabilitation measures that are commonly used. This subtask supports the FHWA Strategic<br />

Goal <strong>of</strong> Optimizing Pavement Performance.<br />

Work Conducted This Quarter<br />

To evaluate the impact <strong>of</strong> moisture on the aging characteristics <strong>of</strong> asphalt binders in terms <strong>of</strong><br />

physical chemical relationships, previously collected data including the rheological <strong>and</strong> chemical<br />

properties <strong>of</strong> eight RTFO- <strong>and</strong> RTFO/PAV-aged SHRP asphalts were analyzed extensively. The<br />

RTFO-aged asphalts were PAV aged for different lengths <strong>of</strong> time in the absence <strong>and</strong> presence <strong>of</strong><br />

water. To further develop an underst<strong>and</strong>ing <strong>of</strong> the relationship between rheological properties<br />

<strong>and</strong> chemical properties mathematically, the complex modulus was plotted against carbonyl<br />

content for different aging times, 140, 240, <strong>and</strong> 480 hours for eight asphalts. The detailed results<br />

were presented in previous quarterly reports (December 31, 2007 <strong>and</strong> March 31, 2008). It was<br />

reported that both the complex modulus <strong>and</strong> carbonyl content increase with increasing aging<br />

time <strong>and</strong> that different asphalts have different sensitivities to aging. The complex modulus<br />

increases more sharply with increasing carbonyl content for some asphalts than for others. A<br />

statistical model, logistic with two parameters, was applied to describe the relationship. This<br />

equation is shown as follows:<br />

7


G<br />

b*<br />

Carbonyl<br />

* = a exp<br />

(2-2.1.1)<br />

Where G* = complex modulus, Pa<br />

a <strong>and</strong> b = coefficients,<br />

Carbonyl = carbonyl content in infrared absorption units.<br />

It was reported earlier that parameter “a” appears to be primarily a function <strong>of</strong> frequency <strong>and</strong><br />

parameter “b” is related to asphalt source.<br />

To determine whether parameter “b” is related to material compatibility, the parameter “b” was<br />

plotted against a compatibility parameter, the Gaestel Index, for eight asphalts as shown in figure<br />

2-2.1.1. It is evident from this figure that the parameter “b” is related to asphalt compatibility in<br />

a monotonic, but non-linear relationship. Detailed information regarding how group<br />

compositions relate to their performance characteristics will be presented at the 45 th annual<br />

Petersen Asphalt Research Conference in Laramie, July 14-16, 2008.<br />

An abstract entitled “Relationship between Group Composition <strong>of</strong> <strong>Asphalts</strong> <strong>and</strong> Their<br />

Performance Characteristics” has been submitted to the “Geohunan International Conference on<br />

Challenges <strong>and</strong> Recent Advances in Pavement Technologies <strong>and</strong> Transportation Geotechnics<br />

(www.geohunan.org)” for consideration for presentation <strong>and</strong> publication.<br />

b-Value<br />

50<br />

40<br />

30<br />

20<br />

AAA-1<br />

AAB-1<br />

AAC-1<br />

AAD-1<br />

AAF-1<br />

AAG-1<br />

AAK-1<br />

AAM-1<br />

ABD<br />

Y=3.351*exp (5.269X) , R 2 =0.81<br />

AAF<br />

AAC<br />

10<br />

AAM<br />

AAK<br />

ABD<br />

AAG<br />

Frequency= 10 rad/s, R<br />

0<br />

0.0 0.1 0.2 0.3 0.4 0.5 0.6<br />

2 =0.81<br />

Gaestel Index<br />

Figure 2-2.1.1. Relationship between b-value <strong>and</strong> Gaestel index for eight different asphalts.<br />

8<br />

AAA<br />

AAD<br />

AAB


To further evaluate how water influences the relationship between complex modulus <strong>and</strong><br />

carbonyl content, the complex modulus was plotted against carbonyl content with respect to<br />

different aging times in the PAV in the presence <strong>of</strong> water for different asphalts, as shown in<br />

figure 2-2.1.2. It appears that the same model, a logistic function with two parameters, can be<br />

applied for the wet oxidized samples. The relationship between the complex modulus <strong>and</strong><br />

carbonyl content for wet oxidized samples is similar to those <strong>of</strong> dry oxidized samples, where the<br />

complex modulus <strong>and</strong> carbonyl content increase with increased aging time, <strong>and</strong> different asphalts<br />

have different sensitivities to oxidative aging.<br />

Complex Modulus, G*, Pa<br />

5e+5<br />

4e+5<br />

3e+5<br />

2e+5<br />

1e+5<br />

0<br />

Control<br />

AAA-1<br />

AAB-1<br />

AAC-1<br />

AAD-1<br />

AAF-1<br />

AAG-1<br />

AAK-1<br />

AAM-1<br />

ABD<br />

Wet Oxidation<br />

AAD-1<br />

AAA-1<br />

0.0 0.1 0.2 0.3 0.4 0.5<br />

Carbonyl Content, a.u.<br />

9<br />

480 hrs<br />

PAV at 80°C<br />

ABD<br />

AAG-1<br />

Figure 2-2.1.2. Relationship between complex modulus <strong>and</strong> carbonyl content for different<br />

asphalts before <strong>and</strong> after PAV aging in the presence <strong>of</strong> water.<br />

Work to be Conducted Next Quarter<br />

● Prepare the manuscript entitled “Relationship between Group Composition <strong>of</strong> <strong>Asphalts</strong><br />

<strong>and</strong> Their Performance Characteristics” for the conference.<br />

● The same approach used for dry oxidized samples will be continually employed to the<br />

wet oxidized samples, <strong>and</strong> parameter values will be compared to determine how water<br />

influences oxidative aging.<br />

● Correlate oxidation master curve parameters (i.e., shift factor <strong>and</strong> rheological index) to<br />

chemical parameters (i.e., FTIR <strong>and</strong> compatibility parameters, Corbett fractions) as well<br />

as low temperature properties (i.e., data from DSC) on the aged asphalts.


● Construct oxidation master curves for asphalts recovered from field core samples <strong>and</strong><br />

correlate master curve parameters (i.e., shift factor <strong>and</strong> rheological index) to chemical<br />

parameters (i.e., carbonyl content <strong>and</strong> compatibility parameters).<br />

● Prepare samples with different levels <strong>of</strong> asphalt film thickness (from thin film <strong>of</strong> 50<br />

microns to st<strong>and</strong>ard film thickness <strong>of</strong> 1/8 inches) <strong>and</strong> subject to PAV aging at pavement<br />

service temperature in the presence <strong>and</strong> absence <strong>of</strong> water to investigate the relationship<br />

between oxidation <strong>and</strong> diffusion.<br />

Problems <strong>and</strong> Solution to Problem<br />

None<br />

10


SUBTASK 2-2.2. SUPPORT OF THE MECHANISTIC-EMPIRICAL PAVEMENT<br />

DESIGN GUIDE<br />

Statement <strong>of</strong> Problem<br />

The global aging system is a series <strong>of</strong> models that attempt to predict the change in binder<br />

viscosity in hot-mix asphalt (HMA) pavement with time <strong>and</strong> depth. These models are an integral<br />

part <strong>of</strong> the NCHRP 1-37A Mechanistic-Empirical Pavement Design Guide. The models were<br />

developed from an extensive database consisting <strong>of</strong> capillarity viscosity, <strong>and</strong> penetration <strong>and</strong><br />

s<strong>of</strong>tening point measurements converted to viscosity, over a broad range <strong>of</strong> temperatures [Mirza<br />

<strong>and</strong> Witczak 1995]. Christensen <strong>and</strong> Bonaquist [2006] modified the global aging system models<br />

to make use <strong>of</strong> rational rheological measurements <strong>and</strong> binder master curve parameters while<br />

maintaining consistency with the original models. They noted in regard to a part <strong>of</strong> their study,<br />

which employed the global aging system, that the results should be considered approximate,<br />

since “there are many questions concerning the accuracy <strong>of</strong> the global aging system.”<br />

WRI, also in 2006, evaluated how well the GAS models predicted the actual aging that occurred<br />

at the FHWA/WRI Arizona validation site [Farrar et al. 2006]. One <strong>of</strong> the main findings <strong>of</strong> the<br />

study was that oxidative aging occurring at the Arizona site was substantially greater than<br />

predicted by the GAS models, particularly in the top 13 mm <strong>of</strong> the pavement. Al-Azri et al.<br />

[2006] in an extensive investigation <strong>of</strong> selected Texas highway pavements concerning<br />

unmodified binder aging reported the level <strong>of</strong> hardening reached in pavement binders<br />

significantly exceeded estimated values calculated by the global aging system both at the<br />

pavement surface <strong>and</strong> at 125 mm below the surface.<br />

More specifically, the global aging system was developed based on laboratory tests <strong>of</strong><br />

conventional “S” binders. These linear or straight-line Class “S” asphalts were non-modified,<br />

<strong>and</strong> the authors <strong>of</strong> the global aging system (Mirza <strong>and</strong> Witczak) advised the global aging models<br />

should not be used for modified asphalts, Class “W” (waxy) or Class “B” (blown) asphalts. The<br />

S, B, <strong>and</strong> W designations were originally developed by Heukelom [1973] who observed the<br />

linear <strong>and</strong> nonlinear relationships <strong>of</strong> asphalts using the Bitumen Test Data Chart developed in the<br />

1960’s.<br />

In addition: (1) the global aging system does not address photo-oxidation effects on the surface,<br />

(2) the air void adjustment factor is based on limited data <strong>and</strong> considered “optional,” (3) the<br />

global aging system does not take into account sealer/rejuvenator, chip seals, or open graded<br />

friction course effects on pavement aging, <strong>and</strong> (4) the global aging system is not based on<br />

“fundamental” binder properties.<br />

Approaches<br />

The accuracy <strong>and</strong> possible modifications to the global aging system are being studied at WRI by<br />

analyses <strong>of</strong> data <strong>and</strong> materials from the FHWA/WRI validation sites <strong>and</strong> laboratory aging<br />

(RTFO <strong>and</strong> PAV). Analytical tests applied include infrared spectroscopy <strong>and</strong> dynamic shear <strong>and</strong><br />

bending beam rheometry <strong>of</strong> unaged <strong>and</strong> aged materials. Master curves <strong>of</strong> unaged <strong>and</strong> aged<br />

asphalts <strong>and</strong> age based shift factors are being studied to quantify age related changes.<br />

11


Goals<br />

The purpose <strong>of</strong> this subtask is tw<strong>of</strong>old: (1) to improve the general underst<strong>and</strong>ing <strong>of</strong> the<br />

physicochemical changes that occur with time <strong>and</strong> depth in asphalt pavement due to aging, <strong>and</strong><br />

(2) to recommend specific modifications to the global aging system to improve its accuracy <strong>and</strong><br />

exp<strong>and</strong> its application to modified asphalts.<br />

Support <strong>of</strong> FHWA Strategic Goals<br />

The work conducted in this subtask supports FHWA’s ultimate pavement research <strong>and</strong><br />

development goal <strong>of</strong> providing performance-based models <strong>and</strong> tools to facilitate effective<br />

management <strong>of</strong> the national highway infrastructure. Exp<strong>and</strong>ing our underst<strong>and</strong>ing <strong>of</strong> how<br />

pavements age in terms <strong>of</strong> time <strong>and</strong> depth is critical to developing accurate models to predict<br />

rutting <strong>and</strong> cracking in rehabilitation <strong>and</strong> new construction projects, <strong>and</strong> to allow better timing <strong>of</strong><br />

preventive maintenance strategies that are cost effective in mitigating pavement aging.<br />

Background<br />

Over the last several decades, there have been many studies on oxidative aging in asphalt<br />

pavements; however, very few have dealt with how oxidative aging varies with depth. Of the<br />

few studies available on the subject, perhaps the most well known is by Coons <strong>and</strong> Wright<br />

[1968]. In this study <strong>and</strong> other related studies [Mirza <strong>and</strong> Witczak 1995; Houston et al. 2005;<br />

Farrar et al. 2006; Al-Azri et al. 2006], the method used to make the evaluation involved slicing<br />

the cores parallel to the pavement surface. In these studies the thickness <strong>of</strong> the slices varied from<br />

as little as 1.6 mm (1/16 inch) in the upper 6.3 mm (1/4 inch) <strong>of</strong> the core in the Coons <strong>and</strong><br />

Wright study, to 25 mm (1 inch) in the upper area <strong>of</strong> the core in the Houston et al. study. How<br />

the cores were sliced may have determined the conclusions reached. For example, the Coons<br />

<strong>and</strong> Wright study showed an increase <strong>of</strong> roughly 50% in the viscosity <strong>of</strong> asphalt recovered from<br />

the top 6.3-mm slice versus asphalt recovered from the next lower 6.3-mm slice. In addition, a<br />

very fine film at the surface <strong>of</strong> the core had a higher viscosity than the average viscosity in the<br />

top 6.3 mm. On the other h<strong>and</strong>, the Houston et al. study found that the viscosity did not vary<br />

significantly with depth. Since different pavement designs, binders, <strong>and</strong> ages were used in these<br />

studies, the results for both may be valid. However, it is also possible that the 25 mm thickness<br />

<strong>of</strong> the slices used in the Houston et al. study masked a viscosity pr<strong>of</strong>ile that would have been<br />

apparent if thinner slices had been examined.<br />

The opposing conclusions are not surprising <strong>and</strong> point out the difficult nature <strong>of</strong> not only<br />

measuring aging in HMA, but developing a global aging system to predict aging. An extensive<br />

literature review indicates the global aging system as developed by Mirza <strong>and</strong> Witczak is the<br />

only system or set <strong>of</strong> models attempting to predict binder aging in HMA pavement on a “global”<br />

basis. Research to date suggests that while the system is an excellent starting point it is at best<br />

“semi-global” <strong>and</strong> given that it is an integral part <strong>of</strong> the MEPDG, research to improve its<br />

predictive capability is clearly warranted.<br />

12


Work Conducted This Quarter<br />

Introduction<br />

Both liquid-cell <strong>and</strong> photoacoustic (PA) infrared spectroscopy experiments are described for<br />

aged <strong>and</strong> unaged asphalt samples. In addition, aggregates <strong>and</strong> mixed aggregate-asphalt samples<br />

(mastics) are examined using PA spectroscopy. PA spectroscopy is unique in several ways to<br />

other more traditional infrared sampling techniques such as attenuated total reflectance (ATR).<br />

For example, the PA technique requires little sample preparation <strong>and</strong> has low sensitivity to<br />

surface condition [McClell<strong>and</strong> et al. 2002]. This is a very important concept that makes it<br />

possible to examine photo-oxidation <strong>of</strong> the pavement surface without use <strong>of</strong> solvents.<br />

Also, an evaluation <strong>of</strong> pavement distress at the FHWA/ WRI Arizona validation site <strong>and</strong> related<br />

rheological parameters is presented.<br />

Data presented here were collected under the <strong>Fundamental</strong> <strong>Properties</strong> <strong>of</strong> <strong>Asphalts</strong> <strong>and</strong> <strong>Modified</strong><br />

<strong>Asphalts</strong>, <strong>III</strong>, FHWA project (DTFH61-07-D-00005) <strong>and</strong> to a small extent under the Asphalt<br />

Surface Aging Prediction (ASAP) project (DTOS59-07-H-0006) for the Research <strong>and</strong> Innovative<br />

Technology Administration (RITA). The projects are complimentary in terms <strong>of</strong> achieving a<br />

better underst<strong>and</strong>ing <strong>of</strong> how asphalt pavement ages, but the ASAP project is specifically directed<br />

towards providing a tool <strong>and</strong> methodologies for determining the appropriate time for the<br />

preemptive application <strong>of</strong> rejuvenating surface treatments.<br />

Experimental<br />

Asphalt Aging<br />

Aging protocols, for this project, included the rolling-thin-film oven (RTFO, AASHTO T-240)<br />

method <strong>and</strong> the pressurized aging vessel (PAV, AASHTO R28) method.<br />

Rheology<br />

The complex shear modulus <strong>and</strong> phase angle <strong>of</strong> the original <strong>and</strong> laboratory aged asphalts were<br />

measured using a Rheometrics Model RDA II Dynamic Analyzer (a research-grade dynamic<br />

shear rheometer, DSR). DSR measurements for the SHRP asphalts were performed at 25˚C <strong>and</strong><br />

60˚C <strong>and</strong> a frequency range <strong>of</strong> 0.1 to 100 radians/sec at each temperature. DSR measurements<br />

for the Arizona validation site asphalts were performed at 10 degree intervals over a temperature<br />

range <strong>of</strong> 0 to 80°C <strong>and</strong> a frequency range <strong>of</strong> 0.1 to 100 radians/second.<br />

FTIR Instrument<br />

The infrared spectrometer used for this survey was a Perkin-Elmer Spectrum One ® equipped<br />

with a deuterated triglycine sulfate (DTGS) detector. The spectrometer was controlled by a<br />

desktop PC running Spectrum v5.0.1 s<strong>of</strong>tware. Available optical path difference (OPD)<br />

velocities were 0.1, 0.2, 0.5, 1.0 <strong>and</strong> 2.0 cm/sec.<br />

13


Transmission – liquid-cell technique<br />

Liquid-cell samples were prepared as solutions (50 mg asphalt per one ml solvent) <strong>and</strong> placed<br />

into a 0.99-mm (path length) KBr cell before infrared spectra collection. A single beam<br />

background spectrum <strong>of</strong> the liquid cell with solvent was acquired daily prior to acquiring single<br />

beam sample spectra. The solvent used was carbon disulfide (CS2). All data were collected at a<br />

resolution <strong>of</strong> 4 cm -1 , spectral range 600 – 4000 cm -1 , <strong>and</strong> each spectra was derived from 32 coadded<br />

scans.<br />

Photoacoustic technique<br />

An MTEC Model 300 photoacoustic detector (acquired under RITA contract) coupled to the<br />

Spectrum One FTIR spectrometer as shown in figure 2-2.2.1 was used to acquire absorption<br />

spectra. All data were collected in rapid-scan (non-phase modulation) mode at a resolution <strong>of</strong><br />

8 cm -1 <strong>and</strong> the spectral range was 600–4000 cm -1 . The number <strong>of</strong> scans co-added depended on<br />

scanning speed. To improve the signal-to-noise ratio, in some cases a series <strong>of</strong> spectra were<br />

obtained <strong>and</strong> averaged where each spectrum corresponded to a relatively short measurement<br />

time. OPD velocity, amplifier gain setting, <strong>and</strong> number <strong>of</strong> scans are reported with each<br />

spectrum. Carbon black (MTEC Photoacoustics) was used as the normalization reference. The<br />

PA detector was purged with helium. A 1-mm deep sample cup <strong>and</strong> brass sample holder (MTEC<br />

Photoacoustics) were used to contain <strong>and</strong> insert the samples into the MTEC acoustic chamber.<br />

Photoacoustic - background<br />

Figure 2-2.2.1. MTEC 300 Photoacoustic spectrometer coupled to a<br />

Perkin Elmer FTIR Spectrum One.<br />

Since PA spectroscopy is a somewhat uncommon FTIR technique a more detailed description<br />

has been included in this report PA spectroscopy can probe surface composition over a range <strong>of</strong><br />

selectable sampling depths from several micrometers to more than 100 micrometers (samples<br />

14


greater than 100 micrometers are considered “thermally thick”). PAS directly measures IR<br />

absorption by sensing absorption induced heating <strong>of</strong> the sample within an experimentally<br />

controllable sample depth below the samples surface [McClell<strong>and</strong> et al. 2002].<br />

The PA sampling (thermal diffusion) depth for a homogeneous sample is conventionally<br />

expressed as [Rosencwaig <strong>and</strong> Gersho 1976]<br />

L π<br />

1<br />

2 = ( D / f )<br />

(2-2.2.1)<br />

where D is the thermal diffusivity, a measure <strong>of</strong> heat propagation speed, in cm 2 /sec, <strong>and</strong> f is the<br />

infrared-intensity modulation frequency in hertz. The equation for thermal diffusivity (D) is<br />

D = k/ρCp (2-2.2.2)<br />

where k is the thermal conductivity in W/mK, ρ is density in kg/m 3 , <strong>and</strong> Cp is the specific heat<br />

in J/kg K. From the literature it is estimated for an asphalt at 25˚C, typical values for ρ, k <strong>and</strong> Cp<br />

are 1030 kg/m 3 , 0.75 W/m K <strong>and</strong> 920 J/kg K, respectively. Frequency modulation ( f ) can be<br />

expressed as a function <strong>of</strong> V the optical path difference (OPD) velocity <strong>of</strong> the interferometer<br />

mirror in cm/sec <strong>and</strong> ν the wave number in cm -1 .<br />

f = Vν<br />

(2-2.2.3)<br />

Combining equations 2-2.2.1 <strong>and</strong> 2-2.2.3 results in<br />

1<br />

2<br />

L = ( D / πVν<br />

)<br />

(2-2.2.4)<br />

The sampling depth is inversely proportional to the square root <strong>of</strong> the modulation frequency <strong>and</strong><br />

therefore low modulation frequency or OPD velocity will result in PA signals from within the<br />

sample, while high OPD velocities are nearer the surface region. The dependence <strong>of</strong> the<br />

sampling depth on wavenumber (ν) means that it varies across a spectrum. For asphalt <strong>and</strong> the<br />

thermal diffusivity estimated above, at an OPD velocity <strong>of</strong> 0.5 cm/s the thermal diffusion length<br />

varies from 29 µm at 600 cm -1 to 11 µm at 4000 cm -1 . Thermal diffusion lengths at various<br />

OPD velocities <strong>and</strong> selected wave numbers from 600 to 4000 cm -1 corresponding to areas <strong>of</strong><br />

significant asphalt absorbance peaks are listed in table 2-2.2.1.<br />

Depth variation across the spectrum can be overcome by using a step-scan spectrometer<br />

employing phase modulation photoacoustic measurements [Jones <strong>and</strong> McClell<strong>and</strong> 1996;<br />

Drapcho et al. 1997]. Step-scan, as described by Drapcho et al. is the case where the “average<br />

optical retardation is adjusted stepwise, with a secondary, single-frequency phase (optical path<br />

difference) modulation applied. The data are collected when the average optical retardation is<br />

kept constant; thus the velocity <strong>and</strong> wavenumber dependence <strong>of</strong> the Fourier frequency is<br />

removed, <strong>and</strong> a constant probing depth across the spectrum for each phase modulation (PM)<br />

frequency is achieved.”<br />

15


Materials<br />

Table 2-2.2.1. Sampling depth due to varying OPD velocity at selected wavenumbers.<br />

SHRP <strong>Asphalts</strong><br />

Strategic Highway Research Program (SHRP) asphalts: AAB-1, AAC-1, AAD-1, <strong>and</strong> AAM-1<br />

were used in this study. The four asphalts were RTFO/PAV aged at 60 <strong>and</strong> 80˚C for various<br />

times (20, 144, 240, <strong>and</strong> 480 hours).<br />

Arizona Validation Site <strong>Asphalts</strong><br />

The Arizona WRI/FHWA validation site is located on the south bound lane <strong>of</strong> US 93,<br />

approximately 50 miles north <strong>of</strong> Wickenburg, Arizona at about milepost 153. The contractor’s<br />

asphalt <strong>and</strong> asphalt from three other sources were used to construct the site. The four asphalts<br />

are generically identified here as AZ1-1 thru AZ1-4. AZ1-1 was produced from a West Texas<br />

intermediate sour blend; AZ1-2 from a Venezuelan crude; AZ1-3 from a Rocky Mountain blend;<br />

<strong>and</strong> AZ1-4 from a Canadian crude. AZ1-1, AZ1-3, <strong>and</strong> AZ1-4 were classified as performance<br />

grade PG 76-16, <strong>and</strong> AZ1-2 as PG 76-22.<br />

The four asphalts were analyzed for the presence <strong>of</strong> polyphosphoric acid <strong>and</strong> polymers using<br />

nuclear magnetic resonance (NMR) spectroscopy Neither PPA nor SBS polymer was detected in<br />

the asphalt samples as indicated by phosphorus NMR <strong>and</strong> proton NMR analyses. The detection<br />

limits for phosphorous <strong>and</strong> polymers using this method were estimated to be 0.1 <strong>and</strong> 0.5 wt %,<br />

respectively.<br />

Aggregates<br />

Wavenumber (v), cm -1<br />

Aggregates used for PA analysis are referred in this report as RA, RD (note RA <strong>and</strong> RD are<br />

SHRP aggregates), <strong>and</strong> Arizona aggregate. RA aggregate consists primarily <strong>of</strong> granite with a<br />

minor amount <strong>of</strong> basalt. Major RA elemental oxides are SiO2 (71%) <strong>and</strong> Al2O3 (16%). Major<br />

mineral components are quartz (55%), K-feldspar (25%), <strong>and</strong> plagioclase (10%). RD is a<br />

16<br />

Sampling Depth L, µm<br />

OPD velocity (V), cm/s<br />

0.1 0.2 0.5 1.0 2.0<br />

4000 25 18 11 8 6<br />

3000 29 20 13 9 6<br />

1600 40 28 18 13 9<br />

1457 42 29 19 13 9<br />

1376 43 30 19 14 10<br />

870 54 38 24 17 12<br />

720 59 42 26 19 13<br />

600 65 46 29 20 14


limestone. Major RA elemental oxides are CaO (39%), <strong>and</strong> SiO2 (17%). Major mineral<br />

components are calcite (61 %), quartz (7.4%), <strong>and</strong> organics (5%) [Robl 1991].<br />

FTIR-PA analysis was performed on RD <strong>and</strong> RA aggregate batched to the gradation shown in<br />

table 2-2.2.2.<br />

FTIR-PA analysis on the Arizona aggregate was performed on the minus 2.36 mm plus<br />

0.425 mm aggregate from the combined Arizona plant mix aggregate. Lime (approximately 1%<br />

by mass <strong>of</strong> the aggregate) may have been added before aggregate samples were collected on the<br />

project.<br />

Laboratory prepared mastic<br />

Table 2-2.2.2. Gradation for fine, dense graded mastic.<br />

Sieve Size, mm Percent Passing<br />

1.19 100<br />

0.600 75<br />

0.300 50<br />

0.150 30<br />

0.075 20<br />

Mastics were prepared by mixing aged AAD-1 asphalt (RTFO/PAV 80˚C, 480 hours) with RD<br />

<strong>and</strong> RA aggregate (see table 2-2.2.2 for mastic gradation). Four relative concentrations <strong>of</strong> each<br />

mastic were prepared: 20% aggregate/80% asphalt, 50% aggregate/50% asphalt, 80%<br />

aggregate/20% asphalt, <strong>and</strong> 90% aggregate/10% asphalt.<br />

Laboratory aged compacted specimens<br />

Four 75-mm diameter by 120 mm tall compacted specimens were received from North Carolina<br />

State University (NCSU, Dr. Y. Richard Kim). The specimens had been cored from larger 150mm<br />

diameter gyratory compacted specimens. A 9.5-mm nominal maximum sized aggregate was<br />

used to fabricate the specimens. The aggregate was a limestone obtained from a quarry in North<br />

Carolina. SHRP AAD-1 was used as the binder (6.2% by mass <strong>of</strong> the compacted specimen).<br />

Results <strong>and</strong> Discussion<br />

Asphalt Aging <strong>and</strong> FTIR Spectroscopy<br />

In this study, we report a Fourier-Transform Infrared (FTIR) spectroscopic survey <strong>of</strong> eight<br />

asphalts in unaged <strong>and</strong> laboratory aged conditions as well as several common aggregates used in<br />

road construction <strong>and</strong> combined asphalt/aggregate samples prepared in the laboratory. FTIR<br />

photoacoustic (FTIR-PA), <strong>and</strong> transmission liquid-cell techniques were employed to perform the<br />

survey. Results from the survey will be used here <strong>and</strong> in later studies to: (1) compare FTIR-PA<br />

<strong>and</strong> liquid cell techniques; (2) explore <strong>and</strong> confirm previously observed relationships between<br />

17


asphalt chemical composition <strong>and</strong> rheological parameters; (3) determine the feasibility <strong>of</strong><br />

obtaining asphalt chemical composition from infrared spectra <strong>of</strong> combined laboratory<br />

asphalt/aggregate mastic <strong>and</strong> from the actual pavement surface;<br />

Liquid-Cell FTIR Spectra<br />

Traditionally, the mid-infrared carbonyl (~ 1700 cm -1 ) <strong>and</strong> sulfoxide (~ 1030 cm -1) ) absorbance<br />

b<strong>and</strong>s have been used to gage oxidation in asphalt. Although, as demonstrated last quarter, there<br />

appear to be important areas such as the region from 1150 to 1250 cm-1 shown in figure 2-2.2.2<br />

that appear to correlate to rheological properties.<br />

However, until these regions <strong>of</strong> the mid-infrared spectrum <strong>of</strong> asphalt are better understood <strong>and</strong><br />

investigated, the key relationship that allows the complex shear modulus (G*) <strong>of</strong> the binder<br />

component <strong>of</strong> pavement to be measured using infrared spectra is the correlation between the<br />

change in Log G* <strong>and</strong> the change in concentration <strong>of</strong> oxidation products, measured here as the<br />

carbonyl content. For dynamic shear rheometry tests performed at 25°C <strong>and</strong> 60°C, the qualities<br />

<strong>of</strong> the correlations are shown in figures 2-2.2.3 <strong>and</strong> 2-2.2-4. The fit data include all aged<br />

asphalts: RTFO, PAV at 60°C, PAV at 80°C, <strong>and</strong> PAV at 100°C. R-squared values for DSR<br />

tests at 25°C average 0.94, <strong>and</strong> for tests at 60°C they average 0.97. In field use, it is anticipated<br />

that rheological master curves can be applied to show the stiffness <strong>of</strong> the asphalt at all in-service<br />

temperatures <strong>and</strong> loading frequencies.<br />

Absorbance<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

2000<br />

Carbonyl<br />

1800<br />

1600<br />

Solvent<br />

1400<br />

Figure 2-2.2.2. Changes in infrared spectra caused by oxidative aging.<br />

18<br />

1200<br />

Wave Number, cm -1<br />

AAB-1 neat<br />

AAB-1 RTFO<br />

AAB-1 PAV 100°C 20hrs<br />

AAB-1 PAV 80°C 480hrs<br />

1150 - 1250 cm -1 Region<br />

1000<br />

800<br />

600


Log G* at 10 rad/s<br />

8<br />

7<br />

6<br />

5<br />

25°C<br />

y = 7.349x + 5.626<br />

R 2 = 0.923<br />

19<br />

y = 3.440x + 5.645<br />

R 2 = 0.962<br />

y = 4.145x + 6.050<br />

R 2 = 0.924<br />

y = 2.202x + 6.205<br />

R 2 = 0.946<br />

AAB-1<br />

AAC-1<br />

AAD-1<br />

AAM-1<br />

0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35<br />

Carbonyl Content, au<br />

Figure 2-2.2.3. Change in Log G* measured at 25°C as a function <strong>of</strong> carbonyl content<br />

for SHRP asphalts.<br />

Log G* at 10 rad/s<br />

6<br />

5<br />

4<br />

3<br />

2<br />

60°C<br />

y = 10.758x + 3.214<br />

R 2 = 0.974<br />

y = 7.081x + 3.459<br />

R 2 = 0.970<br />

y = 4.430x + 3.593<br />

R 2 = 0.972<br />

y = 4.945x + 2.823<br />

R 2 = 0.958<br />

AAB-1<br />

AAC-1<br />

AAD-1<br />

AAM-1<br />

0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35<br />

Carbonyl Content, au<br />

Figure 2-2.2.4. Change in Log G* measured at 60°C as a function <strong>of</strong> carbonyl content<br />

for SHRP asphalts.


A summary <strong>of</strong> aging, FTIR, <strong>and</strong> rheology data for the SHRP asphalts is provided in tables<br />

2-2.2.3a <strong>and</strong> 2-2.2.3b. Results for G* are reported at an angular frequency <strong>of</strong> 10 rad/s. Tables<br />

2-2.2.4a <strong>and</strong> 2-2.2.4b contain similar data for the laboratory aged field samples.<br />

Table 2-2.2.3a. Experiment status matrix for SHRP asphalt aging <strong>and</strong> testing.<br />

Sample 1<br />

Solution FTIR<br />

Analysis<br />

(corr. peak ht.)<br />

PA FTIR<br />

Analysis<br />

(total peak ht.)<br />

Carbonyl 2 , au 2 Carbonyl 2 , au 2<br />

20<br />

DSR - G*, 10 rad/s<br />

(Pa)<br />

DSR - δ, 10 rad/s<br />

25°C 60°C 25°C 60°C<br />

AAB Unaged 0.0100 0.2600 4.42E+05 1.42E+03 65.396 86.627<br />

AAB RTFO only 0.0102 0.4600 8.71E+05 2.59E+03 62.247 84.455<br />

AAB 100°C 20hrs 0.0675 0.7600 2.44E+06 1.13E+04 49.625 77.173<br />

AAB 60°C 20hrs 0.0243 0.4800 1.22E+06 4.19E+03 58.791 83.008<br />

AAB 60°C 144hrs 0.0489 0.7700 2.11E+06 7.19E+03 53.063 80.805<br />

AAB 60°C 240hrs 0.0666 0.8000 2.64E+06 9.61E+03 51.798 79.349<br />

AAB 60°C 480hrs 0.1057 0.9600 3.41E+06 1.38E+04 47.911 77.342<br />

AAB 80°C 20hrs 0.0300 0.5500 1.51E+06 1.90E+04 3<br />

57.035 81.825<br />

AAB 80°C 144hrs 0.1255 0.8900 4.43E+06 2.23E+04 43.732 73.66<br />

AAB 80°C 240hrs 0.1725 1.0300 4.42E+06 3.30E+04 41.255 68.246<br />

AAB 80°C 480hrs 0.2394 1.300 1.05E+07 1.74E+05 31.707 53.048<br />

AAC Unaged 0.0061 0.2700 1.21E+05 4.00E+02 78.400 89.106<br />

AAC RTFO only 0.0288 0.4200 4.23E+05 8.15E+02 70.570 87.702<br />

AAC 100°C 20hrs 0.1242 0.8100 1.43E+06 3.56E+03 54.355 82.707<br />

AAC 60°C 20hrs 0.0370 0.4500 2.17E+06 3<br />

1.22E+03 50.007 86.220<br />

AAC 60°C 144hrs 0.1079 0.7700 1.15E+06 2.17E+03 59.246 85.369<br />

AAC 60°C 240hrs 0.1384 0.8200 1.42E+06 2.60E+03 57.439 84.805<br />

AAC 60°C 480hrs 0.1810 0.9900 1.75E+06 4.18E+03 52.582 82.285<br />

AAC 80°C 20hrs 0.0828 0.6000 8.82E+05 1.81E+03 60.944 86.087<br />

AAC 80°C 144hrs 0.1804 0.8800 2.06E+06 5.86E+03 49.516 80.677<br />

AAC 80°C 240hrs 0.2340 1.0000 2.72E+06 7.27E+03 45.453 77.738<br />

AAC 80°C 480hrs 0.3014 1.2400 4.20E+06 2.64E+04 38.004 67.481<br />

1 All samples displaying temperatures <strong>and</strong> times are PAV aged.<br />

2 au = absorption units


Table 2-2.2.3b. Experiment status matrix for SHRP asphalt aging <strong>and</strong> testing.<br />

Sample 1<br />

Solution FTIR<br />

Analysis<br />

(corr. peak ht.)<br />

PA FTIR<br />

Analysis<br />

(total peak ht.)<br />

21<br />

DSR - G*, 10 rad/s<br />

(Pa)<br />

DSR - δ, 10 rad/s<br />

carbonyl, au 2 carbonyl, au 2 25°C 60°C 25°C 60°C<br />

AAD Unaged 0.0411 0.4900 1.49E+05 1.22E+03 69.344 84.784<br />

AAD RTFO only 0.0461 0.6000 5.01E+05 3.30E+03 63.524 79.601<br />

AAD 100°C 20hrs 0.0722 0.8300 1.82E+06 1.47E+04 52.812 69.739<br />

AAD 60°C 20hrs 0.0457 0.6400 8.11E+05 4.13E+03 59.913 76.413<br />

AAD 60°C 144hrs 0.0605 0.6100 1.24E+06 7.74E+03 57.096 74.954<br />

AAD 60°C 240hrs 0.0762 0.7300 1.82E+06 1.35E+04 54.473 72.31<br />

AAD 60°C 480hrs 0.1008 0.7900 2.70E+06 1.92E+04 51.152 70.142<br />

AAD 80°C 20hrs 0.0571 0.6300 1.05E+06 6.65E+03 58.468 75.594<br />

AAD 80°C 144hrs 0.1129 0.9100 3.70E+06 2.93E+04 47.444 66.497<br />

AAD 80°C 240hrs 0.1419 1.0000 5.87E+06 6.03E+04 42.331 57.757<br />

AAD 80°C 480hrs 0.2335 1.2000 1.60E+07 4.52E+05 30.153 40.798<br />

AAM Unaged 0.0137 0.3300 1.06E+06 2.76E+03 60.03 85.363<br />

AAM RTFO only 0.0353 0.5300 1.51E+06 5.08E+03 55.686 82.066<br />

AAM 100°C 20hrs 0.1397 --- 3.88E+06 2.21E+04 41.238 71.16<br />

AAM 60°C 20hrs 0.0580 0.6200 2.17E+06 7.01E+03 49.662 80.485<br />

AAM 60°C 144hrs 0.1277 0.7300 3.39E+06 1.48E+04 43.456 75.459<br />

AAM 60°C 240hrs 0.1602 0.8700 3.99E+06 2.05E+04 42.116 72.573<br />

AAM 60°C 480hrs 0.2087 0.9600 4.52E+06 2.56E+04 39.265 70.381<br />

AAM 80°C 20hrs 0.0865 0.6600 2.57E+06 9.40E+03 47.562 78.233<br />

AAM 80°C 144hrs 0.2199 0.9200 4.59E+06 4.06E+04 38.106 66.248<br />

AAM 80°C 240hrs 0.2667 1.0600 6.13E+06 4.83E+04 34.856 62.808<br />

AAM 80°C 480hrs 0.3335 1.3100 8.05E+06 1.33E+05 29.889 52.363<br />

1 All samples displaying temperatures <strong>and</strong> times are PAV aged.<br />

2 au = absorption units


Table 2-2.2.4a. Experiment status matrix for laboratory aged field asphalts.<br />

Sample 1<br />

Solution FTIR<br />

Analysis<br />

(total peak ht.)<br />

PA FTIR<br />

Analysis<br />

(total peak ht.)<br />

22<br />

DSR - G*, 10 rad/s<br />

(Pa)<br />

DSR - δ, 10 rad/s<br />

carbonyl, au 2 carbonyl, au 2 20°C 60°C 20°C 60°C<br />

AZ1-1 Unaged 0.11 0.58<br />

AZ1-1 RTFO only<br />

AZ1-1 100°C 20hrs<br />

0.23 7.27e+06 25919 38.889 70.521<br />

AZ1-1 60°C 96hrs 0.30 9.27E+06 56422 34.988 64.158<br />

AZ1-1 60°C 192hrs 0.33 9.92E+06 64523 33.925 62.779<br />

AZ1-1 60°C 336hrs 0.38 1.31E+07 71467 30.98 61.335<br />

AZ1-1 60°C 504hrs 0.39 1.39E+07 87137 30.146 60.441<br />

AZ1-1 80°C 96hrs 0.35 0.93 1.22E+07 66571 31.871 61.538<br />

AZ1-1 80°C 210hrs 0.45 1.02 1.70E+07 1.56E+05 27.801 53.363<br />

AZ1-1 80°C 336hrs 0.47 1.29 1.97E+07 2.03E+05 25.947 49.852<br />

AZ1-1 80°C 504hrs 0.53 1.32 2.05E+07 3.39E+05 24.137 45.539<br />

AZ1-2 Unaged 0.05 0.43<br />

AZ1-2 RTFO only 0.11 6.02E+06 23590 48.359 69.509<br />

AZ1-2 100°C 20hrs<br />

AZ1-2 60°C 96hrs 0.15 1.05E+07 32391 42.678 66.671<br />

AZ1-2 60°C 192hrs 0.18 1.10E+07 49491 41.997 64.307<br />

AZ1-2 60°C 336hrs 0.21 1.19E+07 50063 41.004 64.084<br />

AZ1-2 60°C 504hrs 0.24 1.39E+07 89396 39.185 59.675<br />

AZ1-2 80°C 96hrs 0.21 0.54 1.17E+07 48744 40.306 63.837<br />

AZ1-2 80°C 210hrs 0.32 0.79 2.09E+07 1.65E+05 32.224 53.32<br />

AZ1-2 80°C 336hrs 0.45 1.07 3.09E+07 3.24E+05 27.928 46.597<br />

AZ1-2 80°C 504hrs 0.52 1.35 4.56E+07 1.94E+06 21.097 28.292<br />

1 All samples displaying temperatures <strong>and</strong> times are PAV aged.<br />

2 au = absorption units


Table 2-2.2.4b. Experiment status matrix for laboratory aged field asphalts.<br />

Sample 1<br />

Solution FTIR<br />

Analysis<br />

(total peak ht.)<br />

PA FTIR<br />

Analysis<br />

(total peak ht.)<br />

23<br />

DSR - G*, 10 rad/s<br />

(Pa)<br />

DSR - δ, 10 rad/s<br />

carbonyl, au 2 carbonyl, au 2 20°C 60°C 20°C 60°C<br />

AZ1-3 Unaged 0.04 0.43<br />

AZ1-3 RTFO only<br />

AZ1-3 100°C 20hrs<br />

0.12 1.12E+07 25076 46.103 75.457<br />

AZ1-3 60°C 96hrs 0.19 1.96E+07 67533 38.522 69.777<br />

AZ1-3 60°C 192hrs 0.22 2.17E+07 77308 37.022 68.479<br />

AZ1-3 60°C 336hrs 0.26 2.73E+07 1.04E+05 34.2 67.606<br />

AZ1-3 60°C 504hrs 0.29 2.71E+07 1.52E+05 33.643 63.44<br />

AZ1-3 80°C 96hrs 0.25 0.64 2.86E+07 90512 34.11 67.19<br />

AZ1-3 80°C 210hrs 0.37 0.85 4.32E+07 2.40E+05 28.27 58.962<br />

AZ1-3 80°C 336hrs 0.47 1.16 5.10E+07 4.84E+05 25.288 51.064<br />

AZ1-3 80°C 504hrs 0.53 1.39 5.85E+07 1.54E+06 21.564 40.257<br />

AZ1-4 Unaged 0.05 0.39<br />

AZ1-4 RTFO only 0.11 1.40E+07 30451 44.626 77.583<br />

AZ1-4 100°C 20hrs<br />

AZ1-4 60°C 96hrs 0.19 2.50E+07 83266 36.66 71.775<br />

AZ1-4 60°C 192hrs 0.23 2.73E+07 92662 35.033 70.424<br />

AZ1-4 60°C 336hrs 0.29 3.49E+07 1.25E+05 32.66 68.536<br />

AZ1-4 60°C 504hrs 0.32 3.78E+07 2.10E+05 30.861 64.494<br />

AZ1-4 80°C 96hrs 0.24 0.79 2.97E+07 86300 34.134 70.153<br />

AZ1-4 80°C 210hrs 0.39 0.88 5.22E+07 3.42E+05 27.01 58.226<br />

AZ1-4 80°C 336hrs 0.49 1.13 5.70E+07 5.44E+05 24.19 53.537<br />

AZ1-4 80°C 504hrs 0.52 1.31 6.60E+07 1.63E+06 20.777 41.164<br />

1 All samples displaying temperatures <strong>and</strong> times are PAV aged.<br />

2 au = absorption units<br />

FTIR-PA Spectra<br />

Since the strong C-H stretching absorption in the region from 2800 to 3100 cm -1 appeared<br />

truncated below a velocity <strong>of</strong> 0.5 cm/sec, asphalt PA spectra were collected at an OPD velocity<br />

<strong>of</strong> 0.5 cm/sec. All PA <strong>and</strong> liquid-cell spectra are available for review; two are shown here<br />

(figures 2-2.2.5 <strong>and</strong> 2-2.2.6) for inspection. Functional groups apparent in the unaged <strong>and</strong> aged<br />

liquid-cell spectra are also apparent in PA spectra as shown in figure 2-2.2.7.


(a)<br />

(b)<br />

PA Intensity (arbitray units)<br />

PA Intensity (arbitray units)<br />

7.42<br />

7.0<br />

6.5<br />

6.0<br />

5.5<br />

5.0<br />

4.5<br />

4.0<br />

3.5<br />

3.0<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.03<br />

4000.0 3600 3200 2800 2400 2000 1800<br />

cm-1<br />

1600 1400 1200 1000 800 600.0<br />

5.00<br />

4.8<br />

4.6<br />

4.4<br />

4.2<br />

4.0<br />

3.8<br />

3.6<br />

3.4<br />

3.2<br />

3.0<br />

2.8<br />

2.6<br />

2.4<br />

2.2<br />

2.0<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.03<br />

1800.0 1700 1600 1500 1400 1300 1200 1100 1000 900 800 700.0<br />

cm-1<br />

Figure 2-2.2.5. PA spectra changes on oxidation <strong>of</strong> a thermally thick (1 mm ±) sample <strong>of</strong> AAD-1<br />

(unaged, RTFO, <strong>and</strong> various PAV aging times at 80˚C). (a) Blue: Unaged. Green: RTFO only.<br />

Black: RTFO/PAV 20hours. Brown: RTFO/PAV 144 hours. Pink: RTFO/PAV 240 hours.<br />

Red: RTFO/PAV 480 hours. OPD velocity 0.5 cm/sec, 512 co-added scans, gain (4).<br />

(b) exp<strong>and</strong>ed view - fingerprint region.<br />

24


(a)<br />

(b)<br />

PA Intensity (arbitray units)<br />

6.64<br />

6.0<br />

5.5<br />

5.0<br />

4.5<br />

4.0<br />

3.5<br />

3.0<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.06<br />

4000.0 3600 3200 2800 2400 2000 1800<br />

cm-1<br />

1600 1400 1200 1000 800 600.0<br />

PA Intensity (arbitray units)<br />

5.00<br />

4.8<br />

4.6<br />

4.4<br />

4.2<br />

4.0<br />

3.8<br />

3.6<br />

3.4<br />

3.2<br />

3.0<br />

2.8<br />

2.6<br />

2.4<br />

2.2<br />

2.0<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.06<br />

1800.0 1700 1600 1500 1400 1300 1200 1100 1000 900 800 700.0<br />

cm-1<br />

Figure 2-2.2.6. PA spectra changes on oxidation <strong>of</strong> a thermally thick (1 mm ±) sample <strong>of</strong> AZ1-4<br />

asphalt (unaged, <strong>and</strong> various PAV aging times at 80˚C). (a) Blue: Unaged. Black: RTFO/PAV 96<br />

hours. Brown: RTFO/PAV 210 hours. Pink: RTFO/PAV 336 hours. Red: RTFO/PAV 504 hours.<br />

OPD velocity 0.5 cm/sec, 512 co-added scans, gain (4). (b) exp<strong>and</strong>ed view - fingerprint region.<br />

25


Figure 2-2.2.7. Comparison <strong>of</strong> PA <strong>and</strong> liquid-cell spectra in the fingerprint region. Top: PA<br />

spectra unaged AAD-1, 512 co-added scans, gain (4), <strong>and</strong> OPD velocity 0.5 cm/sec. Second from<br />

top: PA spectra after RTFO/PAV aging (480 hours at 80˚C), 512 co-added scans, gain (4), OPD<br />

velocity 0.5 cm/sec. Third from top: Liquid-cell transmission spectra AAD-1 after RTFO/PAV<br />

aging (480 hours at 80˚C). Bottom: Liquid-cell transmission spectra unaged AAD-1.<br />

Comparison (PA vs. Liquid Cell)<br />

A comparison <strong>of</strong> carbonyl peak height from liquid cell <strong>and</strong> PA spectra <strong>of</strong> the SHRP asphalts<br />

PAV aged at 80˚C is displayed in figure 2-2.2.8. The correlation (R 2 = 0.81) reveals a<br />

reasonably linear relationship between the PA <strong>and</strong> liquid-cell FTIR techniques. PA signal to<br />

noise ratio is much more sensitive to the number <strong>of</strong> co-added scans than the liquid cell technique.<br />

A significantly higher R 2 could have likely been achieved had a greater number <strong>of</strong> co-added<br />

scans been used to acquire the PA spectra. Time constraints limited the number <strong>of</strong> co-added<br />

scans to 512 for this survey.<br />

26


PA - Carbonyl total peak ht. (1700 cm-1)<br />

1.400<br />

1.300<br />

1.200<br />

1.100<br />

1.000<br />

0.900<br />

0.800<br />

0.700<br />

0.600<br />

0.500<br />

y = 2.42x + 0.53<br />

R 2 = 0.81<br />

0.400<br />

0.000 0.050 0.100 0.150 0.200 0.250 0.300 0.350<br />

Liquid Cell - Carbonyl corr. peak ht. (1700 cm-1)<br />

Figure 2-2.2.8. Comparison SHRP asphalts – Carbonyl peak height at 1700 cm -1 , PA vs. Liquid<br />

cell, without unaged or RTFO asphalts. Note: Liquid cell carbonyl represents the corrected peak<br />

height, <strong>and</strong> PA carbonyl represents the total peak height.<br />

SHRP RD aggregate (limestone, CaCO3)<br />

PA spectra <strong>of</strong> RD (limestone) aggregate at several OPD velocities are presented in figure 2-2.2.9.<br />

Spectral artifacts seem evident at lower OPD velocity, particularly at 0.1 cm/sec, where the<br />

broad absorbance b<strong>and</strong> at 1430 cm -1 appears to have sharp side lobes not corrected by the strong<br />

apodization function selected for this analysis. Also, the broad absorption at 1430 cm -1 appears<br />

to be inverted.<br />

Limestone is a sedimentary rock composed <strong>of</strong> more than 50% carbonate minerals. There are two<br />

common, naturally occurring polymorphs <strong>of</strong> calcium carbonate (CaCO3): aragonite <strong>and</strong> calcite.<br />

Aragonite’s crystal lattice differs from that <strong>of</strong> calcite, resulting in a different crystal shape. The<br />

b<strong>and</strong>s at 713 cm -1 , 876 cm -1 , 1075 cm -1 , <strong>and</strong> 1430 cm -1 , which appear most distinct in the top<br />

spectrum, represent carbonate vibrations.<br />

The b<strong>and</strong>s at about 1800 cm -1 , 2510 cm -1 <strong>and</strong> 2900 cm -1 are overtone <strong>and</strong> combination b<strong>and</strong>s<br />

[Logodi et al. 2001; Vassallo et al. 1992]. The lack <strong>of</strong> a sharp b<strong>and</strong> around 1070–1080 cm -1 in<br />

27


figure 2-2.2.9 suggests that calcite is predominate in the RD aggregate since calcite is infrared<br />

inactive in this area [Logodi et al 2001].<br />

The b<strong>and</strong> at about 2510 cm -1 has been used to quantify amounts <strong>of</strong> residual limestone in lime<br />

[Norton <strong>and</strong> McClell<strong>and</strong> 1996] <strong>and</strong> Logodi et al. [2001] have used the out-<strong>of</strong>-plane bending b<strong>and</strong><br />

at 876 cm -1 to quantify the amount <strong>of</strong> limestone in cement blends. Logodi mentions that the<br />

b<strong>and</strong> at 876 cm -1 has also been used to quantify the amount <strong>of</strong> CaCO3 in CaCO3 /Ca(OH)2<br />

mixtures <strong>and</strong> to study the kinetics <strong>of</strong> the decomposition <strong>of</strong> CaCO3 to CaO. Arnold et al. [2006]<br />

has used the b<strong>and</strong> at 3640 cm -1 to evaluate the presence <strong>of</strong> lime in asphalt pavement.<br />

Figure 2-2.2.9. PA spectra <strong>of</strong> SHRP RD (limestone) aggregate at different OPD velocities. Top:<br />

OPD velocity 2.0 cm/sec, average <strong>of</strong> six spectra (128 co-added scans per spectrum), gain (7).<br />

Second from top: OPD velocity 1.0 cm/sec, average <strong>of</strong> four spectra (256 co-added scans per<br />

spectrum), gain (7). Third from top: OPD velocity 0.5 cm/sec, one spectra (256 co-added<br />

scans), gain (7). Bottom: OPD velocity 0.1 cm/sec, one spectra (64 co-added scans).<br />

28


SHRP RA (granite)<br />

PA spectra <strong>of</strong> RA aggregate at two OPD velocities (1 cm/sec <strong>and</strong> 2 cm/sec) are presented in<br />

figure 2-2.2.10. The major RA elemental oxide is SiO2 (71%), which has an intense peak around<br />

1100 cm -1 [Smith 1996]. The major mineral component in RA aggregate is quartz (56%).<br />

Quartz exhibits an intense broad b<strong>and</strong> around 1100 cm -1 from the SiO2 <strong>and</strong> also shows sharper<br />

b<strong>and</strong>s at 795, 775 <strong>and</strong> 690 cm -1 [Jackson 1998].<br />

Figure 2-2.2.10. PA spectra <strong>of</strong> SHRP RA (granite) aggregate at two OPD velocities. Top: OPD<br />

velocity 1.0 cm/sec, average <strong>of</strong> 4 spectra (256 co-added scans per spectrum), gain (7). Bottom<br />

OPD velocity 2.0 cm/sec, average <strong>of</strong> 5 spectra (128 co-added scans per spectrum), gain (7).<br />

29


For comparison, PA spectra <strong>of</strong> RD, RA, Arizona aggregate, <strong>and</strong> reagent grade calcium carbonate<br />

are displayed in figure 2-2.2.11. Visually the Arizona aggregate appears to be primarily<br />

siliceous, but its exact mineralogy has not been determined. The Arizona aggregate spectrum is<br />

not well defined, but appears to contain SiO2 based on the broad b<strong>and</strong> between 1100 <strong>and</strong> 1200<br />

cm -1 . The relatively intense peak at about 3630 cm -1 is probably OH stretching related to lime<br />

which may have been introduced to the aggregate during construction.<br />

Figure 2-2.2.11. PA spectra – comparison <strong>of</strong> RD, RA, Arizona aggregate <strong>and</strong> reagent grade<br />

calcium carbonate. Top: RD aggregate, OPD velocity 2.0, average <strong>of</strong> 6 spectra (128 co-added<br />

scans per spectrum), gain (7). Second from top: Reagent grade calcium carbonate, 0.5 OPD<br />

velocity (32 co-added scans per spectrum), gain (4). Third from top: RA aggregate, OPD<br />

velocity 1.0, average <strong>of</strong> 4 spectra (256 co-added scans per spectrum), gain (7). Bottom: Arizona<br />

aggregate, OPD velocity 0.5, average <strong>of</strong> 4 spectra (256 co-added scans per spectrum), gain (7).<br />

30


Mastics<br />

Several asphalt mastics were prepared consisting <strong>of</strong> SHRP RD aggregate <strong>and</strong> aged AAD-1<br />

asphalt, <strong>and</strong> SHRP RA aggregate <strong>and</strong> aged AAD-1 asphalt. The mastics were prepared with<br />

varying relative amounts <strong>of</strong> the two components. Details on the preparation can be found in the<br />

experimental section <strong>of</strong> this report. Figures 2-2.2.12 <strong>and</strong> 2-2.2.13 show the infrared spectra <strong>of</strong><br />

the two mastics.<br />

Figure 2-2.2.12. PA spectra RD/AAD-1 asphalt mastic. Top: 20% RD/80% AAD-1. Second<br />

from top: 50% RD/50% AAD-1. Third from top: 80% RD/20% AAD-1.<br />

Bottom: 90% RD/10% AAD-1. Variable number <strong>of</strong> separate scans averaged<br />

(256 co-added scans per spectrum) OPD velocity 1.0 cm/sec, gain (7).<br />

31


Figure 2-2.2.13. PA spectra RA/AAD-1 asphalt mastic. Top: 20% RA/80% AAD-1. Second<br />

from top: 50% RA/50%AAD-1. Third from top: 80%RA/20%AAD-1.<br />

Bottom: 90%RA/10%AAD-1. Variable number <strong>of</strong> separate scans averaged<br />

(256 co-added scans per spectrum), OPD velocity 1.0 cm/sec, gain (7).<br />

Spectral subtraction can be performed to obtain the spectrum <strong>of</strong> the aged AAD-1 asphalt. An<br />

example is illustrated in figure 2-2.2.14. The top spectrum in the figure is 20% RD<br />

aggregate/80% aged AAD-1 asphalt mixture, the middle spectrum is pure RD aggregate, the<br />

bottom spectra is the difference spectrum. The difference spectrum was obtained by using the<br />

limestone b<strong>and</strong> at 1800 cm -1 , which is common to both the mixed spectra <strong>and</strong> the reference<br />

spectra, <strong>and</strong> subjectively adjusting the scaling factor (0.64 in this case) until the b<strong>and</strong> appeared<br />

flat.<br />

32


4000.0 3600 3200 2800 2400 2000 1800 1600 1400 1200 1000 800 600.0<br />

cm-1<br />

Figure 2-2.2.14. Spectral subtraction. Top: PA spectrum <strong>of</strong> the mixture <strong>of</strong> 20% RD<br />

aggregate/80% aged AAD-1 asphalt. Middle: PA spectrum <strong>of</strong> pure RD aggregate,<br />

OPD velocity 1.0 cm/sec, average <strong>of</strong> four spectra (256 co-added scans per spectrum), gain (7).<br />

Bottom: The difference spectrum.<br />

The difference spectrum from figure 2-2.2.14 is compared to the spectrum <strong>of</strong> pure aged AAD-1<br />

asphalt in figure 2-2.2.15. The spectra are similar, but there are differences. Differences can be<br />

attributed to a number <strong>of</strong> factors such as sample position within the acoustic chamber <strong>and</strong> b<strong>and</strong><br />

overlap. Additional FTIR methods to separate the spectral components <strong>of</strong> mixtures are being<br />

investigated such as measuring two component mixture spectra at different relative<br />

concentrations, <strong>and</strong> using the ratio <strong>of</strong> the spectra to obtain subtraction coefficients [Hirschfeld<br />

1976; Koenig et al. 1977].<br />

33


4000.0 3600 3200 2800 2400 2000 1800 1600 1400 1200 1000 800 600.0<br />

cm-1<br />

Figure 2-2.3.15. Comparison <strong>of</strong> pure asphalt PA spectrum to difference spectrum. Top: PA<br />

spectrum <strong>of</strong> pure aged AAD-1 asphalt, OPD velocity 0.5 cm/sec, 512 co-added scans, <strong>and</strong> gain<br />

(4). Bottom: PA difference spectrum aged AAD-1 from figure 2-2.2.14.<br />

Laboratory aged compacted specimens<br />

In this study, cores received from NCSU were evaluated by FTIR to determine the extent <strong>of</strong><br />

short- <strong>and</strong> long-term aging that had occurred in the laboratory. See the “Materials’ section <strong>of</strong><br />

this report for details in regard to their preparation. Dynamic modulus <strong>and</strong> direct tension tests<br />

were performed on the specimens at NCSU before shipping to WRI.<br />

The loose mix, prior to compaction, had been short-term aged in a forced draft oven at 135˚C for<br />

four hours. The long-term aging was performed in a forced draft oven at 85˚C at different times<br />

(2, 4 <strong>and</strong> 8 days). The cores were labeled as follows:<br />

• STA (short term aging) – Loose uncompacted mixture conditioned at 135˚C for four<br />

hours <strong>and</strong> then compacted.<br />

34


• LTA1 (long term aging Level 1): The loose mix aging procedure was the same as STA<br />

except the compacted specimen (75-mm diameter by approximately 120-mm tall) was<br />

conditioned at 85˚C for two days.<br />

• LTA2 (long term aging Level 1): The loose mix aging procedure was the same as STA<br />

except the compacted specimen (75-mm diameter by approximately 120-mm tall) was<br />

conditioned at 85˚C for four days.<br />

• LTA3 (long term aging Level 1): The loose mix aging procedure was the same as STA<br />

except the compacted specimen (75-mm diameter by approximately 120-mm tall) was<br />

conditioned at 85˚C for eight days.<br />

In this analysis, the effects <strong>of</strong> short- <strong>and</strong> long-term laboratory aging were measured by first<br />

splitting open the STA <strong>and</strong> LTA3 cores <strong>and</strong> removing small samples <strong>of</strong> mastic from the edge <strong>and</strong><br />

center <strong>of</strong> the open face <strong>of</strong> the cores as illustrated in figure 2-2.2.16. Asphalt from the small<br />

mastic samples was recovered <strong>and</strong> evaluated by FTIR<br />

Figure 2-2.2.16. Core face (typical) - mastic sample locations (STA <strong>and</strong> LTA3).<br />

The FTIR analysis was performed using PA <strong>and</strong> liquid cell infrared spectroscopy. The PA<br />

results were inconclusive due to insufficient signal to noise ratio in the region <strong>of</strong> interest<br />

(approximately 1680 cm -1 to 1720 cm -1 ). It appeared that the number <strong>of</strong> co-added scans (512 in<br />

this case) at an OPD velocity <strong>of</strong> 0.5 cm/sec was insufficient. Liquid cell spectra <strong>of</strong> the extracted<br />

asphalt is shown in figure 2-2.2.17.<br />

35


(a)<br />

(b)<br />

PA intensity (arbitrary units)<br />

2.00<br />

1.9<br />

1.8<br />

1.7<br />

1.6<br />

1.5<br />

1.4<br />

1.3<br />

1.2<br />

1.1<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.00<br />

2000.0 1900 1800 1700 1600 1500 1400 1300<br />

cm-1<br />

1200 1100 1000 900 800 700 600.0<br />

Figure 2-2.2.17. Liquid-cell transmission spectra. (a) Red: spectrum <strong>of</strong> asphalt recovered from<br />

the edge <strong>of</strong> LTA3. Green: spectrum <strong>of</strong> asphalt recovered from the center <strong>of</strong> LTA3.<br />

Blue: spectrum <strong>of</strong> asphalt recovered from the center <strong>of</strong> STA. (b) exp<strong>and</strong>ed view 1770 cm -1 to<br />

1840cm -1 . Spectra were normalized to the umbrella bending b<strong>and</strong> at 1376 cm -1 .<br />

36


The increase in carbonyl at the center <strong>of</strong> LTA3 compared to the center <strong>of</strong> STA suggests<br />

significant aging occurred during the eight-day oven aging period. The apparent carbonyl<br />

gradient from the center to the edge in LTA3 suggest the oxidation reaction is diffusion<br />

controlled <strong>and</strong> that the core has increased anisotropy due to aging.<br />

Since the cores were prepared using AAD-1 asphalt, the observed differences in carbonyl content<br />

were further quantified by estimating G* as shown in figure 2-2.2.18. The asphalt at the center<br />

<strong>of</strong> STA had a G* <strong>of</strong> 10,600 Pa (at 60˚C <strong>and</strong> 10 radians/sec). The estimated G* at the center <strong>and</strong><br />

edge <strong>of</strong> LTA3 was 17,000 <strong>and</strong> 27,500 Pa, respectively. The complex shear modulus <strong>of</strong> AAD-1<br />

(at 60˚C <strong>and</strong> 10 radians/sec) after st<strong>and</strong>ard PAV aging (20 hours, 300 psi) was 14,700 Pa., which<br />

is somewhat less than the G* at the center <strong>of</strong> LTA3, suggesting the eight-day oven aging effect is<br />

more severe than the st<strong>and</strong>ard PAV test.<br />

Since differences in aging in the presence <strong>of</strong> aggregate at atmospheric pressure compared to<br />

aging without aggregate at high pressure are not well understood, these results must be<br />

considered preliminary. Some aggregates may have a catalytic effect on oxidation that is not<br />

currently accounted for in the st<strong>and</strong>ard PAV test.<br />

Log G* , Pa, 60C, 10 rad/sec<br />

6.00<br />

5.50<br />

5.00<br />

4.50<br />

4.00<br />

3.50<br />

2.75E+04 Pa<br />

1.70E+04 Pa<br />

1.06E+04 Pa<br />

37<br />

y = 6.92x + 1.74<br />

R 2 = 0.99<br />

Linear (PAV 60C<br />

<strong>and</strong> 80C points)<br />

3.00<br />

0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6<br />

Normalized Carbonyl (total peak ht) Absorbance<br />

Unaged<br />

RTFO<br />

PAV 100C<br />

PAV 60C<br />

PAV 80C<br />

STA center<br />

LTA3 center<br />

LTA3 edge<br />

Figure 2-2.2.18. Estimated G* <strong>of</strong> recovered asphalt from the center <strong>of</strong> STA <strong>and</strong> the center <strong>and</strong><br />

edge <strong>of</strong> LTA3. Correlation plot <strong>of</strong> G* versus carbonyl peak height is based on AAD-1 60˚C <strong>and</strong><br />

80˚C PAV’ed asphalt. Note: the infrared spectra were normalized to the umbrella bending b<strong>and</strong><br />

at 1376 cm -1 . Unaged, RTFO <strong>and</strong> PAV 100˚C points were not included in the linear regression.


Arizona FHWA/WRI validation site<br />

One <strong>of</strong> the objectives <strong>of</strong> this study is to validate the accuracy <strong>of</strong> laboratory fundamental material<br />

property tests using field performance data. Table 2-2.2.5 shows field pavement performance<br />

data that was collected in 2006 after 5 years <strong>of</strong> service. The distress survey was performed by<br />

using the Distress Identification Manual for the Long-Term Pavement Performance Program (1).<br />

It can be seen that asphalts from different sources perform quite differently in the field pavement.<br />

Asphalt AZ1-1 has developed the most cracking after 5 years <strong>of</strong> pavement service, based on nonwheel<br />

path longitudinal cracking.<br />

Table 2-2.2.5. Distress measurements <strong>of</strong> four asphalts after five years pavement service.<br />

Asphalt<br />

AZ1-1<br />

AZ1-2<br />

AZ1-3<br />

AZ1-4<br />

2006 Survey = 5 years in Service<br />

Fatigue Cracking,<br />

Longitudinal Cracking, m<br />

m 2 Wheel Path Non-Wheel Path<br />

3.9<br />

0.0<br />

0.0<br />

3.3<br />

0.0<br />

14.5<br />

0.4<br />

28.1<br />

10.0<br />

1.80<br />

0.6<br />

0.0<br />

0.0<br />

0.0<br />

4.4<br />

5.7<br />

38<br />

120.6<br />

275.7<br />

59.8<br />

32.8<br />

6.3<br />

4.5<br />

111.8<br />

22.8<br />

Transverse<br />

Cracking, m<br />

In a rheological context the term “dynamic measurement” usually refers to an experiment in<br />

which either, the stress or the strain varies harmonically with time. Materials under an action <strong>of</strong><br />

shear stress always produce shear strain. The phase angle <strong>and</strong> the amplitude ratio <strong>of</strong> the stress to<br />

the strain depend on the material <strong>and</strong> can be regarded as material properties although, in general,<br />

both will vary with frequency. These two frequency-dependent parameters are required to<br />

determine the strain for a given stress, such as the amplitude <strong>of</strong> the component <strong>of</strong> strain which is<br />

in phase with the stress or the amplitude <strong>of</strong> the component <strong>of</strong> strain out <strong>of</strong> phase with the stress.<br />

Instead <strong>of</strong> investigating the conventional storage modulus (G*cosδ) <strong>and</strong>/or loss modulus<br />

(G*sinδ), the complex modulus <strong>and</strong> phase angle were used together in a unique manner. It is<br />

believed that the phase angle is an important parameter for characterizing the flow properties <strong>of</strong><br />

asphalt. The phase angle indicates the level <strong>of</strong> viscoelasticity that exists in the asphalt. It is<br />

always prudent to have a certain level <strong>of</strong> viscous flow behavior in an oxidized asphalt to provide<br />

for the relaxation <strong>of</strong> stress. In other words, an asphalt exhibiting a higher strain to failure after<br />

aging is more resistant to thermal or fatigue cracking than an asphalt binder with a lower strain to<br />

failure at the same aging condition. It is well recognized that as asphalts undergo oxidative<br />

aging, their flow behavior becomes more complex or non-Newtonian. Since the level <strong>of</strong> age<br />

hardening is <strong>of</strong> primary concern when comparing aged asphalts, it is reasonable to assume that<br />

the lower the phase angle at the same stiffness, the more susceptible an asphalt becomes to<br />

fatigue cracking. With this in mind, the complex modulus (G*) versus phase angle (δ), based on<br />

master curve at a reference temperature <strong>of</strong> 40°C for all four asphalts with respect to different<br />

field aging times from neat asphalt to 4 years old top half inch, are plotted <strong>and</strong> presented in<br />

8.0<br />

4.8<br />

0.0<br />

0.0<br />

0.0<br />

0.0<br />

1.6<br />

23.4


figure 2-2.2.19 to show the correlation <strong>of</strong> rheological properties to field cracking performance. It<br />

can be seen that different asphalts oxidized to the same stiffness have different phase angles.<br />

The higher phase angle at the same oxidation stiffness indicates the asphalt has better flow<br />

properties. Based on figure 2-2.2.19, asphalt AZ1-1 has the lowest (smallest) phase angle at the<br />

same oxidation stiffness, indicating this asphalt will tend to become more brittle earlier than the<br />

other three asphalts when oxidized to the same stiffness. From this figure, one can see that the<br />

ranking (cracking potential) from laboratory rheological data agrees well with field cracking<br />

performance. In other words, asphalt AZ1-1 has developed the most cracking so far, followed by<br />

asphalt AZ1-2, asphalt AZ1-4, <strong>and</strong> the least tendency to crack is asphalt AZ1-3.<br />

Complex Modulus, G*, Pa<br />

1e+7<br />

8e+6<br />

6e+6<br />

4e+6<br />

2e+6<br />

0<br />

Master Curve at 10 rad/s<br />

4th Year<br />

30 40 50 60 70 80<br />

Phase Angle, degree<br />

39<br />

Neat<br />

AZ1-1<br />

AZ1-2<br />

AZ1-3<br />

AZ1-4<br />

Figure 2-2.2.19. Complex modulus versus phase angle with respect to<br />

different aging times for four asphalts.<br />

To further elucidate how phase angle relates to the flow properties <strong>of</strong> asphalt binder, the phase<br />

angle was plotted against with reduced frequency (master curves) (figure 2-2.2.20) <strong>and</strong><br />

temperature (figure 2-2.2.21) for original asphalt, asphalt extracted from 4-year pavement, <strong>and</strong><br />

asphalt subjected to laboratory PAV aging at different aging times. A typical phase angle master<br />

curve plot is shown in figure 2-2.2.20. The phase angle decreases as a sigmoid shape as<br />

frequency increases. Figure 2-2.2.21 shows a similar plot for phase angle versus temperature.<br />

All solid symbols along with solid lines represent all laboratory aging at different aging times.<br />

The circular solid symbol represents the control sample, neat asphalt without any aging. Square,<br />

diamond, inverted triangle, <strong>and</strong> triangle solid symbols represent 96 hours (red), 192 hours<br />

(green), 336 hours (orange), <strong>and</strong> 504 hours (light blue), respectively, in PAV at 60°C. The<br />

hollow symbols along with dash lines all represent 4-year old shoulder pavement at different<br />

slices, top half inch (dark blue), 2 nd half inch (red), 3 rd half inch (dark purple) <strong>and</strong> bottom half<br />

inch (dark cyan). Careful examination shows that the data plotted in figure 2-2.2.21 are


essentially a mirror image <strong>of</strong> the data plotted in figure 2-2.2.20, suggesting that the high<br />

frequency in the dynamic shear test is equivalent to the inverse <strong>of</strong> temperature, <strong>and</strong> that high<br />

temperatures represent low frequencies.<br />

Phase Angle, degree<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0 hr<br />

96 hrs<br />

192 hrs<br />

336 hrs<br />

504 hrs<br />

4 yrs/Top Slice<br />

4 yrs/2nd Slice<br />

4 yrs/3rd Slice<br />

4 yrs/Bottom Slice<br />

0<br />

1e-8 1e-6 1e-4 1e-2 1e+0 1e+2 1e+4 1e+6 1e+8<br />

Reduced Frequency, rad/s.<br />

40<br />

AZ1-1 Lab Aging at 60°C<br />

Ref. Temp. =40C<br />

Figure 2-2.2.20. Phase angle as a function <strong>of</strong> frequency for asphalt AZ1-1<br />

Phase Anlge at 10 rad/s, degree<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

0 hr<br />

96 hrs<br />

192 hrs<br />

336 hrs<br />

504 hrs<br />

4-Year/Top Slice<br />

4-Year/2nd Slice<br />

4-Year/3rd Slice<br />

4-Year/Bottom Slice<br />

10<br />

AZ1-1 Lab PAV Aging at 60°C<br />

0<br />

-60 -40 -20 0 20 40 60 80 100<br />

Temperature, °C<br />

Figure 2-2.2.21. Phase angle as a function <strong>of</strong> temperature for asphalt AZ1-1<br />

before <strong>and</strong> after lab <strong>and</strong> field aging.


The majority <strong>of</strong> the results were presented in a journal article that will be published in the<br />

proceedings <strong>of</strong> the 52 nd annual conference <strong>of</strong> Canadian Technical Asphalt Association in<br />

November 2008 in Saskatoon, Saskatchewan.<br />

Conclusion<br />

The FTIR spectroscopic survey presented here provides a framework for further FTIR analysis<br />

<strong>of</strong> complex pavement mastics including surface samples consisting <strong>of</strong> photo-oxidized asphalt<br />

<strong>and</strong> aggregate with varying mineralogy. The results <strong>of</strong> this analysis suggest asphalt spectral<br />

components can be removed from combined spectra <strong>of</strong> asphalt <strong>and</strong> aggregate. Asphalt PA<br />

spectra compared favorably to liquid-cell spectra. Signal to noise ratios were lower for the PA<br />

technique, however the PA signal to noise ratio can be significantly improved with additional coadded<br />

scans. It remains an open question as to whether FTIR pavement surface spectra will be<br />

predictive <strong>of</strong> the bulk mechanical properties <strong>of</strong> the pavement.<br />

The complex shear modulus <strong>of</strong> AAD-1 (at 60˚C <strong>and</strong> 10 radians/sec) after st<strong>and</strong>ard PAV aging<br />

(20 hours, 300 psi) was significantly less than the complex shear modulus <strong>of</strong> the asphalt<br />

recovered from the edge <strong>and</strong> center <strong>of</strong> the LTA3 core (8 days oven aging at 85˚C).<br />

At the Arizona field validation site, after five years in-service test sections constructed with<br />

asphalts <strong>of</strong> similar performance grade but from different sources show significant differences in<br />

the levels <strong>of</strong> pavement distress. Asphalt AZ1-1 has developed the most cracking based on nonwheel<br />

path longitudinal cracking.<br />

Work to be Conducted Next Quarter<br />

• Investigations next quarter <strong>and</strong> probably several quarters after will include:<br />

o Continue FTIR <strong>and</strong> rheological analysis <strong>of</strong> the validation site asphalts <strong>and</strong><br />

validation site pavement material to investigate photo-oxidation at the surface <strong>and</strong><br />

the extent <strong>of</strong> oxidation below the surface.<br />

o Evaluate the variation in carbonyl moieties between field <strong>and</strong> PAV aged asphalts<br />

due to, for example, aggregate catalytic effects.<br />

o Exp<strong>and</strong> the chemometric calibration data set to improve <strong>and</strong> verify the strength <strong>of</strong><br />

the observed statistical correlations reported last quarter.<br />

o Explore the kinetic differences observed with regard to asphalt source.<br />

o Evaluate solvent defined fractions to assess the contribution <strong>of</strong> these components<br />

to physicochemical changes in asphalt due to aging.<br />

o Apply the recently developed Hirsch model to evaluate the potential for<br />

prediction <strong>of</strong> E* from carbonyl <strong>and</strong> other infrared derived indices.<br />

o Perform a preliminary investigation to evaluate the efficacy <strong>of</strong> infrared<br />

spectroscopy in evaluating steric <strong>and</strong> physical hardening. Infrared analysis may<br />

show differences in amorphous (rapidly quenched) asphalt compared to more<br />

crystalline asphalt (cooled slowly). Infrared spectra may also show subtle<br />

41


differences in the degree <strong>of</strong> hydrogen bonding in undisturbed asphalt at different<br />

temperatures <strong>and</strong> time intervals.<br />

● Construct oxidation master curves for asphalts that were laboratory aged in PAV at<br />

different durations <strong>and</strong> correlate master curve parameters (i.e., shift factor <strong>and</strong> rheological<br />

index) to chemical parameters (i.e., carbonyl content <strong>and</strong> compatibility parameters) <strong>and</strong><br />

further to the field pavement performance.<br />

● The same samples that were laboratory aged in PAV at 60°C will be subjected to the<br />

same PAV aging test at higher aging temperatures to develop aging kinetic master curves<br />

to further assist the revised mechanistic empirical pavement design guide.<br />

Problems <strong>and</strong> Solution to Problem<br />

There were no problems under this task for this quarter.<br />

References<br />

Al-Azri, N. A., S. H. Jung, K. M. Lunsford, A. Ferry, J. A. Bullin, R. R. Davison, <strong>and</strong> C.<br />

J.Glover, 2006, Binder Oxidative Aging in Texas Pavements: Hardening Rates, Hardening<br />

Susceptibilities, <strong>and</strong> the Impact <strong>of</strong> Pavement Depth. Paper accepted for presentation at the 2006<br />

TRB January conference, Washington, D.C., 2006.<br />

Arnold, T. S., M. Rozario-Ranasinghe, <strong>and</strong> J. Youtcheff, 2005, Determination <strong>of</strong> Lime in Hot<br />

Mix Asphalt. Paper submitted to Transportation Research Board, November 2005.<br />

Bell C. A. , J. J. Fellin, <strong>and</strong> A. J. Wieder, 1994, Field Validation <strong>of</strong> Aging Procedures for<br />

Asphalt-Aggregate Mixtures. Journal <strong>of</strong> the Association <strong>of</strong> Asphalt Paving Technologists, 63:<br />

45-80.<br />

Christensen, D. W., <strong>and</strong> R. F. Bonaquist, 2006, Volumetric Requirements for Superpave Mix<br />

Design, NCHRP 567. National Cooperative Highway Research Program.<br />

Coons, R. F., <strong>and</strong> P. H. Wright, 1968, An Investigation <strong>of</strong> the Hardening <strong>of</strong> <strong>Asphalts</strong> Recovered<br />

from Pavements <strong>of</strong> Various Ages. Journal <strong>of</strong> Association <strong>of</strong> Asphalt Paving Technologists, 37:<br />

510-528.<br />

Drapcho, D. L., R. Curbelo, E. Y. Jiang, R. A. Crocombe, <strong>and</strong> W. J. McCarthy, 1997, Digital<br />

Signal Processing for Step-Scan Fourier Transform Infrared Photoacoustic Spectroscopy.<br />

Applied Spectroscopy, 51 (4): 453-460.<br />

Farrar, M. J., P. M. Harnsberger, K. P. Thomas, <strong>and</strong> W. Wiser, 2006, Evaluation <strong>of</strong> Oxidation in<br />

Asphalt Pavement Test Sections after Four Years <strong>of</strong> Service. Proc., International Conference on<br />

Perpetual Pavement, September 2006, Columbus, Ohio.<br />

Heukelom, W., 1973, An Improved Method <strong>of</strong> Characterizing Asphaltic Bitumens with the Aid<br />

<strong>of</strong> their Mechanical <strong>Properties</strong>. Journal <strong>of</strong> Association <strong>of</strong> Asphalt Paving Technologists, 42: 67-<br />

98.<br />

42


Hirschfeld, T., 1976, Computer Resolution <strong>of</strong> Infrared Spectra <strong>of</strong> Unknown Mixtures. Analytical<br />

Chemistry, 48 (4): 721-723.<br />

Houston, W. N., M. W. Mirza, C. E. Zapata, <strong>and</strong> S. Raghavendra. National Cooperative<br />

Highway Research Program. Environmental Effects in Pavement Mix <strong>and</strong> Structural Design<br />

Systems, NCHRP 9-23 Preliminary Draft Final Report Part 1, September 2005.<br />

Jackson, K. D. O., 1998, A Guide to Identifying Common Inorganic Fillers <strong>and</strong> Activators Using<br />

Vibrational Spectroscopy. Accessed from The Internet Journal <strong>of</strong> Vibrational Spectroscopy, 2<br />

(3): http://www.ijvs.com/volume2/edition3/section3.html#jackson<br />

Jones, R. W., <strong>and</strong> J. F. McClell<strong>and</strong>, 1996, Quantitative Depth Pr<strong>of</strong>iling <strong>of</strong> Layered Samples<br />

Using Phase-Modulation FT-IR Photoacoustic Spectroscopy. Applied Spectroscopy, 50 (10):<br />

1258-1263.<br />

Jones, R. W., <strong>and</strong> J. F. McClell<strong>and</strong>, 2002, Quantitative Depth Pr<strong>of</strong>iling Using Saturation-<br />

Equalized Photoacoustic Spectra. Applied Spectroscopy, 56 (4): 409-418.<br />

Koenig, J. L., L. D’Esposito, <strong>and</strong> M. K. Antoon, 1977, The Ratio Method for Analyzing Infrared<br />

Spectra <strong>of</strong> Mixtures. Applied Spectroscopy, 31 (4): 292-295.<br />

Legodi, M. A., D. de Waal, <strong>and</strong> J. H. Potgieter, 2001, Quantitative Determination <strong>of</strong> CaCO3 in<br />

Cement Blends by FT-IR. Applied Spectroscopy, 55 (3): 361-365.<br />

McClell<strong>and</strong>, J. F., R. W. Jones, <strong>and</strong> S. J. Bajic, 2002, FT-IR Photoacoustic Spectroscopy, chapter<br />

from H<strong>and</strong>book <strong>of</strong> Vibrational Spectroscopy, J. M. Chalmers <strong>and</strong> P. R. Griffiths, eds., John<br />

Wiley & Sons, Ltd.<br />

Mirza, M. W., <strong>and</strong> M. W. Witczak. “Development <strong>of</strong> a Global Aging System for Short <strong>and</strong> Long<br />

Term Aging <strong>of</strong> Asphalt Cements. Journal <strong>of</strong> Association <strong>of</strong> Asphalt Paving Technologists, Vol.<br />

64, 1995, pp. 393-431.<br />

National Cooperative Highway Research Program. Development <strong>of</strong> the 2002 Guide for the<br />

Design <strong>of</strong> New <strong>and</strong> Rehabilitated Pavement Structures. NCHRP 1-37A, Final Report, National<br />

Research Council, 2004.<br />

Norton, G. A., <strong>and</strong> J. F. McClell<strong>and</strong>, 1997, Rapid Determination <strong>of</strong> Limestone Using<br />

Photoacoustic Spectroscopy. Minerals Engineering, 10 (2): 237-240.<br />

Robl, T. L., D. Milburn, G. Thomas, J. Groppo, K. O’Hara, <strong>and</strong> A. Haak, 1991, The SHRP<br />

Materials Reference Library Aggregates: Chemical, Mineralogical, <strong>and</strong> Sorption Analyses,<br />

SHRP-A/UIR-91-509, Strategic Highway Research Program, National Research Council,<br />

Washington, DC.<br />

43


Rosencwaig, A., <strong>and</strong> A. Gersho, 1976, Theory <strong>of</strong> the photoacoustic effect with solids. Journal <strong>of</strong><br />

Applied Physics, 47 (1), 64-69.<br />

Smith, B. C., 1996, <strong>Fundamental</strong>s <strong>of</strong> Fourier Transform Infrared Spectroscopy. CRC Press,<br />

Boca Raton, FL.<br />

St<strong>and</strong>ard Practice for Accelerated Aging <strong>of</strong> Asphalt Binder Using a Pressurized Aging Vessel<br />

(PAV), AASHTO Designation: R28-02, America Association <strong>of</strong> State Highway <strong>and</strong><br />

Transportation Officials, Washington, D.C.<br />

Strategic Highway Research Program, 1993, Distress Identification Manual for the Long-Term<br />

Pavement Performance Project, SHRP-P-338, National Research Council, Washington, DC.<br />

Vassallo, A. M., P. A. Cole-Clarke, L. S. K. Pang, <strong>and</strong> A. J. Palmisano, 1992, Infrared Emission<br />

Spectroscopy <strong>of</strong> Coal Minerals <strong>and</strong> their Thermal Transformations. Applied Spectroscopy, 46<br />

(1): 73-78.<br />

44


SUBTASK 2-3. NANOTECHNOLOGY: AFM ANALYSIS OF ASPHALT THIN-FILM<br />

MICROSTRUCTURE PHENOMENOLOGY<br />

Task Manager: T. Pauli<br />

Personal: W. Grimes, J. Miller, J. Beiswenger<br />

Statement <strong>of</strong> Problem<br />

Asphalt pavements are known to fail over time by a combination <strong>of</strong> different mechanisms. The<br />

modes <strong>of</strong> failure <strong>of</strong> asphalt pavements that are commonly sited are embrittlement due primarily<br />

to steric <strong>and</strong> oxidative aging, fatigue cracking due primarily to loading cycles <strong>and</strong> moisture<br />

(traffic), rutting due primarily to densification <strong>and</strong> plastic flow, thermal cracking due primarily to<br />

low temperature embrittlement, <strong>and</strong> formation <strong>of</strong> potholes due primarily to breakdown <strong>of</strong> the<br />

sub-base structure. It is further contended in the pavement community that all <strong>of</strong> the<br />

aforementioned modes <strong>of</strong> failure are some how influenced by environmental conditions like<br />

seasonal temperature swings <strong>and</strong> the presence <strong>of</strong> water. Why do pavements constructed to the<br />

same specifications <strong>and</strong> subjected to similar environmental conditions <strong>and</strong> traffic loading fail at<br />

different rates by different failure modes when different materials (e.g., asphalts <strong>and</strong> aggregates<br />

derived from different sources) are used to construct the pavement?<br />

Approach<br />

In this subtask we have asked the question why pavements constructed with asphalt derived from<br />

different crude sources, if all other variables were to be kept the same, perform differently in<br />

terms <strong>of</strong> moisture compounded fatigue resistance. For example, pavement cracking is <strong>of</strong>ten<br />

observed to form distinct patterns during the lifespan <strong>of</strong> the pavement. In many other fields <strong>of</strong><br />

material science, metallurgy for example, pattern forming cracking has been successfully<br />

correlated to the formation <strong>of</strong> microstructural grain boundaries which originate in these materials<br />

during casting [Cappelli et al. 2008; Bian <strong>and</strong> Taheri 2008]. This same pattern cracking<br />

phenomena can also be applied to paving materials [Robertson et al. 2005, 2006]. The approach<br />

will be to develop quick <strong>and</strong> inexpensive experimental techniques derived from other fields <strong>of</strong><br />

materials nano-science. Results from these tests can then be combined with chemo-mechanical<br />

models <strong>of</strong> asphalt-aggregate composite materials to predict pavement performance.<br />

Goal<br />

The goal <strong>of</strong> this work is to gain a more fundamental underst<strong>and</strong>ing <strong>of</strong> the composition <strong>of</strong> asphalt<br />

concrete paving materials <strong>and</strong> how it relates to pavement performance (specifically fatigue<br />

cracking/self healing compounded by the presence <strong>of</strong> moisture).<br />

Support <strong>of</strong> FHWA Strategic Goals<br />

This work plan supports the following FHWA focus areas. Pavement Design <strong>and</strong> Analysis: This<br />

work will provide a more fundamental underst<strong>and</strong>ing <strong>of</strong> the physico-chemical nature <strong>of</strong> asphaltbinder<br />

<strong>and</strong> chemo-mechanical properties <strong>of</strong> mastics as they relate to pavement performance.<br />

Optimum Pavement Performance: A more fundamental underst<strong>and</strong>ing <strong>of</strong> the physico-chemical<br />

45


nature <strong>of</strong> asphalt-binder will lead to better characterization <strong>and</strong> development <strong>of</strong> modified asphalts<br />

based on better materials selection criteria. Environmental Stewardship: A more fundamental<br />

underst<strong>and</strong>ing <strong>of</strong> the physico-chemical nature <strong>of</strong> asphalt-binder will lead to better<br />

characterization <strong>of</strong> warm mix asphalts <strong>and</strong> recycled asphalt pavement technologies <strong>and</strong> should<br />

ultimately result in refinement <strong>of</strong> the present technologies as well as lend insight into developing<br />

other types <strong>of</strong> “green” technology.<br />

Introduction/Hypotheses<br />

K<strong>and</strong>hal <strong>and</strong> Chakaraborty [1996] have suggested, based on calculations <strong>of</strong> gradation void-space<br />

volumetrics <strong>of</strong> aggregate in common pavement mix designs, that the average film thickness <strong>of</strong><br />

asphalt in a pavement has a dramatic effect upon both the resilient modulus <strong>and</strong> viscosity <strong>of</strong> a<br />

mix. As the average film thickness is decreased from around 15 microns to 3 microns, both the<br />

resilient modulus <strong>and</strong> the viscosity are observed to increase almost exponentially. K<strong>and</strong>hal <strong>and</strong><br />

Chakaraborty [1996] also suggest that the optimum film thickness <strong>of</strong> asphalt in a mix is between<br />

8-10 microns with thinner films producing brittle mixes <strong>and</strong> thicker films being susceptible to<br />

increased rutting. These observations suggest that “molecular ordering” <strong>of</strong> asphalt thin-films in<br />

mastic occurs, <strong>and</strong> that this molecular ordering becomes more pronounced with decreasing the<br />

asphalt film thickness. Thus, the potential for crack initiation in most “composite” materials is<br />

shown to be related to the development <strong>of</strong> grain boundaries. These boundaries are caused by<br />

molecular ordering phenomena corresponding to the development <strong>of</strong> individual material phases<br />

within these types <strong>of</strong> systems [Cappelli et al. 2008; Bian <strong>and</strong> Taheri 2008]. Grain boundaries<br />

have been observed in (1) crystals, the cutting <strong>of</strong> diamonds for example, with fracturing<br />

occurring along lattice plane flaws <strong>of</strong> the crystal, (2) the crack-flaws that form in ice cubes as a<br />

function <strong>of</strong> cooling rate <strong>and</strong> impurities in the water, or (3) the development <strong>of</strong> dislocations in<br />

binary metal alloy solidification processes. Micro structuring <strong>of</strong> chemically distinct phases in<br />

asphalt “binder”, particularly at asphalt-aggregate interfaces, which originate at the micro scale<br />

due to the heterogeneous nature <strong>of</strong> these materials could contribute to the formation <strong>of</strong> such grain<br />

boundaries that are crack initiation points.<br />

Crack initiation <strong>and</strong> propagation in metals <strong>and</strong> crystalline materials generally falls within the<br />

category <strong>of</strong> brittle fracture. However, polymers are more amorphous by nature <strong>and</strong> fracturing<br />

becomes more ductile which renders the system more difficult to model. One may hypothesize<br />

that the large number <strong>of</strong> different petro-organic-type molecules (asphalt) interact with<br />

themselves <strong>and</strong> with the mineral surfaces (aggregate <strong>and</strong> fines) to form distinct chemical phases<br />

that somehow determine the viscoelastic properties <strong>of</strong> the binder in the pavement. These<br />

properties can then be used to determine the tendency <strong>of</strong> a pavement to fracture <strong>and</strong> self-heal,<br />

which is further compounded by variations in temperature as well as moisture.<br />

In this subtask, uniform asphalt films are studied at near-molecular (< 5-nm), nano (1-100-nm),<br />

meso (10’s-100’s-nm) <strong>and</strong> microscopic (100’s-1000’s-nm) scales. Scanning probe microscopy<br />

techniques (i.e., atomic force microscopy) are employed to correlate the kinetic <strong>and</strong><br />

thermodynamic processes <strong>of</strong> phase transformation phenomena <strong>of</strong> different chemical moieties<br />

present in asphalt with the materials’ performance in roadways. <strong>Asphalts</strong>, which differ by crude<br />

source, have historically been defined by chemists by their compositionally unique chemical<br />

classes <strong>of</strong> molecules. This is a simplified approach given that petroleum asphalt is potentially<br />

46


comprised <strong>of</strong> tens <strong>of</strong> thous<strong>and</strong>s <strong>of</strong> different types <strong>of</strong> molecules that otherwise would need to be<br />

characterized if a totally fundamental approach to the problem were to be considered.<br />

It has also been observed in previous experiments [Robertson et al. 2005, 2006] that different<br />

asphalts tend to form self-ordered microstructural “features” to different degrees with different<br />

crude sources. This was demonstrated at the asphalt-air free surface interface when prepared as<br />

solvent spin-cast, thin-film coatings with thicknesses ranging from 3 microns to as thin as 150<br />

nm. Generally speaking, structures with clearer phase boundaries are observed as the thickness<br />

<strong>of</strong> these “ultra-thin” films is decreased below 2 microns [Robertson et al. 2005]. Thicker films<br />

tend to exhibit larger structures, particularly the bumble bee shaped structures that appear to<br />

“grow” to a limiting size in 10’s <strong>of</strong> microns thick films with less distinguishable interface<br />

boundaries. Inevitably, the formation <strong>of</strong> these microstructures in very thin films <strong>of</strong> asphalt leads<br />

directly to the development <strong>of</strong> interfacial grain boundaries between chemically different phases<br />

<strong>of</strong> materials. These interfacial grain boundaries then constitute discontinuities in the film that<br />

could lead to fracture initiation in actual pavement structures whether they were to rapidly<br />

develop under normal paving conditions or gradually development over time.<br />

The current experimental approaches should not be considered a direct representation or<br />

simulation <strong>of</strong> pavement structures. Contrarily, in the present research, asphalts are prepared as<br />

thin-films that are similar in magnitude to the average thicknesses <strong>of</strong> asphalt films estimated for<br />

pavements (e.g., 5-15 microns or 8 to 10-microns on average [K<strong>and</strong>hal <strong>and</strong> Chakaraborty 1996;<br />

K<strong>and</strong>hal et al. 1998]). In most cases they are prepared as ultra-thin-films, representing<br />

“theoretical” slices <strong>of</strong> asphalt very close to <strong>and</strong> in contact with an aggregate interface. These<br />

systems are then investigated in order to determine the kinetics <strong>of</strong> microstructure formation as a<br />

function <strong>of</strong> film-thickness, temperature fluctuation, <strong>and</strong> method <strong>of</strong> film preparation (e.g., solvent<br />

spin-coating techniques). This approach has also been adopted due to the difficulty <strong>of</strong><br />

experimentally observing asphalt-aggregate interfacial interactions. As a result, thin-film<br />

experimentation on ideal systems combined with computation simulations will be needed to<br />

adequately study these types <strong>of</strong> systems in order to make recommendations as to how to prolong<br />

the lifecycle <strong>of</strong> pavements due to fatigue.<br />

Much <strong>of</strong> the motivation behind the work that has been proposed in this subtask stems from the<br />

desire to know why <strong>and</strong> how the microstructures that have been observed in these materials, as<br />

observed by atomic force microscopy [Loeber et al. 1996; Pauli <strong>and</strong> Grimes 2003], could<br />

contribute to pavement performance. Consequently, it is hypothesized that the thermo-kinetic<br />

processes <strong>of</strong> phase transformations (i.e., molecular order-disorder kinetics) are anticipated to<br />

directly influence the rheological properties <strong>of</strong> these materials at the mastic thin-film interfaces<br />

in actual pavement structures. These processes include; wax melting-crystallization, potential<br />

for flocculation <strong>of</strong> asphaltenes, chromatographic interactions <strong>of</strong> polar <strong>and</strong> aromatic molecules<br />

with mineral aggregates/filler surfaces, <strong>and</strong> others that are influenced by fluctuations in<br />

temperature <strong>and</strong> shear rate. An improved underst<strong>and</strong>ing <strong>of</strong> the correlations between these phase<br />

transformations <strong>and</strong> rheological properties will help explain the nature <strong>of</strong> fatigue damage <strong>and</strong><br />

subsequent self-healing phenomena brought on during rest periods. In other words, how does the<br />

asphalt chemical composition lead to the order-disorder transitions that produce micro<br />

structuring at interfacial boundaries <strong>and</strong> thin-film regions? Also, how does this ultimately lead<br />

47


to crack initiation brought on by fatigue, followed by pavement failure, as a function <strong>of</strong><br />

environmental conditions <strong>and</strong> loading cycles?<br />

It has also been hypothesized that the process <strong>of</strong> combining asphalt with aggregate at high<br />

temperatures, placing this composite material in a roadway, <strong>and</strong> then allowing it to cool <strong>and</strong> cure<br />

may be modeled as a solidification process. Solidification processes are generally considered to<br />

be non-equilibrium thermodynamic processes where fluxes in composition <strong>and</strong> thermal gradients<br />

are driven by both mechanical <strong>and</strong> thermodynamic forces (e.g., chemical potential, surface<br />

tension <strong>and</strong> pressure differentials), thereby constituting mechanisms for order-disorder<br />

transitions that result in the formation <strong>of</strong> microstructural grain boundaries. These order-disorder<br />

transitions could also be responsible for cohesive failures when, for example, subjected to<br />

temperature fluctuations, shear <strong>and</strong> compressive loading, <strong>and</strong> exposure to moisture. Thus, a<br />

thorough investigation <strong>of</strong> the “chemical-action” <strong>of</strong> asphalt binder as it exists in pavements is<br />

warranted in order to define the interfacial boundary conditions <strong>and</strong> kinetic driving mechanisms<br />

that describe the formation <strong>and</strong> destruction <strong>of</strong> microstructural phases in asphalt when adopted by<br />

continuum mechanical modeling approaches used to simulate fatigue damage.<br />

Background<br />

In this subtask a somewhat simplified view has been taken to describe the composition <strong>of</strong> asphalt<br />

thin-films in a “pavement structure.” This is particularly true at the mastic level, which may be<br />

viewed as if one were building s<strong>and</strong>castles with asphalt in place <strong>of</strong> water as the binding agent to<br />

adhere <strong>and</strong> maintain structural form to the system. While constructing such a system, many<br />

different forces can be considered that could contribute to the strength <strong>of</strong> the system. These<br />

forces may include the capillarity <strong>and</strong> adhesion between asphalt <strong>and</strong> aggregate fine particles, the<br />

viscoelastic response <strong>of</strong> the asphalt thin-films between aggregate particles to compression <strong>and</strong><br />

shear forces, <strong>and</strong> breakdown <strong>of</strong> both the cohesive strength <strong>of</strong> the asphalt <strong>and</strong> the adhesive bonds<br />

between asphalt <strong>and</strong> aggregate in the presence <strong>of</strong> moisture. For the sake <strong>of</strong> the present<br />

discussion, figure 2-3.1 depicts a volume element slice <strong>of</strong> an 8-μm thick asphalt thin-film<br />

adhering three aggregate fine particles together in a mastic sub-structure <strong>of</strong> a pavement structure.<br />

If the presence <strong>of</strong> the aggregate particles in this system have an influence on the ordering <strong>of</strong> the<br />

asphalt molecules at the interface (e.g., first monolayer, first 200-nm, first 1.0-μm, etc) at what<br />

distance away from this interface would long-range molecular ordering subside? At what<br />

effective distance away from this interface would the asphalt layer have properties similar to a<br />

bulk phase state? In the present scenario, asphalt “molecules” present at the interface would<br />

experience the strongest interactions with the aggregate surface. Four microns out into the film,<br />

approximately halfway between two <strong>of</strong> these particles, the asphalt should exhibit properties more<br />

closely resembling that <strong>of</strong> asphalt in the “bulk” state.<br />

Based on AFM surface imaging it has been observed that thicker thin-films <strong>of</strong> asphalt (>4.0 μm)<br />

do not differ that much in appearance from films that are 10’s <strong>of</strong> microns or even 1.0-mm in<br />

thickness. As these films are decreased to


to the aggregate fine particle interface influence the flow properties <strong>of</strong> the molecules farther<br />

away from the interface <strong>and</strong> the adhesive properties <strong>of</strong> the molecules in direct contact with the<br />

interface?<br />

Figure 2-3.1. Pictorial <strong>of</strong> a volume element slice <strong>of</strong> an 8μm thick film adhering three<br />

Aggregate fine particles together in a mastic sub-structure <strong>of</strong> a pavement structure D > d = 8μm.<br />

49


Work Conducted This Quarter<br />

A Topical Report, anticipated to be delivered early in the next quarter, summarizes findings from<br />

the past two to three years <strong>of</strong> research. This report describes the present state <strong>of</strong> practice for the<br />

investigation <strong>of</strong> the compositional nature <strong>of</strong> asphalts <strong>and</strong> aggregates, as studied by atomic force<br />

microscopy <strong>and</strong> presents modeling approaches for integrating asphalt/aggregate physichemical<br />

properties into continuum-damage models <strong>of</strong> pavement performance. Finally, this report will<br />

suggest how work elements from the ARC work plan; ARC Work Element Subtask M1b-2-<br />

Work <strong>of</strong> Adhesion at Nano-Scale using AFM, ARC Work Element Subtask M2a-2-Work <strong>of</strong><br />

Cohesion Measured at Nano-Scale using AFM, Work Element Subtask F1d-7-Coordinate with<br />

AFM Analysis, <strong>and</strong> Work Element F3a-Asphalt Microstructural Modeling [Western Research<br />

Institute 2008], will be coordinated with this task.<br />

Work Plan Next Quarter<br />

• Research conducted to date in this subtask has been primarily concerned with defining<br />

<strong>and</strong> measuring compositional properties <strong>of</strong> material thin-films which exhibit <strong>and</strong>/or cause<br />

crazing phenomena. This is thought to be a result <strong>of</strong> material discontinuity<br />

(heterogeneity) caused by variations in wax <strong>and</strong> asphaltene content in the film. It is<br />

hypothesized that this heterogeneity may be considered an indicator <strong>of</strong> cracking <strong>and</strong><br />

embrittlement in asphaltic materials. Thus, a natural avenue for future research would<br />

then be to conduct measurements at micron <strong>and</strong> nano-scale that focus more on<br />

physical/rheological properties <strong>of</strong> thin films related to the stiffness <strong>and</strong> embrittlement<br />

propensity, in addition to continuing compositional/chemical characterization <strong>of</strong> these<br />

materials, in order to draw correlations between compositional <strong>and</strong> rheological properties.<br />

• In the next quarter, a revised SARA separation method will be employed to separate<br />

asphalt into saturate, aromatic, polar aromatic, <strong>and</strong> asphaltene fractions. In this<br />

procedure, iso-octane asphaltene/maltene separations will be conducted, <strong>and</strong> the maltenes<br />

from this separation will be further separated employing the modified SARA separation<br />

procedure. Based on the information gained from these studies, a revised method for<br />

ASTM 4124 will be submitted to the ASTM D4 committee for approval.<br />

• A solidification stage is being assembled as part <strong>of</strong> the atomic force microscope by<br />

employing both a heating stage <strong>and</strong> a cooling stage fabricated along with micrometer<br />

positioning devices, figure 2-3.2. This apparatus will be used to impart a thermal<br />

gradient across an asphalt thin-film resulting in a moving solidification front (liquidsolid)<br />

interface that may be monitored in time with AFM. Furthermore, an automated<br />

wetting apparatus will also be assembled to better control the spin casting procedure <strong>and</strong><br />

to observe <strong>and</strong> quantify lubrication dynamics (see figure 2-3.3). Finally, metrology <strong>and</strong><br />

nanoindentation accessories have been added to the existing AFM equipment to enhance<br />

capabilities to include micro/nano-rheological testing <strong>of</strong> material thin-films.<br />

50


a. Sample, b. cold stage, c. hot stage, d. micro positioner, e. micro positioner, ΔV- thermal<br />

controller, ΔT-digital thermoter.<br />

f. digital positioning stage on which the solidification stage sits<br />

Figure 2-3.2. Diagram <strong>of</strong> a proposed solidification stage to be assembled<br />

to work with existing AFM equipment<br />

51


a. robotic arm <strong>and</strong> nano-syringe pump for b. digital camera microscope for imaging<br />

precision dispensing <strong>of</strong> solutions on spinning spreading solution drop glass slides<br />

c. Spin caster<br />

Figure 2-3.3. Diagram <strong>of</strong> a proposed dynamic wetting apparatus to better control, measure <strong>and</strong><br />

quantify wetting <strong>of</strong> asphalt on glass slides.<br />

52


Problems <strong>and</strong> Solutions to Problems<br />

An upgrade to the present AFM equipment, which includes a solidification stage, a dynamic<br />

wetting apparatus <strong>and</strong> metrology <strong>and</strong> nanoindentation capabilities will lead to the capability <strong>of</strong><br />

measuring micro <strong>and</strong> nano-rheological properties <strong>and</strong> flow properties <strong>of</strong> asphalt thin-film<br />

materials, in addition to material composition. It is anticipated that these upgrades will lead to<br />

very rapid methods <strong>of</strong> analysis <strong>of</strong> asphalt binders that predict, individually, compositional <strong>and</strong><br />

rheological properties, as well as study how these properties are intimately related to the binding<br />

action <strong>of</strong> pavement grade asphalts.<br />

References<br />

ASTM D4124-01, 2002, St<strong>and</strong>ard Test Method for Separation <strong>of</strong> Asphalt into Four Fractions.<br />

Annual Book <strong>of</strong> ASTM St<strong>and</strong>ards, Road <strong>and</strong> Paving Materials; Vehicle-Pavement Systems,<br />

Section 4, vol. 04.03. ASTM International, West Conshohocken, PA, 821-829.<br />

Bian, L., <strong>and</strong> F. Taheri, 2008, Fatigue fracture criteria <strong>and</strong> microstructures <strong>of</strong> magnesium alloy<br />

plates. Materials Science <strong>and</strong> Engineering A, 48774–85.<br />

Cappelli, M. D., R. L. Carlson, <strong>and</strong> G. A. Kardomateas, 2008, The transition between small <strong>and</strong><br />

long fatigue crack behavior <strong>and</strong> its relation to microstructure. International Journal <strong>of</strong> Fatigue,<br />

30: 1473–1478.<br />

K<strong>and</strong>hal, P. S., <strong>and</strong> S. Chakaraborty, 1996, Effect <strong>of</strong> Asphalt Film Thickness on Short <strong>and</strong> Long<br />

Term Aging <strong>of</strong> Asphalt Paving Mixtures. NCAT Report No. 96-01.<br />

K<strong>and</strong>hal, P. S., K. Y. Foo, <strong>and</strong> R. B. Mallick, 1998, A Critical Review <strong>of</strong> VMA Requirements in<br />

Superpave. NCAT Report No. 98-1.<br />

Loeber, L., O. Sutton, J. Morel, J.-M. Valleton, <strong>and</strong> G. Muller, 1996, New direct observations <strong>of</strong><br />

asphalts <strong>and</strong> asphalt binders by scanning electron microscopy <strong>and</strong> atomic force microscopy.<br />

Journal <strong>of</strong> Microscopy, 182(1): 32-39.<br />

Pauli, A. T., <strong>and</strong> W. Grimes, 2003, Surface Morphological Stability Modeling <strong>of</strong> SHRP<br />

<strong>Asphalts</strong>. American Chemical Society Division <strong>of</strong> Fuel Chemistry Preprints, 48(1): 19-23.<br />

Robertson, R. E., K. P. Thomas, P. M. Harnsberger, F. P. Miknis, T. F. Turner, J. F. Branthaver,<br />

S-C. Huang, A. T. Pauli, D. A. Netzel, T. M. Bomstad, M. J. Farrar, J. F. McKay, <strong>and</strong> M.<br />

McCann. “<strong>Fundamental</strong> <strong>Properties</strong> <strong>of</strong> <strong>Asphalts</strong> <strong>and</strong> <strong>Modified</strong> <strong>Asphalts</strong> II, Final Report, Volume<br />

I: Interpretive Report,” Federal Highway Administration, Contract No. DTFH61-99C-00022,<br />

Chapters 1-4 submitted for publication, November 2005.<br />

Robertson, R. E., K. P. Thomas, P. M. Harnsberger, F. P. Miknis, T. F. Turner, J. F. Branthaver,<br />

S-C. Huang, A. T. Pauli, D. A. Netzel, T. M. Bomstad, M. J. Farrar, D. Sanchez, J. F. McKay,<br />

<strong>and</strong> M. McCann. “<strong>Fundamental</strong> <strong>Properties</strong> <strong>of</strong> <strong>Asphalts</strong> <strong>and</strong> <strong>Modified</strong> <strong>Asphalts</strong> II, Final Report,<br />

53


Volume I: Interpretive Report,” Federal Highway Administration, Contract No. DTFH61-99C-<br />

00022, Chapters 5-7 submitted for publication, March 2006.<br />

Western Research Institute, 2008, Asphalt Research Consortium Annual Work Plan for Year 2,<br />

April 1, 2008-March 31, 2009. Prepared for Federal Highway Administration, FHWA Contract<br />

No. DTFH61-07-H-00009, January 2008.<br />

54


SUBTASK 2-4. LOW-TEMPERATURE PROPERTIES<br />

Statement <strong>of</strong> Problem<br />

The contributions <strong>of</strong> asphalt source (or chemical composition) <strong>and</strong> environmental conditions to<br />

pavement low-temperature performance are not completely understood. Although the current<br />

purchase specifications seem to prevent early failure, there is little confidence in any prediction<br />

<strong>of</strong> long-term behavior.<br />

Approach<br />

Use differential scanning calorimetry <strong>and</strong> stress relaxation rheometry to determine asphalt<br />

component influences on low-temperature properties. This effort will validate the use <strong>of</strong> lowtemperature<br />

stress relaxation tests for determining asphalt mechanical properties.<br />

Use a similar approach to determine the influence <strong>of</strong> oxidative aging on low-temperature<br />

properties.<br />

Goal<br />

The goal <strong>of</strong> this work is to provide the low-temperature information to support the development<br />

<strong>of</strong> an advanced fundamental underst<strong>and</strong>ing <strong>of</strong> how chemical types, reactions, <strong>and</strong> structures<br />

determine the physical properties <strong>of</strong> asphalts <strong>and</strong> the long-term performance <strong>of</strong> asphalt<br />

pavements.<br />

Support <strong>of</strong> FHWA Strategic Goals<br />

The work conducted in this subtask to better underst<strong>and</strong> low-temperature asphalt behavior<br />

supports the FHWA strategic goal that addresses optimizing pavement performance through the<br />

development <strong>of</strong> longer lasting asphalt pavements, thus decreasing the dem<strong>and</strong> for reconstruction<br />

<strong>of</strong> pavements.<br />

Work Conducted This Quarter<br />

According to the Contract DTFH61-07-D-00005 documents, Year-2 work on this topic must be<br />

shifted to Subtask 3-7. No Task 3 research has been authorized.<br />

55


SUBTASK 2-5. MODIFIED ASPHALTS (continuing on Year-1 funds)<br />

Statement <strong>of</strong> Problem<br />

Asphalt modifiers are <strong>of</strong>ten added to meet binder purchase specifications. Common modifiers<br />

include polymers to improve rutting resistance, lime <strong>and</strong> antistrip additives to mitigate moisture<br />

damage, polyphosphoric acid for rutting resistance, <strong>and</strong> reclaimed asphalt pavement (RAP) for<br />

reducing costs. New modifiers are also being introduced to enable warm-mix asphalt (WMA)<br />

technologies where mix <strong>and</strong> compaction temperatures can be substantially reduced. One <strong>of</strong> these<br />

modifiers is water (typically used to foam the asphalt), which improves the workability <strong>of</strong><br />

binders at mixing <strong>and</strong> laydown temperatures <strong>and</strong> appears to have little effect afterwards,<br />

although there is some evidence that entrapped moisture may increase stripping [Hurley <strong>and</strong><br />

Prowell 2005]. In addition, the long-term effectiveness <strong>of</strong> some modifiers under highway<br />

conditions is not known. Often, the mechanism <strong>of</strong> action <strong>of</strong> a modifier is understood only in an<br />

empirical sense <strong>and</strong> the effective treatment levels <strong>and</strong> economical treatment levels may differ. It<br />

is also important to point out that modifier interactions may reduce effectiveness <strong>and</strong> waste<br />

resources.<br />

Approaches<br />

The sensitivity <strong>of</strong> PPA-modified asphalts to environmental factors is being determined using<br />

laboratory PAV aging tests on modified <strong>and</strong> unmodified asphalts in the presence <strong>and</strong> absence <strong>of</strong><br />

moisture in the oven. Analytical tests applied include spectroscopic (FTIR) <strong>and</strong> rheologic (DSR)<br />

<strong>of</strong> the aged materials. Master curve <strong>and</strong> shift factors are used to quantify changes.<br />

Nuclear magnetic resonance techniques are being applied to study the reactions between<br />

phosphorous-containing additives including antistrips <strong>and</strong> PPA in asphalts.<br />

Goals<br />

The goal <strong>of</strong> this research is to develop a detailed description <strong>of</strong> the actions <strong>of</strong> modifiers in<br />

asphalts while varying environmental conditions. Where feasible, efforts are directed toward<br />

developing relationships for predicting long-term behavior from initial laboratory tests.<br />

Support <strong>of</strong> FHWA Strategic Goals<br />

The work conducted in this subtask supports the FHWA strategic goal that addresses<br />

environmental stewardship <strong>and</strong> safety. State agencies need a better underst<strong>and</strong>ing <strong>of</strong> how <strong>and</strong><br />

when to use additives, such as recycled asphalt pavement (RAP), polymers <strong>and</strong>/or<br />

polyphosphoric acid (PPA), to improve the performance <strong>of</strong> asphalt pavements. Obviously, better<br />

performing <strong>and</strong> longer lasting asphalt pavements, which incorporate the use <strong>of</strong> RAP, lead to the<br />

utilization <strong>of</strong> less asphalt, thus decreasing the dem<strong>and</strong> for reconstruction <strong>of</strong> pavements <strong>and</strong><br />

incidentally safer roads.<br />

57


Background<br />

The addition <strong>of</strong> polyphosphoric acid (PPA) to asphalt binders is known to affect a number <strong>of</strong><br />

chemical <strong>and</strong> physical properties <strong>of</strong> the asphalts. However, the mechanisms by which this<br />

happens are not well understood. Several possible mechanisms have been advanced, <strong>and</strong> it has<br />

been noted that the mechanisms <strong>of</strong> action may be asphalt source related [Baumgardner et al.<br />

2005]. In 2005, WRI initiated an investigation (subtask 14-4 <strong>of</strong> FHWA Contract DTFH61-99C-<br />

00022) to elucidate the mechanisms <strong>of</strong> action <strong>of</strong> polyphosphoric acid (PPA) in asphalt. SHRP<br />

asphalts ABD, AAD-1, <strong>and</strong> AAM-1 were modified with 1.5 mass percent <strong>of</strong> polyphosphoric acid<br />

(105 percent). These modified asphalts were then used to verify observations reported by others.<br />

These observations included: (1) an increase in the high-temperature PG grade, (2) a slow<br />

increase in the complex modulus upon storage in a heated tank, (3) a decrease in chemical aging,<br />

as defined by the amount <strong>of</strong> carbonyl-containing compounds produced, <strong>and</strong> (4) an increase in the<br />

gel character <strong>of</strong> the modified asphalt, i.e. the modified asphalt becomes less compatible. The<br />

work conducted at WRI not only confirmed these observations but also provided additional<br />

insight into the mechanism <strong>of</strong> action <strong>of</strong> PPA. In general, it was concluded that, to a certain<br />

extent, the impact <strong>of</strong> PPA modification on the rheological properties is related to the sol-gel<br />

character <strong>of</strong> the asphalt. While a concentration <strong>of</strong> 1.5 wt % is considered too high for paving<br />

applications, this was considered a reasonable concentration to look for chemical changes in<br />

asphalt brought about by PPA. Others [Falkiewicz <strong>and</strong> Grzybowski 2004] have used up to 2 wt<br />

% PPA to look for changes in asphalt behavior with PPA modification. Nonetheless, our study<br />

indicates that PPA-modified asphalt binders should (1) increase early resistance <strong>of</strong> the pavement<br />

to rutting by increasing initial stiffness <strong>and</strong> (2) extend the useful life <strong>of</strong> the pavement by<br />

improving the low-temperature flow properties. These improved properties should result in both<br />

reduced fatigue cracking <strong>and</strong> reduced low temperature cracking [Huang et al. 2008].<br />

Based on current experiments <strong>and</strong> previous studies, it appears that there is a relationship between<br />

physical properties (rheological) <strong>and</strong> a chemical property (carbonyl content) for unmodified<br />

asphalt binders with respect to long-term oxidative aging. However, addition <strong>of</strong> PPA to asphalts<br />

disturbs the linear relationship between physical <strong>and</strong> chemical properties <strong>of</strong> asphalt binders with<br />

respect to their long-term aging.<br />

Based on NMR spectroscopy <strong>and</strong> modified black plots, addition <strong>of</strong> PPA into asphalt binders does<br />

not appear to significantly change the internal structure <strong>of</strong> asphalt binders by chemical reaction<br />

(covalent). Another observation <strong>of</strong> the WRI work, using 31 P NMR, was the apparent hydrolysis<br />

<strong>of</strong> polyphosphoric acid to phosphoric acid by residual water in asphalt [Miknis <strong>and</strong> Thomas<br />

2008]. This was indicated in 31 P NMR spectra <strong>of</strong> PPA modified asphalts that showed the<br />

disappearance <strong>of</strong> phosphorous resonances in the middle <strong>and</strong> terminal groups <strong>of</strong> phosphate chains<br />

over time, with only a phosphoric acid resonance remaining.<br />

Work Conducted This Quarter<br />

During this quarter, work continued on the rheological analysis <strong>of</strong> the unmodified <strong>and</strong> acidmodified<br />

asphalts that were PAV aged at 60°C in the presence <strong>of</strong> water.<br />

58


Figure 2-5.1 shows the complex modulus <strong>of</strong> PPA modified asphalt ABD after PAV aging in the<br />

presence <strong>of</strong> water as a function <strong>of</strong> frequency <strong>and</strong> a constructed oxidation master curve for the<br />

same asphalt at 30°C. As seen from figure 2-5.1, the rheological properties <strong>of</strong> the asphalt binder<br />

are dependent on aging time. As aging time is increased an increase in the complex modulus is<br />

observed. This phenomenon is similar to the temperature dependency <strong>of</strong> asphalt binder, where<br />

the complex modulus increases as temperature decreases.<br />

Complex Modulus, Pa<br />

1e+9<br />

1e+8<br />

1e+7<br />

1e+6<br />

1e+5<br />

1e+4<br />

1e+3<br />

1e+2<br />

0 hr<br />

96 hrs<br />

186 hrs<br />

260 hrs<br />

326 hrs<br />

Master Curve<br />

ABD/PPA, Moist<br />

1e+1<br />

PAV at 60°C<br />

1e+0<br />

1e-6 1e-4 1e-2 1e+0 1e+2 1e+4 1e+6 1e+8<br />

Reduced Frequency, rad/s<br />

59<br />

30°C Data<br />

Figure 2-5.1. Complex modulus as a function <strong>of</strong> frequency for different aging times along with<br />

the constructed master curve for PPA modified asphalt ABD.<br />

The Christensen-Anderson-Marasteanu (CAM) model used for shifting time-temperature data for<br />

regular master curves was modified <strong>and</strong> then applied to construct the aging master curves<br />

[Christensen <strong>and</strong> Anderson 1992; Ferry 1961; Zeng et al. 2001]. The equation is shown below.<br />

*<br />

Gg<br />

G * ( ω , t)<br />

=<br />

(2-5.1)<br />

ωc<br />

k −1/<br />

k<br />

[ 1+<br />

( ) ] '<br />

ω<br />

Where t = aging time,<br />

G*g = glass complex modulus,<br />

ω c = location parameter <strong>of</strong> frequency,<br />

k = shape parameter, dimensionless,<br />

ω’ = reduced frequency, is defined as aA(t)ω,<br />

aA = aging shift factor, dimensionless.


All different aging times were shifted to zero aging time. The aging shift factor is the amount <strong>of</strong><br />

shift <strong>of</strong> the complex modulus at a given aging time to the reference aging time (zero aging time<br />

in this case) to form a single curve. The values <strong>of</strong> these shift factors can be considered to be<br />

modulus (stiffness) changes with respect to the modulus at the reference aging time, <strong>and</strong> give an<br />

indication <strong>of</strong> how the properties <strong>of</strong> a material change with aging time.<br />

Figure 2-5.1 also shows the constructed complex modulus oxidation master curve for asphalt<br />

ABD mixed with 1.5% PPA with respect to different aging times. A well-defined curve can be<br />

seen in the figure indicating that a similarity does exist between the effects <strong>of</strong> aging <strong>and</strong><br />

temperature. The logarithm <strong>of</strong> the aging shift factor is the amount <strong>of</strong> shift for the complex<br />

modulus at a given aging time to the reference aging time to form a single curve. Figure 2-5.2<br />

shows the shift factor as a function <strong>of</strong> PAV aging time for asphalt ABD <strong>and</strong> PPA-modified ABD<br />

aged in the absence <strong>and</strong> presence <strong>of</strong> water.<br />

As seen from figure 2-5.2, the curve <strong>of</strong> the aging shift factor is similar to that <strong>of</strong> viscosity-aging<br />

kinetic curves for typical asphalts <strong>and</strong> is shear rate independent. On oxidation, asphalts show an<br />

initial rapid rate <strong>of</strong> viscosity increase followed by a slower rate <strong>of</strong> viscosity increase. This<br />

behavior has been attributed to the formation <strong>of</strong> sulfoxide <strong>and</strong> carbonyl products in the oxidation<br />

process <strong>of</strong> asphalt binder when exposed to atmospheric oxygen in the pavement. As seen from<br />

figure 2-5.2, addition <strong>of</strong> PPA to asphalts increases the amount <strong>of</strong> stiffness change with respect to<br />

the stiffness at the reference aging time, indicating that addition <strong>of</strong> PPA has changed the aging<br />

characteristics <strong>of</strong> asphalt binders. In addition, aging in the presence <strong>of</strong> water reduces the amount<br />

<strong>of</strong> stiffness change with respect to the stiffness at the reference aging time for neat unmodified<br />

asphalts. However, addition <strong>of</strong> water in the pressure aging oven increases the amount <strong>of</strong><br />

stiffness change for PPA modified asphalt binders.<br />

Figures 2-5.3 <strong>and</strong> 2-5.4 show similar types <strong>of</strong> plots, aging shift factors versus PAV aging times<br />

in the presence <strong>and</strong> absence <strong>of</strong> water, for asphalts AAD-1, AAM-1, <strong>and</strong> their PPA mixtures,<br />

respectively. Again, it can be seen that addition <strong>of</strong> PPA to asphalts increases the amount <strong>of</strong><br />

stiffness change due to PAV aging. As seen from both figures 2-5.3 <strong>and</strong> 2-5.4, the aging shift<br />

factors <strong>of</strong> PPA modified asphalts after PAV aging for 326 hours in the presence <strong>of</strong> water are<br />

similar to those <strong>of</strong> the same PPA modified asphalts at the same aging time in the absence <strong>of</strong><br />

water. This suggests that PPA modified asphalts AAD-1 <strong>and</strong> AAM-1 aged at 326 hours in the<br />

PAV in the presence <strong>of</strong> water have similar complex modulus to the same materials that were<br />

subjected to the same aging time in the absence <strong>of</strong> water. However, it is uncertain if the reduced<br />

complex modulus at the end <strong>of</strong> aging time <strong>of</strong> 326 hours in the presence <strong>of</strong> water is due to the<br />

presence <strong>of</strong> moisture during PAV aging or experiment error. Nonetheless, more experiments<br />

need to be conducted to verify this observation.<br />

During the quarter, work also continued on the study <strong>of</strong> the behavior <strong>of</strong> phosphorous containing<br />

additives in asphalt by using NMR techniques. In particular, NMR work continued on<br />

characterizing the behavior <strong>of</strong> phosphate ester <strong>and</strong> amine antistrip agents in asphalt. In addition,<br />

some exploratory measurements were made to determine the feasibility <strong>of</strong> using 31 P NMR to<br />

determine moisture in asphalt.<br />

60


Log a A<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

ABD, Dry<br />

ABD, Moist<br />

ABD/PPA, Dry<br />

ABD/PPA, Moist<br />

PAV at 60°C<br />

0.0<br />

0 100 200 300 400<br />

Time in PAV, hrs<br />

61<br />

58°C Data<br />

Figure 2-5.2. Aging shift factor as a function <strong>of</strong> PAV aging times for asphalt ABD <strong>and</strong> its PPA<br />

mixtures before <strong>and</strong> after PAV aging in the presence <strong>and</strong> absence <strong>of</strong> water.<br />

Log a A<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

AAD-1, Dry<br />

AAD-1, Moist<br />

AAD-1/PPA, Dry<br />

AAD-1/PPA, Moist<br />

PAV at 60°C<br />

0.0<br />

0 100 200 300 400<br />

Time in PAV, hrs<br />

58°C Data<br />

Figure 2-5.3. Aging shift factor as a function <strong>of</strong> PAV aging times for asphalt AAD-1 <strong>and</strong> its<br />

PPA mixtures before <strong>and</strong> after PAV aging in the presence <strong>and</strong> absence <strong>of</strong> water.


Log a A<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

AAM-1, Dry<br />

AAM-1, Moist<br />

AAM-1/PPA, Dry<br />

AAM-1/PPA, Moist<br />

PAV at 60°C<br />

0.0<br />

0 100 200 300 400<br />

Time in PAV, hrs<br />

62<br />

58°C Data<br />

Figure 2-5.4. Aging shift factor as a function <strong>of</strong> PAV aging times for asphalt AAM-1 <strong>and</strong> its<br />

PPA mixtures before <strong>and</strong> after PAV aging in the presence <strong>and</strong> absence <strong>of</strong> water.<br />

1H NMR <strong>of</strong> Phosphate Ester Antistrips<br />

Additional NMR measurements were made on the phosphate ester antistrip agents that were<br />

acquired last quarter. The phosphate ester antistrip agents were Innovalt W <strong>and</strong> KAO Gripper.<br />

Innovalt W is a phosphated 2-ethyl hexanol antistrip (molecular formula: C8H19O4P, <strong>and</strong><br />

C16H35O4P2), with ingredients: phosphated 2-ethyl hexanol (C8H18O) (>82%, 2-ethyl hexanol<br />


20<br />

these resonances were incorrectly assigned, <strong>and</strong> now have been tentatively identified as due to<br />

the protons in phosphoric acid. Phosphoric acid is used in preparation <strong>of</strong> the phosphate esters<br />

<strong>and</strong> so these resonances would not be unexpected. When D2O was added, these resonances<br />

disappeared due to rapid exchange <strong>of</strong> the OH proton with the deuterium in heavy water (D2O),<br />

i.e.,<br />

P-O-H + D2O P-O-D + DOH<br />

New resonances appeared at 8.1 <strong>and</strong> 5.2 ppm. The resonance at 8.1 ppm was not identified; the<br />

one at 5.2 ppm has been tentatively identified as water (HOD).<br />

Innovalt W in H2O<br />

Innovalt W in D2O<br />

15<br />

10<br />

Figure 2-5.5. 1 HNMR spectra <strong>of</strong> Innovalt W <strong>and</strong> Kao Gripper in D2O <strong>and</strong> in H2O.<br />

31 P NMR <strong>of</strong> Phosphate Ester Antistrips<br />

5<br />

0<br />

-5<br />

-10<br />

20<br />

In a previous quarterly report (Oct.-Dec., 2007) 31 P NMR measurements were made on binder<br />

from four cells taken from the MN Road validation site. Three <strong>of</strong> the four cells used PPA<br />

(115%) <strong>and</strong> a phosphate ester antistrip agent, Innovalt W. The 31 P NMR spectra did not show<br />

the presence <strong>of</strong> the phosphate ester antistrip in the binder.<br />

However, questions have been raised about whether PPA might react with the phosphate ester<br />

antistrip when added to asphalt. Consequently, a NMR investigation has begun to look into this<br />

matter. Initially, 31 P NMR spectra <strong>of</strong> the neat antistrips, <strong>and</strong> antistrips with PPA, but in the<br />

absence <strong>of</strong> asphalt were acquired (figure 2-5.6). The spectra <strong>of</strong> the neat antistrips show a main<br />

peak at ~0.7 ppm <strong>and</strong> a minor one at ~0.2 ppm. The minor resonance is better resolved in the<br />

63<br />

KAO Gripper in H2O<br />

KAO Gripper in D2O<br />

Innovalt W neat KAO Gripper neat<br />

Proton Chemical Shift, ppm<br />

15<br />

10<br />

Proton Chemical Shift, ppm<br />

5<br />

0<br />

-5<br />

-10


KAO Gripper. The minor resonance was not assigned, but is assumed to be due to a dimer <strong>of</strong> the<br />

2-ethyl hexanol phosphate formed during preparation <strong>of</strong> the antistrip. The spectra <strong>of</strong> the Innovalt<br />

W <strong>and</strong> KAO gripper show partial hydrolysis <strong>of</strong> the PPA caused by residual water which is<br />

present in the antistrip agents. This is shown by the increase <strong>of</strong> the phosphoric acid resonance at<br />

~1.7 ppm relative to the phosphate ester resonance at ~0.7 pm with time <strong>of</strong> setting.<br />

There do not appear to be any reactions between the Innovalt W <strong>and</strong> KAO Gripper phosphate<br />

ester antistrip agents <strong>and</strong> PPA. Presumably, there would not be any such reactions when the<br />

antistrip <strong>and</strong> PPA were added to asphalt. To test this assumption, Innovalt W <strong>and</strong> KAO Gripper<br />

antistrips were added to asphalt AAM followed by PPA (115%) addition (figure 2-5.7). Both the<br />

antistrips <strong>and</strong> PPA were added at a level <strong>of</strong> 1 wt %. The purpose <strong>of</strong> these experiments was<br />

tw<strong>of</strong>old: (1) to determine if 31 P NMR can detect signals at the 1% level in a reasonable amount<br />

<strong>of</strong> time, <strong>and</strong> (2) to determine if there might possibly be reactions between the PPA <strong>and</strong> antistrip<br />

in asphalt.<br />

In the case <strong>of</strong> the phosphate ester antistrip agents, there is a single resonance at a chemical shift<br />

(~1.7 ppm) near that <strong>of</strong> phosphoric acid (~1.4.ppm). In both cases spectra with reasonable<br />

signal-to-noise, S/N, ratios were obtained after 3 h <strong>of</strong> signal averaging. When added to asphalt<br />

in the presence <strong>of</strong> PPA, the ester antistrip resonance could not be resolved from that <strong>of</strong><br />

phosphoric acid at the resolution <strong>of</strong> the WRI magnet. Under high resolution conditions, it is<br />

expected that these resonances could be resolved.<br />

Innovalt W in<br />

PPA (115%)<br />

At 99 h<br />

Innovalt W in<br />

PPA (115%)<br />

at 0 h<br />

Innovalt W neat<br />

30 25 20 15 10 5 0 -5 -10 -15 -20 -25 -30 -35 -40 -45 -50 30 25 20 15 10 5 0 -5 -10 -15 -20 -25 -30 -35 -40 -45 -50<br />

Phosphorous Chemical Shift, ppm Phosphorous Chemical Shift, ppm<br />

Figure 2-5.6. 31 P NMR spectra <strong>of</strong> phosphate ester antistrips, neat <strong>and</strong> with PPA (115%)<br />

at different times.<br />

64<br />

KAO Gripper in<br />

PPA (115%)<br />

At 360 h<br />

KAO Gripper<br />

In PPA (115%)<br />

At 0 h<br />

KAO Gripper neat


100<br />

1% Innovalt W in AAM-1<br />

1% PPA(115%) <strong>and</strong><br />

1% Innovalt W in AAM-1<br />

1% PPA(115%) in AAM-1<br />

75<br />

50<br />

25<br />

Figure 2-5.7. 31 P NMR spectra <strong>of</strong> Innovalt W <strong>and</strong> KAO Gripper antistrip agents in<br />

PPA modified asphalt AAM-1.<br />

31P NMR Moisture determination<br />

0<br />

Some exploratory work was performed to determine if 31 P NMR could be used to determine the<br />

residual moisture content <strong>of</strong> asphalt. In part, this work was initiated because previous work had<br />

shown that PPA reacts with residual water in asphalt to form orthophosphoric acid. However,<br />

the extent <strong>of</strong> hydrolysis cannot be predicted apriori because the amount <strong>of</strong> residual water in<br />

asphalts is generally not known. A simple method for determining moisture may be useful in<br />

providing more quantitative data about the reaction.<br />

The 31 P NMR method is based on the use <strong>of</strong> phosphorous compounds that react with water to<br />

form anhydrides. For the moisture determinations, diphenylphosphinic chloride (Ph2POCl), here<br />

referred to as DPPC, was used as the derivatizing agent [Dadey et al. 1988; Wroblewski et al.<br />

1991a,b]. DPPC reacts with water in the presence <strong>of</strong> base (pyridine) to form the anhydride,<br />

Ph2P(O)OP(O)Ph2 ( 1,1,3,3-tetraphenyldiphosphoxane 1,3-dioxide) according to Scheme I,<br />

2Ph2POCl + H2O + Ph2P(O)OP(O)Ph2 + 2<br />

N<br />

-25<br />

-50<br />

Scheme I<br />

-75<br />

-100<br />

65<br />

100<br />

KAO Gripper in AAM-1<br />

1% PPA(115%) <strong>and</strong><br />

1% KAO Gripper in AAM-1<br />

1% PPA(115%) in AAM-1<br />

Phosphorous Chemical Shift, ppm Phosphorous Chemical Shift, ppm<br />

75<br />

50<br />

25<br />

N +<br />

H Cl -<br />

0<br />

-25<br />

-50<br />

-75<br />

-100


Dadey et al. [1988] proposed Scheme II for the reaction <strong>of</strong> DPPC with water to form<br />

diphenylphosphinic acid. We are not sure which scheme more correctly describes the reaction.<br />

However, more recent work seems to agree with Scheme I [Hatzakis et al. 2008].<br />

Other phosphorous compounds can be used to derivatize the labile hydrogen (OH) such as in<br />

phenols <strong>and</strong> carboxylic acids [Wroblewski et al. 1988; Dadey et al. 1988; Lensink <strong>and</strong> Verkade<br />

1990]. Dadey et al. [1988] also proposed Scheme <strong>III</strong> for the reaction <strong>of</strong> DPPC with phenols. We<br />

have not yet verified the reaction with phenols.<br />

2Ph2POCl + H2O + Ph2P(O)OH +<br />

OH<br />

2Ph2POCl + +<br />

N<br />

Scheme II<br />

N<br />

Scheme <strong>III</strong><br />

66<br />

Ph2P(O)O +<br />

In practice, weighed amounts <strong>of</strong> an internal st<strong>and</strong>ard, [MePh3P]I (methyltriphenylphosphoniumiodide),<br />

were placed into oven dried NMR tubes. These tubes were then sealed with rubber septa<br />

to prevent moisture intrusion. A “stock solution” was then made by dissolving Ph2POCl in<br />

CDCl3. An exact amount <strong>of</strong> stock solution was then added to the sealed tubes via syringe. NMR<br />

data for the stock solution in the sealed tubes was collected to determine how much water was<br />

present in the system. Weighed samples <strong>of</strong> asphalt were then dissolved in pyridine, added to the<br />

NMR tubes described above, <strong>and</strong> allowed to react for five minutes at which point NMR spectra<br />

were then acquired. A blank <strong>of</strong> neat pyridine was tested for water content in the same way the<br />

asphalt samples were tested. All solvents were dried over 4 Å sieves.<br />

The 31 P NMR method was applied to asphalts AAB-1, AAD-1, <strong>and</strong> ABD-1. Previously, residual<br />

moisture in these asphalts was determined by a commercial laboratory using the Karl Fischer<br />

titration method. 31 P NMR spectra are shown in figure 2-5.8 along with a spectrum <strong>of</strong> the blank.<br />

The spectra clearly show resonances formed by the reaction <strong>of</strong> the tagging agent with the water<br />

in asphalt. Although the peak in the blank spectrum <strong>of</strong> the water reaction product was<br />

subtracted, the NMR results were not in close agreement with those obtained by Karl Fischer<br />

titration (0.46% NMR vs. 0.37% Karl Fischer for ABD; 0.88% NMR vs. 0.37 Karl Fischer for<br />

AAD; 1.28% NMR vs. 0.31% Karl Fischer for AAB). The reasons for the discrepancies are<br />

being investigated.<br />

N +<br />

H<br />

Cl -<br />

N +<br />

H Cl-<br />

L -


75<br />

Pyridine Blank<br />

ABD-1<br />

AAD-1<br />

AAB-1<br />

Tagging Reagent<br />

Figure 2-5.8. 31 P NMR spectra <strong>of</strong> three asphalts illustrating the use <strong>of</strong> a phosphorous tagging<br />

agent for moisture determination.<br />

Work to be Conducted Next Quarter<br />

50<br />

Water Reaction<br />

Product<br />

● Continue the study to elucidate the mechanism <strong>of</strong> action <strong>of</strong> PPA in asphalt. The use <strong>of</strong><br />

antistrip agents in conjunction with PPA modification <strong>of</strong> asphalts will be investigated.<br />

PPA will be added to asphalt containing amine <strong>and</strong> phosphate ester antistrip agents to<br />

determine if any reactions occur that might negate the use <strong>of</strong> PPA in combination with<br />

the antistrips.<br />

● NMR measurements ( 31 P, 13 C, <strong>and</strong> 1 H) will be made on asphalts containing antistrip<br />

agents to determine the feasibility <strong>of</strong> using NMR to analyze for the antistrips in asphalt at<br />

concentrations used in paving applications.<br />

● Continue the study <strong>of</strong> multiple modifiers in asphalt. Changes in the rheological <strong>and</strong><br />

spectroscopic (FTIR <strong>and</strong> 31 P <strong>and</strong> 13 C NMR), properties on PAV aged samples will be<br />

monitored.<br />

● Prepare a manuscript on this topic <strong>and</strong> submit it to an international conference or journal<br />

for consideration for presentation <strong>and</strong>/or publication.<br />

67<br />

25<br />

Phosphorous Chemical Shift, ppm<br />

Internal St<strong>and</strong>ard<br />

0


Problems <strong>and</strong> Solution to Problem<br />

No problems were encountered during this quarter.<br />

References<br />

Baumgardner, G. L., J-F. Masson, J. R. Hardee, A. M. Menapace, <strong>and</strong> A. G. Williams, 2005,<br />

Polyphosphoric Acid <strong>Modified</strong> Asphalt: Proposed Mechanisms, Proceeding <strong>of</strong> the Association<br />

<strong>of</strong> Asphalt Paving Technologists, 74: 283-305.<br />

Christensen , D.W., <strong>and</strong> D. A. Anderson, 1992, Interpretation <strong>of</strong> Dynamic Mechanical Test Data<br />

for Paving Grade Asphalt. Journal <strong>of</strong> the Association <strong>of</strong> Asphalt Paving Technologists, 61: 67-<br />

116,<br />

Dadey, E. J., S. L. Smith, <strong>and</strong> B. H. Davis, 1988, Determination <strong>of</strong> Hydroxyl Group<br />

Concentration in Coal Liquids by 31 P NMR. Energy <strong>and</strong> Fuels, 2: 326-332.<br />

Falkiewicz, M., <strong>and</strong> K. Grzybowski, 2004, “Polyphosphoric Acid in Asphalt Modification.”<br />

Presented at the Pavement Performance Prediction Symposium, Cheyenne, Wyoming, June 23-<br />

25, 2004.<br />

Ferry, J. D., 1961, Viscoelastic <strong>Properties</strong> <strong>of</strong> Polymers. John Wiley & Sons, New York.<br />

Hatzakis, E., <strong>and</strong> P. Dais, 2008, Determination <strong>of</strong> Water Content in Olive Oil by 31 P NMR<br />

Spectroscopy. J. Agric. Food Chem., 56: 1866-1872.<br />

Huang, S-C., T. F. Turner, F. P. Miknis, <strong>and</strong> K. P. Thomas, 2008, Long-Term Aging<br />

Characteristics <strong>of</strong> Polyphosphoric Acid <strong>Modified</strong> <strong>Asphalts</strong>. Journal <strong>of</strong> Transportation Research<br />

Record (in press).<br />

Hurley, G. C., <strong>and</strong> B. D. Prowell, 2005, Evaluation <strong>of</strong> Aspha-Min® Zeolite for Use in Warm<br />

Mix Asphalt. NCAT Report 05-04, National Center for Asphalt Technology, Auburn University,<br />

June 2005, 33 pp.<br />

Innophos: http://www.innophos.com/msdslib.asp?sort=3&selCountry=us<br />

Lensink, C., <strong>and</strong> J. G. Verkade, 1990, 31 P NMR Spectroscopic Analysis <strong>of</strong> Labile Hydrogen<br />

Functional Groups: Identification with a Dithiaphospholane Reagent. Energy <strong>and</strong> Fuels, 4: 197-<br />

201.<br />

Miknis, F. P., <strong>and</strong> K. P. Thomas, 2008, NMR analysis <strong>of</strong> polyphosphoric acid-modified<br />

bitumens. Road Materials <strong>and</strong> Pavement Design, 9: 59-72.<br />

Wroblewski, A. E., C. Lensink, R. Markuszewski, <strong>and</strong> J. G. Verkade, 1988, 31 P NMR<br />

Spectroscopic Analysis <strong>of</strong> Coal Pyrolysis Condensates <strong>and</strong> Extracts for Heteroatom<br />

Functionalities Possessing Labile Hydrogen. Energy <strong>and</strong> Fuels, 2: 765-774.<br />

68


Wroblewski, A. E., C. Lensink, <strong>and</strong> J. Verkade, 1991a, 31 P NMR Spectroscopy for Labile<br />

Hydrogen Group Analysis: Toward Quantification <strong>of</strong> Phenols in a Coal Condensate. Energy <strong>and</strong><br />

Fuels, 5: 491-496.<br />

Wroblewski, A., K. Reinartz, <strong>and</strong> J. G .Verkade, 1991b, Moisture Determination <strong>of</strong> Argonne<br />

Premium Coal Extracts by 31 P NMR Spectroscopy. Energy <strong>and</strong> Fuels, 5: 786-791.<br />

Zeng, M., B. Hussain, Z. Huachun, <strong>and</strong> P. Turner, 2001, Rheological Modeling <strong>of</strong> <strong>Modified</strong><br />

Asphalt Binders <strong>and</strong> Mixtures. Journal <strong>of</strong> the Association <strong>of</strong> Asphalt Paving Technologists, 70:<br />

403-441.<br />

69


SUBTASK 2-6: VALIDATION SITE MONITORING<br />

Statement <strong>of</strong> Problem<br />

Laboratory tests <strong>and</strong>/or models to predict the performance <strong>of</strong> asphalt pavement materials have<br />

always needed a connection between the original materials <strong>and</strong> actual performance in the field,<br />

especially comparative performance, in order to calibrate or validate the test or model results.<br />

Approach<br />

The approach being used is to construct field sections where different asphalt sources <strong>of</strong> the<br />

same Performance Grade are used at the same location so that comparative pavement<br />

performance can be directly evaluated <strong>and</strong> core samples can be obtained as the comparative<br />

pavements age in service. Original materials were collected at the time <strong>of</strong> construction <strong>and</strong> are<br />

available for testing. The sites being monitored, using LTPP monitoring protocols, were<br />

constructed during the previous FHWA contract, “<strong>Fundamental</strong> <strong>Properties</strong> <strong>of</strong> <strong>Asphalts</strong> <strong>and</strong><br />

<strong>Modified</strong> <strong>Asphalts</strong> II”.<br />

Goal<br />

The goal <strong>of</strong> this subtask is to provide accurate <strong>and</strong> documented performance <strong>of</strong> different asphalt<br />

sources that also includes asphalt-aggregate interactions, to be used to determine the chemical<br />

<strong>and</strong> physical property differences that are important to pavement performance.<br />

Support <strong>of</strong> FHWA Strategic Goals<br />

This subtask supports the FHWA Strategic Goal <strong>of</strong> Optimizing Pavement Performance.<br />

Work Conducted this Quarter<br />

The annual monitoring <strong>of</strong> the Kansas comparative pavement validation site was conducted in<br />

May 2008. The site is still performing rather well after six years in service; however, there is<br />

some distress being noted. There is a small amount <strong>of</strong> fatigue cracking <strong>and</strong> longitudinal cracking<br />

(both non-wheelpath <strong>and</strong> in the wheelpath) that is occurring in two <strong>of</strong> the sections. Two core<br />

samples were obtained by KDOT personnel to investigate one <strong>of</strong> the fatigue cracks. The core<br />

samples revealed that the cracking is top-down cracking <strong>and</strong> only goes through the 40 mm top<br />

lift. The four comparative asphalt sources are located in the bottom 60 mm <strong>of</strong> the pavement with<br />

a 60 mm binder course <strong>and</strong> a 40 mm surface course placed above. There were no apparent<br />

indications in the two core samples that there was any problem with the bottom lift.<br />

Work to be Conducted Next Quarter<br />

It is planned to perform the annual monitoring <strong>of</strong> the Wyoming Highway 216 sections <strong>and</strong> the<br />

Nevada I-15 sections in August <strong>and</strong> September 2008, respectively.<br />

71


TASK 3. OTHER RESEARCH ACTIVITIES (IDIQ)<br />

During this quarter the White Papers that had been prepared in the previous quarter were<br />

approved by FHWA. A task order contract for mechanical testing <strong>of</strong> validation site cores is<br />

awaiting final signatures <strong>and</strong> will be the first Task 3 work element approved.<br />

73


TASK 4. INFORMATION DEPLOYMENT<br />

SUBTASK 4-1. PUBLICATIONS, PRESENTATIONS, NEWSLETTERS, FLYERS AND<br />

BROCHURES<br />

Presentations<br />

The University <strong>of</strong> Oklahoma (O.U.) <strong>and</strong> the Oklahoma Department <strong>of</strong> Transportation (ODOT).<br />

Shin-Che Huang <strong>and</strong> Ray Robertson gave seminars to O.U. <strong>and</strong> ODOT on the FHWA research<br />

being conducted at WRI. These were on April 9 (O.U.) <strong>and</strong> April 10 (ODOT), 2008. Shin-Che<br />

<strong>and</strong> Ray were also asked to critique several ongoing <strong>and</strong> planned O.U. pavement performance<br />

research projects. ODOT requested information on the toluene-ethanol asphalt extraction<br />

method developed during SHRP <strong>and</strong> on the <strong>Modified</strong> German Rolling Flask aging method<br />

developed by WRI <strong>and</strong> AAT. This information was sent to ODOT upon return to WRI.<br />

Conferences <strong>and</strong> Meetings Attended<br />

The Association <strong>of</strong> Asphalt Paving Technologists Annual Meeting, Philadelphia, April 27-30,<br />

2008. Shin-Che Huang attended the meeting.<br />

Training<br />

Rheobit formal course in rheology. During the week <strong>of</strong> June 2-6, 2008, WRI employed Drs.<br />

David Anderson <strong>and</strong> Ge<strong>of</strong>f Rowe to teach the Rheobit rheology course at WRI. This was open<br />

to all WRI Transportation Technology people <strong>and</strong> a few spaces were made available (at<br />

appropriate share <strong>of</strong> cost) to outsiders from state DOTs.<br />

Newsletter<br />

Volume 3, Number 1 <strong>of</strong> the Transportation Technology e-Transfer newsletter was released in<br />

April 2008. The next Transportation Technology e-Transfer newsletter (Vol. 3, No. 2) is<br />

planned for August 2008. It will cover the 2008 Petersen Conference <strong>and</strong> Pavement<br />

Performance Prediction Symposium (July 14-18, 2008), WRI’s 25 Year history, <strong>and</strong> continue to<br />

cover WRI’s work on the MEPDG.<br />

75


SUBTASK 4-2. WEBSITE MAINTENANCE<br />

The Western Research Institute web site <strong>and</strong> the www.petersenasphaltconference.org site were<br />

updated in preparation for the WRI- <strong>and</strong> FHWA-sponsored 2008 Petersen Asphalt Research<br />

Conference <strong>and</strong> the Pavement Performance Prediction Symposium to be held in Laramie,<br />

Wyoming, in July 2008.<br />

77


SUBTASK 4-3. RESEARCH DATABASE—CONTRACTOR TEAM ACTIVITIES<br />

This database is to include recent, ongoing <strong>and</strong> planned research <strong>and</strong> outreach activities. More<br />

specifically, research problem statements, timelines, research results, contacts <strong>and</strong> relationships<br />

to other studies are to be posted. It is to be designed in a format similar to the (portl<strong>and</strong>)<br />

concrete pavements database found at http://www.cproadmap.com//research/search.aspx. WRI<br />

has interpreted this to mean that WRI is to develop its own such database <strong>and</strong> the COTR has<br />

agreed to this. It is the opinion <strong>of</strong> the COTR that this database is not to have restricted access as<br />

with the concrete pavements database. This effort is being accomplished as follows: As each set<br />

<strong>of</strong> work plans has been accepted by FHWA (with revisions if necessary), <strong>and</strong> as quarterly reports<br />

are submitted, each has been posted at www.asphaltmodelsetg.org . Outreach activities can be<br />

seen at www.westernresearch.org . Click Transportation Technology <strong>and</strong> choose from among 18<br />

outreach activities listed in the left column. Although these two sites do not exactly duplicate the<br />

portl<strong>and</strong> concrete database format, the two web sites do provide all <strong>of</strong> the information required<br />

for this subtask.<br />

79


SUBTASK 4-4. RESEARCH IN PROGRESS DATABASE (Two parts)<br />

The effort in this subtask is to incorporate various types <strong>of</strong> data from this project into two<br />

databases. The contract requires listings in the Transportation Research Board (TRB) Research<br />

in Progress (RIP) database <strong>and</strong> the FHWA R&D Project Tracking System.<br />

The TRB RIP Database<br />

On the next two pages is a copy <strong>of</strong> the TRB RIP entry for this project. Note that the URL for<br />

this project is given on the page <strong>of</strong> this entry. The general URL for the TRB RIP is<br />

http://rip.trb.org/browse/additions.asp?days=7<br />

The FHWA R&D Project Tracking System<br />

Since the beginning <strong>of</strong> this project, every quarter (<strong>and</strong> sometimes more frequently) WRI has<br />

inquired as to whether the FHWA R&D Project Tracking System was developed <strong>and</strong> ready for<br />

use. Each time through June 20, 2008, FHWA (TFHRC) responded that the system is not yet<br />

ready for use.<br />

Although (obviously) WRI has not seen the format for the FHWA R&D Project Tracking<br />

System, from the title <strong>of</strong> this part <strong>of</strong> Subtask 4-4, it appears to call for the same information that<br />

is already reported in Subtask 4-3. If this is the case it raises the question <strong>of</strong> whether the same<br />

information should be reported twice.<br />

81


SUBTASK 4-5. SUPPORT OF THE MECHANISTIC-EMPIRICAL PAVEMENT<br />

DESIGN GUIDE<br />

In this subtask WRI is to supply relevant models, materials <strong>and</strong> materials data to the NCHRP<br />

9-30A Project Administrator, Mr. Harold Von Quintus, to further develop or contribute to the<br />

Mechanistic-Empirical Pavement Design Guide. This process will begin as new data <strong>and</strong> models<br />

are available from research on the FHWA-approved work plans for this contract. Mr. Von<br />

Quintus has requested <strong>and</strong> has been supplied with a copy <strong>of</strong> Chapters 1 <strong>and</strong> 6 <strong>of</strong> the DTFH61-<br />

99C-00022 Final Report (which is in press). Mr. Von Quintus also requested that 600 pounds <strong>of</strong><br />

loose mix be shipped to him from each section <strong>of</strong> all future WRI validation sites.<br />

85


SUBTASK 4-6. SEMI-ANNUAL MEETINGS<br />

Work Conducted this Quarter<br />

The topic for the 2008 P3 Symposium will be “Warm Mix <strong>and</strong> Recycled Asphalt Pavements.”<br />

This meeting is scheduled for July 16-18, 2008 in the Washakie Center on the University <strong>of</strong><br />

Wyoming Campus. The Symposium links researchers, highway <strong>of</strong>ficials, producers <strong>and</strong> others<br />

with a need to underst<strong>and</strong> how asphalts may perform in a given application over time.<br />

All <strong>of</strong> the preliminary reservations have been made for the venue, parking, hotel room blocks,<br />

food, etc. The full program is set <strong>and</strong> available for viewing at<br />

www.petersenasphaltconference.org.<br />

Work to be Conducted Next Quarter<br />

During the next quarter we will hold the Symposium <strong>and</strong> begin the reservation process for next<br />

year.<br />

Work for the 2009 Annual Project Review in Washington, DC, will not begin until September or<br />

October <strong>of</strong> this year. It is believed that it will remain at the Marriott Wardman Park directly<br />

following TRB 2009.<br />

Support <strong>of</strong> FHWA Strategic Goals<br />

The work conducted in this subtask supports the FHWA strategic goal to optimize pavement<br />

performance.<br />

87

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!