+ All Categories
Home > Documents > Exploring desiccation cracks in soils using a 2D profile laser device

Exploring desiccation cracks in soils using a 2D profile laser device

Date post: 23-Dec-2016
Category:
Upload: marcin
View: 214 times
Download: 0 times
Share this document with a friend
14
RESEARCH PAPER Exploring desiccation cracks in soils using a 2D profile laser device Marcelo Sanchez Alvis Atique Sewon Kim Enrique Romero Marcin Zielinski Received: 22 November 2012 / Accepted: 26 August 2013 / Published online: 12 October 2013 Ó Springer-Verlag Berlin Heidelberg 2013 Abstract The study of desiccation cracks in soils has been a subject of increasing attention in recent research. This paper presents the use of a 2D profile laser that is coupled with a motion controller (that allows scanning the overall surface of a drying soil) and electronic balance (to measure the water loss). The aim is to accurately track the three most relevant variables associated with the behavior of soils during desiccation: volume change, water loss and evolving crack network’s morphology. The paper presents the methodology to obtain a digital model of the soil using the experimental setup described above. The main results of a natural soil subjected to drying are presented and discussed, including evolution of cracks aperture; evolution of cracks depth, surface contour levels (at different times); and evolution of volume change. It is shown that the pro- posed methodology provides very useful information for studying the behavior of soils subjected to desiccation. Keywords Drying cracks Laser scanner Soil desiccation Volume shrinkage 1 Introduction The presence of desiccation cracks strongly affects hydraulic and mechanical properties of soils. Desiccation cracks act as a preferential path for water flow and pollu- tant transport (e.g., [5, 12, 15, 37, 41]); they also affect the performance of landfill cover and clay liners (e.g., [1, 11, 28, 45, 55]). In earth embankments, cracking reduces strength and may lead to seepage and percolation problems. The presence of desiccation cracks can also trigger land- slides. Deep vertical cracks during prolonged and intense draught increases infiltration capacity of the soil, which in turn mobilizes the shrinkage or swelling potential of deeper soils. Vital infrastructure affected by soil cracking includes levees, road embankments and engineered barriers (e.g., [7, 13, 19, 21, 42]). The formation and propagation of cracks is a highly complex phenomenon due to the strong coupling between the hydraulic and mechanical behavior of soils. Water loss during evaporation induces a rise in capillary forces. The water transfer process in drying soils is basically controlled by the hydraulic properties of the soil mass, which in turn affects the mechanical behavior, because the soil tends to contract under increasing suction [38]. During shrinkage, the changes in the stress state tend to re-accommodate the soil particles and, at a certain point of the drying process, crack generation begins. The actual physical mechanisms involved in the crack initiation and its subsequent propa- gation are the current topics of active research and open discussions (e.g., [4, 8, 17, 21, 24, 26, 31, 32, 36, 38, 43, 48, 49]). Current experimental investigations in this area pri- marily concentrate on the study of slurries prepared in plates and subjected to drying under controlled laboratory conditions (e.g., [24, 33, 36, 38]). In those tests, quite M. Sanchez (&) S. Kim M. Zielinski Zachry Department of Civil Engineering, Texas A&M University, College Station, TX 77843-3136, USA e-mail: [email protected] A. Atique M. Zielinski Department of Civil and Environmental Engineering, University of Strathclyde, Glasgow, UK E. Romero Department of Geotechnical Engineering and Geosciences, Universitat Polite `cnica de Catalunya, Barcelona, Spain 123 Acta Geotechnica (2013) 8:583–596 DOI 10.1007/s11440-013-0272-1
Transcript
Page 1: Exploring desiccation cracks in soils using a 2D profile laser device

RESEARCH PAPER

Exploring desiccation cracks in soils using a 2D profile laser device

Marcelo Sanchez • Alvis Atique • Sewon Kim •

Enrique Romero • Marcin Zielinski

Received: 22 November 2012 / Accepted: 26 August 2013 / Published online: 12 October 2013

� Springer-Verlag Berlin Heidelberg 2013

Abstract The study of desiccation cracks in soils has

been a subject of increasing attention in recent research.

This paper presents the use of a 2D profile laser that is

coupled with a motion controller (that allows scanning the

overall surface of a drying soil) and electronic balance (to

measure the water loss). The aim is to accurately track the

three most relevant variables associated with the behavior

of soils during desiccation: volume change, water loss and

evolving crack network’s morphology. The paper presents

the methodology to obtain a digital model of the soil using

the experimental setup described above. The main results

of a natural soil subjected to drying are presented and

discussed, including evolution of cracks aperture; evolution

of cracks depth, surface contour levels (at different times);

and evolution of volume change. It is shown that the pro-

posed methodology provides very useful information for

studying the behavior of soils subjected to desiccation.

Keywords Drying cracks � Laser scanner � Soil

desiccation � Volume shrinkage

1 Introduction

The presence of desiccation cracks strongly affects

hydraulic and mechanical properties of soils. Desiccation

cracks act as a preferential path for water flow and pollu-

tant transport (e.g., [5, 12, 15, 37, 41]); they also affect the

performance of landfill cover and clay liners (e.g., [1, 11,

28, 45, 55]). In earth embankments, cracking reduces

strength and may lead to seepage and percolation problems.

The presence of desiccation cracks can also trigger land-

slides. Deep vertical cracks during prolonged and intense

draught increases infiltration capacity of the soil, which in

turn mobilizes the shrinkage or swelling potential of deeper

soils. Vital infrastructure affected by soil cracking includes

levees, road embankments and engineered barriers (e.g., [7,

13, 19, 21, 42]).

The formation and propagation of cracks is a highly

complex phenomenon due to the strong coupling between

the hydraulic and mechanical behavior of soils. Water loss

during evaporation induces a rise in capillary forces. The

water transfer process in drying soils is basically controlled

by the hydraulic properties of the soil mass, which in turn

affects the mechanical behavior, because the soil tends to

contract under increasing suction [38]. During shrinkage,

the changes in the stress state tend to re-accommodate the

soil particles and, at a certain point of the drying process,

crack generation begins. The actual physical mechanisms

involved in the crack initiation and its subsequent propa-

gation are the current topics of active research and open

discussions (e.g., [4, 8, 17, 21, 24, 26, 31, 32, 36, 38, 43,

48, 49]).

Current experimental investigations in this area pri-

marily concentrate on the study of slurries prepared in

plates and subjected to drying under controlled laboratory

conditions (e.g., [24, 33, 36, 38]). In those tests, quite

M. Sanchez (&) � S. Kim � M. Zielinski

Zachry Department of Civil Engineering, Texas A&M

University, College Station, TX 77843-3136, USA

e-mail: [email protected]

A. Atique � M. Zielinski

Department of Civil and Environmental Engineering, University

of Strathclyde, Glasgow, UK

E. Romero

Department of Geotechnical Engineering and Geosciences,

Universitat Politecnica de Catalunya, Barcelona, Spain

123

Acta Geotechnica (2013) 8:583–596

DOI 10.1007/s11440-013-0272-1

Page 2: Exploring desiccation cracks in soils using a 2D profile laser device

accurate measurements of the water loss during desiccation

can be gathered by means of high-precision balances. The

measurement of the volume change under such conditions

is more complicated. Conventional instruments used in the

laboratory to measure displacements, such as contact

devices (e.g., linear variable differential transformer,

LVDT) and non-contact sensors (e.g., Hall effect, Eddy

current), are not suitable for these kinds of experiments due

to the perturbation they may induce in stress field (e.g.,

[36]). Digital image analyses are generally used to study

the nature of soil crack propagation and also to obtain

useful parameters to characterize the morphology of the

crack network (e.g., [25, 30, 46, 47, 50–53]). A rough

estimation of the vertical settlement can be obtained from

image analyses from one camera. More accurate mea-

surements will require more advanced systems such us

two- and three-component setups (e.g., [9]); however, they

are not very popular. In a two-component setup, two

cameras are installed at different angles; in a three-com-

ponent setup, three cameras are installed at different

angles. The post-processing of the data gathered from these

kinds of systems is quite complex.

This work is aimed at studying the simultaneous chan-

ges in water content, volume change and cracks pattern

during drying by using a methodology based on a non-

contact device. To precisely track the concurrent changes

in the hydraulic and mechanical fields is a crucial aspect in

a highly coupled problem as soil drying and cracking. The

adopted non-contact device is a 2D laser scanner that

allows tracking the changes in volume of the whole sample

with high resolution. It is possible to gather the evolution

of key local variables associated with the desiccation

cracks, such as cracks aperture, cracks depth and crack

geometry. The 3D representation of the soil sample is

obtained after compiling subsequent linear (2D) profile

data. The methodology presented hereafter to study the

desiccation process in soils is novel and allows gathering

information for some variables that are difficult to measure

with other typical techniques currently used in this kind of

problems. In the following sections, the proposed device

and methodology are presented in detail, alongside the

main results obtained from a drying test of a natural soil

from Indonesia.

2 Experimental methodology

2.1 Soil

The ‘‘Bengawan-Solo’’ natural soil from Surabaya, Indo-

nesia, has been studied in this research. It is an organic silt

of high plasticity. The soil fractions of the sample inves-

tigated are as follows: 30 % sand, 57 % silt and 13 % clay.

The organic content is 8 %. The following Atterberg’s

limits were determined: liquid limit = 54 %; plastic

limit = 36 %; and shrinkage limit = 24.2 %. The soil

activity is 1.4. Powdered X-ray diffraction tests were

undertaken to determine the clay and non-clay minerals

present in the soil. Non-clay minerals present include

quartz, calcite and feldspar (plagioclases), while the dom-

inant clay minerals are montmorillonite (88 % of the clay

fraction) and kaolinite. The soil particle density is around

2.73 Mg/m3. More information about this soil can be found

elsewhere (e.g., [14]).

2.2 Volume change in drying soils

Desiccation experiments focused on soil cracking phe-

nomena are generally performed on slurry samples pre-

pared in either rectangular or circular plates (e.g., [25, 33,

36, 47]). Current methods/techniques to estimate the vol-

ume change under such conditions present severe limita-

tions. For example, non-contact devices are desired to

prevent any perturbation in the stress field that may alter

the natural evolution of the test. Therefore, LVDTs are not

appropriate for this kind of study. Furthermore, the initially

fully saturated soil cannot support the weight of the LVDT

sensor head, which settles (or penetrates) in the soil mass

invalidating the measurements. Non-contact magnetic

sensors (e.g., [10, 20]) require the inclusion of reflecting

elements on the soil surface, which would significantly

alter the stress state in its vicinity, affecting the crack

formation process and the subsequent crack propagation.

Peron et al. [36] used a setup based on calipers, which

required a large number of manual (local) measurements.

Furthermore, those devices (i.e., LVDTs, calipers and non-

contact magnetic sensors) provide local measurements

only, from which it is difficult to obtain an accurate esti-

mation of the volume change in drying soils due to soil

curling [23] and the presence of desiccation cracks.

Digital image processing has also been used to measure

displacements and volume changes of specimens in con-

ventional triaxial tests (e.g., [2, 3, 16, 27, 34, 54]). In

general terms, it was shown that the results obtained from

digital and manual measurements matched well. However,

the measurement of displacements from digital images is

not very precise from only one camera. Moreover, the use

of image processing technique to estimate volume change

in drying soils has a number of limitations. For example,

the shadow (in the ‘‘crack valleys’’) makes it difficult to

detect the actual (3D) geometry of the cracks from a digital

image (i.e., crack depth and variation in crack aperture with

depth are difficult to measure).

Laser scanners have been used with success for mea-

suring the volume of solids [56] and in soil and rock testing

to measure displacements (e.g., [18, 39, 40]). However, its

584 Acta Geotechnica (2013) 8:583–596

123

Page 3: Exploring desiccation cracks in soils using a 2D profile laser device

application to study the process of desiccation cracks in

soils has not been done yet. The use of this device for

studying soil drying is presented in the following section.

2.3 Setup description

The desiccation process in soils is a strongly coupled

hydromechanical problem. Thus, an accurate measurement

of the volume change induced by the water loss is a crucial

component when studying the behavior of drying soils. The

setup proposed in this research was designed to gather

simultaneous information during desiccation associated with

these two variables: water loss and volume change.

The basic experimental setup consists of the following

main components: (1) compact 2D/3D scanner; (2) laser

motion controller; (3) electronic balance with the soil sample

on it; (4) relative humidity and temperature sensor; (5) digital

camera; and (6) a computer with data acquisition system

(DAS). Figure 2 presents a schematic representation of this

setup, and the main components are briefly described below.

A 2D/3D laser scanner ScanControl 2700-100 from

Micro-Epsilon [29] was used in this research. This device is

based on the triangulation principle for a two-dimensional

acquisition of a height profile of various target surfaces. A

laser line is generated by special lenses and projected onto the

target surface. The working principles are illustrated in

Fig. 1. A high-quality optical system projects the diffusely

reflected light of this laser line back onto a highly sensitive

sensor matrix. The controller integrated into the sensor head

uses this matrix image to calculate the position along the laser

line (y-axis) and the vertical distance (z-axis). Figure 2 pre-

sents a schematic representation of this setup, and the main

components are briefly described below.

The laser unit was fixed to the wall using a frame designed

to allow positioning the laser at different heights (see Fig. 2).

To obtain a 3D representation of the sample, the laser was

coupled with a motion controller (model OWIS PS-10) that

moves horizontally (x-axis) at a constant speed perpendicular

to the laser line. This is a single axis position controller with a

two-phase stepper motor including open-loop functionality

which has an USB interface to communicate with the com-

puter. The software tool OWISoft was used to configure and

control this unit. The soil profiles were transferred to the

computer as an array of points and were processed using the

software ICONNECT from Micro-Epsilon [29]. Each of

these profiles consists of a certain number of calibrated

measuring points, including additional information such as

intensity, time and counter information.

Z-axis resolution corresponds to the elevation resolution

(i.e., the one concerning the distance from the surface to

the sensor). Points in the surface point cloud separated by

this amount, or more, can be identified by the scanner. The

laser setup used in this research has a z-axis resolution of

15 lm with a reference line of 320 points. This means that

the sensor is able to measure the difference in a vertical

direction up to 15 lm height. This resolution is obtained

when the ‘‘laser line’’ is adopted equal to ‘‘manufactured

specified standard measuring range’’ [29]. Y-axis resolution

corresponds to the resolution along the laser line (Fig. 1).

Higher resolution means a greater amount of acquired data

along the laser line, hence more information about the

surface. The adopted laser has a predefined parameter for

this purpose, which is in the form of points/profile. The

width of the laser line depends on the actual distance from

the sensor to the target. For example, for surface–laser

distance of 400 mm, the laser line width is 100 mm (i.e.,

the manufactured standard measuring range). X-axis reso-

lution corresponds to the resolution normal to laser line

axis. This resolution depends mainly on the number of

profiles scanned per second (profiles/s); higher pro-

file(s) gives higher x-axis resolution.

A high-capacity precision weighing balance (Mettler

Toledo model SB-8001) was used to gather the water loss

during soil desiccation. It features 8.1 kg of maximum

weighing capacity, which offers 1 9 10-4 kg readability.

The balance was connected to a computer and programmed

(using Windmill software) for continuous monitoring.

A data logger (Model RH520) connected to a computer

was used for relative humidity (RH) and temperature

(T) measurements. The RH measurement range is from 10

to 95 % and T from -28 to 60 �C. The basic accuracy is

3 % RH and 1 �C. A 12.1 MP digital camera was mounted

Light source

Receiver

Fig. 1 3D view showing the laser scan, laser line, soil sample and

adopted reference system. It can be observed how the system projects

the diffusely reflected light of this laser line back onto a highly

sensitive sensor matrix

Acta Geotechnica (2013) 8:583–596 585

123

Page 4: Exploring desiccation cracks in soils using a 2D profile laser device

on the laboratory stand to take images of the sample during

desiccation. A desktop computer was used for data acqui-

sition and device control.

2.4 Procedure

The desiccation tests were performed in a small room under

controlled temperature and relative humidity conditions.

The room has no windows, and the lighting was strictly

controlled to facilitate constant conditions during testing.

The soil sample was prepared close to its liquid limit and

allowed to dry in a circular desiccation plate (diameter

9.689 E-02). A glass plate with a smooth bottom and a

height of 1.292 E-02 was used for this test for preparing the

sample. The moisture content was determined before the

test (57.7 %) and after it (6.2 %). The temperature and

relative humidity of the room were continuously monitored

via the humidity data logger. The average temperature

during the test was 19.5 �C (±0.30 �C), and the average RH

was 37 % (±4 %). Figure 3 presents the time evolution of

RH and temperature in the room during the test.

The sample was placed on the top of the balance that

was programmed to record continuous water loss. The laser

was attached to the motion control unit for scanning on the

move. The laser was operated to move at a speed of

3.88 9 10-3 m/s, which allowed taking linear measure-

ments at 38-lm increments (i.e., distance between profiles

in the ‘‘X’’-direction). Moreover, with a standard laser line

width of 100 mm, it produces a 312.5 lm distance between

points in the ‘‘Y’’-direction. The sample was scanned at

time intervals depending on the crack appearance in the

sample. The whole experimental setup remained untouched

until the end of the test, allowing for the accurate recording

of experimental data during the test without perturbation of

the sample position and condition.

The 3D data obtained with the laser scanner represents

regularly distributed set of points. The scanned data was

processed using the Surfer� software [44] to create a uni-

formly spaced grid from the uniformly spaced vectors (X, Y,

Z). To achieve a high definition, the data sets were

10

15

20

25

30

Tem

pera

ture

(o C

)

Temperature

0 10 20 30 40 50 60

Time (hours)

0

20

40

60

80

100

Rel

ativ

e hu

mid

ity

(%)

Relative humidity

Fig. 3 Time evolution of room temperature and relative humidity

during test

Data acquisition pad for motion

controller

Laser scan

Soil specimen

Electronic balance

Motion controller

Digital camera

Motion bracket

PC

Mold

Fig. 2 Top right photograph of the experimental setup. Bottom left schematic representation showing the experimental setup for the laser 2D/3D

scanner. The digital camera is not part of the laser setup; it is used just for taking photographs of the cracked soil to compare them with the laser outputs

586 Acta Geotechnica (2013) 8:583–596

123

Page 5: Exploring desiccation cracks in soils using a 2D profile laser device

additionally densified in both the x- and y-directions during

gridding. This is a standard feature of Surfer which facilitates

the creation of a very fine grid. Equal spacing of 40 lm in x-

and y-axis was used. The surface was then interpolated at the

points specified by uniformly spaced grid using natural

neighbor linear interpolation method. It calculates the

weighted average of neighboring observations using weights

determined by Voronoi polygon concepts [35]. Nearest-

neighbor algorithm is useful for converting regularly spaced

(X, Y, Z) dense data sets to Surfer grid files. The final output of

this step is a ‘‘digital model of the soil,’’ in which the grid

points (with coordinates X, Y, Z) represent the volume of the

soil sample. Digital models are obtained for the different

times of the analysis.

Once the digital model of the soil is obtained following the

procedure outlined above, different pieces of software can be

used to post-process the experimental data and extract useful

information related to cracked soils. For example, to calculate

the changes in volume during drying, two different approa-

ches have been used in this research: (1) the software Surfer

and (2) an in-house program coded in Matlab. As for method

(1), a module in Surfer allows the calculation of the volume

below the interpolated surface. Volume change during two

different times can be then calculated as the difference

between the corresponding volumes. Details about Surfer can

be found elsewhere (e.g., [44]). The Matlab code (2) is

described in the following section. It is worth mentioning that

practically identical results in terms of volume change have

been obtained with these two approaches.

3 Data processing

3.1 Preliminary study on solids of known volume

Before presenting the results obtained from the drying test

in the soil sample, the studies to check the accuracy of the

proposed setup for volume measurement are briefly intro-

duced. A Teflon ring was scanned first. The volume of the

cylinder is 9.68371E-06 m3. In order to challenge the

capabilities of the device, a plastic brick (i.e., Lego type)

was also tested. This piece is highly complex and provides

the opportunity to explore the performance of the proposed

setup when a number of vertical faces and ‘‘valleys’’ have

to be measured. The volume of the brick is 3.66E-05 m3

and was measured by the water displacement method (BS

1,377:1,975). Before measuring its volume by this method,

the void space at the bottom of the brick was filled with a

plastic filler. Figure 4a, b presents the models of these two

pieces obtained with the laser scan following the procedure

explained in the previous section. The calculated volume

for the ring is 9.6308E-06 m3. Comparing this volume with

the measured one, the difference is around 0.55 %. As for

the Lego brick, the volume found was 3.6917E-05 m3.

Hence, the volumetric accuracy per unit area is

1.17 9 10-2 cm3/cm2, which is higher than the volumetric

resolution estimated considering only z-axis resolution

(1.5 9 10-3 cm3/cm2). Assuming as the reference volume

the one calculated via the water immersion method, the

error associated with the volume measurement using this

device is around 0.87 %.

3.2 Soil sample

Figure 5a presents the digital images of the desiccated soils

at three different times: (a) 4 h, (b) 6 h and (b) 24 h.

Figure 5b presents the corresponding 3D graphical models

of the soil obtained with Surfer. From the figures, the

capability of the proposed device to capture the actual

pattern observed in the tests is apparent; only a few small

fissures are not perfectly reproduced by the model.

Once the digital model of the soil is generated, it can be

used to explore other features/properties of drying soils

difficult to study using other available techniques. Figure 6

Fig. 4 3D models obtained after scanning the samples (of known volume) and post-processing them using Surfer: a plastic brick (Lego type) and

b Teflon ring

Acta Geotechnica (2013) 8:583–596 587

123

Page 6: Exploring desiccation cracks in soils using a 2D profile laser device

presents the position of two typical sections analyzed in

this work. For example, Fig. 7 presents the evolution of a

typical cross section (i.e., x = 46.398 mm) during drying.

Key information associated with changes in soil properties

and characteristics during drying can be obtained from

these figures, for example, volume change, crack depth,

crack aperture and porosity changes. In this research a

program coded in Matlab was developed to extract this

information. In the following paragraphs, a brief descrip-

tion of this program is presented.

Figure 8a presents the flow diagram associated with the

developed code. The first step is to read the matrix asso-

ciated with the geometry of the sample (i.e., X, Y and Z),

which represents the digital model of the soil for a given

time ‘‘t’’ (with ‘‘t’’ varying between ‘‘1,’’ first scanning, and

‘‘k,’’ final scanning). The data is organized by profiles (as

scanned) of constant coordinate ‘‘X.’’ Figure 8b shows a

schematic representation of the cracked soil at a given time

‘‘t.’’ This figure also indicates the coordinate system

adopted in this analysis for a horizontal plane. The distance

between consecutive profiles (Dh) is 0.038 mm; therefore,

a total of 2,550 profiles were processed per each time step.

In this work, a distinction has been made between crack

and gap. The gap corresponds to the void space generated

during desiccation between soil and container. Crack (as

usual) is the void space developed in the drying soil mass.

Figure 8c presents a typical cross section (at constant ‘‘X’’),

identifying the space related to gaps and cracks. The

coordinate system in a vertical plane is also showed in this

picture.

After reading the digital model for a given ‘‘t,’’ the

variables to be used in this step are initialized, and a profile

‘‘i’’ is analyzed (with ‘‘i’’ varying between 1 and ‘‘m’’; with

m = 2,550). Each profile is identified by a (fixed) coordi-

nate ‘‘X.’’ The proposed algorithm calculates the slope

(‘‘h’’) between two consecutive (‘‘Y’’) points along the

profile (Fig. 8c). At the beginning of the test h % 0 (i.e.,

Fig. 7, profile at t = 0); as drying progresses (non-

homogenous), volume changes take place and desiccation

Fig. 5 Images of the soil sample at three different times: 4 h, 6 h and 24 h. a Digital images; b 3D mode after scanning the sample and

processing it with Surfer

X

Y

Δh = 0.038 (mm)

x=81.510

x=46.398

Fig. 6 Scheme of the cracked soil sample indicating the two sections

analyzed in more detail in Figs. 7 and 9

588 Acta Geotechnica (2013) 8:583–596

123

Page 7: Exploring desiccation cracks in soils using a 2D profile laser device

cracks start to appear at different stages of the test (i.e.,

Fig. 7, profile at t = 9 h and subsequent ones). Changes in

the gradient of the soil profile allow identifying the position

of the cracks. The proposed algorithm checks the slope

between adjacent points (from left to right). When an

abrupt change in the gradient is detected, the position

‘‘YCSb’’ associated with the beginning of a crack is stored

(where ‘‘Y’’ is related to the position of the crack, ‘‘C’’

refers to crack, ‘‘S’’ refers to start, and ‘‘b’’ identifies crack

number in the analyzed profile, i.e., b = 1, … n, where n is

0

4

8

12

16

Ele

vati

on (

mm

)

0

4

8

12

16

Ele

vati

on (

mm

)

0

4

8

12

16

Ele

vati

on (

mm

)

0

4

8

12

16

Ele

vati

on (

mm

)

-60 -40 -20 0 20 40 60

Y distance to the centre (mm)

0

4

8

12

16

Ele

vati

on (

mm

)

-60 -40 -20 0 20 40 60

Y distance to the centre (mm)-60 -40 -20 0 20 40 60

Y distance to the centre (mm)

t = 0

t = 2 hours

t = 3 hours

t = 4 hours

t = 5 hours

t = 6 hours

t = 7 hours

t = 9 hours

t = 11 hours

t = 13 hours

t = 15 hours

t = 24 hours

t = 21 hours

t = 18 hours

t = 16 hours

Section ( x = 46.398)

Fig. 7 Evolution of typical cross section (x = 46.398 mm) during drying, showing the section profiles at different times

Digital profile at time ( X, Y, Z ) t = 1, k

Variable initialization

Read digital model profiles (Xi , Yi, Zi ) i = 1, m

Select profile [ Xi =fix ] ( Xi , Yj, Zj ) j = 1, n

Crack Find ( XCi , YCi, Zci ) Find and save crack coordinates

Crack Starting Point (XCSβ, YCSβ, ZCSβ) Crack Ending Point (XCEβ, YCEβ, ZCEβ)

Find crack depth [ YCBβ (Xi , Yj) ]

Calculate total area, crack area [ Ai= trapz (YCi, YCBi ) ]

Calculate total volume, crack volume

[ Vi = ×h ]

END

( No )

( Yes )

Time = t

Xi+1 < Xm

Save crack profiles [ XCi , YCi, ZCi, Di , Ai, Vi ]

t = k

( Yes )

( No ) YCSn YCEn YCSβ YCEβYCS1 YCE1

Yl Y1

YCB1

YCBβYCBn

Yj

YGS1 YGE1

GapCrack 1

Crack β Crack n

(Xi, Y1 )

i = 1

i = m

(Xi, Yl )

Time = t

Δh = 0.038 (mm)

X

Y

Y

Z

YCS1

θ

(a) (b)

(c)

Fig. 8 a Flow diagram indicating the basic steps of the post-process program coded in MATLAB; b top view (schematic) showing the cracked

sample and the adopted reference system; c typical cross section identifying the characteristic points used in the analysis and the reference system

Acta Geotechnica (2013) 8:583–596 589

123

Page 8: Exploring desiccation cracks in soils using a 2D profile laser device

the maximum number of cracks in a profile). In this work it

has been considered that a crack starts when the change in

slope is equal to or higher than 40�. After detecting YCSb,

the algorithm will be tracking points down the slope

associated with a lateral face of the crack (see zoom in

Fig. 8c). A change in the sign of the gradient implies that

the bottom of the crack has been reached and the corre-

sponding position is stored (identified as YCBb, where ‘‘B’’

is related to the bottom of the crack). Afterward, the

algorithm will track points that are in the other lateral face

of the crack, moving up in Z coordinate. At some point, a

sudden change in the gradient will be detected, now toward

subhorizontal slopes (i.e., small h). This change in slope

corresponds to the end of the crack ‘‘YCEb’’ (where ‘‘E’’

refers to the end of crack). Gaps are detected in a similar

way, but in this case the start (YGSb) or end (YGEb) position

is coincident with the sample container (where ‘‘G’’ refers

to gap). Figure 9 presents the results for another profile

(X = 81.510 mm), showing the time evolution of that cross

section during drying. The position of that section is pre-

sented in Fig. 6.

The procedure to identify the cracks outlined above

corresponds to cracks with a relatively simple geometry

(see, e.g., Fig. 7, t = 11 h, Fig. 8c). However, sometimes

the geometry of the crack is more complex and could

present two (or more) valleys (see, e.g., Fig. 7, t = 21 h).

The algorithm developed in this work is able to deal with

such complexities; details about these aspects can be found

in [22].

Once all the cracks (and gaps, if any) in a profile are

found, the program calculates the area of each crack (and

gap), and also the total area of the cross section that cor-

responds to cracks (and gaps). It is also possible to calcu-

late the volume associated with the cracks when the area

related to them in two consecutive profiles (separated by a

distance Dh in the X-direction) is known. For further post-

processing, all the information associated with the profile is

stored by the program (i.e., YCSb, YCBb, YCEb, YGSb, YGBb,

YGEb, cracks-area, gaps-area, etc.). The analysis described

above is repeated for all the profiles (i.e., i = 1, m) and for

all the time steps studied (i.e., t = 1, k). Figure 8a illus-

trates the steps involved in the analysis.

4 Results and discussion

The information collected by the program can be post-

processed to learn about key aspects of soils subjected to

drying. For example, volume change during shrinkage can

be associated with:

(a) vertical displacement of the top surface (indicated as

‘‘settlement’’ hereafter);

(b) lateral shrinkage (designated as ‘‘gap’’); and

(c) cracks.

The proposed methodology is capable of quantifying

these three different components of soil shrinkage. Fig-

ure 10a presents schematically these three components of

the shrinkage for a generic cross section. Figure 10b shows

how these components of the shrinkage varied with time

for the profile located at x = 46.398 mm. It is observed

-60 -40 -20 0 20 40 60

Y distance to the centre (mm)

0

4

8

12

16

Ele

vatio

n (m

m)

Time (hours)0

2

7

13

21

24

Section ( x = 81.510 )

Fig. 9 Evolution of typical cross section (x = 81.510 mm) during

drying, showing the section profiles at different times

Gap

Crack 2 Crack 3

Area of crack

Crack 1

Area of Gap

Area of Settlement

Depth

Aperture

0 2 4 6 8 10 12 14 16 18 20 22 24 26

Elapsed Time (hours)

0

100

200

300

400

Are

a ch

ange

(m

m2 )

ShrinkageSection ( x=46.398)

Total

Settlement

Total crack

Crack 1

Crack 2

Crack 3

Gap

(a)

(b)

Fig. 10 a Scheme showing the different components of soil shrink-

age, b time evolution of the different components of soil shrinkage for

a typical section (x = 46.398 mm)

590 Acta Geotechnica (2013) 8:583–596

123

Page 9: Exploring desiccation cracks in soils using a 2D profile laser device

that at the beginning, the shrinkage is associated with

settlement only. The first crack appears around 4 h after

drying; the other ones appear around 3 h later. The major

shrinkage is associated with settlement. Only about one-

third of the total shrinkage is related to cracks and gaps.

The program also allows learning about the changes in

crack aperture and depth. In fact, in its current version, the

code calculates the crack aperture projected on a (vertical)

plane with X = constant. This value is identified AX in this

paper. Figure 11 presents the variation in depth and AX (at

the top of the crack) for the sections presented in Figs. 7

and 9. Section at x = 46.398 mm shows that crack 1

reached the maximum depth, in a very short period of time

(around 1 h), and then remains relatively constant. Around

two-thirds of the aperture of this crack developed also in a

short period of time (around 1 h as well), and the other

one-third opened almost gradually in time, up to around

20 h, after that the aperture and depth remain fairly

0 2 4 6 8 10 12 14 16 18 20 22 24 26

Elapsed time (hours)

X= 81.510Crack 1Crack 2Crack 3

X= 81.510Crack 1Crack 2Crack 3

0 2 4 6 8 10 12 14 16 18 20 22 24 26

Elapsed time (hours)

0

2

4

6

8

10

12

AX (

mm

)

X= 46.398Crack 1Crack 2Crack 3

0

2

4

6

8

10

12

14

16

18D

epth

(m

m)

X= 46.398Crack 1Crack 2Crack 3

Fig. 11 Time evolution of depth crack and aperture for two typical sections (x = 46.398 and x = 81.510 mm). The value AX corresponds to the

projection of the aperture in a vertical plane with constant ‘‘x’’ (in this case x = 46.398 and x = 81.510 mm)

Acta Geotechnica (2013) 8:583–596 591

123

Page 10: Exploring desiccation cracks in soils using a 2D profile laser device

constant. Crack 2 developed following a similar pattern to

one explained above. Crack 3 developed also in a similar way

but delayed in time. As for the section located at

x = 81.850 mm, the behavior changes slightly with respect

to the one discussed previously, cracks appeared earlier,

and the opening of the cracks aperture was more gradual in

time.

The behavior described above is perhaps better repre-

sented by the plots shown in Fig. 12. It can be observed

that for both sections, at the beginning, the increase in

depth and AX happened simultaneously, then the depth

remains practically constant, and the aperture continued

increasing up to reaching the steady state after around 18 h

of test. Note that the decrease in depth observed in crack 1

section at x = 81.510 (Fig. 12) at advanced stages of the

drying, around 16th h (Fig. 9), is related to the significant

settlement observed near this crack at that time.

As mentioned in Sect. 2, by integrating the information

from subsequent profiles, the corresponding changes in

volume can be calculated. This is one of the main advan-

tages of the methodology used in this research. Figure 13

presents the volume changes during drying for the three

components of the shrinkage discussed above.

The advantage of knowing the volume changes asso-

ciated with each component is that it allows for the

calculation of the variations in the void ratio (or porosity)

that effectively take place during shrinkage. At a given

time, the total volume associated with the soil sample is

the sum of four main (volume) components: pores

(Vpores), solid (Vsolid), cracks (Vcracks) and gaps (Vgaps).

The volume of pores corresponds to voids in the intact

(dried) soil mass. The net volume of the dried soil is

obtained as the sum of the volume of voids plus the

volume of solid. Figure 14a presents the changes in

porosity. Two different values are plotted: total and net

porosities defined as follows:

Total porosity : nT

¼ Volcracks þ Volgaps þ Volpores

Volcracks þ Volgaps þ Volpores þ Volsolids

ð1Þ

AX (mm)

X= 81.510Crack 1Crack 2Crack 3

0 2 4 6 80 2 4 6 8

AX (mm)

0

2

4

6

8

10

12

Dep

th (

mm

)

X= 46.398Crack 1Crack 2Crack 3

Fig. 12 Concurrent changes in depth and crack aperture for two typical sections (x = 46.398 and x = 81.510 mm). The value AX corresponds to

the projection of the aperture in a vertical plane with constant ‘‘x’’ (in this case x = 46.398 and x = 81.510 mm)

0 2 4 6 8 10 12 14 16 18 20 22 24 26

Elapsed time (hours)

0

5

10

15

20

25

Vol

ume

chan

ge (

cm3 )

ShrinkageTotalSettlementCrackGap

Fig. 13 Changes in the volume of the soil sample during drying

592 Acta Geotechnica (2013) 8:583–596

123

Page 11: Exploring desiccation cracks in soils using a 2D profile laser device

Net porosity : nN ¼Volpores

Volpores þ Volsolids

ð2Þ

It can be observed that both porosities are identical until

the first crack/gap appeared (after approximately 7th h of

drying). Both porosities are associated with relevant

properties of the soil mass. Average properties can be

obtained from the global porosity (i.e., global density),

while properties associated with the intact soil (i.e., the

portions of soil without cracks) can be estimated/updated

from the net porosity, as discussed below. Figure 14b, c

presents the evolution of pores void ratio (ep) and discon-

tinuities void ratio (eD), defined as follows:

Pores void ratio : ep ¼Volpores

Volsolids

ð3Þ

Discontinuities void ratio : eD ¼Volcracks þ Volgaps

Volsolids

ð4Þ

The balance (Fig. 2) has been used to measure the

concurrent water loss that takes place during drying. By

combining information about volume change and water

loss, it is possible to learn about the changes in degree of

saturation during desiccation (Fig. 14d). The inspection of

Fig. 14c, d shows that the onset of crack generation

(around 7th hour) takes place when the soil is practically

fully saturated. This was confirmed in the previous

investigations (e.g., [38]).

Figure 15 represents the simultaneous changes in

moisture and dry density that take place in the soil during

desiccation. It can be observed that at the beginning, the

loss of water is accompanied by a volume change of the

sample up to water content near 25 % (which is quite close

to the shrinkage limit of the soil); then (as expected) drying

continues with practically no changes in volume. To cal-

culate the dry density presented in Fig. 15, the ‘‘net vol-

ume’’ was considered as follows:

Dry density : qd ¼Masssolids

Volsolids þ Volpores

ð5Þ

Detailed information about the shrinkage evolution and

how settlements and cracks are distributed in the soil

sample during drying can easily be obtained from the

digital model of the soil. For example, Fig. 16b presents

the contour levels at three different times, from which it is

possible to study how settlements and cracks evolve during

drying. Figure 16a presents the corresponding digital

1.00

1.20

1.40

1.60

e P

0.00

0.05

0.10

0.15

0.20

0.25

0.30

e D

0 10 20 30 40 50 60Time (hours)

0.00

0.20

0.40

0.60

0.80

1.00

0.48

0.52

0.56

0.60

0.64P

oros

ity

NetTotal

(a)

(b)

(c)

(d)Deg

ree

of s

atur

atio

n

Fig. 14 Time evolution: a porosity (total and net), b pores void ratio

(ep), c discontinuities void ratio (eD) and d degree of saturation

65 60 55 50 45 40 35 30 25 20 15 10 5 0

Water content ( % )

1.50

1.40

1.30

1.20

1.10

1.00

Dry

den

sity

( g

/cm

3)

Drying

Fig. 15 Simultaneous changes in dry density and water content

during drying

Acta Geotechnica (2013) 8:583–596 593

123

Page 12: Exploring desiccation cracks in soils using a 2D profile laser device

pictures. Figure 16c shows 3D cuts of the sample, illus-

trating the volume changes (i.e., settlement, cracks and

gaps) during shrinkage. The initial level (red transparent

surface on the top) is included as a reference.

The results and plots presented above are just examples

of information that can be gathered using the setup and

methodology adopted in this research. It has been shown

that relevant data related to the geometry of the cracks can

be obtained. Such kind of information is very valuable to

gain a better understanding of the drying process in soils

and also to model properly its behavior. For example,

changes in porosity of the intact soil can be used to esti-

mate the changes in its permeability, by using well-known

permeability laws (e.g., Kozeny’s law [6]). The changes in

crack geometry can also be tracked and then used to

develop models accounting for the effect of cracks on flow

and evaporation. Such developments are being carried out

by the authors, but they are outside the scope of this

contribution.

5 Conclusions

A methodology to study the behavior of drying soils based

on a 2D/3D laser scan is presented in this paper. The

proposed device combines the laser scanner with a com-

puterized motion controller and an electronic balance. All

the setup components are connected to a computer to

gather all the relevant experimental information via a data

acquisition system. The paper also presents the proposed

methodology to process the scanned volume and to obtain

the digital model of the soil from the data file containing

the X, Y, Z coordinates of the scanned points.

Very valuable information can be obtained from the

digital model of the soil, such as volume change, distri-

bution of settlements on the soil surface, crack aperture and

morphology in depth. Scans at different times allow

learning about the volume changes at the onset of crack

formation and how the phenomena of crack propagation

evolve in time, in both soil surface and depth. Experimental

Fig. 16 Images of the soil sample at three different times: 4, 6, and 24 h. a Digital images (included for reference), b contour maps and c 3D cuts

of the sample, illustrating the volume changes (i.e., settlement, cracks and gaps) during shrinkage showing the initial level (red transparent

surface on the top) (color figure online)

594 Acta Geotechnica (2013) 8:583–596

123

Page 13: Exploring desiccation cracks in soils using a 2D profile laser device

results obtained from a test in a natural soil have been

presented and discussed.

The experimental outcomes show that the technique

based on laser scanning is able to gather very relevant data

associated with the behavior of drying soils, more infor-

mation than those typically gathered from the current

methods. It is in summary a promising experimental

technique that could assist in advancing of current

knowledge.

Acknowledgments The discussions with Drs Rebecca Lunn and

Minna Karstunen are highly appreciated. The fifth author would like

to acknowledge the financial support provided by the EU-funded

project RISMAC (Grant agreement: PIOF-GA-2009-254794).

References

1. Albright WH, Benson CH, Gee GW, Abichou T, McDonald EV,

Tyler SW, Rock SA (2006) Field performance of a compacted

clay landfill final cover at a humid site. J Geotech Geoenviron

Eng 132(11):1393–1403

2. Alshibli KA, Al-Hamdan MZ (2001) Estimating volume change

of triaxial soil specimens from planar images. Comput Aided Civ

Infrastruct Eng 16(6):415–421

3. Alshibli KA, Sture S (2000) Shear band formation in plane strain

experiments of sand. J Geotech Geoenviron Eng 126(6):495–503

4. Amarasiri AL, Kodikara JK, Costa S (2011) Numerical modelling

of desiccation cracking. Int J Numer Anal Meth Geomech

35(1):82–96

5. Armstrong A, Matthews A, Portwood A, Addiscott T, Leeds-

Harrison P (1994) Modelling the effects of climate change on the

hydrology and water quality of structured soils. Springer,

Heidelberg

6. Bear J (1988) Dynamics of fluids in porous media. Dover pub-

lications, Mineola, NY

7. Berney I, Ernest S, Hodo WD, Peters JF, Myers TE, Olsen RS,

Sharp MK (2008) Assessment of the effectiveness of clay soil

covers as engineered barriers in waste disposal facilities with

emphasis on modeling cracking behavior. DTIC Document

8. Bremand F, Trabelsi H, Jamei M, Guiras H, Hatem Z, Romero E,

Sebastia O (2010) Some investigations about the tensile strength

and the desiccation process of unsaturated clay. In: EPJ web of

conferences. EDP Sciences

9. Brossard C, Monnier J, Barricau P, Vandernoot F, Le Sant Y,

Champadal F, Le Besnerais G (2009) Principles and applications

of particle image velocimetry. Aerosp Lab J 1:1–11

10. Clayton C, Khatrush S (1986) A new device for measuring local

axial strains on triaxial specimens. Geotechnique 36(4):593–597

11. Daniel DE, Wu YK (1993) Compacted clay liners and covers for

arid sites. J Geotech Eng 119(2):223–237

12. Drumm E, Boles D, Wilson G (1997) Desiccation cracks result in

preferential flow. Geotech News 15(2):22–25

13. Dyer MR, Utili S, Zielinski M (2009) Field study into fine des-

iccation fissuring at thorngumbald. Proc ICE Water Manage

162(3):221–232. doi:10.1680/wama.2009.162.3.221

14. El Mountassir G, Sanchez M, Romero E, Soemitro R (2011)

Behavior of a compacted fill used to construct flood embank-

ments. Proc ICE Geotech Eng 164(3):195–210. doi:10.1680/

geng.10.00055

15. Flury M, Fluhler H, Jury WA, Leuenberger J (1994) Suscepti-

bility of soils to preferential flow of water: a field study. Water

Resour Res 30(7):1945–1954

16. Gachet P, Geiser F, Laloui L, Vulliet L (2007) Automated digital

image processing for volume change measurement in triaxial

cells. ASTM Geotech Test J 30(2):98–103

17. Hallett P, Newson T (2005) Describing soil crack formation using

elastic–plastic fracture mechanics. Eur J Soil Sci 56(1):31–38

18. Hong ES, Lee IM, Lee JS (2006) Measurement of rock joint

roughness by 3d scanner. ASTM Geotech Test J 29(6):482–489

19. Jackson P, Northmore K, Meldrum P, Gunn D, Hallam J,

Wambura J, Wangusi B, Ogutu G (2002) Non-invasive moisture

monitoring within an earth embankment—a precursor to failure.

NDT and E Int 35(2):107–115

20. Jardine R, Symes M, Burland J (1984) The measurement of soil

stiffness in the triaxial apparatus. Geotechnique 34(3):323–340

21. Jones G, Zielinski M, Sentenac P (2012) Mapping desiccation

fissures using 3-D electrical resistivity tomography. J Appl

Geophys 8:39–51. doi:10.1016/j.jappgeo.2012.06.002

22. Kim S (2013) Experimental study of soils subjected to drying,

PhD Thesis. Texas A&M University, College Station

23. Kodikara J, Nahlawi H, Bouazza A (2004) Modelling of curling

in desiccating clay. Can Geotech J 41(3):560–566

24. Lakshmikantha M, Prat P, Ledesma A (2006) An experimental

study of cracking mechanisms in drying soils. In: Proceedings of

the 5th international conference on environmental geotechnics.

Thomas Telford, London

25. Lakshmikantha M, Prat P, Ledesma A (2009) Image analysis for

the quantification of a developing crack network on a drying soil.

ASTM Geotech Test J 32(6):505–515

26. Lakshmikantha M, Prat PC, Ledesma A (2012) Experimental

evidence of size effect in soil cracking. Can Geotech J

49(3):264–284

27. Macari EJ, Parken J, Costes NC (1997) Measurement of volume

changes in triaxial tests using digital imaging techniques. ASTM

Geotech Test J 20(1):103–109

28. Melchior S (1997) In situ studies of the performance of landfill

caps(compacted soil liners, geomembranes, geosynthetic clay

liners, capillary barriers). Land Contam Reclam 5(3):209–216

29. Micro-Epsilon. http://www.micro-epsilon.co.uk/glossar/iconnect.

html. Accessed 20 Nov 2012

30. Miller CJ, Mi H, Yesiller N (2007) Experimental analysis of

desiccation crack propagation in clay liners. J Am Water Resour

Assoc 34(3):677–686

31. Morris PH, Graham J, Williams DJ (1992) Cracking in drying

soils. Can Geotech J 29(2):263–277

32. Musielak G, Sliwa T (2012) Fracturing of clay during drying:

modelling and numerical simulation. Transp Porous Med

95(2):465–481

33. Nahlawi H, Kodikara J (2006) Laboratory experiments on des-

iccation cracking of thin soil layers. Geotech Geol Eng

24(6):1641–1664

34. Obaidat MT, Attom M (1998) Computer vision-based technique

to measure displacement in selected soil tests. ASTM Geotech

Test J 21(1):31–37

35. Okabe A, Boots B, Sugihara K, Chiu SN (2009) Spatial tessel-

lations: concepts and applications of Voronoi diagrams, vol 501.

Wiley Series in Probability and Statistics, Wiley, Hoboken, New

Jersey

36. Peron H, Hueckel T, Laloui L, Hu L (2009) Fundamentals of

desiccation cracking of fine-grained soils: experimental charac-

terisation and mechanisms identification. Can Geotech J

46(10):1177–1201

37. Rayhani MHT, Yanful E, Fakher A (2007) Desiccation-induced

cracking and its effect on the hydraulic conductivity of clayey

soils from Iran. Can Geotech J 44(3):276–283

38. Rodrıguez R, Sanchez M, Ledesma A, Lloret A (2007) Experi-

mental and numerical analysis of desiccation of a mining waste.

Can Geotech J 44(6):644–658

Acta Geotechnica (2013) 8:583–596 595

123

Page 14: Exploring desiccation cracks in soils using a 2D profile laser device

39. Romero E, Facio J, Lloret A, Gens A, Alonso E (1997) A new

suction and temperature controlled triaxial apparatus. In: Pro-

ceedings of the international conference on soil mechanics and

foundations engineering. Society of soil mechanics and founda-

tion engineering. AA Balkema Publisher, pp 185–188

40. Romero Morales E (1999) Characterisation and thermo-hydro-

mechanical behaviour of unsaturated Boom clay: an experimental

study, PhD Thesis. Technical University of Catalunya

41. Rounsevell M, Evans S, Bullock P (1999) Climate change and

agricultural soils: impacts and adaptation. Clim Change

43(4):683–709

42. Sentenac P, Jones G, Zielinski M, Tarantino A (2012) An

approach for the geophysical assessment of fissuring of estuary

and river flood embankments: validation against two case studies

in England and Scotland. Environ Earth Sci 69(6):1939–1949

43. Shin H, Santamarina J (2011) Desiccation cracks in saturated

fine-grained soils: particle-level phenomena and effective-stress

analysis. Geotechnique 61(11):961–972

44. Software G (2012) Surfer (Golden Software). http://www.

goldensoftware.com/products/surfer/surfer.shtml. Retrieved 20

Nov 2012

45. Southen JM, Rowe RK (2005) Laboratory investigation of geo-

synthetic clay liner desiccation in a composite liner subjected to

thermal gradients. J Geotech Geoenviron Eng 131(7):925–935

46. Tang C, Shi B, Liu C, Zhao L, Wang B (2008) Influencing factors

of geometrical structure of surface shrinkage cracks in clayey

soils. Eng Geol 101(3):204–217

47. Tang CS, Cui YJ, Shi B, Tang AM, Liu C (2011) Desiccation and

cracking behaviour of clay layer from slurry state under wetting–

drying cycles. Geoderma 166(1):111–118

48. Tang GX, Graham J (2000) A method for testing tensile strength

in unsaturated soils. ASTM Geotech Test J 23(3):377–382

49. Trabelsi H, Jamei M, Zenzri H, Olivella S (2011) Crack patterns

in clayey soils: experiments and modeling. Int J Numer Anal

Meth Geomech 36(11):1410–1433. doi:10.1002/nag.1060

50. Velde B (1999) Structure of surface cracks in soil and muds.

Geoderma 93(1):101–124

51. Velde B (2001) Surface cracking and aggregate formation

observed in a Rendzina soil, La Touche (Vienne) France. Geo-

derma 99(3):261–276

52. Vogel HJ, Hoffmann H, Leopold A, Roth K (2005) Studies of

crack dynamics in clay soil: II. A physically based model for

crack formation. Geoderma 125(3):213–223

53. Vogel HJ, Hoffmann H, Roth K (2005) Studies of crack dynamics

in clay soil: I. Experimental methods, results, and morphological

quantification. Geoderma 125(3):203–211

54. White D, Take W, Bolton M, Munachen S (2001) A deformation

measurement system for geotechnical testing based on digital

imaging, close-range photogrammetry, and PIV image analysis.

In: Proceeding of the international conference on soil mechanics

and geotechnical engineering. AA Balkema Publisher,

pp 539–542

55. Yesiller N, Miller C, Inci G, Yaldo K (2000) Desiccation and

cracking behavior of three compacted landfill liner soils. Eng

Geol 57(1):105–121

56. Zhang X, Morris J, Klette R (2005) Volume measurement using a

laser scanner. Communication, and Information Technology

Research (CITR) Computer Science Department, The University

of Auckland

596 Acta Geotechnica (2013) 8:583–596

123


Recommended