Jgein 21 00005

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Geosynthetics International, 2022, 29, No.

Laboratory evaluation of different geosynthetics for


water drainage
Y. Guo1, C. Lin2, W. Leng3 and X. Zhang4
1
Assistant Professor, School of Civil Engineering, Changsha University of S&T, Changsha, Hunan,
410114, China; Graduate Research Assistant, Department of Civil Engineering, Central South University,
Changsha, Hunan, 410075, China, Email: yipguo@csust.edu.cn, haiyao19@csu.edu.cn
2
Assistant Professor, School of Transportation Science and Engineering, Harbin Institute of Technology,
Harbin, Heilongjiang, China, 150090, Email: linchuang@hit.edu.cn
3
Professor, Department of Civil Engineering, Central South University, Changsha, Hunan, 410075,
China, Email: wmleng@csu.edu.cn
4
Professor, Department of Civil, Architectural, and Environmental Engineering, Missouri S&T,
135 Butler Carlton Hall, 1401 N. Pine Street, Rolla, MO 65409-0030, USA, Email: zhangxi@mst.edu
(corresponding author)

Received 01 February 2020, accepted 29 March 2021, published 01 April 2022

ABSTRACT: During road construction, soils are often compacted at the optimum water content to
achieve maximum dry density and best performance. After construction is completed, the soil water
content in the field will inevitably increase with time due to capillary rise, rainfall infiltration, and other
factors. Conventional drainage systems rely on geomaterials or geosynthetics with large pores to drain
gravity (or free) water but cannot drain out capillary water. The excess water in the road system causes
pavement deterioration under repetitive traffic load. Recently, two new types of geosynthetics were used
as drainage materials. However, most of the field tests were inconclusive due to complicated site
conditions and soil nonuniformity. The relative performances of these drainage geosynthetics and their
working mechanisms were largely unclear. In this study, laboratory tests were conducted to quantify the
cumulative amount of water drained under different drainage situations. The volumetric water content
of soils was monitored by moisture sensors and the water contents of soils under different drainage
situations were evaluated and compared. Finally, the working mechanisms of different drainage
materials were discussed.

KEYWORDS: Geosynthetics, Drainage belt, Wicking geotextile, Capillary water, Water retention
curve, Unsaturated soil

REFERENCE: Guo, Y., Lin, C., Leng, W. and Zhang, X. (2022). Laboratory evaluation of different
geosynthetics for water drainage. Geosynthetics International, 29, No. 3, 254–269. [https://doi.org/
10.1680/jgein.21.00005]

Cooperative Highway Research Program (NCHRP)


1. INTRODUCTION
study estimated that excess water reduces the life expect-
Water within pavement layers is a principal cause of ancy of pavement systems by more than half (Christopher
pavement deterioration (Cedergren 1994; Christopher and and McGuffey 1997). Government transportation engin-
McGuffey 1997; Henry and Holtz 2001). Specific pro- eers in cold regions have credited a minimum of half of
blems associated with water are stripping of asphalt road maintenance expenditures to the effects of freezing
pavement, joint displacement in concrete pavements, and thawing (Henry and Holtz 2001). Lin et al. (2018)
reduction in pavement strength caused by excess water in carried out a series of tests to investigate the effect of water
the base course layers, shrinking and swelling of subgrade content on the variations of resilient modulus and
materials caused by water content changes, and frost permanent deformation for base course material (AB3)
heave and thaw weakening caused by capillary water flow in Kansas and found that a small amount of water content
beneath pavements. Water-related problems are thus variation could lead to a significant reduction in the
responsible for decreased pavement life, increased costs resilient modulus and increment in the permanent
for maintenance, and increased pavement roughness, and deformation.
they occur to some extent throughout all regions and Nonwicking geosynthetics refer to those geosynthetics
climates of the United States. A recent National without wicking fibers and existing drainage systems and
1072-6349 © 2022 Thomas Telford Ltd 254

Downloaded by [ Indian Institute Of Technology - Bombay] on [07/03/23]. Copyright © ICE Publishing, all rights reserved.
Laboratory evaluation of different geosynthetics for water drainage 255

that mainly rely on gravity to drain water and cannot work The wicking geotextile is a dual-function woven geotextile
under unsaturated conditions. Geosynthetics are com- for drainage and reinforcement purposes. The wicking
monly used for reinforcement and drainage purposes. fabric yarns are made of hydrophilic, hygroscopic 4DG
Extensive field and laboratory experimental works have TM nylon fibers with multichannel deep grooves for
been conducted to investigate the working mechanism drainage purposes, as shown in Figure 1(b). The poly-
of geosynthetics for reinforcement and drainage purposes. propylene fiber yarns have a high tensile strength and are
A large number of geosynthetic products and design used for reinforcement purposes. Detailed description of
guidelines have also been proposed (Gray and Al-Refeai the wicking geotextile can be found in section 2.1.
1986; Athanasopoulos 1993; Ghosh and Madhav 1994; Laboratory tests were carried out to evaluate the effec-
Gobel et al. 1994; Zornberg and Mitchell 1994; Cancelli tiveness of the wicking geotextile in reducing soil moisture
et al. 1996; Perkins 1999; Som and Sahu 1999; McGown (Wang et al. 2017; Guo et al. 2019). Zhang and Belmont
et al. 2004; Stulgis 2005; Pathak and Alfaro 2010). When (2009) reported that the wicking geotextile showed a better
used for reinforcement purposes, geosynthetics are often drainage performance compared with nonwicking geo-
used within unsaturated soils that are compacted at the textiles. In addition, Lin and Zhang (2018a) carried out a
optimum water content to achieve the maximum dry series of wetting and drying tests to evaluate the drainage
density. After the construction is complete, the soil water performance of the wicking geotextile with different types
content in the field will inevitably increase with time due of soils and found that both gravitational and capillary
to capillary rise, rainfall infiltration, and other factors. water could be drained out by the wicking geotextile.
When used for drainage purposes, nonwicking geosyn- Moreover, Lin and Zhang (2018b) also proposed a
thetics function under saturated conditions and rely on ‘bio-wicking’ system to further improve the drainage
large pores to drain gravitational (or ‘free’) water. Between efficiency of the wicking geotextile. Furthermore, the
the full saturation and optimum water content, there is a wicking geotextile had been successfully used to mitigate
large amount of capillary water that cannot be drained by frost heave for Alaskan pavements and a roadway installed
nonwicking geosynthetics. (Ling et al. 1993; Mitchell and with the wicking geotextile successfully eliminated the
Zornberg 1995; Tan et al. 2001; Iryo and Rowe 2003, frost heave problem (Zhang et al. 2014; Lin et al. 2017).
2005; Benjamin et al. 2007; Garcia et al. 2007; Noorzad Some of the tests showed promising results; however,
and Mirmoradi 2010; Raisinghani and Viswanadham others are inconclusive due to complicated site conditions
2011; Bouazza et al. 2013; Portelinha et al. 2013). Most and the nonuniformity of the soil in the field. Therefore, it
existing drainage systems can only drain ‘free’ water is impossible to compare the drainage performances of
but not capillary water (Bathurst et al. 2009; Christopher different geosynthetics and their working mechanisms.
et al. 2010). Increasing water content can lead to The objectives of this study were to evaluate and
significant reductions in the resilient modulus and incre- compare working mechanisms of different geosynthetics.
ments in permanent deformations. Failing to drain Laboratory model tests were conducted to quantify the
capillary water is a major cause of geotechnical failures amount of water drained under four different drainage
and pavement distresses (Mitchell and Zornberg 1995; situations, including a poorly graded sand as the control
Koerner and Soong 2001; Stulgis 2005; Yoo and Jung case (without any geosynthetics),and the same sand
2006; McCartney and Zornberg 2010; Zornberg et al. with three different types of geosynthetics, including
2010). a drainage belt, a wicking geotextile, and a modified
Some researchers attempted to use new types of geo- wicking geotextile. The soil volumetric water contents
synthetics to drain capillary water for better performance. were monitored by moisture sensors installed at different
For example, Henry et al. (2002) proposed a geocomposite locations of a testing box. In addition, the cumulative
capillary barrier drain (GCBD) to limit moisture changes amount of water collected from each testing box was also
in pavements. The GCBD use a hydrophilic fiberglass determined. Finally, the working mechanisms of different
with a higher specific surface area to absorb water and geosynthetics as drainage materials were discussed.
transport capillary water under a low suction level.
Similar to a conventional drainage system, the GCBD
relies on gravity to generate a hydraulic gradient to drain 2. MATERIALS AND METHODS
capillary water. It only works within a small suction range,
which is less than 6 kPa (equivalent to 0.6 m water head). 2.1. Test materials
Leng et al. (2017) used plastic sheets with Ω-shaped The soil used in the study was Missouri river sand and the
grooves (hereafter ‘drainage belt’) to drain water in the grain size ranged from 1.19 mm to 0.42 mm. The specific
roadside slopes along the railway system in China, as gravity (Gs) of the soil is 2.65, and the Cu and Cc values are
shown in Figure 1(a). Field applications indicated that the 1.42 and 0.93, respectively. According to the Unified Soil
drainage belt could effectively lower the groundwater table Classification System (USCS) (ASTM D2487), the soil is
without any clogging. Guo et al. (2018) also demonstrated classified as poorly graded sand (SP). In this paper, sand
the drainage efficiency of the drainage belt via laboratory was selected on purpose to demonstrate that even a
tests. Recently, a woven geotextile with wicking fibers coarse-grained soil, conventionally used as a good
(hereafter ‘wicking geotextile’) was proposed to drain drainage material, can still hold a large amount of
both gravitational and capillary water under both satu- capillary water under unsaturated conditions and part of
rated and unsaturated conditions, as shown in Figure 1(b). the water cannot be drained by gravity. In addition, most
Geosynthetics International, 2022, 29, No. 3

Downloaded by [ Indian Institute Of Technology - Bombay] on [07/03/23]. Copyright © ICE Publishing, all rights reserved.
256 Guo, Lin, Leng and Zhang

Centreline A weft yarn set

0.3 1.5
Weft yarn Wicking fibre Wrap yarn

2 (c)
1

(a)

um
0
–5
30
Wrap yarn
Wicking fibres Weft yarn
5–12 um

5–12 um

2 mm

A weft yarn set

(b)

Centerline A weft yarn set

Additional wicking fibres


(e)

Additional wicking fibres

(d)

Figure 1. Drainage materials used in laboratory tests: (a) drainage belt (unit in mm); (b) wicking geotextile; (c) schematic plot of weaving
structure of the wicking geotextile; (d) modified wicking geotextile; and (e) schematic plot of weaving structure of the modified wicking
geotextile (c and e were modified from Lin et al. 2018)

of the base course material used in road construction can geotextile (Figures 1(b) and 1(c)), and a modified wicking
be classified as gravel with sand, which are typically geotextile (Figures 1(d) and 1(e)). The drainage belt
considered to be good drainage materials, and the resilient comprised a 2 mm-thick plastic sheet with Ω-shaped
modulus is significantly influenced by soil moisture drainage grooves on one side of the sheet as shown in
content. Figure 1(a). The drainage grooves have circular shapes
Three different drainage materials were used in this with inner diameters of approximately 1.0 mm and the
study, including a drainage belt (Figure 1(a)), a wicking width between two adjacent grooves is 1.5 mm. The
Geosynthetics International, 2022, 29, No. 3

Downloaded by [ Indian Institute Of Technology - Bombay] on [07/03/23]. Copyright © ICE Publishing, all rights reserved.
Laboratory evaluation of different geosynthetics for water drainage 257

opening size for drainage is about 0.3 mm. Owing to a less than 50% passing the No. 200 sieve, the O95 of the
relatively small opening size, soil particles with large sizes wicking geotextile shall be smaller than 0.6 mm. For this
will not clog the drainage paths while ‘free’ water can study, the soil used in the test is a SP and the O95 of the
be drained out through the grooves (Guo et al. 2018). wicking geotextile is 0.195 mm, which meets the
Note that the drainage belt can only drain ‘free’ water, AASHTO criterion. In summary, the wicking geotextile
while the capillary water cannot be drained by this type of meets all the filtration criteria.
geosynthetic. Figure 1(d) shows the images of a modified wicking
Figure 1(b) shows the wicking geotextile. The wicking geotextile. An additional new wicking fiber yarn was
geotextile is a dual-function woven geotextile for artificially knitted into the wicking geotextile. The
reinforcement and drainage purposes. Similar to other wicking fiber yarn was new and the grooves within each
nonwicking geotextiles, the polyethylene yarns have high wicking fiber were in a good shape. The modified wicking
tensile strength and are used for reinforcement purposes. geotextile had four rather than three wicking fiber yarns
More importantly, the wicking fiber yarns are made of within a weft yarn set when compared with the wicking
hydrophilic, hygroscopic 4DG TM nylon fibers for geotextile, as shown in Figures 1(d) and 1(e).
drainage purposes. As shown in Figure 1(b), the average
diameter of a wicking fiber is between 30 μm and 50 μm 2.2. Experimental setup
and the average opening of the grooves is between 5 μm Figures 2(a) and 2(b) show the schematic plot and
and 12 μm (Lin and Zhang 2018a, 2018b). The multi- configuration of the testing box. The frame of the testing
channel shaped cross-section has a high shape factor and box was built with aluminum alloy to provide a strong
a great number of channels per fiber (specific surface support for testing materials. The sidewalls were made of
area = 3650 cm2/g), which gives the wicking fabric a transparent acrylic to facilitate observations of moisture
great potential for maximizing the capillary action and migrations within the testing box. The dimensions of the
transporting water under an unsaturated environment. testing box were 132 cm × 60 cm × 20 cm (length × height ×
Figure 1(c) shows the schematic plot of the profile of the width). In order to compare the drainage efficiency of
wicking geotextile, which comprises a double layered different geosynthetics at the same time, the testing box was
fabric formed from a single weave. The fabric comprises a divided into two sections, the dimensions for each of the
weft yarn in the second layer, a weft yarn in the first layer, sections were 66 cm × 60 cm × 20 cm (length × height ×
wicking fibers woven in the weft direction of the fabric, width), as shown in Figure 2(a). In all tests, the soil was
and a warp yarn interweaving the first and second weft compacted to a height of 48 cm with an initial void ratio of
yarns and the wicking fibers as shown in Figure 1(c). 0.64. The corresponding saturated permeability of the soil
The weft yarn is a high modulus tape comprising an was 6.5 × 10−4 cm/s. Two valves were installed at the bottom
admixture of polypropylene and a polypropylene/ethylene of the left and right side of the box for supplying water.
copolymer for reinforcement purposes. The specifications Meanwhile, there were two openings (20 cm × 0.2 cm
of the wicking geotextile are listed in Table 1. Note that (W × H )) located at each side of the sidewalls of
water is capable of flowing through the wicking geotextile the test apparatus with an elevation of 10 cm from the
at a rate of 1222 l/min/m2. To meet the filtration criterion, bottom. These openings were designed to install the
the wicking geotextile has to provide an adequate drainage materials. Rectangular grids with dimensions
permeability in the cross-plane direction while also of 3 cm × 3 cm were marked on the front acrylic wall
providing a proper soil retention capability. The saturated to facilitate visual observations of the drainage performance.
cross-plane hydraulic conductivity of the wicking geotex-
tile can be determined based upon the permittivity value 2.3. Instrumentation
provided in Table 1. The saturated cross-plane hydraulic METER Environment® EC-5 Moisture sensors were
conductivity is 3.8 × 10−4 m/s, which is equivalent to that installed in the testing box to monitor the drainage
of a coarse sand. In other words, the wicking geotextile performance of different materials. In total, three layers
can provide an adequate cross-plane permeability. On the of soil moisture sensors were installed at depths of 10 cm,
other hand, the wicking geotextile is also capable of soil 24 cm, and 38 cm from the bottom, as shown in
retention. According to AASHTO (1991), for soil with Figure 2(a). The first layer of sensors (10 cm from the
bottom) was installed at the level of the opening and
immediately above the geosynthetics (if there was). In each
Table 1. Wicking geotextile specification test section, there were three sensors in each layer. As a
result, in total nine sensors were installed in each test
Mechanical properties Test method Unit Minimum section as shown in Figure 2(a). Within each layer, the
average
spacing between two adjacent sensors was 28 cm, and the
roll value
sensors at the edges were 5 cm away from the sidewall.
Apparent opening size ASTM D4751 mm 0.4 Note that the accuracy of the sensor was ±2% according
(AOS) to the manufacture’s manual. No corrections were made
Pore size (O50) ASTM D6767 Microns 85 to EC-5 sensors and all the reported data were raw
Pore size (O95) ASTM D6767 Microns 195
volumetric moisture from the EC-5 sensors. After the
Permittivity ASTM D4491 s−1 0.24
Flow rate ASTM D4491 L/min/m2 1222 test had been completed, 54 samples were taken from
each testing box. The volumetric water content was back
Geosynthetics International, 2022, 29, No. 3

Downloaded by [ Indian Institute Of Technology - Bombay] on [07/03/23]. Copyright © ICE Publishing, all rights reserved.
258 Guo, Lin, Leng and Zhang

66 66
5
20 28 28
5 12 15 12 15 12

10
M7 M8 M9
14
60 M4 M5 M6
14
M1 M2 M3
10

Opening
Geosynthetics
Box separator Geosynthetics
Moisture content sensor
Sand
Geosynthetics Valve
Sampling point
Water pump Water tank

(a)

Moisture content sensors


Test box

Location of drainage materials Opening


Opening

Valves

Water supply pipes

Switch of pump
Geosynthetics
Geosynthetics
Water tank

Balance
Balance

(b)

Figure 2. Test apparatus: (a) schematic plot (unit in cm); and (b) configuration

calculated based upon the measured gravimetric water system without any geosynthetics. It is important to point
content and the specific gravity of the sand. It was found out again that sand is purposely selected to demonstrate
that the monitored and calculated results matched well that a coarse-grained soil can still hold a large amount of
with each other, indicating that sensor readings were capillary water under unsaturated conditions. T2 was
reasonable and accurate. This can be considered as an an efficient drainage material in removing ‘free’ water
indirect calibration. The data acquisition system was and was commonly used in dealing with the poor
composed of a Campbell Scientific® CR1000 datalogger drainage condition of existing railways in China (Guo
together with an AM16/32 multiplexer. The soil volu- et al. 2018). Lin and Zhang (2018a, 2018b) reported
metric water contents were monitored and recorded every that the wicking geotextile (T3) could be used to drain
2 min. both gravitational and capillary water. T4 was used to
demonstrate the influence of the number of wicking fiber
2.4. Experimental design yarns on the drainage efficiency. Each test had three
Drainage tests were performed in the testing box under replicates to ensure the consistency of the test and
four different conditions, as shown in Figure 2. Each representative test results are reported in the following
time two materials could be compared side by side. Three sections.
groups of comparison tests were performed, including
the sand without any drainage material (T1), with a 2.5. Testing procedures
drainage belt (T2), with a wicking geotextile (T3), and All tests were carried out in a controlled temperature and
with a modified wicking geotextile (T4). T1 was used as a humidity room. The relative humidity varied from 48% to
control case to evaluate the performance of a drainage 56%, with an average value of 52% and the temperature
Geosynthetics International, 2022, 29, No. 3

Downloaded by [ Indian Institute Of Technology - Bombay] on [07/03/23]. Copyright © ICE Publishing, all rights reserved.
Laboratory evaluation of different geosynthetics for water drainage 259

varied from 19°C to 22°C, with an average value of 3. RESULTS AND DISCUSSIONS
21°C. Each test had three different stages. At the
saturating stage, the soils in the entire testing system In this section, discharge capacity, volumetric water
were saturated. During this stage, water was slowly content, gravimetric water content variations, and the
supplied to the testing box through two valves at the water retention curves of different testing materials are
bottom of the box (Figure 2) and gradually saturated soils discussed.
from the bottom to the top. The water flow rate was
3.1. General visual observations for water drainage
carefully controlled to prevent an upward seepage force
from disturbing the tested soil. The soil were considered During the test, the water drainage process was visually
fully saturated after a 10 cm water head was ponding observed and summarized as follows. In general, water
above the soil surface. In order to achieve a full saturation drained out of the system within the first 2 h for T1 and
for T1, the opening on the sidewall was blocked in T1 and T2 when the flow of ‘free’ water dominated. However,
no water was allowed to drain out of the system. Since after that the water flow quickly reduced as the soil
water can be drained out from the openings in T2, the became unsaturated after 2–3 h, no measurable amount
drainage belt was lifted up to minimize the amount of of water could be collected from T1 and the draining
water flowing out so that the sand in T2 could be process ceased at this moment. In comparison, ‘free’ water
gradually saturated. For T3 and T4, the wicking geotex- could be discontinuously collected from T2 for 6838 min
tiles could be left as they were during the saturation (4.7 days), indicating that T2 was partially working under
process, as shown in Figure 2(b). unsaturated condition. In contrast, water could be
After the saturation process was completed, the valves continuously collected from T3 and T4 for approximately
at the bottom of the testing box were shut down and the 4.9 days and then no visible water could be collected after
second stage of free draining started. Water was allowed to that. However, the wicking geotextile was still damp 10
be drained out of the system and the amount of ‘free’ days after the test, indicating that the wicking geotextile
water was collected during the draining process. Then, the was still functional and was able to drain both ‘free’ and
top of testing box was covered by a plastic wrap to prevent capillary water.
any possible water loss through the evaporation process. Figure 3 shows representative photos for tests on T1,
The drainage materials (if there was) were naturally hung T2, T3, and T4 after 10 days of drainage before being
downwards under the gravitational influence. The water disassembled. As discussed previously, the testing box
collection process started when water above the soil allowed testing of two different materials side by
surface was at a level of 3 cm. The collected water was side. Figures 3(a)–3(c) show the comparisons on T1–T2,
weighed continuously using a digital balance until no free T2–T3, and T3–T4, respectively. To facilitate obser-
water flowed out from the system. This process normally vations, locations of the drainage materials and wetting
took about six days. fronts were also labeled by solid and dotted lines,
The second stage of the test was completed when no respectively. Figure 3 clearly indicates that the geosyn-
visible water drops could be observed. At the third stage of thetic drainage materials showed different performances.
draining capillary water, the T2, T3, and T4 (if there was) When there were no drainage materials (T1 with sand
were elevated to the horizontal direction and exposed to a only), the final wetting front was about 8 cm above the
room atmosphere to allow further water loss through elevation of the opening, indicating that the capillary rise
evaporation. A strip of the geosynthetic was elevated to a could be as high as 8 cm. For T2 and T3, the final wetting
horizontal direction and a chair was used to hold the fronts were 5 and 2 cm above the opening elevation after
exposed end of the geosynthetic in its horizontal position. 10 days of drainage, respectively. For T4, the final wetting
The third stage was completed when all the geotextiles front was 2 cm below the opening elevation, indicating
were dried to a distance of 20 cm to the openings. This that the modified wicking fabric can siphon water from
process normally took about another 4.9 days. During the the underlying soil.
entire three draining stages, soil volumetric water contents
were monitored every 2 min. 3.2. Variations of volumetric water content at
After all three stages were completed, the testing different locations
system was disassembled and soils at different locations Figure 4(a) shows variations of volumetric water content
were carefully sampled to measure their gravimetric with time in a log scale at sensor M8, which is located at
water contents using the oven method. Solid points 10 cm from the soil surface. The soil was initially dry with
in Figure 2(a) show the locations of the sampling a volumetric water content of nearly zero for all cases.
process. From the top to the bottom, a total of After water was supplied to saturate the soil, the
nine layers of soils were sampled with an interval of volumetric water contents of the soil quickly increased
6 cm. In each layer, six locations were selected for to a value ranging between 0.355 and 0.362. When the soil
sampling and the sampling distance between two adjacent was allowed to drain, the volumetric water contents
locations were 12, 15, 12, 15, and 12 cm (Figure 3a), quickly dropped to approximately 0.074 in less than 2 h.
respectively. As a result, a total of 54 samples were taken Among the four tests, water was drained faster in T1 (sand
for each test section and a total of 108 gravimetric water only) and T2 (drainage belt) compared with that in T3 and
contents were measured for the whole testing box for two T4. It seemed that T4 was the slowest in draining water.
test sections. The volumetric water contents in all tests kept dropping
Geosynthetics International, 2022, 29, No. 3

Downloaded by [ Indian Institute Of Technology - Bombay] on [07/03/23]. Copyright © ICE Publishing, all rights reserved.
260 Guo, Lin, Leng and Zhang

Wetting front
Opening

8 cm
5 cm
T2 (drainage belt) T1 (control)
10 cm Opening

(a)

Wetting front
Opening Opening
2 cm
5 cm
T2 (drainage belt)
10 cm T3 (wicking geotextile)

(b)

Wetting front
Opening 2 cm Opening
2 cm

T3 (wicking geotextile)
T4 (modified wicking geotextile)
10 cm

(c)

Figure 3. Wetting fronts after 10 days of drainage: (a) drainage belt versus control case; (b) drainage belt versus wicking geotextile; and
(c) modified wicking geotextile versus wicking geotextile

slowly after 2 h and nearly all reached the same value of drainage belt (T2) was similar to T1, it was clear that more
0.067 after 10 days of free drainage. water can be drained by T2. In less than 2 h, the
Figure 4(b) shows variations of volumetric water volumetric water content dropped from a saturation of
content with time at sensor M5, which was at 24 cm 0.369 to 0.083 and the final water content was 0.074. In
from the soil surface. When the soil was allowed to drain, comparison, water drainage in T3 and T4 was much
water in T1 drained much faster than that in other cases. slower. It took about 0.69 days for the volumetric water
In about 30 min, the volumetric water content content to drop from saturation of 0.369 to approximately
dropped from 0.369 to 0.15 and remained relatively 0.074 and remained nearly constant after that. After 10
constant after that. The draining process after 2 h days, the volumetric water content in T4 was 0.082,
was very slow and at the end of 10 days, the final water slightly higher than that in T2 and T3 at the same
content was 0.111. Although water drainage at M5 in the location.
Geosynthetics International, 2022, 29, No. 3

Downloaded by [ Indian Institute Of Technology - Bombay] on [07/03/23]. Copyright © ICE Publishing, all rights reserved.
Laboratory evaluation of different geosynthetics for water drainage 261

0.40 sensors were slightly above the geosynthetics. As can be


Saturation Drainage process Sand (T1)
0.35 process Drainage belt (T2) seen from Figure 4(c), the volumetric water contents in T1
Wicking geotextile (T3)
Volumetric water content (m3/m3)

0.30 0.362 Modified wicking geotextile (T4) and T2 remained relatively constant during the whole
M8 testing period except for a slight decrease within the
0.25
first two hours. Visual observations during the testing
M5
0.20
period and disassembling process also confirmed that the
0.15 M2 soil was fully saturated. The saturated volumetric water
0.10
0.067
content of the soil was 0.354 for T1 and this value will be
0.05 used as a reference for further discussions. Although water
2 hours 1 day 7 days
drainage at M2 in the drainage belt (T2) was similar to
0
10 100 1000 10 000 that of T1, it was clear that more water can be drained by
Testing time (min) T2. Volumetric water contents in T3 and T4 varied with
(a) time similar to that in T1 and T2 for a few days at the
0.40
Saturation Drainage process Sand (T1)
beginning and remained saturated. After 5.5 and 2.5 days,
Volumetric water content (m3/m3)

0.35 process Drainage belt (T2)


Wicking geotextile (T3) respectively, the volumetric water contents in T3 and T4
Modified wicking geotextile (T4)
0.30 0.369 started to decrease quickly, indicating the soil was losing
M8

0.25
water. After 10 days, the volumetric water content in T3
M5
and T4 was 0.28 and 0.124, respectively. The trends did
0.20
M2 not flatten out even at the end of 10 days. Although there
0.15
0.111 was no visible water droplet during this drainage stage,
0.10 0.082
visual observations indicated the exposed strips in T3 and
0.05 0.074 T4 remained damp after 10 days. This fact implies that
2 hours 1 day 7 days
0 water can still be drained out of the system if the testing
10 100 1000 10 000
period lasts for a longer time. Moreover, the volumetric
Testing time (min)
(b)
water content in T4 was lower than that for T3 and the
drainage performance of the modified wicking geotextile
0.40
Saturation process Drainage process 0.354 was better than that of the wicking geotextile. There are
Volumetric water content (m3/m3)

0.35 two possible explanations for the performance differences


0.372 0.348
0.30 in T3 and T4. First, there are 96 wicking fibers in T3 and
0.25 0.28 128 wicking fibers in T4. Because the wicking fiber had a
0.20
2 hours 1 day 7 days large number of channels per fiber, more wicking fibers
resulted in a greater potential for maximizing capillary
0.15 M8

0.124
action and more water being transported under an
0.10 M5
Sand (T1) unsaturated environment. Second, the SEM image from
Drainage belt (T2)
0.05 M2
Wicking geotextile (T3) a previously published paper (Lin et al. 2017) showed that
Modified wicking geotextile (T4)
0 permanent deformation was observed under the overlap-
10 100 1000 10 000
Testing time (min)
ping area in T3. In contrast, the additional wicking fiber
(c) yarn was artificially woven into the geotextile in T4 and
did not suffer from any mechanical force. This fact might
0.40
Saturation process Drainage process be another contributor that caused such a performance
0.35
difference in T3 and T4.
Volumetric water content (m3/m3)

0.30 0.30 Figure 4(d) shows the comparisons of the M2 sensor for
0.25 0.28 repeated tests on T3 and T4. After 10 days, the volumetric
0.20
2 hours 1 day 7 days water contents in T3 and T4 were 0.30 and 0.163,
0.163
respectively. Although the volumetric water contents in
0.15 M8
0.124
T3 and T4 were not the same (Figure 4(c)), the
0.10 M5
T3 phenomenon that the water content in T3 was less than
T4
0.05 M2
Repeated test on T3 that of T4 did not change, which indicated that the
Repeated test on T4
0 experiment could be repeatable.
10 100 1000 10 000
Testing time (min) 3.3. Cumulative amount of water
(d)
Figure 5(a) shows the variations of the cumulative amount
Figure 4. Monitored volumetric water content variations with time of water collected from stages 2 and 3. The cumulative
under different testing conditions: (a) top (M8); (b) middle (M5); amount of water collected from T1 was 17.34 l within 2 h,
(c) bottom (M2); and (d) bottom (M2) (repeated test for M2) slightly increasing to 17.67 l at 4620 min (3.2 days), and
no measurable amount of water could be collected
afterward. This phenomenon indicated that the drainage
Figure 4(c) shows the variations of volumetric water of free water stopped at 2 h and T1 was not able to drain
content with time at sensor M2, which was exactly at the capillary water. For T2, which was installed with a
level of the openings. If there were geosynthetics, the drainage belt, the cumulative amount of water collected
Geosynthetics International, 2022, 29, No. 3

Downloaded by [ Indian Institute Of Technology - Bombay] on [07/03/23]. Copyright © ICE Publishing, all rights reserved.
262 Guo, Lin, Leng and Zhang

24 cumulative amount of water for T2, T3, and T4 was


2 hours 1 day
9.45%, 14.54%, and 21.51%, respectively, more than for
Cumulative amount of water (L)

20
T1. The high-to-low rank of the overall drainage per-
16 formance was T4, T3, T2, and T1.
Figure 5(b) shows an enlargement of Figure 5(a) for the
12 time ranging from 0 to 200 min. The drainage of ‘free’
8
water was fast at the beginning of the test for all materials
Sand (T1) and the slopes of collected water versus time curves (flow
Drainage belt (T2)
4 Wicking geotextile (T3) rates) were steep. At the beginning of the test (0–25 min),
Modified wicking geotextile (T4)
the cumulative amount of ‘free’ water from T1 (17.03 l)
0
0 1000 2000 3000 4000 5000 6000 7000 8000 was higher than those from T2 (17.0 l), T3 (8.60 l), and
Texting time (min) T4 (10.01 l). This was mainly attributed to its larger
(a) cross-sectional area that could be used for ‘free’ water
drainage. For all the other materials, part of the openings
24
2 hours
was occupied by the drainage materials. As a result,
Cumulative amount of water (L)

20 although the area of the opening on the sidewalls of the


box was the same, the equivalent drainage areas for all
16 other geosynthetics were smaller compared with T1. After
12
about 26 min, the water collected in T2 (17.04 l) was the
same as that collected in T1 and exceeded the cumulative
8 amount of water by T1 after that. At the same time, the
Sand (T1) cumulative amount of water in T3 and T4 was 8.85 l and
4 Drainage belt (T2) 10.24 l, respectively. This was mainly attributed to the
Wicking geotextile (T3)
0 Modified wicking geotextile (T4) larger pores in the drainage belt (compared with T3 and
0 50 100 150 200
T4) and its direct contact with the soil from inside
Texting time (min)
(compared with T1), which could facilitate the water
(b) drainage. After 128 and 105 min, the cumulative amount
of water by T3 and T4 were the same as that collected by
Figure 5. Cumulative amount of water under different testing T1. T3 and T4 had the same cumulative amount of water
conditions: (a) 0–8000 min; (b) 0–200 min compared with T2 at 164 and 141 min, respectively.

3.4. Gravimetric water content after 10 days


from the system was 17.72 l within 2 h, slightly increasing In addition to monitoring the variations in terms of
to 18.1 l after one day, and finally ending up with 19.3 l at volumetric water content, gravimetric water contents were
6838 min (4.7 days). Note that starting from 5358 min determined at the end of the test by sampling the soil and
(3.7 days), no continuous water flow was observed at the oven-drying the samples during the disassembling process.
drainage opening. The cumulative amount of water was The gravimetric water content contours were to observe
discontinuous, implying that the drainage belt was the water content distributions below the drainage open-
partially functional in draining water out of the system. ings where sensors could not cover. For each test, 54
The total cumulative amount of water from T2 was 9.5% sampling points were selected (Figure 2(a)) in each test
higher than that from T1, indicating that the drainage belt section with 6 cm per lift and 12–15 cm apart.
was more effective than the control case in draining ‘free’ Figure 6 shows the gravimetric water content contours
water. For T3, which was installed with the wicking for T1–T4 after 10 days. The dash lines represent
geotextile, the cumulative amount of water was 17.2 l elevations of drainage openings for T1 or locations of
within 2 h, gradually increasing to 19.6 l after 1 day, and drainage materials for T2–T4. The saturated gravimetric
eventually reached 20.2 l at 4.9 days. The cumulative water content of the soil is 27% and will be used as a
amount of water from T3 was 14.5% higher than that reference for further discussions. As shown in Figure 6(a),
for T1 and 4.7% higher than T2. This means that the the gravimetric water content was lower than 3% for soil
wicking geotextile was able to drain out more water if with elevations of 24 cm and above. For soils between
the testing time was long enough. In other words, the 10 cm and 24 cm, gravimetric water contents decreased
overall drainage performance of the wicking geotextile with increasing elevations. The gravimetric water content
was better than that of the drainage belt. For T4 which contour for T2 was shown in Figure 6(b). The gravimetric
was installed with the modified wicking geotextile, the water content was lower than 3% for soils with elevations
cumulative amount of water was 17.5 l within 2 h, of 22 cm and above. The soil gravimetric water contents
increasing to 20.8 l after 1 day, and reaching 21.5 l at between 10 cm and 22 cm were lower in T2 than in T1.
4.9 days. The additional wicking fiber yarn did not aim to Test results indicated that the average gravimetric water
increase its drainage efficiency for ‘free’ water. The content in T2 was lower than in T1 and more water could
additional wicking fiber yarn was effective in draining be drained out for the system installed with a drainage
capillary water, resulting in 6% more water collected from belt. Figure 6(c) shows the gravimetric water content
T4 than in T3 after 4.9 days. In summary, the total contour for T3, which was installed with a wicking
Geosynthetics International, 2022, 29, No. 3

Downloaded by [ Indian Institute Of Technology - Bombay] on [07/03/23]. Copyright © ICE Publishing, all rights reserved.
Laboratory evaluation of different geosynthetics for water drainage 263

48 27

24
40
21

32 18
Vertical (cm)

15
3
32
3
6 12
9
12 15 6
18 9 12 9
16
21 15
18 21
24 24
27 27 6
8
3

0 0
0 11 22 33 44 55 66 0 11 22 33 44 55 66
Horizontal (cm) Horizontal (cm)
(a) (b)

48 27

24
40
21

32 18
Vertical (cm)

15
32
12

3
16 9
6 3
9 6
12 9 6
15 12 15
8 18 21 18
24 21
24 3

0 0
0 11 22 33 44 55 66 0 11 22 33 44 55 66
Horizontal (cm) Horizontal (cm)
(c) (d)

Figure 6. Gravimetric water content contours after the test: (a) control case (T1); (b) with drainage belt (T2); (c) with wicking geotextile
(T3); and (d) with modified wicking geotextile (T4) (unit in %)

geotextile. In contrast with T1 and T2, the 3% water a wicking geotextile. Since all the soil underlying the
content contour line for T3 moved 6 cm downward to an wicking geotextile (modified wicking geotextile) was
elevation of 18 cm. Moreover, the soils underlying the under unsaturated conditions, the effective range of the
wicking geotextile were unsaturated and the highest water siphon effect was at least 10 cm.
content was 24%. This phenomenon indicated that the
wicking geotextile was able to wick the water from the 3.5. Water retention curves for different materials
underlying soil through a siphoning effect. Capillary The water storage capacity of a material could be
water could be transported along the wicking geotextile demonstrated by the relationship between water content
and eventually wicked out of the system. Figure 6(d) and suction, referred to as the Water Retention Curve
shows the gravimetric water content contour for T4, which (WRC). Several techniques have been performed to
was installed with a modified wicking geotextile. The 3% obtain the WRCs for the materials used in this study,
water content contour line moved further downward to including capillary rise test (suction ≤ 10 kPa), pressure
14 cm, Meanwhile, gravimetric water content distri- plate test (suction ≤ 1500 kPa), and salt concentration
butions below the wicking geotextile for T4 were also test (suction > 1500 kPa). To reduce the complexities and
lower than that for T3, in which the soil elevation was variations of the test results, only drying WRC curves are
between 6 cm and 10 cm. For the soil elevation lower than presented in this paper. Figure 7 presents WRCs for sand
6 cm, the water content was 24%, which was the same as (T1), drainage belt (T2), wicking geotextile (T3), modified
in T3. In conclusion, the drainage performance of a wicking geotextile (T4), wicking fibers, and T3 without
modified wicking geotextile was better than that of wicking fibers. The test results were fitted using Fredlund
Geosynthetics International, 2022, 29, No. 3

Downloaded by [ Indian Institute Of Technology - Bombay] on [07/03/23]. Copyright © ICE Publishing, all rights reserved.
264 Guo, Lin, Leng and Zhang

and Xing’s (1994) equation and the regression curves are 54.51%, respectively) when compared with a sand (27%).
also shown in Figure 7. The WRC of the tested sand was Owing to the existence of wicking fibers, both a wicking
flat in the suction range of 0–1 kPa and then showed a geotextile and a modified geotextile could hold a certain
sharp reduction to nearly zero as the suction value amount of water up to a suction value of 600 kPa.
increased from 1 kPa to 4 kPa. Essentially, the drainage Table 2 lists the values of the model parameters for
belt could not hold any water under unsaturated con- different materials for Fredlund and Xing (1994)’s
ditions, and its water content became zero when the equation. The air-entry values for the sand, the drainage
suction value was less than 0.2 kPa. On the other hand, belt, and the geotextile without wicking fibers were
the wicking fibers had a saturation water content of 80%. 1.65 kPa, 0.13 kPa, and 0.80 kPa, respectively. In com-
Its water content significantly decreased for suctions parison, the wicking geotextile and the modified geotex-
ranging between 1 kPa and 30 kPa. Owing to the tiles have two different levels of pore sizes, resulting in two
existence of multiple micro-channels, the water content different air-entry values. The lower air-entry values are
of wicking fibers could be 11.36% when the suction was related to pore sizes in the macro-channels of the wicking
800 kPa. For comparison purposes, the WRC for a fibers. The air-entry values (AEVs) were 0.88 and
wicking geotextile without wicking fibers was also 0.89 kPa for the wicking geotextile and the modified
determined using the same wicking geotextile except wicking geotextile, respectively, when suction values were
that the wicking fibers were removed. The WRC of a less than 50 kPa. Air-entry values were 254 and 300 kPa
wicking geotextile without wicking fibers was similar to for the wicking geotextile and the modified wicking
that of sand but with two minor differences: (1) the geotextile, respectively, if suction values were greater
saturation water content was slightly higher, and (2) the than 50 kPa. The structural differences of the geosyn-
water content quickly reduced to zero within the suction thetics and their water retention characteristics had
range of 0.3–3 kPa, indicating that a wicking geotextile significant influences on their drainage performances.
without wicking fibers had a slightly larger pore size and Figure 8 shows suction distributions with elevations in
the corresponding air-entry value. The WRCs for a T1–T4 after 10 days calculated from the WRCs in Figure 7
wicking geotextile and a modified geotextiles had and water content distributions as shown in Figure 6. The
similar shapes. Both of them had a much higher ability drainage materials were installed at an elevation of 10 cm
to hold water under saturated conditions (52.8% and (horizontal dashed line). The water table for T1 and T2 was

90
Sand (T1)
80 Drainage belt (T2)
Wicking geotextile (T3)
70 Modified wicking geotextile (T4)
Gravimetric water content (%)

Wicking fibers
60 T3 without wicking fibers

50

40

30

20

10

0
0.1 1 10 100 1000 10 000 100 000
Suction (kPa)

Figure 7. WRCs of different testing materials

Table 2. WRC parameters for different drainage materials

Parameter Sand Drainage Wicking geotextile (T3) Modified wicking geotextile (T4) Wicking fibers T3 without
(T1) belt (T2) wicking fibers
Suction Suction Suction Suction Suction Suction
≤ 50 kPa > 50 kPa ≤ 50 kPa > 50 kPa ≤ 300 kPa > 300 kPa

ws (%) 27 9 52.8 5 54.51 8 80.99 11.36 50.58


ψr (kPa) 2.4 1 7 600 8 800 23 800 3
af (kPa) 1.65 0.13 0.88 254 0.89 300 1.76 300 0.8
nf 7.37 20 2.19 1.62 0.65 1.15 1.8 0.3 3
mf 1.69 30 0.99 1.03 3.68 0.86 0.78 0.2 2

Geosynthetics International, 2022, 29, No. 3

Downloaded by [ Indian Institute Of Technology - Bombay] on [07/03/23]. Copyright © ICE Publishing, all rights reserved.
Laboratory evaluation of different geosynthetics for water drainage 265

48
Sand (T1)
Drainage belt (T2)
Wicking geotextile (T3)
Modified wicking geotextile (T4)
36

Steady state
Elevation (cm)
24

12 Location of drainage material

0
–1 0 1 2 3 4 5
Suction (kPa)

Figure 8. Comparisons of suction distributions with elevations in T1–T4 after the test

at 10 cm and soil suctions were linearly distributed at a zone, air bubbles were occluded and the air phase was
steady state. As shown in Figure 8, neither T1 nor T2 had discontinuous. As the soil became further desaturated, the
reached a steady state within 10 days. The positive pore hydraulic conductivity of the soil significantly decreased
water pressure for T1 and T2 indicated that the soil was due to the existence of air bubbles within voids. This zone
submerged in water with elevations less than 10 cm. For the corresponded to a suction range from AEV to the residual
soil between 10 cm and 42 cm, suction distributions suction on the WRC. Moving upward to the residual zone,
deviated from the steady state for both T1 and T2, the air phase became continuous while the water phase
indicating that the water flow was in a transient state. For became coated outside soil particles, making it very
the soil with elevations higher than 42 cm, a steady state difficult for water to flow. On the WRC, this zone referred
has been achieved. In comparison, suction distribution to the curve where suction values were greater than the
curves for T3 and T4 were positioned at the right side of T1 residual suction.
and T2, indicating the soils in T3 and T4 were drier. In Figure 9(a) showed the drainage process for T1, which
other words, the drainage performance of a (modified) was the control case without installation of any drainage
wicking geotextile was better than that of a drainage belt. material. Water could be drained out via the drainage
In addition, the soils underlying geotextiles were under opening at the left side of the system. The area of the
unsaturated conditions in T3 and T4. This phenomenon opening for T1 was the biggest among all four cases,
indicated that a wicking geotextile was able to absorb water resulting in the highest cumulative volume of collected
from the underlying soil via a siphon effect and the effective water for T1 within the first 12 min (refer to Figure 5(b)).
range was at least 10 cm. Visual observations showed that The soil below the elevation of the opening remained
the wicking geotextile was still damp at the end of the test, saturated throughout the test and three distinct zones built
indicating that the wicking geotextile was still working. up as excess water was drained out of the system. From the
Given a longer testing time, the wicking geotextile was WRC of sand, the AEV was 1.2 kPa indicating that air
expected to drain more water out of the system and bubbles could be easily entered into voids and block the
advantages of a wicking geotextile against a drainage belt water flow. Therefore, the cumulative volume of water for
would be more obvious. T1 only increased from 17.15 l in 32 min to 17.67 l in
4620 min (Figure 5). According to Figure 6(a), the
3.6. Working mechanisms for different materials average water content for T1 after 10 days was 10.81%,
Based upon the laboratory test results, working mech- which was the highest of the all cases. In summary, T1
anisms of different types of drainage materials could relied on the soil itself to drain ‘free’ water and was not
be schematically explained as in Figure 9. During the able to drain capillary water.
laboratory test, the representative soil profile could be Figure 9(b) shows the working mechanism for T2,
divided into four zones as shown in Figure 9(a): saturated which was installed with the drainage belt. At the
zone, capillary zone, transition zone, and residual zone. beginning of the test, the cumulative amount of collected
Below the water table, the soil was saturated, with the water for T2 was smaller than that for T1 because the
degree of saturation being 100%. When entering the opening on the sidewall was partially blocked by the
capillary zone, the degree of saturation was close to 100% drainage belt. As the overlying soil became desaturated,
and the pore water pressure was negative. This zone the advantages of the drainage belt were observed. In the
corresponded to a suction range from 0 kPa to the AEV lateral direction, grooves within the drainage belt became
on the WRC (Figure 7) where geomaterials were able to the shortest drainage path for water to flow from right to
hold a large amount of water. For a soil at the transition left. The ‘free’ water would preferably flow into the
Geosynthetics International, 2022, 29, No. 3

Downloaded by [ Indian Institute Of Technology - Bombay] on [07/03/23]. Copyright © ICE Publishing, all rights reserved.
266 Guo, Lin, Leng and Zhang

Air Residual zone


Water
Particle Transition zone

Capillary zone

Saturated zone

(a)

Residual zone

Transition zone

Capillary zone

Drainage belt
Saturated zone

(b)

Residual zone

Suction > 53.7 MPa


(RH < 56%)
Transition zone

Wicking geotextile
(modified wicking geotextile) Capillary zone

Saturated zone

(c)

Figure 9. Working mechanisms for different drainage materials: (a) control case (T1); (b) drainage belt (T2); and (c) (modified) wicking
geotextile (T3 and T4)

grooves first and then laterally flowed out of the system. This explanation was validated by the observation
As shown in Figure 5(b), the cumulative amount of during the test of which measurable amount of water
collected water from T2 surpassed T1 in approximately could be discontinuously collected from T2 from
30 min after the test. The AEV of the drainage belt was 4569 min to 6838 min (Figure 5a). It was important to
small (0.2 kPa) and air could easily enter into the grooves note that the soil underlying the drainage belt was still
to block the water flow when suction was higher than the saturated throughout the test and no capillary water could
AEV. Meanwhile, the water within the transition and be drained out by the drainage belt. According to
capillary zones gradually flowed downward under the Figure 6(b), the average gravimetric water content for T2
influence of gravitational force. The pore water pressure after 10 days was 9.86%.
for the soil immediately above the drainage belt would Figure 9(c) shows the working mechanism for T3/T4 in
gradually build up from a negative value (suction) to a which the (modified) wicking geotextile were installed. In
positive value as the degree of saturation continued to contrast to the drainage belt, the wicking fibers had
increase. At a critical moment when the positive pore multichannel cross-sections (Figure 1(b)) that were able to
water pressure was higher than the AEV of the drainage hold and transport water under unsaturated conditions.
belt, the air within the grooves would drain out. At the beginning of the test, the drainage performance of
Geosynthetics International, 2022, 29, No. 3

Downloaded by [ Indian Institute Of Technology - Bombay] on [07/03/23]. Copyright © ICE Publishing, all rights reserved.
Laboratory evaluation of different geosynthetics for water drainage 267

the wicking geotextile was not as effective as the drainage tested sand could be reduced from 10.81% to 9.86%
belt in terms of draining ‘free’ water due to the relatively by installing a layer of the drainage belt. The average
small pore sizes within the weaving structure. The gravimetric water content was reduced by 8.78%.
cumulative amount of collected water for T3 and T4 (d) In contrast to the drainage belt, the wicking fibers had
was lower than that for T1 and T2 at the beginning of the a much higher AEV (254 kPa), which enabled the
test (Figure 5(b)). As the soil further desaturated, the wicking geotextile to hold and transport water under
advantages of the wicking geotextile to drain capillary unsaturated conditions. The wicking geotextile could
water were observed. Comparison of the AEVs of the soil drain both ‘free’ and capillary water, and the average
(1.65 kPa) and the wicking fibers (254 kPa) (Figure 7) gravimetric water contents could be further reduced to
indicated that the ability of the wicking geotextile to hold 6.89% (with wicking geotextile) and 6.67% (with
water was much stronger than that of the soil. In other modified wicking geotextile), respectively. The
words, the wicking geotextile had much a stronger ability corresponding gravimetric water contents were
to absorb water from the surrounding soils. This fact also reduced by 36.20% and 38.27%, respectively.
explained the observation (Figures 6(c) and 6(d)) that the (e) The wicking geotextile was able to wick water from
soils underlying the (modified) wicking geotextile were both the overlying and the underlying soils. The
unsaturated. It was the siphon effect (caused by different influencing range of the siphon effect was at least
AEVs of the soil and the wicking fiber) that enabled the 10 cm.
wicking geotextile to absorb water from the surrounding (f ) Increasing the amount of wicking fibers would
soil to the grooves within the fibers. In addition, relative enhance its ability to drain capillary water and
humidity (RH) at the exposed end of the wicking therefore increase the total amount of water drained
geotextile was lower than 56% (the corresponding out of the system.
suction >53.7 MPa) while the relative humidity at the (g) The working mechanisms of the different drainage
other end was close to 100% (suction approximately materials could be explained by their WRCs and
0 kPa). The significant suction gradient was the driving depended greatly on their AEVs. A higher AEV
force that continuously wicked the water from the system value represented a higher ability of the geomaterial
in the horizontal direction. Meanwhile, the exposed end to hold and transport water and a higher capability
of the wick geotextile was damp throughout the test and to dehydrate the soil.
water could be continuously vaporized to the ambient
environment via the evaporation process. Because capil- In summary, the drainage belt showed better performance
lary water within the testing apparatus could be drained in draining ‘free’ water but could not drain capillary water.
out, the water tables for T3 and T4 were much lower than The wicking geotextile could drain both gravitational and
those for T1 and T2. According to Figures 6(c) and 6(d), capillary water and showed the best drainage performance
the average gravimetric water contents for T3 and T4 of all the drainage materials.
after the test were 6.89% and 6.67%, respectively. In
summary, the ambient environment worked as a ‘pump’
and the wicking geotextile worked as ‘pipes’ that continu-
ACKNOWLEDGMENTS
ously wicked both gravitational and capillary water out of
the soil and made the soil drier with time. This research was supported by the China Scholarship
Council, China (201706370160). The authors sincerely
thank TenCate Geosynthetics (North American) for
providing the materials used in the study.
4. CONCLUSIONS
This paper studied the drainage performance and working
mechanisms of three types of drainage materials, includ-
NOTATION
ing a drainage belt, a wicking geotextile, and a modified
wicking geotextile. The major conclusions are summar- Basic SI units are given in parentheses.
ized as follows:
AEV air entry value (Pa)
(a) Capillary water within the soils could not be drained AOS apparent opening size (m)
out by conventional drainage systems. For the tested af fitting parameter that is primarily a function of
uniform sand, which was considered as a good air-entry value (Pa)
drainage geomaterial, the gravimetric water content Cc coefficient of gradation (dimensionless)
was still 10.81% 10 days after the test. Cu uniformity coefficient (dimensionless)
(b) The microstructures of the geosynthetics Gs specific gravity (dimensionless)
significantly influenced the drainage performance. H height (m)
(c) The drainage belt was able to drain more ‘free’ water mf fitting parameter that is primarily a function of
out of the system compared with the control case. residual water content (dimensionless)
However, owing to the very small AEV value nf fitting parameter that is primarily a function of
(0.2 kPa), it was not able to drain any capillary rate of water extraction from soil once air-entry
water. The average gravimetric water content for the value has been exceeded (dimensionless)
Geosynthetics International, 2022, 29, No. 3

Downloaded by [ Indian Institute Of Technology - Bombay] on [07/03/23]. Copyright © ICE Publishing, all rights reserved.
268 Guo, Lin, Leng and Zhang

O50 geotextile opening size of which 50% of the pores Christopher, B. R. & McGuffey, V. C. (1997). Pavement Subsurface
are smaller than this size (m) Drainage Systems, NCHRP Synthesis of Highway Practice 239.
Transportation Research Board, National Research Council,
O95 geotextile opening size of which 95% of the pores Washington, DC, USA.
are smaller than this size (m) Christopher, B. R., Schwartz, C. W. & Boudreau, R. L. (2010).
RH relative humidity (dimensionless) Geotechnical Aspects of Pavements: Reference Manual, US
W width (m) Department of Transportation, Federal Highway Administration.
ws saturated water content (dimensionless) Fredlund, D. G. & Xing, A. (1994). Equations for the soil-water
characteristic curve. Canadian Geotechnical Journal, 31, No. 4,
ψr suction at residual condition (Pa) 521–532.
Garcia, E. F., Gallage, C. P. K. & Uchimura, T. (2007). Function of
ABBREVIATIONS permeable geosynthetics in unsaturated embankments subjected to
rainfall infiltration. Geosynthetics International, 14, No. 2, 8999.
AASHTO American Association of State Highway and Ghosh, C. & Madhav, M. R. (1994). Reinforced granular fill-soft soil
system: membrane effect. Geotextiles and Geomembranes, 13,
Transportation Officials 743–759.
ASTM American Society for Testing and Materials Gobel, C. H., Weisemann, U. C. & Kirschner, R. A. (1994). Effectiveness
DG deep groove of a reinforcing geogrid in a railway subbase under dynamic loads.
GCBD geocomposite capillary barrier drain Geotextiles and Geomembranes, 13, No. 1, 91–99.
NCHRP National Cooperative Highway Research Gray, D. H. & Al-Refeai, T. (1986). Behavior of fabric-versus
fiber-reinforced sand. Journal of Geotechnical Engineering, 112,
Program No. 8, 804–820.
SP poorly graded sand Guo, J., Han, J., Zhang, X., et al. (2019). Evaluation of
T1 sand without any drainage material moisture reduction in aggregate base by wicking geotextile using
T2 sand with drainage belt soil column tests. Geotextiles and Geomembranes, 47, No. 3,
T3 sand with wicking geotextile 306–314.
Guo, Y. P., Leng, W. M., Nie, R. S., Zhao, C. Y. & Zhang, X. (2018).
T4 sand with modified wicking geotextile Laboratory evaluation of a new device for water drainage in
TM trademark roadside slope along railway systems. Geotextiles and
USCS Unified Soil Classification System Geomembranes, 46, No. 6, 897–903.
WRC water retention curve Henry, K. S. & Holtz, R. D. (2001). Geocomposite capillary barriers to
reduce frost heave in soils. Canadian Geotechnical Journal, 38,
No. 4, 678–694.
Henry, K. S., Stormont, J. C., Barna, L. A. & Ramos, R. D. (2002).
Geocomposite capillary barrier drain for unsaturated drainage of
pavements. International Conference on Geosynthetics, Nice,
REFERENCES France.
AASHTO (1991). Report on Task Force 25. Joint Committee Report of Iryo, T. & Rowe, R. K. (2003). On the hydraulic behavior of unsaturated
AASHTO-AGC-ARTBA. American Association of State, Highway nonwoven geotextiles. Geotextiles and Geomembranes, 21, No. 6,
and Transportation Officials, Washington, DC, USA. 381–404.
ASTM D2487-11. Standard practice for classification of soils for Iryo, T. & Rowe, R. K. (2005). Infiltration into an embankment
engineering purposes (Unified soil classification system). reinforced by nonwoven geotextiles. Canadian Geotechnical
ASTM International, West Conshohocken, PA, USA. Journal, 42, No. 4, 1145–1159.
ASTM D4751-16. Standard test methods for determining apparent Koerner, R. M. & Soong, T. Y. (2001). Geosynthetic reinforced
opening size of a geotextile. ASTM International, segmental retaining walls. Geotextiles and geomembranes, 19,
West Conshohocken, PA, USA. No. 6, 359–386.
ASTM D6767-14. Standard test method for pore size characteristics Leng, W. M., Guo, Y. P., Nie, R. S., Zhao, C. Y., Cui, Y. Q. & Dong, C.
of geotextiles by capillary flow test. ASTM International, (2017). A new drainage facility and its drainage performance.
West Conshohocken, PA, USA. Journal of the Railway Society, 39, No. 07, 151–158, (in Chinese).
ASTM D4491. Standard test methods for water permeability Lin, C. & Zhang, X. (2018a). Laboratory drainage performance of a new
of geotextiles by permittivity. ASTM International, geotextile with wicking fabric. Journal of Materials in Civil
West Conshohocken, PA, USA. Engineering, 30, No. 11, 04018293.
Athanasopoulos, G. A. (1993). Effect of particle size on the mechanical Lin, C. & Zhang, X. (2018b). A bio-wicking system to dehydrate road
behaviour of sand-geotextile composites. Geotextiles and embankment. Journal of Cleaner Production, 196, 902–915.
Geomembranes, 12, No. 3, 255–273. Lin, C., Presler, W., Zhang, X., Jones, D. & Odgers, B. (2017). Long-term
Bathurst, R. J., Siemens, G. & Ho, A. F. (2009). Experimental performance of wicking fabric in Alaskan pavements. J. Perform.
investigation of infiltration ponding in one-dimensional sand– Constr. Facil, 31, No. 2, D4016005.
geotextile columns. Geosynthetics International, 16, No. 3, 158–172. Lin, C., Zhang, X. & Han, J. (2018). Comprehensive material
Benjamin, C. V., Bueno, B. & Zornberg, J. G. (2007). Field monitoring characterizations for a roadway installed with wicking fabric.
evaluation of geotextile-reinforced soil retaining wall. Geosynthetics Journal of Materials in Civil Engineering, https://doi.org/10.1061/
International, 14, No. 2, 100–118. (ASCE)MT.1943-5533.0002587.
Bouazza, A., Zornberg, J., McCartney, J. S. & Singh, R. M. (2013). Ling, H. I., Wu, J. T. H. & Tatsuoka, F. (1993). Short-term strength
Unsaturated geotechnics applied to geoenvironmental engineering and deformation characteristics of geotextiles under typical
problems involving geosynthetics. Engineering geology, 165, 143–153. operational conditions. Geotextiles and Geomembranes, 11, No. 2,
Cancelli, A., Montanelli, F., Rimoldi, P. & Zhao, A. (1996). Full scale 185–219.
laboratory testing ongeosynthetics reinforced paved roads. In Earth MathWorks (2016). MATLAB user’s Guide, Version R2016a.
Reinforcement, Balkema, Proceedings of the International Symposium MathWorks Inc, South Natick, MA, USA.
on Earth Reinforcement, November, 1996, Fukuoka, Kyushu, Japan, McCartney, J. S. & Zornberg, J. G. (2010). Effect of cyclic wetting and
Ochiai, H., Yaufuku and N., Omine, K., Editors, vol. 1, 573–578. drying on the formation of a capillary break between soil and
Cedergren, H. R. (1994). America’s pavements: world’s longest bathtubs. geosynthetic drainage layers. Canadian Geotechnical Journal, 47,
Civil Engineering, 64, No. 9, 56–58. No. 11, 1201–1213.

Geosynthetics International, 2022, 29, No. 3

Downloaded by [ Indian Institute Of Technology - Bombay] on [07/03/23]. Copyright © ICE Publishing, all rights reserved.
Laboratory evaluation of different geosynthetics for water drainage 269

McGown, A., Khan, A. J. & Kupec, J. (2004). Determining the design Tan, S. A., Chew, S. H., Ng, C. C., Loh, S. L., Karunaratne, G. P. &
parameters for geosynthetics reinforcements subject to multi-stage Delmas Ph Loke, K. H. (2001). Large-scale drainage behavior of
actions using the isochronous strain energy approach. Geosynthetics composite geotextile and geogrid in residual soil. Geotextiles and
International, 11, No. 6, 455–469. Geomembranes, 19, No. 3, 163–176.
Mitchell, J. K. & Zornberg, J. G. (1995). Reinforced soil structures with Wang, F., Han, J., Zhang, X. & Guo, J. (2017). Laboratory tests
poorly draining backfills. Part II: Case histories and applications. to evaluate effectiveness of wicking geotextile in soil
Geosynthetics International, 2, No. 1, 265–307. moisture reduction. Geotextile and Geomembranes, 45, No. 1,
Noorzad, R. & Mirmoradi, S. H. (2010). Laboratory evaluation 8–13.
of the behavior of a geotextile reinforced clay. Geotextiles and Yoo, C. & Jung, H. Y. (2006). Case history of geosynthetic reinforced
Geomembranes, 28, No. 4, 386–392. segmental retaining wall failure. Journal of Geotechnical and
Pathak, Y. P. & Alfaro, M. C. (2010). Wetting-drying behavior of Geoenvironmental Engineering, 132, No. 12, 1538–1548.
geogrid-reinforced clay under working load conditions. Geosynthetics Zhang, X. & Belmont, N. (2009). Use of Mirafi Nylon Wicking Fabric to
International, 17, No. 3, 144–156. Help Prevent Frost Heaving in Alaska Pavement: 1st, 2nd, 3rd, 4th
Perkins,S.W.(1999).Mechanicalresponseofgeosyntheticreinforcedflexible and 5th Progress Reports. Institute of Northern Engineering (INE),
pavements. Geosynthetics International, 6, No. 5, 347–382. Alaska University of Transportation Center (AUTC), University of
Portelinha, F. H. M., Bueno, B. S. & Zornberg, J. G. (2013). Performance Alaska Fairbanks, AK.
of nonwoven geotextile-reinforced walls under wetting conditions: Zhang, X., Presler, W., Li, L., Jones, D. & Odgers, B. (2014). Use of
laboratory and field investigations. Geosynthetics International, 20, wicking fabric to help prevent frost boils in Alaskan pavements.
No. 2, 90–104. J. Mater. Civ. Eng, 385, 728–740, https://doi.org/10.1061/(ASCE)
Raisinghani, D. V. & Viswanadham, B. V. S. (2011). Centrifuge model MT.1943-5533.0000828.
study on low permeable slope reinforced by hybrid geosynthetics. Zornberg, J. G. & Mitchell, J. K. (1994). Reinforced soil
Geotextiles and Geomembranes, 29, No. 6, 567–580. structures with poorly draining backfills. Part I: Reinforcement
Som, N. & Sahu, R. B. (1999). Bearing capacity of a geotextile-reinforced interactions and functions. Geosynthetics International, 1, No. 2,
unpaved road as a function of deformation-a model study. 103–147.
Geosynthetics International, 6, No. 1, 1–17. Zornberg, J. G., Bouazza, A. & McCartney, J. S. (2010). Geosynthetic
Stulgis, R. P. (2005). Full-scale MSE test walls. Proceedings of NAGS capillary barriers: current state of knowledge. Geosynthetics
2005/GRI-19 Conference, Las Vegas, NV, USA (CD-ROM). International, 17, No. 5, 273–300.

The Editor welcomes discussion on all papers published in Geosynthetics International. Please email your contribution to
discussion@geosynthetics-international.com by 15 December 2022.

Geosynthetics International, 2022, 29, No. 3

Downloaded by [ Indian Institute Of Technology - Bombay] on [07/03/23]. Copyright © ICE Publishing, all rights reserved.

You might also like