You are on page 1of 15

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

net/publication/232887103

Accelerated performance testing of concrete pavement with short slabs

Article  in  International Journal of Pavement Engineering · June 2011


DOI: 10.1080/10298436.2011.575134

CITATIONS READS
18 164

3 authors, including:

Victor G Cervantes Armen Amirkhanian

3 PUBLICATIONS   58 CITATIONS   
University of Alabama
20 PUBLICATIONS   328 CITATIONS   
SEE PROFILE
SEE PROFILE

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

Friction and Texture Retention of Concrete Pavements after Diamond Grinding and Grooving View project

All content following this page was uploaded by Armen Amirkhanian on 27 January 2017.

The user has requested enhancement of the downloaded file.


GPAV 575134—19/4/2011—RAJA.S—390217——Style 4

International Journal of Pavement Engineering


Vol. 00, No. 0, 2011, 1–14

Accelerated performance testing of concrete pavement with short slabs


5 60
Jeffery R. Roeslera, Victor G. Cervantesb and Armen N. Amirkhaniana*
a
Department of Civil and Environmental Engineering, University of Illinois Urbana– Champaign, Urbana, USA; bNAVFAC, Naval
Facilities Engineering Service Center, Port Hueneme, USA
(Received 24 September 2010; final version received 22 March 2011)
10 65
A new concept for designing concrete pavements by optimising the slab geometry in order to reduce the slab thickness as
well as to minimise the mechanical load transfer devices has recently been proposed. Theoretically, the reduced slab size
lowers the load and curling-induced tensile stresses and concomitantly a thinner concrete slab can be constructed. Full-scale
test sections were constructed and tested under accelerated pavement loading conditions to validate this design concept
hypothesis. The design and concrete material factors studied in this research were concrete thickness of 9, 15 or 20 cm;
15 70
granular or asphalt concrete base layer; and plain or fibre-reinforced concrete (FRC). A methodology was presented to
convert the channelised traffic loading to ESALs so that comparisons could be made between the various test sections. The
accelerated pavement testing showed that shorter slab sizes can sustain a significant number of overloads and greater number
of ESALs before developing cracking relative to standard jointed concrete pavements. The most prevalent distress observed
was corner cracking which occurred twice as much as longitudinal cracks, whereas only 3 out of 46 cracking distresses were
transverse cracks. The 20 cm concrete slabs on granular base did not experience fatigue cracking for trafficking up to
20 75
51 million ESALs. The 15 cm concrete slabs on granular base began cracking on an average of 11 million ESALs. As
expected, the concrete slabs on asphalt base resisted a significant larger number of ESALs relative to the same concrete
thickness on granular base. The cracking performance of the 9 cm concrete slabs on granular base varied with the stiffness of
the soil. For the 9 cm slab thickness, structural fibres provided a longer fatigue life and extended service life relative to the
plain concrete slabs. Finally, the smaller slab sizes maintained a medium-to-high load transfer efficiency over the
accelerated loading period for all slab thicknesses without the development of any faulting. As expected, these slab systems
25 80
resulted in higher deflections, and, therefore, the granular base and subgrade layers as well as lateral drainage system must
be designed and specified to reduce the rate of permanent deformation and minimise the possibility of support erosion.
Keywords: concrete pavements; short slab geometry; full-scale testing; accelerated testing

30 1. Introduction sizes become larger, the susceptibility of the pavement 85


One factor linked to the favourable performance of system to temperature, moisture curling and excessive
concrete pavements is the slab geometry (Darter et al. joint openings increases.
1985, AASHTO 1993, Smith et al. 1998, ARA 2007). There have been numerous studies over the years on
Jointed reinforced concrete pavements were originally built-in curling, and it has been reported to be quite high in
35 designed to limit the number of joints and increase the some regions due to the local climate conditions (Hveem 90
1951, Larrain 1985, Armaghani et al. 1987, Poblete et al.
pavement smoothness, but this concrete pavement type
1988, Eisenmann and Leykauf 1990, Bustos et al. 1998,
eventually resulted in many intermediate panel cracks and
Yu et al. 1998b, Hansen et al. 2000, Vandenbossche 2003a,
large movements at the joint. Continuously reinforced
2003b, Rao and Roesler 2005, Rojas-Torrico 2008).
concrete pavements were designed to avoid man-made
40 With the presence of large built-in curling, mechanistic – 95
contraction joints by providing continuous reinforcing empirical designs with conventional slab sizes can result in
steel which would produce short crack spacings and larger required thicknesses (ARA 2007, Hiller 2007) and
ultimately, a smooth riding surface. Conventional jointed alternative failure locations (Hiller and Roesler 2005).
plain concrete pavements (JPCP) with 3.7 –5.0 m joint With excessive built-in curling, conventional JPCP may
45 spacing have also produced long service lives in the US not be the most economical pavement system to construct. 100
(Smith et al. 1998, Yu et al. 1998a). With these slab sizes, In an effort to improve the cost-benefit of new rigid
the thicknesses generally range from 15 to 36 cm. For a pavements, a design concept was proposed that would
mechanistic – empirical procedure, the design thickness shorten the slab size to allow only one wheel load per slab
primarily depends on the expected axle weights, number of as well as reduce the curling stresses, yet still provide the
50 load repetitions, concrete strength, slab dimensions, required fatigue life for a given thickness (Covarrubias 105
subgrade support and local climate conditions. As slab and Covarrubias 2008). Previous research on ultra-thin

*Corresponding author. Email: amirkha1@illinois.edu


55 110
ISSN 1029-8436 print/ISSN 1477-268X online
q 2011 Taylor & Francis
DOI: 10.1080/10298436.2011.575134
http://www.informaworld.com
GPAV 575134—19/4/2011—RAJA.S—390217——Style 4

2 J.R. Roesler et al.

whitetopping has shown that smaller slab sizes (1.2 – Three 40 m test sections were constructed contiguously to
1.8 m) improve the performance of this concrete determine the effects of the slab thickness, base stiffness
rehabilitation option (Mack et al. 1997, ACPA 1998, and concrete mixture on the concrete pavement perform-
Vandenbossche and Fagerness 2002, Roesler et al. 2008). ance. A single lane of 3.6 m wide was constructed and
115 With the introduction of this new short slab design saw-cut into 1.8 m slab sizes. Figure 1(a) shows the 14 170
concept, many questions arise regarding the fatigue life of slabs on test Section 1 that were loaded with the APT
the concrete, stability of the base layer and joint behaviour. device. Slabs 5– 11 are the 10 cm slabs, whereas slabs
A full-scale testing plan was developed to validate the 12 –18 are 15 cm thick. These slabs were supported by
viability of the design concept, to provide fatigue data for 21.5 and 14 cm asphalt concrete base, respectively, on a
120 calibrating the design software (Covarrubias 2009, silty – clay soil. Section 2, shown in Figure 1(b), consisted 175
Covarrubias et al. 2010) and to evaluate the performance of 15 and 20 cm of concrete thickness over 15 cm
of the base layer and joints under repeated loading. aggregate base. Slabs 27 –33 are the 15 cm thick slabs and
34 –40 are the 20 cm slab thicknesses. Figure 1(c) is the
layout for Section 3 which consists of 9 cm slab thickness
125 2. Experimental plan over a 15 cm aggregate base. The test variable in this 180
In order to determine the fatigue and joint performance of section is that half of the slabs are plain concrete (slabs
concrete pavements with shorter slab sizes, accelerated 49 –55), while the remaining slabs (slabs 56 – 62) contain
pavement testing (APT) was conducted. The purpose of FRC. The slab numbers not listed above were locations
APT is not to traffic the pavement with similar wheel load where the APT machine’s supports sat on the pavement
130 and repetitions experienced in typical roadways, because surface. There were no dowels or tie bars in any of the test 185
this would take a prohibitive amount of time. The main section and no joint sealants were used.
objective is to overload the test section with the APT The Accelerated Transportation Loading ASsembly
device so that slab failure occurs in a shorter amount of (ATLAS), seen in Figure 2, was utilised to complete the
time and then the results are equated to standard loading response and failure testing of the three test sections using
135 levels and repetitions expected for in-service pavements. the unidirectional testing mode. The ATLAS applied 190

140 195

145 200

150 205

155 210

160 215

Figure 1. Accelerated concrete pavement test section for short-jointed slabs: (a) Section 1 with 10 and 15 cm sections on asphalt
concrete base, (b) Section 2, comparison of 15 and 20 cm sections on an aggregate base and (c) Section 3, comparison of 9 cm section on
165 aggregate base with and without FRC (not to scale). 220
GPAV 575134—19/4/2011—RAJA.S—390217——Style 4

International Journal of Pavement Engineering 3

225 280

230 285

235 290

Figure 2. ATLAS device shown with aircraft tyre.


240 295
approximately 2500 repetitions per day over a 26 m loading was approximately 333 kg/m3 with 25% replacement of
length. For most of the testing, an aircraft tyre was used Type C fly ash. The water-to-cement ratio was 0.42. The
except for trafficking on Section 3 with a wide-based single dolomite coarse aggregate had a maximum aggregate size
tyre (Michelin X One). All of the test sections were exposed of 25 mm. The average flexural strength (ASTM C78)
245 to the effects of ambient conditions and ATLAS loading measured at 90 days was 6.7 MPa. Synthetic macro-fibres 300
except for Section 3 (north) where environmental panels were added at a dosage of 3.6 kg/m3 for the FRC test
enclosed the test section. By combining the results of the Section 3b. This level of fibres was selected to achieve an
full-scale tests with a mechanistic – empirical analysis equivalent flexural strength ratio (R150
150 ) of 30% (Roesler
approach, a design procedure to optimise the slab geometry et al. 2008) based on the measured residual strength (f 150
150 )
250 for a given traffic level, soil support value and climate from ASTM C1609-07. A superplasticiser was also added 305
condition can be developed (Covarrubias 2009, 2010). at the site to promote uniform dispersion of the fibres in the
concrete.

3. Pavement layers and materials


255 The subgrade material supporting the test section was 4. Falling weight deflectometer 310
primarily a brown silty – clay with trace sand. The One month after the construction of the 120 m test section,
maximum dry density (ASTM D1557) for the soil was falling weight deflectometer testing was conducted on the
determined to be 1962 kg/m3 at an optimum moisture north section of slabs (number 1 –65). The centre slab
content of 10.7% and was compacted to a minimum of 95% normalised deflections are plotted vs. the slab number in
260 maximum dry density. Dynamic cone penetration tests Figure 3 and confirms the expected trends. The deflections 315
were conducted on the subgrade shortly after compaction. are smallest when the concrete slabs are supported by the
Correlated CBR values for the subgrade ranged from 4 to stiffer asphalt concrete layer (Section 1) or thicker slabs
20 depending on the spatial location on the 120 m test (15 vs. 20 cm) in Section 2. The highest deflection was
section and in situ soil moisture content. To achieve a more observed on slab 56 which is the transition between the
265 free-draining granular base layer and to minimise the 9 cm plain and FRC section but there was no deflection 320
pumping potential, a graded dolomite material with a difference between the plain and FRC slabs.
maximum aggregate size of 25 mm was utilised. This
material had approximately 6% fines passing the 75 m sieve Table 1. Concrete mixture design.
and a maximum dry density of 2140 kg/m3 at an optimum
Material Quantity (kg/m3)
270 moisture content of 4.8%. Prior to the placement of the 325
granular base, a 400 g/m2 non-woven geotextile was Coarse aggregate 1128
Fine aggregate 720
inserted between subgrade and subbase to avoid pen- Cement 250
etration and mixing of these two unstabilised materials. Fly ash (type C) 83
The concrete mixture design for the test sections is Water 140
275 shown in Table 1. The total cementitious content 330
GPAV 575134—19/4/2011—RAJA.S—390217——Style 4

4 J.R. Roesler et al.

335 390

340 395

345 400

350 405

355 410

Figure 3. Centre slab 40 kN normalised deflections (solid line) and load transfer efficiencies (dashed line) on north lane.

360 Due to the short slab sizes, only 52% of the joints had understood loading terms used in pavement design, such 415
fully propagated cracks by the end of the first month, and as ESALs. In past research efforts with APT, the use of
thus, dominant joints were present on the test sections. In individual wheel loads, number of load repetitions and
fact, with the reduced stresses in shorter slabs, only 54 Miner’s cumulative damage approach have not necessarily
out of 65 of the contraction joints had formed after four resulted in a better cracking predictions relative to ESALs
365 months. Figure 3 shows the load transfer efficiency (Rao and Roesler 2004, Rao 2005, Kohler and Roesler 420
(LTE) across all the joints after the first month. Joints 20, 2006). For this study, the following equation was used to
23, 32, 35, 45 and 55 had an initial LTE less than 80% correlate the number of wheel passes at a given load level
and they all propagated within 2 weeks after casting the and edge offset to the equivalent number of ESALs:
section.
 4:2
370 P 425
ESAL ¼ b * n * ; ð1Þ
40
5. ATLAS load history and repetition conversion
factor where b is a channelised magnification factor approxi-
As shown in Figure 1, almost all trafficking of the ATLAS mately equal to 20 for channelised edge loading on JPCP
375 wheel load (40 – 156 kN) occurred near the north or south (Packard 1984, Zollinger and Barenberg 1989), n is the 430
longitudinal free edge. Trafficking to failure in the wheel number of passes for a given wheel load, P, and the
path occurred on Section 3 (north lane) and was also used exponent 4.2 approximately represents the load equival-
to gather responses on Sections 1 and 2. A summary of the ency factors from the AASHO Road Test and is also
entire loading sequence for each test section is given in used by the California Department of Transportation to
380 Tables 2 –4 for Sections 1– 3, respectively. calculate ESALs (Caltrans 2008). 435
APT typically exposes the test section to loading In order to establish a lateral wander magnification
conditions that will cause significant levels of pavement factor (b) for each section and wheel offset, a finite
damage (cracking) at a relatively lower number of wheel element analysis using ILLISLAB (Khazanovich 1994)
passes. The accelerated pavement damage induced by was required. The modulus of subgrade reaction (k) for
385 overloading must then be related to more commonly each of the test sections, shown in Table 5, was 440
GPAV 575134—19/4/2011—RAJA.S—390217——Style 4

International Journal of Pavement Engineering 5

Table 2. ATLAS loading pattern for Section 1.

10 cm 15 cm
Load [kN] Passes Cumulative ESALs Percent cracked Cumulative ESALs Percent cracked
445 500
Section 1 south (10 cm b ¼ 20, 15 cm b ¼ 20)
40.0 2506 5.01 E4 0.0 5.01 E4 0.0
66.7 2500 4.77 E5 0.0 4.77 E5 0.0
80.1 2500 1.40 E6 0.0 1.40 E6 0.0
93.4 3499 3.85 E6 0.0 3.85 E6 0.0
111 1126 5.50 E6 14.3 5.50 E6 0.0
450 505
111 1211 7.27 E6 28.6 7.27 E6 0.0
124 2300 1.27 E7 71.4 1.27 E7 0.0
137 2200 2.06 E7 71.4 2.06 E7 0.0
151 25 2.07 E7 71.4 2.07 E7 0.0
156 2730 3.71 E7 85.7 3.71 E7 14.3
156 1170 4.41 E7 100.0 4.41 E7 14.3
455 510
147 1642 5.18 E7 100.0 5.18 E7 14.3
147 347 5.35 E7 100.0 5.35 E7 14.3
147 274 5.48 E7 100.0 5.48 E7 14.3
147 62 5.50 E7 100.0 5.50 E7 14.3
156 416 5.75 E7 100.0 5.75 E7 14.3
460 Section 1 north (10 cm b ¼ 20, 15 cm b ¼ 20) 515
40.0 2700 5.40 E4 0.0 5.40 E4 0.0
66.7 2500 4.81 E5 0.0 4.81 E5 0.0
93.4 2230 2.05 E6 85.7 2.05 E6 0.0
93.4 270 2.24 E6 85.7 2.24 E6 0.0
111 136 2.44 E6 85.7 2.44 E6 0.0
465 111 2680 6.35 E6 100.0 6.35 E6 0.0 520
133 4870 2.16 E7 0.0
156 500 2.46 E7 14.3
156 1000 3.07 E7 14.3
156 3500 5.17 E7 28.6
156 2954 6.94 E7 57.1
470 525

backcalculated from ILLISLAB based on the measured allowable repetitions (Na):


deflections from a wheel load of 10 cm offset from the free
 s 4
475 edge. The tensile stresses at the bottom of the mid-slab n
N a ¼ 225; 000 * ; ð2Þ 530
edge as the load was moved laterally in 2.5 cm increments MOR
(0 – 50 cm) were then determined. One significant chal-
lenge in determining the channelised magnification factor where sn is the stress at a given distance from the edge and
(b) was the choice of the fatigue algorithm to translate the MOR is the modulus of rupture of the concrete. As all
480 tensile stresses to allowable number of repetitions. Most of the trafficking was channelised, the magnification factor 535
the concrete fatigue algorithms are extremely sensitive to (b) for each section was calculated using a ratio of the
changes in tensile stresses, and, therefore, only wheel longitudinal edge fatigue damage produced at the
loads adjacent to the slab edge contribute to the fatigue channelised trafficking offset over the fatigue damage
damage. The approach published by Zollinger and produced by a load in the wheel path (45 cm offset):
485 Barenberg (1989) and the Portland Cement Association 0 1 540
ne ðxÞ
(Packard 1984) was attempted with several concrete N allowðxÞ
fatigue algorithms without success for these shorter b ¼ @  A; ð3Þ
ne ðxÞ
concrete slab geometries. N allowð45Þ
A fatigue equation developed by Vesic and Saxena
490 (1969) based on the AASHO Road Test concrete sections where ne is the expected number of passes at some lateral 545
was finally selected because it was based on full-scale offset (x), Nallow is the number of allowable passes at that
traffic with wander, concrete slabs on a granular base layer lateral offset and Nallow(45) is the number of allowable
and the concrete slabs failed under various modes, i.e. passes at 45 cm offset. As expected, this magnification
erosion and cracking. The Vesic and Saxena (1969) fatigue factor is a function of the slab thickness and base type.
495 equation, given next, was used to calculate the number of Table 5 lists the magnification factor used to calculate 550
GPAV 575134—19/4/2011—RAJA.S—390217——Style 4

6 J.R. Roesler et al.

Table 3. ATLAS loading pattern for Section 2.

15 cm 20 cm
Load [kN] Passes Cumulative ESALs Percent cracked Cumulative ESALs Percent cracked
555 610
Section 2 south (15 cm b ¼ 19, 20 cm b ¼ 16)
40.0 5375 1.02 E5 0.0 8.76 E4 0.0
40.0 1912 1.39 E5 0.0 1.19 E5 0.0
66.7 6000 1.12 E6 0.0 9.55 E5 0.0
66.7 150 1.14 E6 0.0 9.75 E5 0.0
89.0 2850 2.70 E6 0.0 2.30 E6 0.0
560 615
89.0 2265 3.94 E6 0.0 3.36 E6 0.0
89.0 5112 6.73 E6 0.0 5.74 E6 0.0
111 6995 1.65 E7 28.6 1.41 E7 0.0
111 2651 2.01 E7 42.9 1.72 E7 0.0
111 1655 2.25 E7 42.9 1.92 E7 0.0
156 75 2.29 E7 57.1 1.96 E7 0.0
565 620
Section 2 north (15 cm b ¼ 18, 20 cm b ¼ 15)
40.0 2850 5.01 E4 0.0 4.39 E4 0.0
40.0 150 5.28 E4 0.0 4.62 E4 0.0
66.7 3187 5.32 E5 14.3 4.66 E5 0.0
66.7 1913 8.20 E5 14.3 7.17 E5 0.0
570 66.7 100 8.35 E5 14.3 7.31 E5 0.0 625
80.1 5000 2.45 E6 14.3 2.15 E6 0.0
93.4 5000 5.54 E6 14.3 4.85 E6 0.0
111 816 6.59 E6 14.3 5.77 E6 0.0
111 1801 8.91 E6 57.1 7.79 E6 0.0
111 640 9.73 E6 57.1 8.51 E6 0.0
575 120 2405 1.40 E7 57.1 1.23 E7 0.0 630
133 200 1.45 E7 57.1 1.27 E7 0.0
156 200 1.56 E7 71.4 1.37 E7 0.0
156 150 1.64 E7 71.4 1.44 E7 0.0
156 32 1.66 E7 85.7 1.45 E7 0.0
111 292 1.69 E7 85.7 1.48 E7 0.0
580 111 152 1.50 E7 0.0 635
156 300 1.64 E7 0.0
156 7070 4.91 E7 0.0
111 2000 5.13 E7 0.0

585 640
ESALs for each test section. During the trafficking of the for all sections was calculated at this offset with a
sections, the wheel was typically offset from the free edge maximum value of 20 (Cervantes 2009). A magnification
of approximately 13 cm, and thus, the magnification factor factor was not used when the wheel load was trafficked
greater than 30 cm from the edge.
590 645
Table 4. ATLAS loading pattern for Section 3.

Percent slabs cracked 6. Slab cracking performance


Cumulative
Load [kN] Passes ESAL PCC FRC The slab’s cracking patterns were documented in order to
determine the failure modes of the various sections and
Section 3 south
595 how they continued to deteriorate with trafficking. A full 650
22.2 2643 4477 0.0 0.0
22.2 2778 9183 0.0 0.0 description of the crack development on each section can be
40.0 3000 69,183 0.0 0.0 found in Roesler and Cervantes (2009). Figures 4–6 are
40.0 309 75,363 0.0 0.0 typical failure patterns for Sections 1–3, respectively, at the
40.0 5875 192,863 57.1 14.3 end of the accelerated load testing. Cracking patterns varied
40.0 2053 233,923 57.1 28.6
600 depending on the location of the wheel load, thickness of 655
53.4 10 234,592 100.0 42.9
the slab and deformation levels in the base layers. Overall,
Section 3 north (wheelpath) a total of 46 slabs developed fatigue cracks out of the 84
40.0 2509 2509 0.0 0.0
40.0 451 2960 85.7 14.3
slabs tested on the three test sections. Corner breaks were
40.0 1585 4545 N/A 28.6 61% of the cracking distresses observed followed by
605 longitudinal cracking at 33% and only 6% transverse 660
GPAV 575134—19/4/2011—RAJA.S—390217——Style 4

International Journal of Pavement Engineering 7

Table 5. Backcalculated modulus of subgrade reaction (k) values and magnification factor (b) for each section.

Slab/base description Section number Section location k (MPa/m) Magnification factor


10 cm/Asphalt 1 North/south 68 20
665 15 cm/Asphalt 1 North/south 136 20 720
15 cm/Granular 2 South 81 19
15 cm/Granular 2 North 27 18
20 cm/Granular 2 South 41 16
20 cm/Granular 2 North 27 15
9 cm/Granular 3 South 41 20
670 9 cm/Granular 3 Northa 14 20 725
a
30 cm offset

cracks. As the maximum calculated tensile stress occurred This equation assumes that a terminal serviceability of 2.5
675 at the bottom of the mid-slab edge under fully supported is reached when cracking is between 0 and 20%, whereas 730
conditions, the appearance of longitudinal and corner 100% cracking has a PSIt of 1.0. Table 6 lists the AASHTO
cracks as the predominant distress confirmed a change in predicted ESALs for the same concrete thickness at
base/subgrade support stiffness had occurred with increased 90% reliability, standard deviation of 0.35, load transfer
load level and repetitions. The support change can be seen coefficient (J) of 3.2 and a drainage coefficient of 0.9. The
680 in Figure 7 which tracks the increase in the slab’s rebound composite k-value, shown in Table 6, was modified from 735
deflection with repetitions for a constant load level. the soil k-value listed in Table 5 for the stabilised base
Table 6 is a summary of the maximum fatigue cracking stiffness and thickness in Section 1. As seen in Table 6, the
measured on each test section vs. the total applied ESALs.
As expected, the thicker slabs and the same thickness slab
685 with a stiffer base layer resisted more ESALs until a 740
similar level of cracking. Figure 8 graphically presents the Slab 27 28 29 30 31 32 33
development of fatigue cracking as a function of ESALs
for several slab thicknesses on both base types.
In order to evaluate the viability of short-jointed slabs
690 with conventional JPCP, the APT traffic results were 745
compared with the allowable traffic from the AASHTO
Slab 34 35 36 37 38 39 40
(1993) design guide. The first step in this comparison is to
relate the observed cracking at the end of the service life to
a terminal serviceability (PSIt). The following equation
695 was developed to estimate PSI as a function of slabs 750
cracked (C) only, because roughness was not measured
during testing. Figure 5. Section 2 (south) cracking performance after
trafficking (15 cm ¼ 22.9 million ESALs; 20 cm ¼ 20 million
ESALs; Slabs 27 – 33 are 15 cm and slabs 34 – 40 are 20 cm).
PSI ¼ 2:875 2 0:01875C; 20 # C # 100: ð4Þ
700 755

Slab 5 6 7 8 9 10 11 Slab 49 50 51 52 53 54 55

705 760

Slab 12 13 14 15 16 17 18 Slab 56 57 58 59 60 61 62

710 765

Figure 4. Section 1 (south) cracking performance after 57.5 Figure 6. Section 3 (south) final cracking performance after
million ESALs (slabs 5 – 11 are 10 cm and slabs 12 –18 are 234,000 ESALs (slabs 49 – 55 are 9 cm plain concrete and slabs
715 15 cm). 56 – 62 are 9 cm FRC). 770
GPAV 575134—19/4/2011—RAJA.S—390217——Style 4

8 J.R. Roesler et al.

775 830

780 835

785 840

790 845

795 850

800 855

805 860
Figure 7. (a) Rebound deflections for the 9 cm plain concrete section for slabs 51, 52 and 53; (b) Rebound deflection of the 9 cm FRC
slab over a granular base (Section 3 south).

Table 6. Maximum fatigue cracking and ESALs for accelerated pavement test sections and AASHTO predicted ESALs at the estimated
810 terminal serviceability. 865

Composite Estimated terminal AASHTO


Thickness Location Base type k-value [MPa/m] Total ESALs Slabs cracked serviceability predicted ESALs
10 cm 1-South Asphalt 190 4.4 E7 100.0% 1.0 3.2 E6
10 cm 1-North Asphalt 190 6.4 E6 100.0% 1.0 3.2 E6
815 15 cm 1-South Asphalt 136 3.7 E7 14.3% 2.5 3.2 E6 870
15 cm 1-North Asphalt 136 6.9 E7 57.1% 1.8 4.2 E6
15 cm 2-South Granular 81 2.3 E7 57.1% 1.8 2.8 E6
15 cm 2-North Granular 27 1.7 E7 85.7% 1.3 1.6 E6
20 cm 2-South Granular 41 2.0 E7 0.0% 2.5 6.8 E6
20 cm 2-North Granular 27 5.1 E7 0.0% 2.5 5.9 E6
820 9 cm Plain 3-Southa Granular 41 2.3 E5 100.0% 1.0 2.4 E5 875
9 cm FRC 3-Southa Granular 41 2.3 E5 42.9% 2.1 2.0 E5
9 cm Plain 3-Northb Granular 14 2.9 E3 85.7% 1.3 9.6 E4
9 cm FRC 3-Northb Granular 14 4.5 E3 28.6% 2.3 8.9 E4
a
Frozen tests sustained 229,000 ESALs without cracking.
825 b 880
Spring thaw testing period.
GPAV 575134—19/4/2011—RAJA.S—390217——Style 4

International Journal of Pavement Engineering 9

885 940

890 945

895 950

900 955

905 Figure 8. Percentage of cracked slabs on aggregate and asphalt base vs. ESALs. 960

only short-jointed section that did not sustain more ESALs testing was between 100 and 80% for the 15 and 20 cm
than the conventional JPCP from AASHTO was Section 3 concrete slabs. This initial response resulted from the
north. The main reason for this performance was the dominant joint effect, i.e. joints that cracked early tended
910 965
testing that was completed during a thaw period of the soil to have a greater crack width than joints that propagated
and base layer, which lead to a premature failure. Overall, later. The LTE for joint 30 and 37 fell below 40% at about
the APT has shown that the fatigue performance of short- 7000 passes but recovered to about 70% at the end of
jointed slabs in terms of ESALs significantly exceeds the testing. The instrumented joints in Section 2 (north and
allowable traffic calculated from the 1993 AASHTO south) converged to an LTE approximately of 70%,
915 970
design guide for conventional JPCP. suggesting that the slabs more evenly distributed the joint
movement between adjacent slabs as the load repetitions
increased. The accelerated pavement test data suggested
7. Joint performance and LTE that the slabs with optimised geometry can have sufficient
joint performance without load transfer devices for low-to-
920 One proposed advantage of concrete slabs with optimised 975
medium traffic volume facilities. Whether the joints will
geometry is the potential for limiting the placement of
continue to function at a high level over a 20-year period
mechanical load transfer devices. The transverse contrac-
given higher speeds and ESAL loadings is still uncertain,
tion joints in the various test sections relied on aggregate
but as noted, several sections did have adequately
interlock as the primary load transfer mechanism. There
performing joints at 50 million ESALs under the APT.
925 was no measurable faulting on any transverse contraction 980
Several potential solutions for extending joint perform-
joints although loading speeds were less than 10 km/h. In
ance and reliability for higher speed and ESAL loadings
order to further support the limited usage of dowels across
may be targeted doweling systems or addition of structural
the transverse contraction joints, the LTE across the joints
fibres.
must be maintained at a certain level throughout the
930 testing. Deflection LTE values were calculated vs. load 985
repetitions for several instrumented joints. The LTE for
Section 3 was above 90% during the entire test for both 8. Size effect on concrete slab fatigue life and
plain and FRC. The LTE for Section 1 was highly overloads
influenced by the thick asphalt base layer. As shown in Researchers have recognised that the specimen size (e.g.
935 Figure 9 for Section 2, the range of LTE at the start of the thickness) affects the concrete’s nominal strength (Bazant 990
GPAV 575134—19/4/2011—RAJA.S—390217——Style 4

10 J.R. Roesler et al.

40.0 kN Loading 66.7 kN Loading 89.0 kN Loading 111 kN Loading


100 60

90
50
995 80 1050

70
40

Temperature (˚C)
Efficiency (%)
60

50 30
1000 1055
40
20
30

20
10
1005 10 1060

0 0
0 5,000 10,000 15,000 20,000 25,000 30,000
Passes

Joint 29 Joint 36
1010 Surface Temp. 1065
Joint 30 Joint 37

Figure 9. Deflection LTE for edge loading (Section 2 south).

1015 1070
and Planas 1998). This translates into thinner specimens capacities, respectively. The interaction between the crack
having a greater nominal strength than thicker concrete growth, concrete fracture properties, slab geometry and
specimens with the same geometric ratios and boundary boundary condition (soil foundation) is the main reason for
conditions. Rao (2005) developed curves based on the discrepancy with simply supported concrete beam
1020 Bazant’s size effect method (Bazant and Kazemi 1990) tests. Furthermore, the 9 cm FRC slab relative to the 15 cm 1075
to account for the decrease in apparent strength as concrete plain concrete slab results demonstrates the size effect and
slab thickness increased to explain full-scale test results of fracture properties effects on the slab capacity. Since the
varying slab thickness. Roesler (2006) also summarised slabs were not geometrically similar as the thickness was
the results of historical beam and slab tests which clearly changed, it is difficult to conclude if there was a strong
1025 demonstrated that the beam flexural strength test under- thickness size effect. However, a correction factor to 1080
estimated the slab concrete flexural strength by a factor of account for the scaling of beam flexural to slab flexural
1.3 –2.8. This ratio was related to the type of concrete strength based on the thickness of the slab has been
material (plain vs. FRC), the thickness of the slab and the
implemented into the short-jointed slab design software
non-dimensional slab size.
(Covarrubias et al. 2010) to explain the slab cracking
1030 The ATLAS tests results support these previous 1085
behaviour differences shown in Figure 10.
findings on the size effect. Figure 10 shows the monotonic
Overloaded axles are the main reason why concrete
load – displacement results for several of the 9 and 15 cm
pavements fail prematurely. The results of the APT and the
slabs extracted from the APT sections and tested in a large-
scale laboratory soil box. The mean peak load capacity for monotonic load tests have positive implications for the
1035 the 9 cm FRC slabs was 78 kN, whereas the 15 cm had a design of concrete pavements in corridors where over- 1090
mean peak load capacity of 110 kN. These peak values loaded vehicles are expected or weight enforcement is
should be considered an upper limit because the support insufficient. During the APT loading, the 20 cm concrete
condition was not subjected to repeated or moving wheel slab on granular base was able to resist over 7000
loads or varying climatic conditions. For the slab tests repetitions of 156 kN wheel load (16 t) and a total of 51.3
1040 shown in Figure 10, the measured concrete beam flexural million ESALs without any slab cracking. Likewise, the 1095
strength was 6.7 MPa which translated into maximum load single wheel load capacity of the 15 cm slabs was
capacity of 28 kN and 75 kN for the 9 cm FRC slab and the approximately 110 kN and the 20 cm slab is anticipated to
15 cm plain concrete slab, respectively. The measured load have an even greater load capacity. The APT and laboratory
capacity of the two slab thicknesses (9 and 15 cm) was testing results suggest that the 20 cm slab thickness with
1045 approximately 2.8 and 1.5 times their beam predicted 1.8 m panel sizes can provide an extended-life concrete 1100
GPAV 575134—19/4/2011—RAJA.S—390217——Style 4

International Journal of Pavement Engineering 11

1105 1160

1110 1165

1115 1170

1120 1175

1125 1180

Figure 10. Monotonic load tests for the 9 cm FRC and 15 cm plain concrete slabs.

1130 1185
pavement as well as sustain a significant number of correlated with the adverse performance of Section 3
overloads. (north).
In the thicker slab sections, the k-value had less of an
influence on the performance of the slab because the
1135 9. Effects of support layer stiffness stresses in the underlying foundation layers were smaller. 1190
The base and subgrade stiffness is an important parameter The 20 cm concrete slabs in Section 2 south and north had
in designing the thickness of concrete pavements with k-values of 41 and 27 MPa/m, respectively, and resisted the
optimised slab geometry as seen in Figure 8. As the slab largest number of ESALs without failure, 19.6 and 51.3
thicknesses are less than conventional JPCP with 4.5 m million ESALs, respectively (see Table 3). The 15 cm slabs
1140 joint spacing, the deformations accumulating in the in Section 2 were somewhat influenced by the soil support 1195
underlying layers are more critical. The trafficking of the layer. The south section had an initial k-value of 81 MPa/m
sections at different times of the year relayed some insight and eventually had four cracked slabs at 22.9 million
into how the support conditions extended or decreased the ESALs. On Section 2 (north), the 15 cm slabs had
fatigue life of the sections. When the trafficking was done experienced six out of seven slabs cracked at 16.5 million
1145 in the winter (January 2008) with a frozen base and ESALs with a k-value of 27 MPa/m. 1200
subgrade layer, 9 cm slabs (Section 3) were able to sustain Section 1 included an asphalt concrete support layer
more than 229,000 ESALs without developing any over the natural subgrade. The backcalculated k-value for
distresses. When this same section was tested in the both the north and south section for the 15 cm slabs was
spring (April 2008), it began showing corner fatigue 136 MPa/m. In terms of cracking performance, the south
1150 cracks at around 75,000 ESALs. Wheel path testing on and north sections had one slab cracked at 37.1 and 24.6 1205
Section 3 (north) during a thaw cycle in March 2009 million ESALs. The 10 cm slabs over asphalt concrete
resulted in significant cracking after 3000 ESALs. The base on Section 1 (north) were much more distressed with
backcalculated k-values for these sections are shown in six out of seven slabs cracked at 2 million relative to 37.1
Table 5, 41 and 14 MPa/m, respectively. This weaker ESALs on the south edge. Although the backcalculated
1155 support condition and likely non-uniformity was directly k-value for both sides (10 cm slab) was approximately 1210
GPAV 575134—19/4/2011—RAJA.S—390217——Style 4

12 J.R. Roesler et al.

68 MPa/m, the concrete slabs lost support due to and erodible material. It is clear from the AASHO Road
trafficking during rainy days without a sufficient drainage Test (HRB 1962) that pumping and failure of concrete
system, thereby resulting in higher distresses on the north pavements can occur with dowelled joints even if the base
set of slabs on Section 1. layer is free draining. In the short-jointed slab design,
1215 higher speed traffic increases the erosion potential along 1270
with the inherently higher slab deflections. In order to
10. Effects of structural fibres on fatigue cracking minimise potential for joint performance or support
and serviceability erosion issues with higher volume traffic designs, efficient
doweling techniques or FRC could be employed, a
Full-scale testing of FRC slabs in the early 1970s by the
1220 drainage system must be installed, the granular base 1275
US Army Corps of Engineers demonstrated the enhanced
should have a minimum CBR, e.g. 80, and geotextile
fatigue life of FRC over plain concrete slabs (Parker 1974,
separation must be placed at the subgrade – base interface.
Rollings 1981). More recent research with application to
Furthermore, research suggests that granular bases with
slab on ground design has shown significant enhancement
non-plastic fines as well as specifying compacted densities
in concrete slab cracking performance by the addition
1225 based on the modified proctor test will greatly improve 1280
of structural fibres (Beckett 1990, Falkner et al. 1995,
the shear strength, reduce the permanent deformation
Roesler et al. 2004). Finally, theoretical analysis of the
potential and lower the anisotropic behaviour of the
flexural capacity of FRC slabs using a fracture-based
aggregate base layer (Tutumluer et al. 2009).
cohesive zone modelling in a finite element framework has
clearly demonstrated that the increased load capacity
1230 1285
fibrous slabs have over plain concrete slabs (Gaedicke 12. Conclusions
2009).
A new concrete pavement thickness design concept based
The APT of 9 cm slabs with and without structural
on optimising the slab dimensions was assessed by APT of
fibres clearly confirmed their benefit to extending the
several full-scale test sections. Three 40 m test sections
fatigue and service life of plain concrete slabs as seen in
1235 were constructed and loaded with an APT device to 1290
Table 5. The deflection response differences can be seen
determine the concrete slab fatigue cracking and joint
when comparing the 9 cm plain concrete with the FRC
performance. The main factors addressed in the test
slabs in Figure 7. The variation in deflection between
sections were slab thickness (9 –10, 15 and 20 cm), base
joints and slabs is much smaller for the FRC section. The
stiffness (granular vs. asphalt concrete base) and concrete
increase in load-carrying capacity afforded by the addition
1240 material type (plain vs. FRC). All slabs had the same 1295
of fibres comes without the need to increase the concrete’s
dimensions of 1.8 m by 1.8 m.
flexural strength (Roesler et al. 2004). The fibres also
APT was completed on all sections until significant
helped extend the service life of the FRC slabs by
cracking distresses and crack deterioration were present.
maintaining vertical and horizontal slab alignments. In
The most prevalent distress on all the sections was corner
addition, structural fibres aided in maintaining high LTE
1245 cracking (61%) followed by longitudinal cracks (33%) and 1300
across the transverse and longitudinal contraction joints
only a few transverse cracks (6%). In most cases,
throughout the concrete pavement’s life. For designs with
significant overloads were required to fail the concrete
thin slabs (, 15 cm) fibres can be used to prevent
sections. The approximate ESALs applied to each section
differential slab horizontal movement due to braking and
were calculated based on lateral wander magnification
tyre friction and can eliminate the need for lateral
1250 based on a cumulative fatigue damage analysis. The 1305
restraining pins to hold tight the longitudinal contraction
results of the APT showed that the 20 cm concrete section
joint. Finally, the severity of the cracking was less on the
over a granular base sustained the greatest number of
FRC section relative to the plain concrete.
ESALs (average of 35 million) without showing any
distress. In fact, this section sustained almost 7400 passes
1255 of a 16 t wheel load without a single slab failure. The 1310
11. Pavement design and material considerations 15 cm slab on an asphalt base sustained over 20 million
The APT results have shown that the fatigue resistance of ESALs before any slab cracking was observed. The 15 cm
the short-jointed slab concept is better to the equivalent concrete slab over the granular base resisted 11 million
thickness of conventional jointed slabs designed by ESALs on average before exhibiting primarily corner
1260 AASHTO (1993; see Table 6). However, there are cracking distress. The 10 cm concrete sections over asphalt 1315
unanswered questions on the long-term performance of concrete withstood almost 4 million ESALs on average
undowelled joints with short slabs, the stability of the before fatigue cracking began developing on the section.
granular base especially under higher traffic volumes, and Finally, the 9 cm concrete slabs on aggregate base began
their collective effects on IRI. Faulting develops as a result developing cracks after 75,000 ESALs when the soil was
1265 of base erosion from high deflections, available moisture in poor condition. The repeated load testing on the 9 cm 1320
GPAV 575134—19/4/2011—RAJA.S—390217——Style 4

International Journal of Pavement Engineering 13

slabs clearly demonstrated that FRC extended the fatigue Bazant, Z.P. and Kazemi, M.T., 1990. Determination of fracture
cracking life of plain concrete and reduced the rate of energy, process zone length and brittleness number from size
crack deterioration. effect, with application to rock and concrete. International
Journal of Fracture, 44, 111– 131.
The support conditions affected the cracking perform-
Bazant, Z.P. and Planas, J., 1998. Fracture and size effect in
1325 ance especially for the thinner sections. The 9 cm concrete concrete and other quasibrittle materials. Florida: CRC 1380
slab on 15 cm aggregate base was tested for over 229,000 Press.
ESALs without any cracking when the base/subgrade was Beckett, D., 1990. Comparative tests on plain, fabric reinforced
frozen. In the spring time testing, this same section began and steel fibre reinforced concrete ground slabs. Concrete, 24
showing signs of fatigue cracking after 75,000 ESALs. (3), 43– 45.
1330 The effect of support conditions on the allowable Bustos, M., et al., 1998. Calibration of performance models for 1385
JPCP using long-term pavement performance database.
repetitions to fatigue cracking could be further seen as Transportation Research Record: Journal of the Transpor-
the 10 and 15 cm concrete slab on an asphalt concrete base tation Research Board, No. 1629, 108– 116.
could take approximately 50 and 2 times more ESALs than Caltrans, 2008. Caltrans highway design manual. California
the concrete slabs on granular base, respectively. Department of Transportation.
1335 The APT demonstrated that the aggregate interlock Cervantes, V.G., 2009. Performance of rigid pavements with 1390
joints can still provide medium-to-high deflection LTE optimized slab geometry. Thesis (MS). University of Illinois
after a significant amount of ESALs. The addition of fibres Urbana – Champaign.
Cervantes, V. and Roesler, J.R., 2009. Performance of concrete
on the 9 cm slab section showed consistently high LTE pavements with optimized slab geometry, Final Report,
across the transverse and longitudinal joints. The climate, FHWA-ICT-09-053, Illinois Center for Transportation,
1340 especially the amount of rainfall, can significantly affect University of Illinois, Urbana, IL, 112 pp. 1395
the performance of slabs with optimised geometry if Covarrubias, J.P., 2009. Optipave v3.015 Software, Santiago,
precautions are not taken. A base layer containing a lower Chile.
percentage of non-plastic fines and high shear strength is Covarrubias, J.P.T. and Covarrubias, J.P.V., 2008. TCP design
for thin concrete pavements. In: 9th International conference
essential for achieving the desired service life. A non-
on concrete pavements, August 17 –21, 2008, San Francisco.
1345 woven geotextile is also a necessity to prevent mixing of Covarrubias, J.P., Roesler, J., and Covarrubias, J.P., 2010. Design 1400
the subgrade and aggregate base layer. Finally, adequate of concrete slabs with optimized geometry. 11th Inter-
lateral drainage is required to avoid lowering the support national symposium on concrete roads, October 13 – 15,
stiffness and unbound material shear strength that could 2010, Sevilla, Spain.
lead to premature failure. Darter, M.I., et al., 1985. Portland cement concrete pavement
1350 evaluation system (COPES). NCHRP Report 277, Transpor- 1405
tation Research Board, National Research Council, Washing-
ton, DC, 175 pp.
Acknowledgements Eisenmann, J. and Leykauf, G., 1990. Simplified calculation of
slab curling caused by surface shrinkage. In: Proceedings of
The authors would like to acknowledge TCPavements Ltd. for the 2nd International workshop on the theoretical design of
their financial support of this research. TCPavements is patented
1355 concrete pavements, Siguenza, Spain. 1410
technology in Chile (44820-2009), the USA (757,1581) and
Falkner, H., Huang, Z., and Teutsch, M., 1995. Comparative
international application PCT/EP2006/064732. The contents of
this report reflect the view of the authors, who are responsible for study of plain and steel fibre reinforced concrete ground
the facts and accuracy of the data presented herein. slabs. Concrete International, 17 (1), 45 – 51.
Gaedicke, C., 2009. Fracture-based method to determine the
flexural load capacity of concrete slabs. Thesis (PhD).
1360 University of Illinois Urbana –Champaign. 1415
Hansen, W., et al., 2000. Validating top– down premature
References
transverse slab cracking in jointed plain concrete pavement.
American Association of State Highway and Transportation Transportation Research Record: Journal of the Transpor-
Officials, USA, 1993. AASHTO Guide for Design of tation Research Board, No. 1809, 52 – 59.
Pavement Structures. Highway Research Board, 1962. The AASHO Road Test – Report
1365 American Concrete Pavement Association, 1998. Whitetopping– 5, Pavement Research. Special Report 61E, Highway Research
1420
state of the practice. Publication EB210P. Skokie, IL:
Board, National Research Council, Washington DC.
American Concrete Pavement Association.
Hiller, J.E., 2007. Development of mechanistic – empirical
ARA, Inc, 2007. Interim mechanistic – empirical pavement
design guide manual of practice. Final Draft. National principles for jointed plain concrete pavement fatigue
Cooperative Highway Research Program Project 1-37A. design. Thesis (PhD). University of Illinois Urbana –
1370 Armaghani, J.M., Larsen, T.J., and Smith, L.L., 1987. Champaign. 1425
Temperature response of concrete pavements. Transpor- Hiller, J.E. and Roesler, J.R., 2005. Determination of critical
tation Research Record: Journal of the Transportation concrete pavement fatigue damage locations using influence
Research Board, No. 1121, 23 – 33. lines. ASCE Journal of Transportation Engineering, 131 (8),
ASTM C1609-07: Standard Test Method for Flexural Perform- 599– 607.
ance of Fiber-Reinforced Concrete (Using Beam with Third- Hveem, F.N., 1951. Slab warping affects pavement joint
1375 Point Loading), 2007. performance. Journal American Concrete Institute, 22 (10). 1430
GPAV 575134—19/4/2011—RAJA.S—390217——Style 4

14 J.R. Roesler et al.

IDOT, 2007. Standard Specifications for Road and Bridge Illinois Center for Transportation Series No. 08-016,
Construction, Illinois Department of Transportation. January University of Illinois, Urbana, IL, June 2008, 181 pp.
2007. Rojas-Torrico, C., 2008. Estudio del Desempeño del Pavimento
Khazanovich, L., 1994. Structural analysis of multi-layered Rı́gido ‘Ancaravi– Huachacalla’ Mediante Análisis Estruc-
concrete pavement systems. Thesis (PhD). University of tural por Elementos Finitos. Thesis (Master). Universidad de
1435 Illinois at Urbana – Champaign, Urbana, Illinois. Tecnica Oruro and Universidad Nacional de Rosario- 1490
Kohler, E. and Roesler, J., 2006. Accelerated pavement testing of Argentina, Bolivia.
extended life continuously reinforced concrete pavement Rollings, R.S., 1981. Corps of engineers design procedures for
sections. Final Report, Transportation Engineering Series rigid airfield pavements. In: Second international conference
No. 141, Illinois Cooperative Highway and Transportation on concrete pavement design, Purdue University, West
Series No. 289, University of Illinois, Urbana, IL, 354 pp. Lafayette, IN.
1440 Larrain, C., 1985. Análisis Teórico-Experimental del Comporta- 1495
Smith, K.D., et al., 1998. Performance of concrete pavements.
miento de Losas de Hormigón de Pavimentos. Thesis (MSc). Volume II– evaluation of in-service concrete pavements.
School of Engineering, Catholic University of Chile. FHWA-RD-95-110, Washington, DC.
Packard, R., 1984. Thickness design for concrete highway and Tutumluer, E., Mishra, D., and Butt, A., 2009. Characterization
street pavements. Skokie, IL: Portland Cement Association. of Illinois aggregates for subgrade replacement and subbase.
Parker, F., 1974. Steel fibrous concrete for airport pavement Illinois Center for Transportation, University of Illinois,
1445 applications. Technical Report 5-74-12, US Army Engineer 1500
Report ICT-090060, 179 pp.
Waterways Experiment Station, Vicksburg, MS. Vandenbossche, J.M., 2003a. Interpreting falling weight
Poblete, M., et al., 1988. Field evaluation of thermal deflectometer results for curled and warped Portland cement
deformations in undoweled PCC pavement slabs. Transpor- concrete pavements. Thesis (PhD). University of Minnesota,
tation Research Record: Journal of the Transportation
Minneapolis, Minnesota.
Research Board, No. 1207, 217– 228.
1450 Vandenbossche, J.M., 2003b. Performance analysis of ultrathin 1505
Rao, S., 2005. Characterizing effective built-in curling and its
whitetopping intersections on US-169: Elk River, Minnesota.
effect on concrete pavement cracking. Thesis (PhD).
Transportation Research Record: Journal of the Transpor-
University of Illinois Urbana Champaign.
tation Research Board, No. 1823, 18 – 27.
Rao, S. and Roesler, J.R., 2004. Cumulative fatigue damage
analysis of concrete pavement using accelerated pavement Vandenbossche, J.M. and Fagerness, A.J., 2002. Performance,
testing results. In: Second international conference on analysis, and repair of ultrathin and thin whitetopping at
1455 accelerated pavement testing, 25 – 29 September 2004, Minnesota road research facility. Transportation Research 1510
Minneapolis. Record: Journal of the Transportation Research Board, No.
Rao, S. and Roesler, J.R., 2005. Characterizing effective built-in 1809, 191– 198.
curling from concrete pavement field measurements. ASCE Vesic, A.S. and Saxena, K., 1969. Analysis of structural behavior
Journal of Transportation Engineering, 131 (4), 320–327. of road test rigid pavements. Highway Research Record No.
Rao, C.B., et al., 2001. Effects of temperature and moisture on 291, Highway Research Board, National Research Council,
1460 the response of jointed concrete pavements. In: Proceedings 156– 158. 1515
of the 7th International conference on concrete pavements, Yu, T., et al., 1998a. Performance of concrete pavements. Volume
9 – 13 September, 2001, Orlando, 23 – 38. III – Improving concrete pavement performance. FHWA-
Roesler, J.R., 2006. Fatigue resistance of concrete pavements. In: RD-95-111, Washington, DC, 302 pp.
6th International DUT – Workshop on fundamental Yu, H.T., et al., 1998b. Analysis of concrete pavement responses
modeling of design and performance of concrete pavements, to temperature and wheel loads measured from instrumented
1465 15 – 16 September 2006, Belgium. slabs. Transportation Research Record: Journal of the 1520
Roesler, J., et al., 2004. Fracture of plain and fiber-reinforced Transportation Research Board, No. 1639, 94 – 101.
concrete slabs under monotonic loading. ASCE Journal of Zollinger, D.G. and Barenberg, E.J., 1989. Proposed mechanistic
Materials in Civil Engineering, 16 (5), 452– 460. based design procedure for jointed concrete pavements.
Roesler, J.R., et al., 2008. Design and concrete material Illinois Cooperative Highway Research Program – 518,
requirements for ultra-thin whitetopping. Final Report, University of Illinois, Urbana, Illinois.
1470 1525

1475 1530

1480 1535

1485 1540

View publication stats

You might also like