Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Modeling Effect of Geocomposite Drainage Layers

on Moisture Distribution and Plastic Deformation


of Road Sections
M. Bahador1; T. M. Evans, A.M.ASCE2; and M. A. Gabr, F.ASCE3

Abstract: The effect of geosynthetic layers on moisture distribution and plastic deformation of paved and unpaved road sections is studied
Downloaded from ascelibrary.org by University of Leeds on 08/15/13. Copyright ASCE. For personal use only; all rights reserved.

using numerical simulations. The geosynthetic layers consisted of, from top to bottom, a transport layer, a geonet, and a nonwoven geotextile
(referred to as a geocomposite capillary barrier drain by previous researchers). Two geotextile types were modeled as the transport layer: woven
fiberglass and nonwoven polypropylene. The numerical models were verified against published results obtained from a soil-geotextile column.
Inclusion of the geosynthetic layers at the interface of the aggregate base course (ABC) and subgrade increased suction in the subgrade and
decreased it in the ABC during a simulated rainfall event. The woven fiberglass geotextile led to higher suctions in the ABC compared with
the nonwoven polypropylene geotextile. The geosynthetic layers decreased the plastic deformation in both paved and unpaved road sections
through combined mechanistic and hydraulic actions. Increasing the thickness of the asphalt and ABC layers decreased the reinforcement effect
of the geotextile while increasing its beneficial hydraulic effect in term of the suction level. In sections with a thinner asphalt layer, the woven
fiberglass, functioning as a transport layer, decreased the plastic deformation of the profile by up to 20% compared with the profile with the-
nonwoven polypropylene geotextile. Increasing the thickness of the asphalt layer, however, reduced this difference to approximately 4%. In
unpaved sections, the inclusion of the woven fiberglass layer decreased the plastic deformation by approximately 24% more than the
profile with nonwoven polypropylene geotextrile, regardless of the aggregate base course thickness used in the analysis. DOI: 10.1061/
(ASCE)GT.1943-5606.0000877. © 2013 American Society of Civil Engineers.
CE Database subject headings: Deformation; Moisture; Pavements; Reinforcement; Subgrades; Suction; Drainage.
Author keywords: Deformation; Geocomposite; Moisture; Pavement; Reinforcement; Subgrade; Suction.

Introduction pavement layers (Heckel 1997). To design an appropriate drain-


age system, water flow under unsaturated conditions and the un-
Moisture is a main cause of deterioration in pavement sections saturated properties of the geomaterials should be considered. An
(Cedergren 1994). Excess moisture in pavement sections can de- interlayer drainage system has been proposed to mitigate moisture
crease the pavement life by more than half (Christopher and changes of soils within pavement sections considering flow under
McGuffey 1997). Damage caused by moisture includes breaking the unsaturated conditions (Henry 1988; Stormont and Zhou 2001;
cementation bond in the stabilized aggregate base course (ABC), Stormont et al. 2009).
separating asphalt from aggregates (stripping), water bleeding and Currently, drainage configuration in pavement sections is based
pumping, shrinking, swelling, and frost heave of the subgrade layer. on using geosynthetic layers at the interface of the ABC and sub-
Moisture can enter the pavement section through rainfall or by upward grade and considering unsaturated conditions [Christopher et al.
water flow caused by capillary forces. Conventional drainage sys- 2000; Henry et al. 2002; Henry and Stormont 2002; Stormont et al.
tems consist of a permeable aggregate base course under the asphalt 2009; K. S. Henry and J. C. Stormont, “Geocomposite capillary
layer and an edge drainage to divert the water toward drainage pipes. barriers drain,” U.S. Patent No. 6,162,653 (2000)]. In this case, the
This drainage system, however, assumes saturated flow and is not moisture characteristic curve (MCC) and in-plane transmissivity of
always effective in maintaining the moisture content of the unbound both geomaterials and geosynthetic layers are considered in the
drainage design. The MCC represents the relationship between the
1
Ph.D. Alumnus, Dept. of Civil, Construction, and Environmental water content of a porous medium and the potential energy of
Engineering, North Carolina State Univ., Raleigh, NC 27695. E-mail: the pore water. The MCC can be measured experimentally for soils
mahdi.bahador@gmail.com (ASTM D6836-02; ASTM 2008a) and geotextiles (Stormont et al.
2
Associate Professor, School of Civil and Construction Engineering, 1997; Lafleur et al. 2000; Knight and Kotha 2001; Stormont et al.
Oregon State Univ., Corvallis, OR 97331. E-mail: matt.evans@oregonstate 2001b; Kuhn et al. 2005; Garcia 2007; Nahlawi et al. 2007). The
.edu MCCs of geotextiles are steeper than most geomaterials, and they
3
Alumni Distinguished Professor, Dept. of Civil, Construction, and have lower air entry values (AEVs) and water entry values (WEVs)
Environmental Engineering, North Carolina State Univ., Raleigh, NC compared with fine-grained soils (Stormont et al. 1997; Stormont
27695 (corresponding author). E-mail: gabr@eos.ncsu.edu
et al. 2001b; Iryo and Rowe 2003). These properties allow geo-
Note. This manuscript was submitted on May 11, 2012; approved on
December 6, 2012; published online on December 10, 2012. Discussion textiles to function as drainage/moisture barrier layers when in
period open until February 1, 2014; separate discussions must be submitted contact with partially saturated soils in pavement sections (Henry
for individual papers. This paper is part of the Journal of Geotechnical 1988; Henry and Holtz 1997; Stockton 2001; Krisdani et al. 2006;
and Geoenvironmental Engineering, Vol. 139, No. 9, September 1, Stormont et al. 2009). Henry and Stormont proposed a system
2013. ©ASCE, ISSN 1090-0241/2013/9-1407–1418/$25.00. termed the geocomposite capillary barrier drain (GCBD) that

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / SEPTEMBER 2013 / 1407

J. Geotech. Geoenviron. Eng. 2013.139:1407-1418.


consisted of a geonet as a capillary barrier sandwiched between two considered for the transport layer: nonwoven polypropylene (NWP)
geotextiles [K. S. Henry and J. C. Stormont, “Geocomposite cap- and woven fiberglass (WF). The moisture distribution in the
illary barriers drain,” U.S. Patent No. 6,162,653 (2000)]. The top pavement section, with various ABC thicknesses, is studied during
geotextile functioned as a transport layer. A fiberglass geotextile that a simulated rainfall event using SIGMA/W. The computed moisture
has a higher AEV and becomes transmissive at higher suctions, distributions are then used in FLAC to investigate the effectiveness
relative to other types of polymeric geotextiles, was recommended of the drainage system, as a function of the asphalt and ABC layer
for use as the transport layer in the GCBD (Stormont and Ramos thicknesses, on reducing the total plastic deformation of the section.
2004). The fiberglass layer diverts water from the ABC toward The reinforcement versus hydraulic effect of the NWP and WF
the edge drain, and the geonet functions as a capillary barrier for geotextiles is separated, and their impact with varying the asphalt
downward water flow into the subgrade under unsaturated conditions. and ABC thicknesses is studied. Given the complexity of the
Stormont and Zhou (2001) modeled a pavement section with a GCBD modeled sections, the results are intended to provide insight into the
at the interface of the ABC and subgrade layers using VS2DHI (Hsieh partitioning of mechanical and hydraulic effects in geosynthetics-
et al. 2000) and showed that the GCBD prevented downward water reinforced pavement profiles and to identify trends in system re-
flow into the subgrade during the assumed rainfall event. sponse based on varying material and section parameters (rather
Downloaded from ascelibrary.org by University of Leeds on 08/15/13. Copyright ASCE. For personal use only; all rights reserved.

In general, an increasing moisture content level decreases the than to provide an exact prediction for seepage and deformation
shear strength of geomaterials (e.g., Fredlund and Rahardjo 1993; Lu magnitudes for a given configuration).
and Likos 2004), and such a reduction in strength should be con-
sidered in pavement design. The Mechanistic-Empirical Pavement
Design Guide (MEPDG) (Applied Research Associates–Trans- Effect of Suction on Soil Strength
portation 2009) uses the resilient modulus of the subgrade as one of
the fundamental parameters for pavement design. It has been In the MEPDG, the resilient modulus of a geomaterial is a function of
reported that increasing water content results in a considerable re- stress state and degree of saturation [National Cooperative Highway
duction in the resilient modulus of the ABC and subgrade soils Research Program (NCHRP) 2000]. The subgrade layer herein was
(Sweere 1990; Dawson et al. 1996; Drumm et al. 1997; Gehling et al. modeled using the Cam-Clay constitutive model because it can
1998; Witczak et al. 2000; Kolisoja et al. 2002). Rada and Witczak capture the effect of soil moisture content and stress state on soil
(1981) concluded that the resilient modulus of granular materials strength. In the Cam-Clay model, the bulk modulus is expressed as a
under saturated conditions can be decreased to one-third of the value function of the mean effective stress through the following equation:
measured at moisture levels below saturation. Tian et al. (1998)
n × p9
measured the resilient modulus of a coarse granular material under K¼ (1)
dry of optimum and wet of optimum water contents and showed that k
the specimens compacted wet of optimum had a 20% lower resilient
modulus compared with that at the optimum water content. Thus, to where K 5 bulk modulus; n 5 specific volume defined as the ratio
maintain the strength of geomaterials in a pavement section, the of total volume to volume of solids; p9 5 mean effective stress
drainage system should function to not only prevent the de- (s19 1 s29 1 s39=3); and k 5 slope of the overconsolidation line in
velopment of positive pore pressure but also to mitigate moisture v 2 Inðp9Þ space.
increases in unsaturated ABC and subgrade layers. Conversely, the The effective stress is calculated as follows:
hydromechanical effect of geosynthetic drainage layers on plastic
deformation in ABC and subgrade layers has not been extensively s9 ¼ ðs 2 ua Þ þ S × ðua 2 uw Þ (2)
studied in the past.
The current design method for reinforced unpaved roads was where s9 5 effective stress; S 5 degree of saturation; ua 5 pore air
proposed by Giroud and Han (2004). In their method, the required pressure; and uw 5 pore water pressure. The degree of saturation
base course thickness of a reinforced unpaved road is calculated may be calculated at any suction using the van Genuchten equation
based on wheel load, number of load cycles, subgrade and base (van Genuchten 1980)
course CBR, and the reinforcing layer properties. The influence of
1 2 Sr
reinforcing layer is accounted for through considering a higher SðhÞ ¼ Sr þ  121=n (3)
value of the bearing capacity factor (Nc 5 5:14 for geotextile and 1 þ ðahÞn
Nc 5 5:71 for geogrid-reinforced unpaved roads) and the effect of
the aperture stability modulus (ASM) on the stress distribution where S 5 degree of saturation; h 5 matric suction; Sr 5 residual
through the reinforced section. Simac et al. (2006) compared the water saturation; and a and n 5 fitting parameters.
required aggregate base course thickness for an unpaved reinforced Accordingly, in the Cam-Clay model, the effects of stress state
road using the methods of Giroud and Han (2004), Giroud and and suction on soil strength are considered using Eqs. (1)–(3).
Noiray (1981), and Bender and Barenberg (1978). For the last two Cam-Clay model parameters are scarcely reported in the litera-
methods, the average tensile strength at 5% strain was used instead of ture for coarse-grained geomaterials because of the high required
ASM as the performance property of the reinforcement layer. It was isotropic pressures in the tests (Atkinson and Bransby 1978). In this
noted that the geotextile reduced the required aggregate base course paper, the ABC was modeled as a Mohr-Coulomb material with
thickness by 34, 24, and 46% after 1,000 cycles of 80-kN axle load suction and stress state effects on strength. Mohr-Coulomb can be
using the aforementioned methods, respectively (Simac et al. 2006). used as a failure criterion in a linearly elastic perfectly plastic model
In this paper, paved and unpaved road sections with various as- in which the elastic strain is computed using Hooks law and plastic
phalt and ABC thicknesses are simulated using SIGMA/W (Krahn strain is computed using the failure envelope and associated flow
2004) and FLAC (Itasca Consulting Group 2008) to study the effect rule. In such a model configuration, decreasing the elastic modulus
of a three-layer drainage system on the hydraulic and mechanical results in reducing the slope of the linear elastic portion of the curve
response of the section layers. The drainage system consists of and increasing the elastic deformations before reaching the yield
a geonet sandwiched between a nonwoven geotextile at the bottom surface. Accordingly, with a higher modulus value, strain energy
and a transport geotextile at the top. Two geotextile types are decreases in the plastic region resulting in a smaller plastic

1408 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / SEPTEMBER 2013

J. Geotech. Geoenviron. Eng. 2013.139:1407-1418.


deformation and therefore a misrepresentation of the effect of The elastic deformation in each element below the centerline of
moisture content on soil stiffness and strength. Fig. 1 shows the the load was calculated by multiplying the thickness of each element
unloading-reloading stress-deformation curves for a numerically by its total elastic strain. The total elastic deformation in the ABC
modeled displacement-controlled triaxial test on a 20-cm-tall was computed by summing the calculated elastic deformations in
specimen with 5-cm radius and three different elastic moduli of each of these elements. The plastic deformation in the ABC was then
10.0, 9.0, and 8.0 MPa. The specimen was first isotropically confined calculated by subtracting the elastic deformation from the total de-
under a cell pressure of 0.1 MPa, and then a vertical displacement of formation of the ABC as computed from FLAC.
5 mm was specified at the top of the specimen. As is evident in Fig. 1, In the Cam-Clay model, bulk modulus is a function of the mean
decreasing the elastic modulus of the specimen decreases the slope effective stress and changes at each step. The total and plastic vol-
of the linear elastic portion of the curve, and consequently, increases ume change is computed in FLAC at each step. To compute the
the elastic deformation and decreases the plastic deformation. To plastic deformations in the subgrade, the elastic volumetric strain in
address this issue, the elastic modulus of the ABC material is not each element was calculated by subtracting the plastic volumetric
changed during the stress-deformation analysis herein. Instead, strain from the total volumetric strain, and the vertical elastic strain
suction is included in the effective principal stress calculations per increment was calculated as follows in each step (based on the
Downloaded from ascelibrary.org by University of Leeds on 08/15/13. Copyright ASCE. For personal use only; all rights reserved.

Eq. (2), which leads to displacing the Mohr’s circles relative to the elasticity theory and axisymmetric loading condition):
origin and the failure envelope and, consequently, accounts for the   
effects of suction and stress state on soil strength. 2
Ds1 2 Dɛ ev × Ki 2 G
3
Dɛ 1 ¼ (5)
2G
Modeling Plastic Deformations
where Dɛ 1 5 vertical elastic strain increment; Ds1 5 vertical stress
To calculate the plastic deformation in MEPDG, elastic strains are
increment; Dɛ ev 5 elastic volumetric strain increment; Ki 5 bulk
first calculated using a linear elastic analysis and then plastic strains
modulus at each step; and G 5 shear modulus.
are calculated as a function of the calculated elastic strain, water
The same approach of subtracting the total elastic deformation
content, and number of load cycles. In the numerical model, the
from the total deformation and used for plastic deformation calcu-
vertical elastic strain increments at each step in all the elements
lations in the ABC layer was used to estimate the plastic deformation
below the centerline of the loaded area were calculated as follows:
in the subgrade layer.
    To verify the formulation described previously, results from
De1 ¼ Ds1 × 1 þ 1 þ ðDs2 þ Ds3 Þ × 1 2 1 (4) a simulated displacement-controlled triaxial test were used. The
3G 9K 9K 6G triaxial test specimen was modeled first as crushed stone using an
elastic-perfectly plastic model with a Mohr-Coulomb failure crite-
where De1 5 vertical elastic strain increment; Ds1 , Ds2 , and Ds3 rion and then as silty sand using Cam-Clay model, and the plastic
5 stress increments; G 5 shear modulus; and K 5 bulk modulus. deformation in each case was calculated using the unloading-
reloading curves. Then, the same triaxial test was simulated, and
the aforementioned approach was used to calculate the plastic de-
formations. The difference between the plastic deformations ob-
tained from each of the two methods was less than 0.05% for both
the Cam-Clay and Mohr-Coulomb models.

Material Properties

Hydraulic Properties
The paved road section was modeled as a four-layer profile: asphalt,
ABC, geocomposite, and subgrade. The hydraulic properties of
these materials are shown in Table 1 Cooley et al. (2002) conducted
a series of experimental and field permeability tests on 23 Superpave
pavement construction projects with different nominal maximum
aggregate size (NMAS) and lift thicknesses. The permeability of
Fig. 1. Unloading–reloading stress-deformation curves of the simu- the Superpave asphalt with a NMAS of 25 mm and lift thickness
lated displacement-controlled triaxial test to NMAS ratio of 4.0, as measured by Cooley et al. (2002), is used
herein in the seepage simulations.

Table 1. van Genuchten Parameters and Saturated Hydraulic Conductivity for Materials Used in the Modeling
Saturated water Residual water Saturated hydraulic
Material content, us content, ur að1=kPaÞ n conductivity (m=s) References
25
Silty sand (subgrade) 0.270 0.0 0.012 1.331 1:02 3 10 EICM
Crushed stone (ABC) 0.239 7:8 3 1022 0.320 2.750 1:30 3 1024 Henry et al. (2001)
WF geotextile 0.754 0.0 2.577 1.680 3:44 3 1023 Stormont and Ramos (2004)
NWP geotextile 0.60 0.0 3.891 6.900 6:60 3 1023 Stormont and Morris (2000)
Geonet 0.850 5:0 3 1023 50.251 2.190 1:00 3 1021 Ramos (2001)
Asphalt 0.13 — — — 1:22 3 1025 Cooley et al. (2002)

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / SEPTEMBER 2013 / 1409

J. Geotech. Geoenviron. Eng. 2013.139:1407-1418.


Downloaded from ascelibrary.org by University of Leeds on 08/15/13. Copyright ASCE. For personal use only; all rights reserved.

Fig. 2. Properties of study materials: (a) moisture characteristic curves; (b) unsaturated hydraulic conductivity curves

The geocomposite drainage layer was modeled as a three-layer Table 2. Elastic Properties of Materials Used in the Model for Various
system: a geonet sandwiched between a nonwoven geotextile at Layers (Data from Brunton et al. 1992; Ramos 2001; Bergado et al.
the bottom and a transport geotextile at the top. The hydraulic 2001; EICM; Pease 2010)
properties of the WF geotextile and geonet were measured by Ramos Elastic modulus Poisson’s Unit weight
(2001), and the hydraulic properties of the NWP geotextile were Materials (MPa) ratio (kN=m3 )
measured by Stormont and Morris (2000). Fiberglass is a hydro-
Asphalt 3:5 3 103 0.35 22.93
philic (low contact angle) material that becomes transmissive at
Crushed stone (ABC) 4:5 3 102 0.3 20.34
higher suctions relative to other types of polymeric geotextiles
WF geotextile 1:18 3 103 0.1 7.40
(Stormont and Ramos 2004). Hydraulic properties of crushed stone
NWP geotextile 1:8 3 102 0.1 1.59
and silty sand were used for ABC and subgrade, respectively [Henry
Silty sand (subgrade) 2:0 3 101 0.4 19.00
et al. 2001; Enhanced Integrated Climate Model (EICM)]. The
MCCs and hydraulic conductivity curves for all materials used in
this study are shown in Fig. 2. In developing these curves, wetting
and drying hysteresis and the effect of soil intrusion into the geo- NWP and WF geotextiles were modeled as linear elastic mate-
textile on hydraulic properties were not considered. rials. The typical elastic modulus for nonwoven needle-punched
geotextiles was reported by Bergado et al. (2001). The elastic
modulus of the WF geotextile was measured as a part of this study
Mechanical Properties
because representative values were not available in the literature. A
The elastic properties of the materials used in this study are pre- special clamping adapter for a standard load frame was necessary for
sented in Table 2. The asphalt was modeled as a linear elastic the WF geotextile because of its high strength and gripping problems
material with properties presented by Brunton et al. (1992) and Pease (ASTM D4595-11; ASTM 2008b). The two ends of specimens 2 cm
(2010). Subgrade and ABC were modeled using the Cam-Clay and wide and 13 cm long were prepared and glued to a metal wedge
Mohr-Coulomb constitutive models, respectively, and their model clamp using a special resin to prevent stress concentration and
properties are presented in Table 3 (Desai and Siriwardane 1984). slipping in the clamp. Strains were measured using an extensometer
The elastic modulus and Poisson ratio of the ABC (crushed stone) with 0.0254-mm (0.001-in.) accuracy. It was noted that the WF
and silty sand along with friction and dilation angle of crushed geotextile failed in the middle of the specimen with no slippage in the
stone were obtained using model calibration with MEPDG. The clamps. The stress-strain curves from two tests are shown in Fig. 3.
model calibration is presented next. The slopes of the initial linear portion of the curves for the two

1410 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / SEPTEMBER 2013

J. Geotech. Geoenviron. Eng. 2013.139:1407-1418.


Table 3. Mohr-Coulomb and Cam-Clay Material Properties for ABC and Subgrade (Data from Desai and Siriwardane 1984)
Friction Dilation angle Slope of Slope of Reference pressure, Specific volume
Materials angle (f) (degrees) NCL (l) OCL (k) M p09 (kPa) at p09ðn0 Þ
Silty sand (subgrade) 31.0a — 0.014 0.0024 1.24 68.95 1.36
Crushed stone (ABC) 40.0 20.0 — — — — —
a
Calculated from M.

specimens were averaged and used as the elastic modulus for


the material (Table 2).

Modeling Approach
Downloaded from ascelibrary.org by University of Leeds on 08/15/13. Copyright ASCE. For personal use only; all rights reserved.

The seepage analysis was performed using SIGMA/W. The pore


water pressure distribution at steady state was then used in FLAC as
the initial hydraulic condition to perform the stress-deformation
analysis. SIGMA/W was not used for stress-deformation analysis
because it computes the effective stress as the subtraction of pore
water pressure from total stress, which significantly overestimates
the effective stress at high negative pore water pressures. Indeed, the
suction stress decreases as water becomes discontinuous, and this
reduction in suction stress must be considered when calculating
effective stress in unsaturated soils (Vanapalli et al. 1996; Lu and
Likos 2004). FLAC, however, uses Eq. (2) to capture suction stress
in a more robust manner. Fig. 3. Measured stress strain curves for the woven fiberglass geotextile
FLAC was not used for a seepage (or coupled hydromechanical)
analysis because of unreasonably long simulation times. The time
step in transient seepage analysis is constant in FLAC and is esti- (Krisdani et al. 2008). The hydraulic properties of the fine sand and
mated as follows: geotextile layer used in the simulation were provided in Krisdani
  et al. (2008) and are shown in Table 4. The geonet was modeled as a
1 , 1 layer with low air entry value (0.2 kPa). The results of the numerical
Dt ¼ L2z min (6)
kw Kw kg Kg simulations are shown in Fig. 4 along with the measured experi-
mental data. Comparative results shown in Fig. 4 indicate a good
where Dt 5 time step; Lz 5 smallest zone size in the simulation; agreement between numerical computations and experimental mea-
Kw 5 bulk modulus of water; Kg 5 bulk modulus of air; kw 5 water surements and serve to validate the applicability of the seepage
saturated conductivity; and ka 5 air saturated conductivity. model.
A relatively high saturated hydraulic conductivity (3:44 3 To calibrate the deformation model, the MEPDG software was
1023 m=s) coupled with the thinness of the geotextile layer (3.2 mm) used to provide results for baseline comparison (Applied Research
resulted in a small time step ( ∼10210 s) and long simulation times (on Associates–Transportation 2009). A profile 100 cm wide and con-
the order of weeks). However, in SIGMA/W, the backward dif- sisting of 12.7 cm asphalt, 25.4 cm ABC, and 100 cm subgrade was
ference scheme is used for time integration. As a result, the solution modeled in FLAC under axisymmetric loading conditions. Hy-
is stable for any time step size (Smith 1971). To avoid numerical draulic and mechanical properties presented in Tables 1–3 for as-
oscillation, the minimum time step is estimated as follows: phalt, crushed stone, and silty sand, respectively, were used, and the
water table was placed at the top of the profile. The elastic modulus
l2 gw ð1 þ yÞð1 2 2yÞ and Poisson’s ratio of the crushed stone and silty sand and the
tmin $ (7) friction and dilation angles of the crushed stone were obtained
12 k Eð1 2 nÞ
through the calibration process. A 690-kPa stress was applied over
where tmin 5 minimum time step; l 5 element size; E 5 modulus of 13.6 cm of the top left corner of the profile, simulating a tire load
elasticity; n 5 Poisson’s ratio; k 5 hydraulic conductivity; and gw of 40 kN under axisymmetric conditions. The same calculation was
5 unit weight of water. The time step is then automatically varied performed with the MEPDG software using one pass of Vehicle
from 0.001 to 1 s based on the results of the previous step using an Class 5 (one single axle with a single tire) to represent the load-
adaptive time-stepping routine, thus significantly reducing overall ing condition used in FLAC. After performing each calculation in
simulation time. MEPDG, the calculated elastic modulus for asphalt, ABC, and
subgrade was used in the FLAC simulations. Multiple simulations
were performed using both approaches with varying elastic moduli,
Model Verification Poisson’s ratios, friction angles, and dilation angles for the crushed
stone to calibrate the plastic deformations calculated in FLAC.
The pore pressure data measured by Krisdani et al. (2008) in a soil- The appropriate material parameters are presented in Tables 2 and 3.
geotextile column were used for model verification. Krisdani et al. These parameters are consistent with the published values for
(2008) performed soil-geotextile column tests on a 1-m fine sand silty sand and crushed stone (Brunton et al. 1992; Evdorides and
column in which a Polyfelt Megadrain 2040 geosynthetic was Snaith 1996; Budhu 1999). Using the parameters in Tables 2 and 3,
embedded 0.51 m from the top of the column. The measured pore the difference between plastic deformations computed in MEPDG
water pressures along the column are shown as symbols in Fig. 4 and FLAC was less than 3%.

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / SEPTEMBER 2013 / 1411

J. Geotech. Geoenviron. Eng. 2013.139:1407-1418.


Seepage Analysis boundary on the left side, zero pressure head at the base, and a
potential seepage boundary condition on the right side. The potential
seepage boundary condition allows drainage of moisture from the
Paved Roads boundary under saturated conditions. This simulates drainage of
A very fine mesh was necessary to represent the geotextile layer moisture from the geocomposite into the edge pipe (which at this
because of its relative thinness, leading to computational difficulties point is assumed to be saturated for simplicity). A constant flux of
caused by the very large conductivity matrices. Such difficulties in 5:29 3 1026 m=s (19 mm=h) was applied on the top of the profile
modeling thin layers of geotextile in a two-dimensional (2D) domain simulating rainfall condition. This flux corresponds to a 6-h rain-
have been reported in previous studies (Park and Fleming 2006; fall with a 50-year return period for Greensboro, North Carolina
Stormont et al. 2009) as well. To reduce the required number of (McDowell and Borchers 2007). The initial condition was a hy-
elements and to model the profile with the three thin layers of the drostatic pore water pressure distribution throughout the profile.
geocomposite in 2D, a profile with 100 cm width was considered. A volumetric water content equal to the air-void content of the
Some of the elements in the geotextile layer were as small as 1.5 mm, Superpave asphalt and a constant permeability corresponding to
such that three nodes were located vertically within the geotextile the field measurements were used for the asphalt layer, and thus, the
Downloaded from ascelibrary.org by University of Leeds on 08/15/13. Copyright ASCE. For personal use only; all rights reserved.

layer. Transition elements were used to increase the size of the mesh thickness of the asphalt layer did not affect the pore pressure dis-
for the remaining model area. The geocomposite was placed at the tributions in ABC and subgrade.
interface of the ABC and 100-cm-thick subgrade. The surface of the Figs. 5(a–c) show the pore pressure contours after 6 h of sim-
profile and the geocomposite layer was sloped 3%. Two types of ulated rainfall for a paved road profile without a geocomposite layer
transport geotextile were considered (WF and NWP). The thick- and profiles with NWP and WF geotextile as the transport layer,
nesses of the geonet, WF, and NWP geotextile were selected as 5.9, respectively. The thickness of the ABC layer in Fig. 5 is 25.4 cm. In
3.2, and 5.9 mm, respectively (Stormont and Morris 2000; Ramos all cases, steady state was reached during the 6-h rainfall, and the
2001; Stormont et al. 2001a). Various thicknesses were considered amount of water flux entering the system was equal to the water
for the ABC and asphalt, representing low, medium, and high leaving the system ( ∼0:019 m3=h). As can be seen from Figs. 5(b and c),
volume road sections, as presented in Table 5. Repetitions of 107 , the inclusion of the geocomposite at the interface of the subgrade
106 , and 105 E18KSAL loads were considered for high, medium, and ABC increased the suction in subgrade by up to 8.0 kPa. The
and low traffic volume, respectively (NCHRP 2004). The minimum hydraulic conductivity of the geotextile transport layer is very low
asphalt required for the ABC thicknesses was then determined using under suction levels higher than its water entry value (WEV). When
the traffic volume and the Asphalt Institute (AI) design method (AI the wetting front reaches the geotextile transport layer, water
1982) for a mean annual air temperature of 15.5°C, which covers a accumulates above the geotextile and decreases the suction in the
major part of the United States. These values are presented in Table 5. ABC as shown in Figs. 5(b and c). When suction decreases to the
The hydraulic properties of the materials used in the simulations WEV of the geotextile, water enters the geotextile and is diverted
are presented in Table 1. The boundary conditions were a no-flow toward the edge drain. The WF geotextile has a higher WEV and
thus becomes transmissive at higher suctions relative to other types
of polymeric geotextiles (Stormont and Ramos 2004). Conse-
quently, the WF geotextile drains more water from the ABC,
resulting in a higher suction in the ABC compared with the NWP
geotextile. This is consistent with findings from previous studies
(Stormont and Zhou 2001; Stormont et al. 2009).

Unpaved Roads
The unpaved road section was similarly modeled, but the asphalt
layer was excluded. Three different ABC thicknesses of 508, 63.5,
and 68.6 cm were selected, corresponding to E18KSAL repetitions
of 104 , 5 3 103 , and 103 cycles, respectively. These were computed
using the standard equation for unpaved roads (Giroud and Noiray
1981)

0:19 × logðNÞ
h0 ¼ (8)
CBR0:63

where h0 5 ABC thickness; and N 5 number of E18KSAL


Fig. 4. Comparison of the measured experimental pore pressures repetitions.
(data points) by Krisdani et al. (2008) and the computed pore pressures To facilitate a better comparison, the pore pressure distributions
in this paper (lines) along the column during drainage along the left side of the profile for the three cases (without geo-
composite, with WF transport layer, and with NWP geotextile

Table 4. Hydraulic Properties of Fine Sand and Geotextile Used in Verification Simulations (Data from Krisdani et al. 2008)
Saturated water Residual water Saturated hydraulic
Materials content, us content, ur a ð1=kPaÞ n conductivity (m=s)
Fine sand 0.41 0.02 0.437 3.5 2:7 3 1024
Filter geotextile 0.834 0.0 1.777 3.578 0.1
Geonet 0.834 0.0 5.0 3.578 0.1

1412 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / SEPTEMBER 2013

J. Geotech. Geoenviron. Eng. 2013.139:1407-1418.


transport layer) and the three ABC thicknesses are plotted in Figs. 6 equilibrate under gravity to simulate initial in situ stresses, and then
and 7 for paved and unpaved road sections, respectively. In all cases, a stress of 690 kPa was applied over 13.6 cm of the top left corner of
the geocomposite layer increases the suction in the subgrade and the profile, simulating an axisymmetric tire load of 40 kN.
decreases the suction in the ABC layer compared with the profile The computed deformations were relatively small ( ∼0:01 mm)
without a geocomposite layer. Also, the geocomposite with the because only one load cycle was modeled. To investigate whether
WF geotextile results in a higher suction (by ∼1 kPa) in the ABC the displacement at the first load cycle can be used as an indication of
layer compared with the NWP geotextile. the deformation trend for higher number of load cycles, a saturated
profile and a dry profile were modeled using MEPDG formulations.
The thicknesses of the ABC and asphalt layers were 254 and
Stress-Deformation Analyses 127 mm, respectively. The total plastic deformations under one load
cycle in the saturated and dry profiles were 0.084 and 0.083 mm,
The geocomposite layer affects the pore pressure distributions in respectively. The simulations were then performed for 106 load
the ABC and subgrade (Figs. 6 and 7) and consequently changes cycles, and the results indicated that the initial deformation dif-
the effective stress and shear strength [Eqs. (1) and (2)]. The geo- ference of 0.001 mm resulted in a 7.4-mm difference in the total
composite layer also functions as a reinforcement layer and affects
Downloaded from ascelibrary.org by University of Leeds on 08/15/13. Copyright ASCE. For personal use only; all rights reserved.

plastic deformation after 106 load cycles. In this paper, deformations


the stress distribution and deformations of the ABC and subgrade. In were computed for one load cycle, and the percent change in de-
this paper, the former effect is termed the hydraulic effect and the formation was used, rather than the absolute values, to compare the
latter is referred to as the mechanical effect. To study the effect of different cases and study the effect of asphalt and ABC thicknesses
asphalt and ABC thicknesses on the contribution of each of these two on the efficacy of the geocomposite in reducing the plastic
components to reducing the plastic deformation, stress-deformation deformation.
analyses of the composite sections were performed. A profile without geocomposite, but with the pore pressure dis-
tribution as if the geocomposite layers were present, was modeled
Paved Roads for each case as well, to separate the hydraulic and mechanical
effects and to find the hydraulic effect of the geocomposite on plastic
The profile used in the stress-deformation analysis is the same as
deformation. The mechanical effect was then calculated as the total
that used for the seepage analysis. The left and right boundaries of
effect minus the hydraulic effect. Fig. 8 shows the percent decrease
the profile were fixed in the lateral direction, and the bottom
in the total (ABC and subgrade) plastic deformation caused by
boundary was fixed in both directions. The profile was allowed to
mechanical and hydraulic effects of the geocomposite layer com-
pared with the profile without a geocomposite for the three different
Table 5. ABC and Asphalt Thickness Used in Paved Road Profile ABC and asphalt thicknesses. Increasing the thickness of the asphalt
Simulations layer decreases the stress level in the geocomposite layer and thus
E18KSAL repetitions reduces its reinforcement effect. As shown in Fig. 6, the geo-
5
composite layer decreases the suction in the ABC and increases the
ABC thickness (cm) 10 106 107 suction in the subgrade. A thicker asphalt layer, however, decreases
Asphalt thickness (cm) the stress level in the ABC, and thus the negative through decreasing
suction hydraulic effect of the geocomposite on the ABC is miti-
15.2 19.1 26.9 38.1
gated. Fig. 8 also shows that the mechanical effect of the WF is more
25.4 17.1 25.1 36.3
significant than that of the NWP geotextile for smaller asphalt
45.7 11.9 20.1 34.3
thicknesses because of the higher elastic modulus of the WF

Fig. 5. Pore pressure contours in the profile: (a) without geocomposite; (b) with WF geotextile transport layer; (c) with NWP transport layer

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / SEPTEMBER 2013 / 1413

J. Geotech. Geoenviron. Eng. 2013.139:1407-1418.


Downloaded from ascelibrary.org by University of Leeds on 08/15/13. Copyright ASCE. For personal use only; all rights reserved.

Fig. 6. Pore pressure distribution in ABC and subgrade of the paved road section with ABC thickness of (a) 15.2 cm; (b) 25.4 cm; and (c) 45.7 cm

Fig. 7. Pore pressure distribution in ABC and subgrade of the unpaved road section with ABC thickness of (a) 50.8 cm; (b) 63.5 cm; and (c) 68.6 cm

1414 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / SEPTEMBER 2013

J. Geotech. Geoenviron. Eng. 2013.139:1407-1418.


Downloaded from ascelibrary.org by University of Leeds on 08/15/13. Copyright ASCE. For personal use only; all rights reserved.

Fig. 8. Percent decrease in total plastic deformation of profile with geocomposite compared with the profile without geocomposite for ABC thicknesses
of (a) 15.2 cm; (b) 25.4 cm; and (c) 45.7 cm

geotextile. With increasing asphalt thickness, the mechanical effect


of the WF decreases and becomes similar to the NWP geotextile. The
relative decrease in the total plastic deformation considering both the
mechanical and hydraulic contributions of the geocomposite is also
shown in Fig. 8. Increasing the asphalt thickness decreases the total
effect of the geocomposite on plastic deformations in the profile with
the WF geotextile and increases the total effect in the profile with
NWP geotextile. In the profile with NWP geotextile, as the asphalt
thickness increases, the contribution of the hydraulic effect com-
pensates for the decrease in mechanical effect (because of the use of
NWP versus WF), and thus, the total effect of the NWP geotextile on
plastic deformations increases. As a result, the effect of the geo-
composite layers on the total plastic deformation for the NWP and
WF geotextiles become similar with increasing asphalt thickness
(Fig. 8).
It is also apparent from Fig. 8 that the geocomposite can decrease
the total plastic deformation up to 55%. To investigate how much the
ABC thickness can be reduced by inclusion of the geocomposite
layer, plastic deformation in a profile without a geocomposite and
with 45.7 cm ABC was compared with the profiles with the geo-
composite and 15.2, 25.4, and 45.7 cm ABC. Three asphalt
thicknesses of 19.1, 26.9, and 38.1 cm were used. The plastic
deformations of the profiles with geocomposite (dpg ) were divided
by the one for the profile without geocomposite and with 45.7 cm
ABC (dp45:7 ), and the results are shown in Fig. 9. For each of the Fig. 9. Ratio of the plastic deformation of the profile with geo-
asphalt thicknesses, the plastic deformation of the profiles with the composite (dpg ) to the plastic deformation for the 45.7-cm ABC (dp45:7 )
geocomposite was lower than without the geocomposite (with a ratio profile without geocomposite
of between 0.5 and 0.7). In the profile with the thinnest asphalt

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / SEPTEMBER 2013 / 1415

J. Geotech. Geoenviron. Eng. 2013.139:1407-1418.


layer, WF had a larger effect on reducing the plastic deformations, To determine how much of the ABC thickness can be reduced
and increasing the asphalt layer reduced this effect. However, NWP with the inclusion of the geocomposite, the ratio of the plastic de-
geotextile has a larger effect on the plastic deformations in the profile formation of the profile with geocomposite and different ABC
with thicker asphalt. thicknesses (dpg ) was divided by that for the profile without geo-
composite and with the greatest ABC thickness (dp68:6 ). The results
for both the initial and reduced elastic modulus of the WF and NWP
Unpaved Roads
geotextiles are shown in Fig. 12. When the initial elastic modulus is
To simulate an unpaved road, a surface stress of 482.7 kPa (70 psi) used, the plastic deformation with WF and NWP geotextiles is lower
was applied, and a small cohesion of 3.5 kPa was required for the than the one without geocomposite. However, when the elastic
ABC to overcome convergence issues. The rest of the modeling modulus is reduced, the plastic deformation of the profile with NWP
details were the same as the paved road section. Fig. 10 shows geotextile and 50.8 cm ABC becomes more than the one without
the hydraulic and mechanical effect of the geocomposite on total geocomposite and with 68.6 cm ABC. Simulation results for
plastic deformations for all the three ABC thicknesses. Increasing a NWP geotextile with reduced elastic modulus and 58.4 cm (23 in.)
the ABC thickness increases the hydraulic effect and decreases the ABC are also shown in Fig. 12. In this case, the WF and NWP
Downloaded from ascelibrary.org by University of Leeds on 08/15/13. Copyright ASCE. For personal use only; all rights reserved.

mechanical effect of the geocomposite. Fig. 10 also show that the geotextiles with reduced elastic modulus can decrease the required
geocomposite has a higher mechanical effect and lower hydraulic ABC thickness from 68.9 to 50.8 and 58.4 cm, respectively.
effect on the plastic deformation compared with paved sections
(Fig. 8). The reason is that the asphalt layer in paved roads decreases
the stress level in the underlying layers and thus decreases the re- Summary and Conclusions
inforcement effect and the negative hydraulic effect of the geo-
composite. Data in Fig. 10 show that with elimination of the Work in this manuscript presents a numerical study on the effect
mechanical effect, WF and NWP geotextiles still decrease the total of inclusion of a geocomposite drainage layer on the moisture
plastic deformations by an average of 31 and 28%, respectively.
The elastic modulus of geosynthetics usually decreases with
increasing the tensile strain. Bergado et al. (2001) showed that the
elastic modulus of the nonwoven geotextiles decreases to 13.3% of
its initial value (from 1:8 3 102 to 2:4 3 101 MPa) for tensile strains
higher than 2.5%. In unpaved road sections, the induced tensile
strain at the top of the subgrade may reduce the elastic modulus and
therefore the mechanical effect of the geocomposite layer. To study
the effect of this decrease, the elastic modulus of the WF and NWP
geotextiles was reduced to 13.3% of their initial values to 1:6 3 102
and 2:4 3 101 MPa, respectively. Fig. 11 shows the total effect of the
geocomposite on the plastic deformation with the initial and reduced
elastic modulus of the geotextiles. Decreasing the elastic modulus
of the WF and NWP geotextiles decreases their effect on the total
plastic deformations by a maximum of approximately 20% (from
∼60 to 40%) for WF and 7.0% (from ∼36 to 29%) for NWP,
respectively, with a slight dependency on the thickness of the ABC
layer (Fig. 11). The WF geotextile decreases the total plastic de-
formations by approximately 24% more than the profile with NWP
geotextile regardless of the ABC thickness. Fig. 11. Percent decrease in total plastic deformation caused by the
total effect of the geocomposite in the unpaved road section

Fig. 10. Percent decrease in total plastic deformation of profile with Fig. 12. Ratio of the plastic deformation of the profile with geo-
geocomposite compared with the profile without geocomposite in un- composite (dpg ) to the one for the profile without geocomposite and
paved road section with 68.6-cm ABC (dp68:6 )

1416 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / SEPTEMBER 2013

J. Geotech. Geoenviron. Eng. 2013.139:1407-1418.


distribution and plastic deformations of simulated unpaved and boundary conditions and material properties used in the numerical
paved road sections. Simulations were performed for different ABC modeling.
and asphalt thicknesses and investigated the hydraulic versus the
mechanistic contribution of the drainage layer. The following con-
clusions are drawn from the results: References
1. The effect of the suction on soil strength and plastic deforma-
tions using the Mohr-Coulomb model is more appropriately Applied Research Associates–Transportation. (2009). Mechanistic-
incorporated by modifying the failure envelope rather than the empirical pavement design guide (MEPDG), version 1.100, Cham-
paign, IL.
elastic modulus.
Asphalt Institute (AI). (1982). Research and development of the asphalt
2. The seepage analysis showed that during the simulated rainfall institute’s thickness design manual (MS-1), 9th Ed., Lexington, KY.
infiltration, the geocomposite increases the suction in subgrade ASTM. (2008a). “Standard test methods for determination of the soil water
by up to 8 kPa and decreases the suction in ABC by up to 3.6 characteristic curve for desorption using a hanging column, pressure
kPa. The WF geotextile causes about 2 kPa less reduction in extractor, chilled mirror hygrometer, and/or centrifuge.” D6836-02,
suction in ABC compared with NWP geotextile because of its West Conshohocken, PA.
Downloaded from ascelibrary.org by University of Leeds on 08/15/13. Copyright ASCE. For personal use only; all rights reserved.

higher WEV. ASTM. (2008b). “Standard test methods for tensile properties of geotextiles
3. The stress-deformation analysis showed that increasing the by the wide-width strip method.” D4595-11, West Conshohocken, PA.
pavement thickness increases the hydraulic effect and de- Atkinson, J. H., and Bransby, P. L. (1978). The mechanics of soils:
creases the mechanical effect of the geocomposite with regard An introduction to critical state soil mechanics, McGraw Hill, New York.
Bender, D. A., and Barenberg, E. J. (1978). “Design and behavior of soil-
to the total plastic deformations in both the paved and unpaved
fabric-aggregate systems.” Transportation Research Record 671,
road sections. Transportation Research Board, Washington, DC, 64–75.
4. WF and NWP geotextiles can decrease the total plastic de- Bergado, D. T., Youwai, S., Hai, C. N., and Voottiipruex, P. (2001). “In-
formation by up to 55 and 35% in paved road sections and by teraction of nonwoven needle-punched geotextiles under axisymmetric
up to 60 and 30% in unpaved road sections, respectively, loading conditions.” Geotextile Geomembrane, 19(5), 299–328.
depending on the ABC and asphalt thicknesses. Increasing the Brunton, J. M., Armitage, R. J., and Brown, S. F. (1992). “Seven years’
thickness of the asphalt layer decreased the difference between experience of pavement evaluation.” Proc., 7th Int. Conf. on Asphalt
the effect of the WF and NWP geotextile on plastic defor- Pavements, International Society for Asphalt Pavements, Lino Lakes,
mation to less than 5%, as the reinforcement contribution is MN, 17–30.
reduced. In unpaved road sections, WF geotextile decreases Budhu, M. (1999). Soil mechanics and foundations, Wiley, New York.
Cedergren, H. R. (1994). “America’s pavements: World’s longest bathtubs.”
the total plastic deformation by an average of 24% more than
Civil Eng. 64(9), 56–58.
the NWP geotextile, regardless of the ABC thickness, mainly Christopher, B. R., Hayden, S. A., and Zhao, A. (2000). “Roadway base and
because of the hydraulic effect. subgrade geocomposite drainage layers.” Proc., Testing and Perfor-
5. In unpaved road sections, reducing the elastic modulus of the mance of Geosynthetics in Subsurface Drainage, ASTM, West Con-
WF and NWP geotextile by nearly one order of magnitude shohocken, PA 35–51.
decreases their effect on reducing the total plastic deformations Christopher, B. R., and McGuffey, V. C. (1997). NCHRP synthesis of
by a maximum of approximately 20% (from ∼60 to 40%) for highway practice 239: Pavement subsurface drainage systems, National
WF and 7.0% (from ∼36 to 29%) for NWP, respectively, with Research Council, Washington, DC.
a slight dependency on the thickness of the ABC layer. Even Cooley, L. A., Jr., Prowell, B. D., and Brown, E. R. (2002). “Issues per-
then, geocomposite with reduced elastic modulus of WF and taining to the permeability characteristics on coarse-graded Superpave
mixes.” J. Assoc. Asphalt Paving Technologists, 71, 1–29.
NWP geotextile can decreases the required thickness of the
Dawson, A. R., Paute, J. L., and Thom, N. H. (1996). “Mechanical char-
ABC by 26 and 15%, respectively. acteristic of unbound granular materials as a function of condition.”
6. In general, results from this study showed that both the mech- Flexible pavements: Proc., European Symposium Euroflex 1993, A.
anical and hydraulic effects of geocomposite inclusion are Gomes Correia, ed., Rotterdam, Netherlands, 35–45.
significant, particularly when moisture movement under un- Desai, C. S., and Siriwardane, H. J. (1984). Constitutive laws for engineering
saturated conditions is considered. Thus, design approaches materials with emphasis on geologic materials, Prentice Hall, Engle-
that do not consider both effects may be overly conservative. wood Cliffs, NJ.
A robust set of measured data for comparison with results Drumm, E. C., Reeves, J. S., Madgett, M. R., and Trolinger, W. D. (1997).
from numerical modeling was not found in the literature. This is a “Subgrade resilient modulus correction for saturation effects.” J. Geo-
testament to the difficulties associated with making these types of tech. Geoenviron. Eng., 123(7), 663–670.
Enhanced Integrated Climate Model (EICM) 3.2 Beta [Computer soft-
measurements either in the field or in the laboratory. The results
ware]. College Station, TX, Texas Transportation Institute, Texas
from the numerical study presented herein serve to provide un- A&M Univ.
derstanding of the multiphysics aspects of the multilayer drainage Evdorides, H. T., and Snaith, M. S. (1996). “A knowledge based analysis
barrier system with markedly different hydraulic characteristics in process for road pavement condition assessment.” Proc. ICE-Transport,
terms of moisture distribution and associated impact on plastic 117(3), 202–210.
deformation. The assumptions and simplifications made in the Fredlund, D. G., and Rahardjo, H. (1993). Soil mechanics for unsaturated
numerical modeling should, however, be considered in interpret- soils, Wiley, New York.
ing the results. The applied boundary conditions simulate a water Garcia, E. F. (2007). “Function of permeable geosynthetics in unsaturated
table of 100 cm below the drainage system and a drainage pipe embankments subjected to rainfall infiltration.” Geosynthetics Int.,
under saturated flow. Changing the water level will alter the initial 14(2), 89–99.
Gehling, W. Y. Y., Ceratti, J. A., Nunez, W. P., and Rodrigues, M. R. (1998).
and subsequent suction distribution throughout the profile. Having
“A study on the influence of suction on the resilient behavior of soils from
an unsaturated drainage pipe affects the water flow from the southern Brazil.” Proc., 2nd Int. Conf. on Unsaturated Soils, In-
drainage system into the pipe. The change in hydraulic properties ternational Academic Publishers, Beijing, 27–30.
of the drainage system caused by the wetting and drying hysteresis Giroud, J. P., and Han, J. (2004). “Design method for geogrid-reinforced
or soil intrusion into the geotextile was not considered in the unpaved roads. II. Calibration and applications.” J. Geotech. Geo-
analysis. The results presented in this paper reflect the selected environ. Eng., 130(8), 787–797.

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / SEPTEMBER 2013 / 1417

J. Geotech. Geoenviron. Eng. 2013.139:1407-1418.


Giroud, J. P., and Noiray, L. (1981). “Geotextile-reinforced unpaved road soil moisture: Summary of predictive models, ARA, ERES Division,
design.” J. Geotech. Engrg. Div., 107(9), 1233–1254. Champaign, IL.
Heckel, L. (1997). “Performance problems of open-graded drainage layers National Cooperative Highway Research Program (NCHRP). (2004).
under continuously reinforced concrete pavement in Illinois.” Trans- “Guide for mechanistic-empirical design of new and rehabilitated
portation Research Record 1596, Transportation Research Board, pavement structures. Appendix GG-1: Calibration of permanent
Washington, DC, 51–57. deformation models for flexible pavements, ARA, ERES Division,
Henry, K. S. (1988). “Use of geotextiles to mitigate frost heave in soils.” Champaign, IL.
Proc., 5th Int. Conf. on Permafrost, TAPIR Publications, Pasadena, CA, Park, K. D., and Fleming, I. R. (2006). “Evaluation of geosynthetic capillary
1096–1011. barrier.” J. Geotextile Geomembr., 24(1), 64–71.
Henry, K. S., and Holtz, R. D. (1997). “Capillary rise of water in geo- Pease, R. E. (2010). “Hydraulic properties of asphalt concrete.” Ph.D. thesis,
textiles.” Proc., Int. Symp. on Ground Freezing and Frost Action in Univ. of New Mexico, Albuquerque, NM.
Soils, Taylor & Francis, Oxford, U.K., 227–233. Rada, G., and Witczak, M. W. (1981). “Comprehensive evaluation of
Henry, K. S., and Stormont, J. C. (2002). “Geocomposite capillary barrier laboratory resilient moduli for granular material.” Transportation
drain for limiting moisture changes in pavement subgrades and base Research Record 810, Transportation Research Board, Washington,
courses.” NCHRP-IDEA 68 Final Rep., Transportation Research DC, 23–33.
Board, Washington, DC. Ramos, R. D. (2001). “Performance of a fiberglass based geocomposite
Downloaded from ascelibrary.org by University of Leeds on 08/15/13. Copyright ASCE. For personal use only; all rights reserved.

Henry, K. S., Stormont, J. C., Barna, L. A., and Ramos, R. D. (2002). capillary barrier drain.” M.Sc. thesis, Univ. of New Mexico, Albu-
“Geocomposite capillary barrier drain for unsaturated drainage of querque, NM.
pavements.” Proc., 7th Int. Conf. on Geosynthetics, Balkema, Lisse, Simac, M. R., Elton, D. J., and Gale, S. M. (2006). “Discussion of “design
Netherlands, 877–880. method for geogrid-reinforced unpaved roads. I: Development of
Henry, K. S., Stormont, J. C., Ramos, R. D., and Barna, L. (2001). “Geo- design method” by J. P. Giroud and J. Han.” J. Geotech. Geoenviron.
composite capillary barrier drain for limiting moisture changes in pavement Eng., 132(4), 547–549.
subgrades and bases.” Rep. Prepared for the IDEA Program, Transportation Smith, G. D. (1971). Numerical solution of partial differential equations,
Research Board, National Research Council, Washington, DC. Oxford University Press, Oxford, U.K.
Hsieh, P. A., Wingle, W., and Healy, R. W. (2000). “A graphical software Stockton, T. B. (2001). “Performance experimental testing of diversion
package for simulating fluid flow and solute or energy transport in lengths for a geocomposite unsaturated drainage system.” M.Sc. thesis,
variably saturated porous media.” USGS Water-Resources Inves- Univ. of New Mexico, Albuquerque, NM.
tigations Rep. 99-4130, USGS National Center, Reston, VA. Stormont, J., and Zhou, S. (2001), “Improving pavement sub-surface
Iryo, T., and Rowe, R. K. (2003). “On the hydraulic behavior of unsaturated drainage systems by considering unsaturated water flow.” Final Rep.
nonwoven geotextiles.” J. Geotextile Geomembr., 21(6), 381–404. DTFH61, Federal Highway Administration, Washington, DC.
Itasca Consulting Group. (2008). FLAC (Fast Lagrangian Analysis of Stormont, J. C., Henry, K. S., and Evans, T. M. (1997). “Water retention
Continua), Version 6.0, Itasca Consulting Group, Minneapolis. functions of four nonwoven polypropylene geotextiles.” Geosynthetics
Knight, M. A., and Kotha, S. M. (2001). “Measurement of geotextile-water Int., 4(6), 661–672.
characteristic curves using a controlled outflow capillary pressure cell.” Stormont, J. C., Henry, K. S., and Roberson, R. (2009). “Geocomposite
Geosynthetics Int., 8(3), 271–282. capillary barrier drain for limiting moisture changes in pavements:
Kolisoja, P., Saarenketo, T., Peltoniemi, H., and Vuorimies, N. (2002). Product application.” Final Rep., Contract No. NCHRP-113, Trans-
“Laboratory testing of suction and deformation properties of base course portation Research Board, Washington, DC.
aggregates.” Transportation Research Record 1787, Transportation Stormont, J. C., and Morris, C. E. (2000). “Characterization of unsaturated
Research Board, Washington, DC, 83–88. nonwoven geotextiles.” Proc., Advances in Unsaturated Geotechnics,
Krahn, J. (2004). Stress and deformation modeling with SIGMA/W, an en- ASCE, Reston, VA, 529–542.
gineering methodology, GEO-SLOPE International, Calgary, Alberta, Stormont, J. C., and Ramos, R. D. (2004). “Characterization of a fiber-
Canada. glass geotextile for unsaturated in-plane water transport.” Geotech.
Krisdani, H., Rahardjo, H., and Leong, E. C. (2006). “Experimental study of Test. J., 27(2), 214–220.
1-D capillary barrier model using geosynthetic material as the coarse- Stormont, J. C., Ramos, R. D., and Henry, K. S. (2001a). “Geocomposite
grained layer.” Proc., 4th Int. Conf. on Unsaturated Soils, Taylor & capillary barrier drain system with fiberglass transport layer.” Trans-
Francis, Oxford, U.K., 1683–1694. poration Research Record 1772, Transporation Research Board,
Krisdani, H., Rahardjo, H., and Leong, E. C. (2008). “Measurement of Washington, DC, 131–136.
geotextile-water characteristic curve using capillary rise principle.” Stormont, J. C., Ray, C., and Evans, T. M. (2001b). “Transmissivity of
Geosynthetics Int., 15(2), 86–94. a nonwoven polypropylene geotextile under suction.” ASTM Geotech.
Kuhn, J. A., McCartney, J. S., and Zornberg, J. G. (2005). “Impinging flow Test. J., 24(2), 164–171.
over drainage layers including a geocomposite.” Proc., Sessions of the Sweere, G. T. H. (1990). “Unbound granular bases for roads.” Ph.D. thesis,
Geo-Frontiers 2005 Congress, ASCE, Geo-Institute, Reston, VA. Univ. of Delft, Delft, Netherlands.
Lafleur, J., Lebeau, M., Faure, Y. H., Savard, Y., and Kehila, Y. (2000). Tian, P., Zaman, M. M., and Laguros, J. G. (1998). “Variation of resilient
“Influence of matric suction on the drainage performance of polyester modulus of aggregate base and its influence on pavement performance.”
geotextiles.” Proc., 53rd Annual Conf. of Canadian Geotechnical J. Test. Eval., 26(4), 329–335.
Society, The Canadian Geotechnical Society, Richmond, BC, Canada, Vanapalli, S. K., Fredlund, D. G., Pufahi, D. E., and Clifton, A. W. (1996).
1115–1122. “Model for the prediction of shear strength with respect to soil suction.”
Lu, N., and Likos, W. J. (2004). Unsaturated soil mechanics, Wiley, Can. Geotech. J., 33(3), 379–392.
Hoboken, NJ. van Genuchten, M. Th. (1980). “A closed-form equation for predicting
McDowell, H. K., and Borchers, M. (2007). Storm drainage design manual, the hydraulic conductivity of unsaturated soils.” Soil Sci. Soc. Am. J.,
North Carolina Erosion and Sediment Control Planning and Design, 44(5), 892–898.
Storm Water Management Division, Raleigh, NC. Witczak, M. W., Houston, W. N., and Andrei, D. (2000). “Resilient modulus
Nahlawi, H., Bouazza, A., and Kodikara, J. (2007). “Characterisation of as function of soil moisture: A study of the expected changes in resilient
geotextiles water retention using a modified capillary pressure cell.” modulus of the unbound layers with changes in moisture for 10 LTPP
J. Geotextile Geomembr., 25(3), 186–193. sites.” Development of the 2002 Guide for the Development of New and
National Cooperative Highway Research Program (NCHRP). (2000). Rehabilitated Pavement Structures, NCHRP 1-37 A, Interteam Tech-
Guide for mechanistic-empirical design of new and rehabilitated nical Rep. (Seasonal 2), Federal Highway Administration, Washington,
pavement structures. Appendix DD-1: Resilient modulus as function of DC.

1418 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / SEPTEMBER 2013

J. Geotech. Geoenviron. Eng. 2013.139:1407-1418.

You might also like