You are on page 1of 17

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

net/publication/318430967

Predicting thermal cracking of asphalt pavements from bitumen and mix


properties

Article  in  Road Materials and Pavement Design · July 2017


DOI: 10.1080/14680629.2017.1350598

CITATIONS READS
11 1,003

2 authors:

Bagdat Teltayev Boris Radovskiy


Kazakhstan Highway Research Institute 21 PUBLICATIONS   95 CITATIONS   
72 PUBLICATIONS   267 CITATIONS   
SEE PROFILE
SEE PROFILE

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

Development of a Road Asset Management System (RAMS) for Kazakhstan View project

All content following this page was uploaded by Boris Radovskiy on 29 November 2017.

The user has requested enhancement of the downloaded file.


Road Materials and Pavement Design, 2017
https://doi.org/10.1080/14680629.2017.1350598
1
2
3
4
5
6 Predicting thermal cracking of asphalt pavements from bitumen and mix
7 properties
8
9 Bagdat Teltayeva∗ and Boris Radovskiyb
10
a Kazakhstan Highway Research Institute, Almaty, Kazakhstan; b Radnat Consulting, Irvine, CA, USA
11
12
13 (Received 11 October 2016; accepted 24 June 2017 )
14
15 The objective of this study was to develop approximate relationships for prediction of
16 the rheological properties and tensile strength of asphalt concrete with conventional bitu-
17 men binder to address the low-temperature cracking analysis. The parameters of modified
Christensen–Anderson–Marasteanu model for relaxation modulus of bitumen were related
18 with its penetration index and softening point by approximate formulas based on Van Der
19 Poel’s nomograph for determining bitumen stiffness. Hot mix asphalt relaxation function has
20 been described by the Christensen–Bonaquist model. Hot mix asphalt tensile strength as a
21 function of loading time and temperature was related to the bitumen stiffness by equation
Coll:

22 based on Heukelom’s data and the Molenaar-Li empirical formula. The strength of hot mix
asphalt has been also studied by performing tensile tests at constant strain rate and temper-
23 atures. The results were analysed considering the calculated stress increase with temperature
24
CE: NK QA:

drop at various cooling rates. Critical temperature was estimated from comparison of thermal
25 stress and tensile strength of asphalt concrete at constant stress rate.
26 Keywords: asphalt pavement; thermal stress; low-temperature cracking; relaxation modulus;
27 tensile strength; critical temperature
28
29
30
31 1. Introduction
32 Low-temperature cracking is one of the major causes of failure of asphalt pavements in cold
33 weather climates, including much of the United States, Canada, Russia, Ukraine, Kazakhstan,
34 and other countries at extreme northern and southern latitudes. Thermal cracking has been asso-
35 ciated with the volumetric change in the asphalt concrete layer. When the temperature drops,
36 the pavement tends to contract its volume but the thermal strains are unrealised. That causes the
37 thermal stresses in pavement. They build up until they reach the strength of the material, leading
38 to the formation of cracks to relieve these stresses.
39 Monismith, Secor, and Secor (1965) developed a theoretical calculation method for the ther-
40 mally induced stress in asphalt pavement as in an infinite viscoelastic beam based on Humphreys
41 and Martin (1963) solution. Boltzmann’s superposition principle and hereditary integral form
42 of the constitutive equation for linear viscoelastic material was applied to relate time-dependent
43 stresses and strains. This method is currently used for the estimation of critical cracking tem-
44 perature by many researchers. Hills and Brien (1966) devised a simple pseudo-elastic method
45 to compute the thermal stresses in asphalt pavement. The predicted stress was dependent on
46 loading time equal to time interval of numerical integration over which the change of stress
47 was computed. Fromm and Phang (1971) eliminated the dependency on the time interval of
48
49 *Corresponding author. Email: bagdatbt@yahoo.com

© 2017 Informa UK Limited, trading as Taylor & Francis Group


2 B. Teltayev and B. Radovskiy

50 numerical integration by specifying loading times. However, the results of the stress compu-
51 tations were still directly influenced by the loading time specified. Christison, Murray, and
52 Anderson (1972) employed five different methods of stress computation, including the Moni-
53 smith’s and Hills-Brien’s methods, and concluded that a potential of low-temperature cracking
54 for a given pavement can be evaluated if the mix stiffness and strength characteristics are known
55 as a function of temperature and time of loading. Buttlar and Roque (1994) developed the indi-
56 rect tension test (IDT) to measure the creep compliance and tensile strength of asphalt mixtures
57 at low temperatures. The IDT testing of cores is very appropriate for the evaluation of existing
58 pavements. However, several differences between the axial tension and IDT tests raise ques-
59 tions about the interchangeability of the creep compliance and tensile strength values obtained
60 from the two test methods. Bouldin, Dongre, Rowe, Sharrock, and Anderson (2000) consid-
61 ered the thermally induced stress in asphalt binder as in viscoelastic bar and reported that the
62 midpoint of the binder’s glass transition is close to the pavement critical cracking temperature.
63 Based on the study of Bouldin et al. (2000), in AASHTO and ASTM specifications, MP1a-02
64 (AASHTO, 2002), PP 42-07 (AASHTO, 2007), and ASTM D 6816-02 (ASTM, 2002), asphalt
65 mixture thermal stresses were calculated from asphalt binder thermal stresses multiplied by an
66 empirical pavement constant equal to 18. This was done to obtain a critical cracking temperature
67 at which the mixture thermal stress curve and binder strength curve intersect (Moon, Marasteanu,
68 & Turos, 2013). First, the binder tensile strength at a constant strain rate of 3%/min cannot be
69 used as same as the mix tensile strength at different stress rates. Second, only one pavement
70 constant is not sufficient to convert the binder properties to the mix properties. This method
71 cannot reflect the difference in thermal stresses for mixtures prepared with the different aggre-
72 gate source and gradation and the asphalt mixture volumetric (i.e. air void, binder quantity, etc.).
73 For instance, Roy and Hesp (2001) for limited set of binders estimated the variation of pavement
74 “constant” between 3.4 and 16.7. Hu, Zhou, and Walubita (2009), Prieto-Muñoz, Yin, and Buttlar
75 (2013), and Dave, Buttlar, Leon, Behnia, and Paulino (2013) replaced the one-dimensional pave-
76 ment model with the two-dimensional stress analysis of low-temperature cracking implemented
77 within a viscoelastic finite element modelling framework. Farrar, Hajj, Planche, and Alavi (2013)
78 combined the thermal stress restrained and unrestrained specimen tests of asphalt mixture. The
79 measured restrained thermal stress development was very similar to the thermal stress build-up
80 calculated from the Boltzmann’s hereditary integral. Relaxation modulus of asphalt mixture was
81 determined by approximate conversion of complex modulus estimated from the model based on
82 the law of mixtures for composite materials and from the measured complex modulus of the
83 recovered asphalt binder (Farrar et al., 2013).
84 The low-temperature transverse cracking in asphalt pavements is the result of a combination
85 of two distress mechanisms: single-event thermal cracking and thermal fatigue. The single-
86 event thermal cracking as the significant contributor to transverse cracking of pavements was
87 considered in this study. To better understand the ways to reduce low-temperature cracking, a
88 mechanistic model is needed that includes the predicted rheological and fracture properties of
89 asphalt mixture.
90 The objective of our study was to develop the relationships for prediction of the viscoelastic
91 properties and tensile strength of asphalt concrete with conventional bitumen binder as a function
92 of time and temperature and apply them to the low-temperature cracking problem.
93
94
95 2. Stiffness and relaxation modulus of bitumen
96 Stiffness modulus S(t) introduced by Van der Poel (1954) is the ratio of the constant applied
97 uniaxial stress σ0 to the resulting uniaxial strain ε(t) at time t. Based on the test results of 47
98 bitumens, Van der Poel developed a nomograph to determine the stiffness of asphalt cement as
Road Materials and Pavement Design 3

99 a function of temperature T and loading time t (or frequency ω) if the softening temperature Ts
100 and the penetration index PI are known. To calculate binder stiffness, it is convenient to use the
101 BitProps program (free software) based on the scanned Van der Poel’s nomograph developed by
102 G. Rowe and M. Sharrock (Abatech, Inc.).
103 Several researchers proposed the empirical equations for stiffness modulus of bitumen depend-
104 ing on penetration and softening temperature based on Van der Poel’s data. Saal (1955) proposed
105 the relationship between the bitumen stiffness and its penetration for load duration t = 0.4 s and
106 PI = 0. Ullidz and Larsen (1984) proposed a simple equation for S(t) and limited its applica-
107 bility to the duration of loading 0.01 < t < 0.1, range of penetration index − 1 < PI < 1, and
108 temperature range 10°C < T < 70°C. Even in this narrow range of input parameters, the average
109 coefficient of variation of stiffness calculated using the proposed equation from Van der Poel’s
110 data is 40%. Shahin (1977) developed two regression equations based on Van der Poel’s nomo-
111 graph to relate stiffness modulus with time, temperature, penetration index, and softening point:
112 one of them for S < 1 MPa and another for S > 1 MPa. The range of their applicability in terms
113 of input parameters was not specified. Meanwhile, it is unknown what value of stiffness modulus
114 will be and which of these formulas should be used. Moreover, often for the same set of input
115 parameters, the first equation returns S < 1 MPa while the second gives S > 1 MPa and it is not
116 clear which of the results is correct. To our knowledge, the proposed formulas for the bitumen
117 stiffness based on Van der Poel’s data were mostly developed using purely regression techniques
118 without involving any rheological models.
119 In this study as a mathematical model for describing the stiffness of bitumen, Christensen and
120 Anderson (1992) expression was used:
121   β −1/β
122 Eg t
123 S(t) = Eg 1 + , (1)

124
125 where Eg is the uniaxial glassy modulus of binder (MPa), η is the steady-state viscosity (MPa·s),
126 and β is constant.
127 The instantaneous value for longitudinal modulus Eg was obtained by extrapolation of values
128 for stiffness modulus S(T, t) according to Van der Poel at low temperature T and small load
129 duration t for t → 0. With the purpose of extrapolation, we used the model developed to describe
130 the viscoelastic properties of amorphous glass forming polymers (Drozdov, 2001) based on the
131 theory of cooperative relaxation (Adam & Gibbs, 1965). For the asphalt with certain penetration
132 index, we selected the lowest temperature at which using the program BitProps the value of
133 stiffness S can be obtained for t = 0.00005 s. This lowest temperature varied from − 42° for
134 PI = + 2 to 9° for PI = − 3. The stiffness modulus S was obtained at that lowest temperature
135 for a set of eight loading times t = 0.00005, 0.0001, 0.0002, 0.0005, 0.001, 0.002, 0.005, and
136 0.01 s. Then using the Equations (18) and (37) of Drozdov (2001), the value of instantaneous
137 modulus was found by fitting the Van der Poel’s data for stiffness modulus at eight loading
138 times. It was concluded that for asphalts having the penetration index from − 2 to + 3, the
139 average value of instantaneous longitudinal modulus equals to E g = 2460 MPa with the standard
140 deviation of 7%.
141 After fitting Equation (1) to Van der Poel data, the following equations were obtained:
142
143 0.1794
β= , (2)
144 1 + 0.2084PI − 0.00524PI 2
145
146 η = aT Ahrr (T) · η(Tr ) (at T ≤ Ts − 10),
(3)
147 η = aT WLF (T) · η(Tr ) (at T > Ts − 10),
4 B. Teltayev and B. Radovskiy
    
148 12(20 − PI ) 0.2011
η(Tr ) = 0.00124 1 + 71 exp − · exp , (4)
149 5(10 + PI ) 0.11 + 0.0077PI
150
where Tr is reference temperature Tr = (Ts − 10) [°C], and aT (T) is time-temperature superpo-
151
sition function:
152
  
153 3(30 + PI ) 1 1
154 aT Ahrr (T) = exp 11720 · − , (5)
5(10 + PI ) (T + 273) (Ts + 263)
155
156  
2.303(T − Ts + 10)
157 aT WLF (T) = exp − . (6)
(0.11 + 0.0077PI )(114.5 + T − Ts )
158
159 For 1910 points in the range of PI from − 3 to + 2, (T − Ts ) from − 45°C to 10°C, and t from
160 10−4 to 104 s, the standard deviation of stiffness calculated using Equation (1) from the Van der
161 Poel’s data was 14.6% and the coefficient of determination R2 = 0.978.
162 By definition, stiffness modulus S(t) is the inverse of uniaxial creep compliance. A related
163 function is the shear creep compliance:
164    1/β
165 1 Gg t β
166 J (t) = 1+ , (7)
Gg η
167
168 where β is given by Equation (2) and Gg = the glassy modulus in shear; Gg ≈ Eg /3 = 820 MPa
169 if the material is considered isotropic with Poisson’s ratio ν = 0.5, although it should be noted
170 that ν is time dependent (Di Benedetto, Delaporte, & Sauzeat, 2007).
171 There are different methods for converting the shear creep compliance J (t) to relaxation modu-
172 lus G(t). In this study, the modified Hopkins–Hamming algorithm (Tschoegl, 1989) was selected,
173 and relaxation modulus G(t) was approximated by Christensen, Anderson and Marasteanu
174 (CAM) model (Marasteanu & Anderson, 1999):
175
  b −k/b
176 Gg t
177 G(t) = Gg 1 + , (8)
η
178
179
where b and k are constants. It can be shown that the parameter k is a function of b from the
180
following equation (Ferry, 1980):
181  ∞
182 G(t) dt = η, (9)
183 0
184 which can be treated as a definition of zero-shear viscosity η. After substituting Equations (8) to
185 (9) and integration, it is easy to find that exactly k = 1 + b. This leads to the following formula
186 for relaxation modulus at shear:
187    −(1+1/b)
188 Gg t b
189 G(t) = Gg 1 + , (10)
η
190
191 where
192  −1
1 ln(π )
193 b= + −2 . (11)
194 β ln(2)
195 The uniaxial relaxation modulus of binder can be found as Eb (t) ≈ 3G(t), where its shear
196 modulus G(t) is given by Equation (10).
Road Materials and Pavement Design 5

197 Complex modulus of bitumen in shear was determined from creep compliance Equation (7) at
198 a frequency equal to the inverse of the loading time using the Schwarzl and Struik (1968) method
199 in the following form:
200   β −1/β
201 Gg Gg
Gd (ω) = 1+ , δ(ω) = m(ω)π/2, (12)
202 (1 + m(ω)) ηω
203
204 where Gd (ω) is the norm of complex modulus, δ is phase angle, ω is frequency, (x) is the
205 gamma function, and
206 (Gg /ηω)β
207 m(ω) = . (13)
1 + (Gg /ηω)β
208
209 For example, comparison of experimental results (Christensen & Anderson, 1992) at
210 T = 25°C versus calculated from Equation (12) values of Gd (ω) and δ(ω) is shown in Figure 1
211 for SHRP core bitumen AAB-1 with pen 25 = 98 dmm, Ts = 47.8°C, PI = 0 (Mortazavi &
212 Moulthrop, 1993). Tests were performed on dynamic shear rheometer with a range in frequency
213 from 0.1 to 100 rad/s at temperatures − 35, − 25, − 15, − 5, 5, 15, 25, 35, 45, and 60°C. The
214 data at all temperatures were then shifted with respect to time to construct the master curve at
215 reference temperature 25°C (Christensen & Anderson, 1992). Notice that the agreement between
216 the calculated and experimental results for the norm of complex modulus and the phase angle is
217 good.
218 The relaxation spectrum is a useful, fundamental way of characterising the time-dependent
219 behaviour of asphalt binders (Christensen, Anderson, & Rowe, 2016). According to Alfrey’s
220 rule (Ferry, 1980), relaxation spectrum H (τ ) of binder is obtainable in first approximation as a
221 negative slope of the relaxation modulus. Differentiating Equation (10) with respect to ln t leads
222 to
     −(2+1/b)
223 Gg τ b Gg τ b
224 H (τ ) = Gg (1 + b) 1+ . (14)
η η
225
226
227
228

Colour online, B/W in print


229
230
231
232
233
234
235
236
237
238
239
240
241
242
243 Figure 1. An example of comparisons between the model predictions and experimental data for complex
244 modulus and phase angle as a function of frequency: points – measured (Christensen & Anderson, 1992),
245 lines – calculated from Equation (12).
6 B. Teltayev and B. Radovskiy

246
247
Colour online, B/W in print

248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263 Figure 2. The relaxation spectrum for bitumen AAB-1 at 25°C: dashed line – from experimental data
(Reproduced with permission from Christensen, 1992); solid line – calculated from Equation (14).
264
265
266 Relaxation spectrum calculated in Christensen (1992, Figure 3.4) for the same SHRP core
267 bitumen AAB-1 based on measurements of complex modulus is presented in Figure 2 together
268 with calculated from Equation (14). The agreement between them is good. From Equation (14),
269 it is easy to obtain the relaxation time corresponding to the maximum density of spectrum
270  1/b
η b
271 τmax = , (15)
272 Gg 1 + b
273 and its maximum density
274  
b + 1 (2b+1)/b
275 Hmax = Gg b . (16)
276 2b + 1
277 An acceptable estimate for the maximum density of relaxation spectrum of bitumen is Hmax =
278 Gg b/3. It follows that maximum density of spectrum is proportional to the parameter b of CAM
279 model and is independent on temperature of bitumen.
280 One can see that the important rheological properties of bitumen such as stiffness, relax-
281 ation modulus, complex modulus, and relaxation spectrum can be estimated from the pro-
282 posed equations using the simplest standard properties of bitumen: penetration and softening
283 temperature.
284
285
286 3. Stiffness and tensile strength of asphalt concrete
287 To analyse the temperature-induced stresses, the uniaxial relaxation modulus of asphalt concrete
288 Emix (t) is needed. Christensen and Bonaquist (2015) recently improved their model:
289
290 Emix (t) = Pc [Eagg · (1 − VMA) + Eb (t) · VFA · VMA], (17)
291
292 Pc = 0.006 + 0.994[1 + exp(−(0.663 + 0.586 · ln(VFA · Eb (t)/3) − 12.9VMA
293
294 − 0.17 · ln(εs · 106 ))]−1 , (18)
Road Materials and Pavement Design 7

295 where Eb (t) is relaxation modulus of binder (MPa), Eagg is elastic modulus of aggregate (MPa),
296 VMA are voids in mineral aggregate (volume fraction), VFA are voids filled with asphalt (vol-
297 ume fraction), εs = 0.0001 is the standard target strain, and Pc is the contact factor introduced in
298 Christensen and Bonaquist (2015) and defined by Equation (18). The same model can be used
299 for the stiffness modulus of asphalt concrete Smix (t) if Eb (t) is replaced by binder stiffness S(t)
300 defined according to Equation (1).
301 To estimate a critical temperature, the tensile strength of asphalt concrete as a function of
302 temperature is needed. W. Heukelom has presented compelling evidence that the tensile strength
303 of mix is related to the properties of bitumen contained herein (Heukelom, 1966). He gave some
304 data of tensile strength measurements of eight dense-graded mixes carried out at a variety of
305 temperatures and speeds. Heukelom has showed that the relative tensile strength, that is, the
306 tensile strength divided by its maximum value, can be represented by one curve for all mixes
307 tested as a function of the stiffness of bitumen recovered (Heukelom, 1966, Figure 22):
308
309 Tensile Strength
R(S) = Of Mix. (19)
310 Maximum Tensile Strength
311
312 The Heukelom curve can be approximated by equation
313
314 0.774 + 0.039r + 0.141r4.547
R= , (20)
315 1 + 0.026r3.608 · exp(1.245r)
316
317 where r = log(Eg /S), Eg is the uniaxial glassy modulus of binder (MPa), S is the binder stiffness.
318 Molenaar and Li (2014) proposed an empirical equation to estimate the maximal tensile
319 strength of asphalt concrete Ph as a function of mixture stiffness and the volumetric composition.
320 Six dense-graded mixes and one porous mixture with a bitumen of 10/20 to 70/100 penetra-
321 tion and aggregate of maximum grain size from 4 to 32 mm were tested at the Delft University
322 of Technology in a temperature range of 5–35°C at constant tensile strain rate from 0.0001 to
323 0.04/s. Regression analysis was performed to predict the maximum tensile strength of mixtures
324 depending on mixture stiffness Smix at 20°C and asphalt-void ratio (VFA). Molenaar and Li
325 (2014) developed the following regression equation:
326
327 Ph = 0.505 · Smix
0.308
· VFA0.849 . (21)
328
329 The combination of Equation (20) based on the Heukelom curve for relative strength of mix with
330 empirical Equation (21) for the maximal strength of mix (Molenaar & Li, 2014) leads to the
331 following expression for tensile strength f as a function of temperature T and the loading time t:
332
333 0.774 + 0.039r + 0.141r4.547
f = 0.505 · Smix
0.308
· VFA0.849 · . (22)
334 1 + 0.026r3.608 · exp(1.245r)
335
336 where S is the binder stiffness (MPa) defined in Equation (1) as a function of time and tempera-
337 ture, Smix is the mix stiffness at T = 20°C and at fixed loading time of 0.06 s, and f is the tensile
338 strength under rectangular pulse loading of duration t.
339 Equation (22) relates the tensile strength of asphalt concrete f with temperature T and time to
340 failure tf in terms of the binder creep stiffness S(tf ), as Heukelom (1966) suggested. At constant
341 strain rate Vε , the tensile strength fε in the left-hand side of Equation (19) can be expressed as a
342 product of time to failure, the strain rate, and the secant modulus at failure Hmix (tf ). The secant
343 modulus Hmix (t) is related to the relaxation modulus Emix (t) by the following equation (Smith,
8 B. Teltayev and B. Radovskiy

344 1976):

345 1 t
346 Hmix (t) = Emix (t − τ ) dτ . (23)
t 0
347
Thus using Equation (22), one can find the time to failure tf numerically as a root of the following
348
equation:
349
350 0.774 + 0.039r + 0.141r4.547
351 tf Vε Hmix (tf ) = 0.505 · Smix
0.308
· VFA0.849 · , (24)
1 + 0.026r3.608 · exp(1.245r)
352
353 where r = log(Eg /S(tf )).
354 Then, the tensile strength of asphalt concrete fε at constant strain rate Vε can be calculated as
355 fε = tf Vε Hmix (tf ). (25)
356
357 The calculated tensile strength at constant strain rate was compared with test results of direct
358 tensile measurements at Braunschweig Technical University (Stock & Arand, 1993). The main
359 features of testing machine developed at the University of Braunschweig are a very stiff frame,
360 strain transducers for the measurement of the length of the specimen, and a step motor to apply
361 strain to a specimen at a pre-determined rate. The samples were prismatic with a cross section
362 of 40 mm by 40 mm. The specimen length was 160 mm. Seven asphalt concrete mixtures with
363 the same content of different binders of 4.7% by weight of total mix were testedat constant strain
364 rate Vε = 1 · 10−4 /s (1 mm/min) and at temperatures 20, 5, − 10, and − 25°C. The test results
365 for mix with unmodified bitumen B1 having the penetration index PI = − 0.692 and softening
366 temperature Ts = 49°C for the mix volumetric properties VMA = 0.15 and VFA = 0.733 are
367 shown in Figure 3 together with tensile strength curve calculated from Equation (25). The cal-
368 culated maximal strength and the general shape of curve agree with measured strength although
369 the curve looks shifted along the temperature axis by approximately 4–5°C. Possible reasons of
370 shifting will be discussed later.
371 In the Kazakhstan Highway Research Institute, the uniaxial tensile tests were performed on
372 testing system TRAVIS (InfraTest GmbH), which is a version of device developed at the Uni-
373 versity of Braunschweig (Figure 4).This testing system includes a compact test frame integrated
374
375
376
Colour online, B/W in print

377
378
379
380
381
382
383
384
385
386
387
388
389
390
391 Figure 3. Measured tensile strength values at constant strain rate (Stock & Arand, 1993) and those
392 calculated from Equations (24) and (25).
Road Materials and Pavement Design 9

393
394

Colour online, B/W in print


395
396
397
398
399
400
401
402
403
404
405
406
407
Figure 4. Machine used for the mechanical testing in the Kazakhstan Highway Research Institute.
408
409
410 in the temperature chamber. The load is applied to the specimen via heavy-duty screw jack and
411 stepping motor. An electronic load cell is directly attached to the spindle. The equipment is con-
412 trolled by a PC connected to the motor and the transducers. PC is used for measurement data
413 logging, control of the test procedure, and the temperature test chamber.
414 Dense-graded asphalt mixture was prepared with the use of granite aggregate fractions of
415 5–10 mm (20%), 10–15 mm (13%), and 15–20 mm (10%) from the Novo-Alekseevsk rock pit
416 (Almaty region); sand of fraction 0–5 mm (50%) from the plant Asphaltconcrete-1 (Almaty city)
417 and activated filler (7%) from the Kordai rock pit (Zhambyl region). The air-blown bitumen was
418 produced by the Pavlodar petrochemical plant from the crude oil of Western Siberia (Russia).
419 After short-time ageing, the bitumen penetration was 70 dmm (at 25°C), softening temperature
420 Ts = 48°C, and PI = − 0.91.A mixture designated as K2 was prepared with 4.8% of bitumen
421 by weight of aggregates. Mixture was compacted with the Cooper roller compactor (model CRT-
422 RC2S) to the average void content of 3.6%. Rectangular specimens (50 × 50 × 160 mm3 ) were
423 then sawed from these slabs and glued with the epoxy resin to the mounts. Specimens were
424 tested with nominal deformation rate of 1 mm/min (Vε = 1 · 10−4 /s) . Tests were performed at
425 temperatures 20, 10, 0, − 10, − 20, and − 30°. The strength values were obtained from five
426 tensile tests for all temperatures (Figure 5). The average coefficient of variation was of 15.6%.
427 Figure 5 presents the comparison of the measured tensile strength of asphalt concrete samples
428 and the strength of asphalt concrete predicted form Equations (24) and (25) for Vε = 1 · 10−4 /s
429 and the mix volumetric properties VMA = 0.144, and VFA = 0.75 (red curve).The modulus of
430 aggregate of bulk specific gravity Gsb = 2.760 was estimated using the correlation recommended
431 in Christensen and Bonaquist (2015) as Eagg = 7650G1.59 sb = 36, 000MPa.
432 In Figure 5 as before in Figure 3, the calculated maximal strength and the shape of the predicted
433 curve (dashed red) agree with measured tensile strength but the predicted curve is constantly
434 shifted along the temperature axis to higher temperatures even more than in Figure 3. This shift
435 can be caused by various reasons. One of them might be the time-temperature superposition.
436 Obviously, the binder, regardless of the aggregates skeleton, drives the temperature dependency
437 of a bituminous material (Di Benedetto et al., 2011). In this paper, the time-temperature superpo-
438 sition function for the binder aT (T) in Equations (5) and (6) was determined from Van der Poel’s
439 data. Meanwhile, Anderson et al. (1990) found that Van der Poel’s nomograph underpredicted
440 the bitumen stiffness at long loading times and low temperatures. Although the time-temperature
441 sensitivity of bitumen at low temperatures is an important issue and deserves further research
10 B. Teltayev and B. Radovskiy

442
443
Colour online, B/W in print

444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463 Q3 Figure 5. Measured and calculated tensile strength values: dashed red curve – calculated from Equations
464 (24) and (25) at constant strain rate Vε = 1 · 10−4 /s, dotted blue curve – calculated from Equation (22) for
465 average time to failure tf = 40 s.
466
467
468 efforts, this reason does not explain the shift at temperatures T > 0°C on Figure 5 (the dashed
469 red curve).
470 Another possible reason of shift is the effect of machine compliance on specimen strain rates.
471 To facilitate data acquisition, the system TRAVIS was designed to extend the testing time at
472 cold temperatures. As a result, because of machine compliance, the on-specimen strain rate was
473 not constant, given that the cross-head displacement rate 1 mm/min remains constant throughout
474 the test, due to the high stiffness of the material at low temperatures. Actual time to failure was
475 between 40 and 50 s at temperatures 20, 10, 0, − 10–20°C and around 20–25 s at T = − 30°C.
476 To check the validity of this reason, we calculated the tensile strength of the same asphalt mix
477 K2 at loading time 40 s from Equation (22) for PI = − 0.91, Ts = 48°C, VMA = 0.144, and
478 VFA = 0.75 (dotted blue curve in Figure 5). Shift of calculated curve to cold temperatures closer
479 to measured strengths confirms the significant effect of machine compliance on test results. If a
480 true constant-rate-of-deformation test is to be performed on asphalt concrete, feedback to loading
481 unite must be from on-specimen LVDTs and not from the cross-head displacement as was the
482 case in this experiment. This constant-strain loading requirement is difficult to achieve. More-
483 over, although used by some researchers (Bouldin et al., 2000; Christison et al., 1972; Stock
484 & Arand, 1993), the advantages of constant strain-rate test are doubtful of its value for low-
485 temperature cracking problem because the longitudinal strain in asphalt pavement induced by
486 cooling is unrealised, that is, equals to zero up to appearance of the transverse crack. For the
487 single-event low-temperature cracking analysis, the stress-controlled strength test looks more
488 important than strain-controlled test.
489 At constant stress rate Vσ , time to failure equals the tensile strength fσ over the stress rate
490 tf = fσ /Vσ . In that case, the binder stiffness in the right-hand side of Equation (22) depends on
Road Materials and Pavement Design 11

491
492

Colour online, B/W in print


493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
Figure 6. Tensile strength of asphalt concrete as a function of temperature calculated from Equation (26)
509 at different stress rates.
510
511
512 tensile strength and the stress rate. The tensile strength fσ at constant stress rate Vσ can be found
513 numerically from Equation (26) as a function of T and Vσ :
514
515 0.774 + 0.039r + 0.141r4.547
fσ = 0.505 · Smix
0.308
· VFA0.849 · , (26)
516 1 + 0.026r3.608 · exp(1.245r)
517
where r = log(Eg /S(tf = fσ /Vσ )).
518
Figure 6 depicts the tensile strength calculated from Equation (26) for the same as before
519
asphalt mix K2 at different constant stress rates.
520
To illustrate, the mix strength at T = − 2°C and Vσ = 0.05MPa/s equals to 3.1 MPa
521
(Figure 6). In that case, time to failure is 62 s; stiffness of binder at T = 20°C and t = 0.06 s
522
according to Equation (1) is S = 6.02 MPa; stiffness of mix at T = 20°C and t = 0.06 s
523
from Equation (17) equals to Smix = 6040MPa. Parameter r is equal to r = log(2460/S(T =
524
−2◦ C, t = 62 s)) = 2.611, and the right-hand side of Equation (26) returns the tensile strength
525
of 3.1 MPa. A 10-fold increase in stress rate shifts the strength vs. temperature curve to higher
526
temperatures as much as around 7°C (Figure 6).
527
528
529
530 4. Thermal stresses and critical temperature
531 Predicted relaxation modulus Emix and the stress rate dependent tensile strength fσ were used
532 to estimate the critical single event cracking temperature. In this study, the critical cracking
533 temperature (Tcr ) is defined as the temperature at which the asphalt concrete tensile strength at
534 constant stress rate crosses the thermal stress curve.
535 Low-temperature-induced stresses in asphalt pavement were calculated as in an infinite
536 viscoelastic beam resting on frictionless rigid foundation:
537  t  
538 d
σ (t) = − Emix (ξ(t) − ξ(τ )) αmix (T(τ ))T(τ ) dτ . (27)
539 0 dτ
12 B. Teltayev and B. Radovskiy

540 where t is the present time (s), τ is the passed time (s), T(τ ) is the temperature variation with time
541 (°C), αmix is the asphalt mixture coefficient of thermal t contraction (/°C), Emix is the relaxation
542 modulus (MPa) and ξ(t) is the reduced time: ξ(t) = 0 dτ/aT (T(τ )).
543 Coefficient of thermal contraction was assumed constant αmix = 2.5 · 10−5 /◦ C. The proper-
544 ties of binder and mix were taken as before: Ts = 48°C, PI = − 0.91, E agg = 36,000 MPa,
545 VMA = 0.144, VFA = 0.75. The relaxation modulus of asphalt concrete was determined from
546 Christensen–Bonaquist model – Equation (17).
547 A sinusoidal variation between the maximum and minimum temperatures during a day was
548 assumed in the first example (Figure 7).The temperature drops from 5°C to − 35°C and returns
549 to 5°C over 24-hour period. Thermal stress build-up was calculated according to Equation (27).
550 A tensile stress rate was estimated as Vσ (t) = dσ (t)/dt and it varies from zero to 3.8·10−4 MPa/s
551 with the average of about 2·10−4 MPa/s. Tensile strength of asphalt concrete fσ at Vσ = 2 ×
552 10−4 MPa/s according to Equation (26) as a function of T is shown on the same graph. The
553 critical cracking temperature at which the thermal stress curve and asphalt concrete strength
554 curve intersect is Tcr = − 33°C.
555 In the second example, the pavement temperature drops from 5°C to − 35°C at different con-
556 stant cooling rates. As expected, the stress accumulation rate was found dependent on cooling
557 rate (Figure 8). Cooling rate increase greatly increases the thermal stress rate. At the cooling
558 rate of 2°C/h that is similar to the rates experienced by real pavements, the loading rate at the
559 temperature close to − 30°C is around Vσ (t) = 2 · 10−4 MPa/s (Figure 8). Figure 9 is a plot of
560 stress according to Equation (27) at cooling rate of 2°C/h and of asphalt concrete strength ver-
561 sus temperature from Equation (26) at Vσ (t) = 2 · 10−4 MPa/s. The critical temperature is about
562 Tcr = − 34.5°C.
563 Low-temperature cracking of asphalt pavements is a widespread and costly problem in cold
564 regions. Ability to predict the temperature-induced stress in asphalt pavement and asphalt con-
565 crete strength starting from the properties of binder and mix should be very helpful for designing
566 mixes that are more crack resistant.
567
568
569
570
Colour online, B/W in print

571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587 Figure 7. Thermal stress and estimated fracture temperature for sinusoidal temperature variation during
588 a day.
Road Materials and Pavement Design 13

589
590

Colour online, B/W in print


591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609 Figure 8. Thermal stress rate at different cooling rates.
610
611
612

Colour online, B/W in print


613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631 Figure 9. Thermal stress and estimated fracture temperature at constant-rate cooling.
632
633 5. Conclusions
634 The following conclusions can be drawn based on the analysis performed in this study:
635
636 • The important rheological properties of bitumen, such as the creep compliance (Equation
637 (7)), relaxation modulus (Equation (10)), complex modulus (Equation (12)), and relaxation
14 B. Teltayev and B. Radovskiy

638 spectrum (Equation (14)), can be related with its simplest properties such as penetration
639 and softening temperature based on experimental data and rheological models (CA model
640 in this study) using the interrelationships of linear viscoelasticity rather than applying the
641 purely regression techniques. It would be of interest to enlist a more comprehensive rhe-
642 ological model such as the 2S2P1D model (Di Benedetto, Olard, Sauzeat, & Delaporte,
643 2004), for modelling the viscoelastic properties of both bitumen and mixture.
644 • The direct tensile strength of asphalt mixture as a function of temperature can be esti-
645 mated depending on properties of binder and mix volumetric composition from empirical
646 equations (22), (25), and (26). Evaluation of asphalt concrete strength needs a more
647 comprehensive mechanically based approach to address a ductile-to-brittle transition
648 at low temperatures, perhaps based on new theory of materials failure (Christensen,
649 2013).
650 • The feasibility of constant strain-rate test of asphalt mixture is doubtful of its value for low-
651 temperature cracking problem because the longitudinal strain in asphalt pavement induced
652 by cooling is unrealised up to appearance of the transverse crack. Moreover, the constant-
653 strain loading requirement at various temperatures is difficult to achieve due to the testing
654 machine compliance. For low-temperature cracking analysis, the strength test at constant
655 rate of stress loading looks more valuable.
656 • The proposed method to determine the critical single-event cracking temperature as the
657 temperature at which the thermal stress curve intersect the constant-rate strength curve
658 leads to reasonable results.
659
660 Additional research is needed to verify the proposed method to calculate the thermal stress
661 build-up from a single cooling event, to estimate the critical temperature and to extend the
662 application of the proposed method.
663
664
665 Disclosure statement
Q4 666 No potential conflict of interest was reported by the authors.
667
668
Funding
669
Q1
670 This work was supported by Road Committee of Ministry for Investments and Development of the Republic
Q2 of Kazakhstan (Contract No. 36 dated 21.07.2016).
671
672
673 References
674 AASHTO. (2002). Standard specification for performance graded(PG) asphalt binder. AASHTO Provisional
Q5 675 Standards, MP1a-02, Washington, DC.
676 AASHTO. (2007). Determination of low-temperature performance grade (PG) of asphalt binders. AASHTO
677 Provisional Standards, PP 42-07. Washington, DC.
678 Adam, G., & Gibbs, J. H. (1965). On the temperature dependence of cooperative relaxation properties in
glass-forming liquids. The Journal of Chemical Physics, 43, 139–146.
679 Anderson, D., Christensen, D., Dongre, R., Sharma, M., Runt, J., & Jordhal, P. (1990). Asphalt behaviour at
680 low service temperatures. Federal Highway Administration-RD-88-078, U.S. Department Transporta-
681 tion, Washington, D.C., 337p.
682 ASTM. (2002). Standard practice for determining low-temperature performance grade (PG) of asphalt
683 binders D 6816-02. Annual book of ASTM standards, road and paving materials: Vehicle-pavement
systems, 04-03. West Conshohocken, Pennsylvania.
684 Bouldin, M., Dongre, R., Rowe, G., Sharrock, M. J., & Anderson, D. (2000). Predicting thermal cracking
685 of pavements from binder properties-theoretical basis and field validation. Journal of the Association
686 of Asphalt Paving Technologists, 69, 455–488.
Road Materials and Pavement Design 15

687 Buttlar, W. G., & Roque, R. (1994). Development and evaluation of the strategic highway research pro-
688 gram measurement and analysis system for indirect tensile testing at low temperatures. Transportation
689 Research Record, 1454, 163–171.
Christensen, D. W. (1992). Mathematical modelling of the linear viscoelastic behaviour of asphalt cements
690 (thesis). The Pennsylvania State University.
691 Christensen, R. M. (2013). The theory of materials failure. Oxford: Oxford University Press.
692 Christensen, D. W., & Anderson, D. A. (1992). Interpretation of dynamic mechanical test data for paving
693 grade asphalt cements. Journal of the Association of Asphalt Paving Technologists, 61, 67–116.
694 Christensen, D. W., Anderson, D. A., & Rowe, G. W. (2016). Relaxation spectra of asphalt binders and the
Christensen-Anderson rheological model. Journal of the Association of Asphalt Paving Technologists,
695 Meeting Pre-prints, 546–577. Q6
696 Christensen, D. W., & Bonaquist, R. (2015). Improved Hirsch Model for estimating the modulus of hot mix
697 asphalt. Journal of the Association of Asphalt Paving Technologists, 84, 527–557.
698 Christison, J. T., Murray, D. W., & Anderson, K. O. (1972). Stress prediction and low temperature frac-
699 ture susceptibility of asphaltic concrete pavements. Journal of the Association of Asphalt Paving
Technologists, 41, 494–523.
700 Dave, E., Buttlar, W. G., Leon, S. E., Behnia, B., & Paulino, G. H. (2013). IlliTC – low-temperature cracking
701 model for asphalt pavements. Road Materials and Pavement Design, 14(S2), 57–78.
702 Di Benedetto, H., Delaporte, B., & Sauzeat, C. (2007). Three-dimensional linear behavior of bituminous
703 materials: Experiments and modeling. International Journal of Geomechanics, 7, 149–157.
704 Di Benedetto, H., Olard, F., Sauzeat, C., & Delaporte, B. (2004). Linear viscoelastic behavior of bituminous
materials: From binders to mixes. Road Materials and Pavement Design, 5, 163–202.
705 Di Benedetto, H., Sauzeat, C., Bilodeau, K., Buannic, M., Mangiafico, S., Nguyen, Q. T., . . . Van Rompu,
706 J. (2011). General overview of the time-temperature superposition principle validity for materials
707 containing bituminous binder. International Journal of Roads and Airports, 35–52. Q7
708 Drozdov, A. D. (2001). A model for the viscoelastic and viscoplastic responses of glassy polymers.
709 International Journal of Solids and Structures, 38, 8285–8304.
Farrar, M. J., Hajj, E. Y., Planche, J. P., & Alavi, M. Z. (2013). A method to estimate the thermal stress
710 build-up in an asphalt mixture from a single-cooling event. Road Materials and Pavement Design,
711 14(1), 201–211.
712 Ferry, J. D. (1980). Viscoelastic properties of polymers (3rd ed.). New York, NY: Wiley.
713 Fromm, H. J., & Phang, W. A. (1971). Temperature susceptibility control in asphalt cement specifications.
714 Highway Research Record, 350, 30–35.
Haas, R. C. G., & Phang, W. A. (1988). Relationships between mix characteristics and low-temperature
715 pavement cracking. Journal of the Association of Asphalt Paving Technologists, 57, 290–319. Q8
716 Heukelom, W. (1966). Observations on the rheology and fracture of Bitumens and asphalt mixes. Journal
717 of the Association of Asphalt Paving Technologists, 35, 358–399.
718 Hills, J. F., & Brien, D. (1966). The fracture of Bitumens and asphalt mixes by temperature induced stresses.
719 Prepared discussion. Journal of the Association of Asphalt Paving Technologists, 35, 292–309.
Hu, S., Zhou, F., & Walubita, L. F. (2009). Development of a viscoelastic finite element tool for asphalt
720 pavement low temperature cracking analysis. Road Materials and Pavement Design, 10(4), 833–858.
721 Humphreys, J. S., & Martin, C. J. (1963). Determination of transient thermal stresses in a slab with
722 temperature dependent viscoelastic properties. Transactions of the Society of Rheology, 7, 155–170.
723 Marasteanu, M. O., & Anderson, D. A. (1999). Improved model for bitumen rheological characteriza-
724 tion. Paper presented at the Eurobitumen workshop on performance related properties for Bituminous
Binders, Luxembourg.
725 Molenaar, A. A. A., & Li, N. (2014). Prediction of compressive and tensile strength of asphalt concrete.
726 International Journal of Pavement Research and Technology, 7, 324–331.
727 Monismith, C. L., Secor, G. A., & Secor, K. E. (1965). Temperature induced stresses and deformations in
728 asphalt concrete. Journal of the Association of Asphalt Paving Technologists, 34, 248–285.
729 Moon, K. H., Marasteanu, M. O., & Turos, M. (2013). Comparison of thermal stresses calculated from
asphalt binder and asphalt mixture creep tests. Journal of Materials in Civil Engineering, 25(8), 1059–
730 1067.
731 Mortazavi, M., & Moulthrop, J. S. (1993). The SHRP Materials Reference Library. SHRP-A-646.
732 Washington, DC: The Transportation Research Board.
733 Prieto-Muñoz, P. A., Yin, H. M., & Buttlar, W. G. (2013). Two-dimensional stress analysis of low-
734 temperature cracking in asphalt overlay/substrate systems. Journal of Materials in Civil Engineering,
25(9), 1228–1238.
735
16 B. Teltayev and B. Radovskiy

736 Roy, S. D., & Hesp, S. A. M. (2001). Fracture energy and critical crack tip opening displacement: Fracture
737 mechanics-based failure criteria for low-temperature grading of asphalt binders. Paper presented at the
Q9
738 proceedings of Canadian Technical Asphalt Association (pp. 185–212).
Saal, R. N. J. (1955). Mechanical testing of asphaltic bitumen. Paper presented at the Fourth World
739 Petroleum Congress, Rome, Section VI/A, Paper 3, 1–17.
740 Schwarzl, F. R. L., & Struik, C. E. (1968). Analysis of relaxation measurements. Advances in Molecular
741 Relaxation Processes, 1, 201–255.
742 Shahin, M. Y. (1977). Design system for minimizing asphalt concrete thermal cracking. Paper presented at
743 the proceedings of fourth international conference on the structural design of asphalt pavements, Ann
Arbor, University of Michigan (pp. 920–932).
744 Smith, T. L. (1976). Linear viscoelastic response to a deformation at constant rate: Derivation of physical
745 properties of a densely cross linked elastomer. Transactions of the Society of Rheology, 20(1), 103–117.
746 Stock, A. F., & Arand, W. (1993). Low temperature cracking in polymer modified binder. Journal of the
747 Association of Asphalt Paving Technologists, 62, 23–53.
748 Tschoegl, N. W. (1989). The phenomenological theory of linear viscoelastic behaviour. Heidelberg:
Springer-Verlag.
749 Ullidz, P., & Larsen, B. K. (1984). Mathematical model for predicting pavement performance. Transporta-
750 tion Research Record 949, TRB, 45-54.
751 Van der Poel, C. (1954). A general system describing the visco-elastic properties of bitumens and its relation
752 to routine test data. Journal of Applied Chemistry, 4, 221–236.
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784

View publication stats

You might also like