Вы находитесь на странице: 1из 38

CHAPTER II

LITERATURE REVIEW

2.1 INTRODUCTION

Subgrades provide structural stability to pavements by transmitting


superimposed traffic loads safely to the soil strata below. The evaluation of subgrade
strength plays a major role in the design of pavements since traffic loads need to be
transmitted in a manner that the subgrade-deformation is within elastic limits, and the
shear forces developed are within safe limits under adverse climatic and loading
conditions. In this connection, subgrades need to be evaluated for the soil-stiffness
and strength. In addition to this, it is also required to investigate the influence of soil-
properties of the subgrade on pavement performance.

In a number of circumstances, road-engineers come across situations where


subgrades need to be strengthened in order to improve their load-carrying
characteristics. Also, road-engineers adopt alternative construction practices, and new
road-construction materials including the use of recycled aggregates in an effort to
economize road-construction costs (Fleming, 1998). Thus, it is often required to
evaluate the effectiveness of the improved subgrades on a regular basis.

Compacted soil subgrades are used to improve the engineering properties of


existing weaker subgrades. Subgrades for highways and airports need to be evaluated
and monitored for their stiffness, strength, and stability. These need to be designed to
resist stresses due to wheel-loads during the design life. One of the main parameters
used in estimating the stiffness and strength of subgrades is the elastic stiffness
modulus (modulus of resilience) of the pavement layers.

Flexible pavements are traditionally designed based on the CBR approach or


by considering elastic deformations as in the case of Burmister’s (1958) layer theory.
The CBR method of analysis gives more importance to the estimation of the density
of the subgrades and the pavement layers, while the analyses based on quality control
of pavements relies more on the determination of in-situ density and moisture content.
But according to Chen et al. (1999), and Livneh and Goldberg (2001), although
‘density’ is a good indicator of the strength of granular subgrades, it is also necessary

21
to investigate the modulus of resilience of the subgrade, since these measures
represent different natural characteristics.

The plate load test employed to determine the bearing capacity of the
subgrade, is very laborious, and time-consuming since the rate of loading is very slow.
In this test, a static load is applied through a circular plate and deflection is measured
using dial gauges placed over the circular plate. This test is advantageous in
determining the deformation modulus of the subgrade.

However, the invention of non-destructive methods including the FWDs, and


PFWDs, has revolutionized the approach to pavement evaluation. The portable falling
weight deflectometer (PFWD) is a non-destructive testing-device that assists in rapid
estimation of the modulus of resilience of subgrades. The use of non-destructive
approaches to pavement evaluation incorporating the use of PFWDs has gained
popularity in the recent years due to their inherent capability in obtaining quick
estimates of the modulus of resilience of the subgrade in addition to their simplicity of
design. This Chapter provides an overview of investigations on pavement evaluation
and correlation studies using different non destructive equipment.

2.2 HISTORICAL PERSPECTIVE OF SUBGRADE ANALYSIS

Pavement engineers continuously look for ways to improve pavement service


life and performance. Historically, pavement design procedures were empirical. In
many cases, relationships were based on factors such as traffic loading and volumes,
materials, layer configurations and the environment (Mahoney et al. 1991).

Traditional method of subgrade evaluation such as the CBR method has


played a significant role in pavement design. Other traditional approaches used in the
evaluation of sub-grade strength include the use of the vane-shear apparatus, cone-
penetrometers, unconfined compression test, and the Texas triaxial test (Ladner, 1973;
Clegg, 1983; Al-Joulani, 1987). However, these traditional soil tests do not fully
simulate actual loading conditions in the field. Instead, they measure different soil
properties that are related to the strength of the soil.

In view of the shortcomings on the use of traditional methods for pavement


evaluation, and the need to incorporate the use of modulus of resilience in pavement
design, there exists an urgent need to develop and evolve an effective approach to
evaluate sub-grade strength and stiffness using alternative equipment (Al-Amoudi et

22
al. 2002). Currently, falling weight deflectometers (FWD), GeoGauges, dirt-seismic
pavement analyzers (DSPA), and laboratory-based repetitive triaxial tests are being
used to estimate the modulus of elasticity of pavement layers (Nazarian et al. 2002;
Livneh and Goldberg, 2001; Rahim and George, 2002; Sawangsuriya et al. 2002).

Characterization of field material properties using laboratory tests is a usual


procedure adopted as part of the pavement design process. This procedure comprises
two aspects, collection of samples, and their testing. But one of the inherent defects in
this approach is that it is difficult to collect, and test representative samples due to the
large variability of pavement materials. This problem can be overcome by conducting
tests on a large number of random samples that can generate results with good
statistical significance. One of the other problems with this approach is that it is
difficult to simulate field conditions in the laboratory. This difficulty can be overcome
by resorting to in-situ strength. The inherent difficulties in using laboratory testing to
characterize field sub-grade conditions was discussed at the Transportation Research
Board meeting of 1993 (Burnham and Johnson, 1993). It was found that the cost of
pavement construction based on this approach was exorbitant (Ksaibati et al. 1994).

Originally, testing-piles were used as devices for estimation of soil strength


properties. Pile-driving is thus regarded as an earlier method of penetration testing
especially in the 15th century (Burnham and Johnson, 1993). In 1872, Sir Stanford
Fleming proposed a soil investigation method where the resistance to penetration of a
steel rod driven into the soil was taken as a measure of soil strength (Broms and
Flodin, 1988). Later, Nicholaus Goldmann (17th century) of Germany developed the
subsoil penetration testing device called the "ram penetrometer" (Luo et al. 1998).
This is the similar to the present day DCP. Subsequently, Kiinzel (1936) of Germany
developed the "Prufstab", a device that was later used extensively by Paproth (1943).
This device was standardized in 1964, and was named as the "Light Penetrometer"
(German Standard DIN 4094, 2002).

Subsequently, various countries developed their own standard penetration


devices. The DCP was originally fabricated in Australia by Scala (1956). This device
was the adopted as the Central African Standard, and was also modified by Van
Vuuren (1969) in South Africa. It came to be adopted by the United States based on a
number of investigations conducted at the University of Illinois by Marshall
Thompson. The Mn/DOT obtained the specifications from the University of Illinois,

23
and fabricated two DCPs that were used in the studies performed by the Minnesota
Road Research Project -Mn/ROAD (Burnham and Johnson, 1993). The technical
details of the DCP and the testing procedure were well-documented by TRL (1999),
and MnDOT (1996), and were formalized as national standards in USA, and Australia
(ASTM-D-6951-03, 2003; AS1289.6.3.2, 1997). Subsequently, various other
countries developed their own standard penetration devices.

Penetration tests mainly measure the resistance of soil to deformation by


shearing (Gidigasu, 1976). The DCP has been considered to be an effective tool for
assessing in-situ strength and stiffness of pavement and subgrade (Abu-Farsakh et al.
2004). This device can verify both the level and uniformity of compaction for the
layers tested, which makes it an ideal device for quality control in pavement
construction (Chen et al. 2001), in addition to providing soil-investigators an
opportunity to verifying the thickness of the layers tested.

The DCP is one of the most widely used penetration devices for subgrade, and
pavement evaluation. It is a simple low-cost testing device that can be easily
fabricated. It is highly portable, and easy to operate. The DCP can assist road
engineers in performing penetration tests within 15 minutes for each location. It is
also capable of assessing the strength characteristics of underlying hard subgrades,
and can be used in estimating the CBR values of the subgrades (Jahren et al.1999).

Although the modulus of resilience of the subgrade is used as a measure to


determine the required layer thickness of a pavement structure, the approach to
quality control of pavements gives more importance to the determination of in-situ
density and moisture content. But according to Chen et al. (1999) and Livneh and
Goldberg (2001), although ‘density’ is a good indicator of the strength of granular
materials, the density and modulus of resilience of sub-grade are two different natural
characteristics of sub-grades that need to be investigated.

Over the past two decades, a number of modifications have been incorporated
to traditional pavement design techniques to include elastic and/or visco-elastic
theories, along with empirical approaches. These new mechanistic-empirical
procedures enable engineers to examine the stresses, strains, and deflections in the
pavement structure, in addition to establishing empirical relationships between the
mechanistic responses, and the performance of the pavement (Ksaibati et al. 1994).

24
In view of the advantages of the DCP, and its increasing popularity, a number
of studies were conducted using the DCP, and the CBR observations for various field
conditions. This contributed to the development of relationships between mechanistic
responses of the DCP (such as the rate of penetration) to the pavement performance
evaluation using the CBR. A review of the literature shows that a number of CBR and
DCP correlations have been developed using values obtained from both laboratory
and field-testing (Harison, 1987; Webster et al. 1992; Kleyn et al. 1975; Van Vuuren,
1969).

One of the important conclusions based on the American Association of State


Highway Officials (AASHO) Road Test conducted from October 15, 1958 to
November 30, 1960 in Ottawa, Illinois was that the deflection of the pavement surface
resulting from the application of a load is a strong indicator of pavement performance.
It was observed that the subgrade deformations amounted to 60 to 80 percent of the
surface deflections observed in the pavements (HRB, 1962). In the elastic approach to
design of pavements, it is important to evaluate the elastic modulus of resilience of the
subgrades.

Static deflection measurement devices estimate the pavement response to


loads applied with slow moving vehicles or a stationary loading frames (Stoffels and
Lytton, 1987). NDT devices used in the measurement of static deflections include the
Benkelman beam, the California traveling deflectometer, and the LaCroix
deflectometer (Ksaibati et al. 1994). But according to AASHTO (1993), pavement
design may be performed using the modulus of resilience.

A number of studies have proved that the modulus of resilience can be used as
an indicator of pavement strength. Therefore, the modulus of resilience test for
subgrades was considered to be an important part of flexible pavement performance
(Ksaibati et al. 1994). Seed et al. (1962) pioneered the development of the modulus of
resilience test where logical approaches to pavement design and materials testing
were devised using the Stabilometer, and the Resilometer. A modem variation of the
latter equipment is known as the K-Mould (Semmelink and De Beer, 1995) in which
lateral stress is mobilized by an elastic support system with a stiffness which can be
set to simulate a range of in-situ conditions.

25
In 1986, the AASHTO pavement design procedures took steps in the right
direction by allowing the incorporation of non-destructive testing (NDT) approaches
for measurement of pavement deflection and the defining of reliability factors based
on modulus of resilience in the pavement design procedure (Ksaibati et al. 1994). The
FWD, LFWD, and the PFWD, are non-destructive testing devices that can be used to
determine the modulus of resilience of the soil (Alshibli et al. 2005).

Although no standard field tests are specified for the determination of the
modulus of resilience of base-courses and sub-grades for ensuring the quality of road
construction, road transport authorities use the stiffness of pavement layers, and the
modulus of resilience of sub-grades as references along with the information on
density and moisture content (Chen et al. 2005). Thus, in circumstances where
pavements are designed based on traditional techniques or otherwise, it also becomes
necessary to estimate the stiffness or modulus of resilience of the pavement before and
especially after their construction as part of the quality-control (QC) measure and also
for quality assurance (QA) (Chen et al. 2005). Thus, it is possible to link observations
of the modulus of resilience to the mechanistic responses of related studies using the
plate bearing test (PBT), DCP, and the CBR in an effort to develop relationships
useful to the pavement design community.

Non-destructive testing (NDT) technology as applied to pavement evaluation


has made substantial progress in the last two decades. In the past decade, non-
destructive approaches using the ground penetrating radar (GPR), and seismic
techniques have been automated. Currently, the deflection-based non-destructive
pavement evaluation methods, as implemented using Falling Weight Deflectometers
(FWD), are extensively used in day-to-day activities by many highway agencies. The
FWD, and the GPR methods have gained popularity, and are being adopted widely by
a number of highway agencies. The GPR is an investigative tool to check for
pavement anomalies, and layer thicknesses. The following sections present a literature
review on important pavement evaluation approaches using the CBR, DCP, and the
PBT, in addition to non-destructive approaches using the FWD, LFWD, DCP, and the
Geo-gauge. Studies on important correlations developed are also discussed in the
relevant sections.

26
2.3 STUDIES ON CORRELATIONS BETWEEN THE CBR VALUE, AND THE
SOIL PROPERTIES

The California Bearing Ratio (CBR) method is one of the traditional


approaches to pavement evaluation. It provides an estimate of the shear strength of
soil subgrades. It is an indirect measure that compares the strength of the subgrade
material to the strength of standard crushed rock. This method was originally
developed by the California Division of Highways in 1930s to determine the relative
stability of subgrades with respect to that of standard crushed rock base material
(Mak-wai-kin, 2006).

A number of studies have been conducted using the CBR, for the
determination of pavement strength, the results of which are used in pavement design.
Black (1962), De Graft-Johnson and Bhatia (1969), Agarwal and Ghanekar (1970),
Highway Agency (1994), and NCHRP (2001) discussed the effects of soil types and
soil-characteristics on CBR values, while Vuuren (1969), Fine and Remington (1972),
Kleyn (1975), Harison (1987), Webster et al. (1992), and others have contributed
towards the application of the CBR approach to pavement design and evaluation.

Among the various attempts at correlating the California Bearing Ratio (CBR)
values to the grain-size distribution and plasticity of soils, Black (1962) developed a
method of estimating the CBR value for cohesive soils based on the plasticity index
and the liquidity index as shown in Fig. 2.1. In this study, the soil samples were
compacted in 5 layers to maximum dry density at optimum moisture content, and
soaked for 4 days according to the Ghana Standard of compaction. The subgrade
material in the CBR mould was compacted using a 4.5 kg rammer falling through a
drop-height of 450 mm. Each layer was compacted using 25 blows of the rammer.

Fig. 2.1 CBR based on plasticity index for various liquidity indices

27
De Graft-Johnson and Bhatia (1969) developed a correlation relating the CBR
to the plasticity index (Ip), liquid limit (wL), percentage passing 2.4 mm British
Standard sieve (A), and the grading of the subgrade-soil based on the concept of a
suitable index (A/ (wL LogIp )) as shown in Fig. 2.2.

Fig. 2.2 CBR based on suitability index for soaked soils


Agarwal and Ghanekar (1970) attempted to develop a correlation between the
CBR values and the liquid limit, and between the CBR values and the plastic limit or
plasticity index, based on a study on 48 samples of fine-grained soils of India. But no
meaningful correlation was found. However, on further investigation, it was observed
that the CBR values could be correlated to the optimum moisture content (OMC), and
the liquid limit (wL). The CBR values of the samples tested were lesser than 9 percent,
and the standard deviation was found to be 1.8. The relationship developed is given
as,

CBR = 2 – 16 Log (OMC) + 0.07 wL Eq.


2.1

They also developed a correlation between the CBR values for soaked samples
to the ratio between the maximum dry-density and the plasticity index. See Fig. 2.3.

Similarly, the Highway Agency (1994) had proposed to estimate the CBR
values based on the plasticity index for British soils compacted at field moisture-
contents (FMC) as shown in Table 2.1. However, in this study, the precise density,
moisture conditions, and the depth of water tables below the formations, were not
specified.

28
Fig. 2.3 CBR based on ratio of maximum dry density to plasticity index for soaked
laterite-quartz gravels
Table 2.1 CBR Estimation for British Soils Compacted to FMC
Type of Soil Plasticity Index (%) Predicted CBR (%)
Heavy Clay 70 2
Heavy Clay 60 2
Heavy Clay 50 2
Heavy Clay 40 2 to 3
Silty Clay 30 3 to 4
Silty Clay 20 4 to 5
Sandy Clay 10 4 to 5
Sand (Poorly Graded) - 20
Sand (Well Graded) - 40
Sandy Gravel (Well Graded) - 60
Source: Highway Agency, 1994

In the above table, it may be observed that the soil type plays crucial role in
determining the CBR values. In the case of coarse-grained soils, the predicted CBR
values ranged between 20 to 60%, while for fine-grained soils, the CBR values
predicted, ranged between 2 to 5%. Thus, it can be observed that coarse-grained or
cohesion less soil has higher CBR values when compared to fine-grained or cohesive
soils.

In this study, it was also found that the grading of soil samples influence the
CBR values significantly. Poorly graded sand samples were found to have lesser CBR
values of about 20%, while well-graded samples of sand and subgrades gave higher
CBR values of about 40%. Thus it was confirmed that the density of soils and their
strengths depend upon the grading of the soil sample. But the density of soils is also

29
dependent on the moisture content. Thus the CBR can be considered to be dependent
on not only the soil-index properties, but also on the density, and moisture content.

The Guide for Mechanical-Empirical Design of New and Rehabilitated


Pavement Structures (NCHRP, 2001) developed correlations between the CBR values
to the soil index properties, the D60 values (representing the diameter of the sieve used
in grain-size distribution analysis, for which 60% of the particles pass), the fraction of
soil (w in decimal) passing 75 micron sieve or No.200 U.S. sieve, and the plasticity
index (Ip). The expression given below provides the correlation developed for non-
plastic coarse-grained soils. This relationship was derived for D60 values ranging
between 0.01mm and 30mm. For D60 values lesser than 0.01mm, the NCHRP
recommends a CBR value of 5%, and where the D60 values are greater than 30mm, the
CBR value recommended was 95%. The NCHRP (2001) also developed a correlation
for other fine-grained plastic soils with 12% fines as shown in Eq. 2.3.

CBR= 28.09 (D60)0.358 Eq. 2.2


CBR = 75/ [1+0.728(w. Ip)] Eq.
2.3

2.4 STUDIES ON CORRELATIONS BETWEEN CBR AND DCPI

This section provides a brief review of investigations on correlation-studies


performed using the CBR, and the DCP. Vuuren (1969), Kleyn (1975), and Harison
(1983) have reported the results of preliminary studies using the DCP. Livneh (1989)
performed a number of studies on correlations between the observed values for DCP
and CBR and developed 20 relationships, of which the most accepted, are listed
below (Luo et al. 1998);

Log CBR= 3.38 - 0.71 (log DCPI)1.5 Eq. 2.4

Log CBR= 4.66 - 1.32 (log DCPI) Eq.


2.5

CBR= 405.3 (DCPI)-1.259 Eq. 2.6

Log CBR= 2.0 - 1.3 log (DCPI- 0.62) Eq.


2.7

where, DCPI is the penetration index in mm per blow. Eq. 2.4 was developed from
results obtained using a DCP fitted with a 30 degree cone tip, while Eq. 2.5, Eq. 2.6,

30
and Eq. 2.7 were developed based on data obtained using a 60-degree cone tip. The
studies showed that the Penetration Indices (DCPI) could be converted to equivalent
CBR measures of stability and strength.

Extensive studies have been conducted to investigate the correlations between


DCP and CBR enhanced the reliability of the DCP method and its capability in
estimating the CBR values more accurately (Harison 1987; Livneh and Ishai 1988;
Chua 1988; Smith and Pratt 1983; Livneh et al. 1992; Livneh and Livneh 1994; Ese et
al. 1994; and Coonse 1999) as reported by Amini (2003). The most widely accepted
relationships between California Bearing Ratio (CBR) values and the DCP penetration
index (DCPI) measured in mm per blow based on these studies are summarized in
Table 2.2.

Table 2.2 CBR-DCP Relationships


Correlation Equation Material tested Reference
log (CBR)= 2.55 – 1.14 log (DCPI) Granular and cohesive Harison (1987)
log (CBR)= 2.56 – 1.16 log (DCPI) Granular and cohesive Livneh (1989)
log (CBR)= 2.45 – 1.12 log (DCPI) Granular and cohesive Livneh et al. (1992)
log (CBR)= 2.46 – 1.12 log (DCPI) Various soil types Webster et al. (1992)
log (CBR)= 2.62 – 1.27 log (DCPI) Unknown Kleyn (1975)
log (CBR)= 2.44 – 1.07 log (DCPI) Aggregate base course Ese et al. (1995)
log (CBR)= 2.60 – 1.07 log (DCPI) Aggregate base course & NCDOT (Pavement,
cohesive soil 1998)
log (CBR)= 2.53 – 1.14 log (DCPI) Piedmont residual soil Coonse (1999)
Source: Amini (2003)

A visual inspection of Table 2.2 reveals that most of the relationships between
DCPI and CBR have the following form:

Log (CBR) = a + b log (DCPI) Eq.


2.8

where, a = a constant that ranges from 2.44 to 2.60; and b = a constant that
ranges from -1.07 to -1.16.

The U.S. Army Corps of Engineers (Webster et al. 1992) suggested the
following relationship between DCPI and CBR readings:

CBR= (292 / DCPI) 1.12 Eq.


2.9

31
However, based on further investigations conducted at the Waterways
Experiment Station (WES) for soils with CBR values less than 10%, and for fat-clay
soils, Eq.2.10, and Eq. 2.11 respectively were proposed (Webster et al. 1992).

CBR= 1/ (0.017019 DCPI) 2 Eq.


2.10

CBR= 1/ (0.002871 DCPI) Eq. 2.11

Webster et al. (1992) has also reported the development of a relationship


between the CBR value, and the penetration-rate (DCPI) of the DCP, expressed in
mm per blow as in Eq. 2.12 by the U.S. Army Corps of Engineers for a wide range of
granular and cohesive materials. This correlation has been adopted by many
researchers (Livneh, 1989; Webster et al. 1992; Seikmeir et al. 1999) in their
investigations.

Log CBR = 2.465 – 1.12 Log (DCPI) Eq.


2.12

Sood et al. (1996) conducted studies at 27 locations on 3 categories of North


Indian soils namely SM, ML, and CL and developed correlations between field CBR
(FCBR) and DCPI values. The relations developed, along with the R-square values
and the t-test values are shown below:

Soil type: SM
Log (FCBR) = 1.73 - 1.63Log (DCPI) + 1.02(FDD) - 0.06(FMC) Eq. 2.13
R2 = 0.91 [6.35] [1.32] [1.64] [21]

Soil type: ML
Log (FCBR) = 0.45 - 0.30Log (DCPI) + 0.73(FDD)) - 0.02 (FMC) Eq. 2.14
R2 = 0.91 [1.46] [1.81] [2.11] [0]

Soil type: CL
Log (FCBR) = 0.82 - 0.41Log (DCPI) + 0.66 (FDD) - 0.03 (FMC) Eq. 2.15
R2 = 0.86 [1.03] [2.08] [1.57] [0]

A correlation with R2 value of 0.85 was also developed between laboratory


CBR and DCP values of remolded samples based on the limited tests conducted on
un-soaked samples as shown below:

32
CBR = 24.277 – 10.253 Log (DCPI) Eq.
2.16

Gabr et al. (2000) investigated the use of DCP for evaluation of pavement
distress for experiments conducted on subgrades overlaid with aggregate base course
of granite of Los Angeles value of 20%, and developed the expression shown in Eq.
2.17.

log (CBR) = 1.4 – 0.55 log (DCPI) Eq.


2.17

Karunaprema and Edirishinghe (2002) were performed laboratory tests on


Srilankan residual soils by varying the moisture content and the dry density. Based on
these studies, three correlations were established relating the CBR values to the DCPI
values for disturbed un-soaked CBR samples (CBRdu), undisturbed un-soaked CBR
samples (CBRuu), and disturbed soaked CBR samples (CBRds), as expressed in Eq.
2.18, Eq. 2.19, and Eq. 2.20, respectively for DCPI ranging between 7mm, and 75mm
per blow.

Log CBRdu= 2.182 – 0.872 Log DCPI Eq.


2.18

Log CBRuu= 1.145 – 0.336 Log DCPI Eq.


2.19

Log CBRds= 1.671 - 0.577 Log DCPI Eq.


2.20

The above investigations also revealed a strong relationship between the DCP
and CBR for un-soaked samples (CBRu) for DCPI ranging between 11mm, and 386
mm per blow for soft soils. The power model proposed based on this study as in Eq.
2.21, gave the smallest mean-square error (MSE) of 0.112 and the highest coefficient
of determination (R2) of 0.95 for 23 observations.

Log CBRu = 1.966 – 0.667 Log DCPI Eq.


2.21

But the above investigations on Srilankan residual soils failed to developed


similar relationships for soaked soil samples although it is an accepted fact that an
increase in field moisture content will lead to deeper penetrations. However, it was

33
found that there was a significant correlation between the DCPI and the difference
between the un-soaked and the soaked CBR value (CBRu – CBRs). The best-fit model
was the logarithmic model that gave a mean-square error (MSE), of 2.58 and a
coefficient of determination (R2) of 0.89. The relationship formulated was as given
below:

CBRu – CBRs = 25.611 – 11.5 Log DCPI Eq.


2.22

The relationship was derived for DCPI values ranging from 11 to 104
mm/blow. A better correlation was obtained considering the effects of moisture
content (MC) additionally, that yielded a relationship with a coefficient of
determination (R2) of 0.91 for 11 observations for DCPI values in the same range as
follows:

CBRu – CBRs= 67.12–1.48MC–30.64DCPI1/MC Eq. 2.23


Karunaprema and Edirishinghe (2002) also found that the DCPI value
increases with the moisture content and that a significant correlation exists between
them. The logarithmic model was found to fit the data with a mean-square error
(MSE) of 4.51, and a coefficient of determination (R2) of 0.71 for 23 observations for
DCPI values ranging between 11 and 386 mm/blow.

MC = 0.5 + 6.9 Log DCPI Eq.


2.24
The above studies also revealed that when moisture content was used as an
additional independent variable, it was possible to develop a significant correlation
between the dry density (DD) measured in kg/m3, and the DCPI values with a
coefficient of determination (R2) of 0.77, for the above ranges of DCPI values, and for
moisture contents varying between 5 to 20 percent.

DD = 1940.75 - 1783.34[1/ (1 + MC)] – 0.058DCPI Eq.


2.25
Karunaprema, and Edirishinghe (2002) also proved that there existed a
significant correlation between the DCPI and the degree of compaction when the
moisture content was also considered as an independent variable. The correlation
developed had a coefficient of determination (R 2) of 0.87, for the above ranges of
DCPI values, and moisture contents.

34
DD/MDD = 1.126 + 0.064MC + 0.126DCPI 1/MC Eq. 2.26

The above investigations also showed that the ratio between the moisture content, and
the optimum moisture content (OMC) could be correlated to the DCPI values with a
coefficient of determination (R2) of 0.97, for the above ranges of DCPI values, and
moisture contents.

MC/OMC = -0.186 + 0.064MC + 0.126DCPI 1/MC Eq.


2.27

Abu-Farsakh et al. 2004 carried out regression analyses to correlate the CBR
values to the DCPI obtained from field tests. The following non-linear regression
model with an R2 value of 0.93 was obtained.

CBR= 5.1/ (DCPI 0.2-1.41) Eq.


2.28

Abu-Farsakh et al.2005 carried out regression analysis to correlate the DCPI


with CBR values obtained from the laboratory, the field, and the combined data. They
yielded non-linear regression models as given below in Eq. 2.29, Eq. 2.30, and Eq.
2.31 with an R-square value of 0.56, 0.81, and 0.79 respectively.

Log CBR = 2.256 - 0.954 Log (DCPI) Eq.


2.29

CBR = 1.03+(2600/(DCPI1.84-7.35)) Eq.


2.30

CBR = 1161.1/DCPI1.52 Eq.


2.31

2.5 STUDIES ON CORRELATION BETWEEN CBR, CLEGG IMPACT


VALUE (CIV), AND OTHER TESTS

Al-Amoudi et al. (2002) performed investigations on the estimation of


strength of compacted soils of Saudi Arabia. Two relationships of R-square values of
0.76, and 0.85, for GM and SM soils respectively, relating CBR and Clegg Hammer
Impact Values (CIV) were developed as given below:

CBR = 0.861 (CIV)1.136 Eq.


2.32

35
CBR = 1.3577 (CIV)1.011 Eq.
2.33

Rosyidi et al. (2006) developed a method called ‘spectral analysis of surface


wave’ (SASW) in order to estimate the stiffness of the pavement foundation. The
study produced relationships between the CBR value and the shear wave velocity (VS)
measured in m/s, and relationships between CBR values and the maximum shear
modulus (G) measured in MPa, as in Eq. 2.34 and Eq. 2.35. The correlation between
CBR values and the maximum shear modulus (G) was observed to be high with an R-
square value of 0.947.

CBR = 0.0006 (VS) 1.99 Eq.


2.34
CBR = 0.266 (G) 1.0027 Eq. 2.35

2.6 STUDIES ON CORRELATIONS BETWEEN MODULUS OF ELASTICITY


AND CBR OBSERVATIONS

A number of investigations were conducted using the CBR equipment on the


determination of in-situ strength characteristics of pavement layers. The CBR
approach was extensively used in testing granular soil-samples. Nascimento and
Simoes (1957) developed relationships between the modulus of strength (which is the
product of the modulus of subgrade reaction and the diameter of the loading plate),
and the CBR values.

Huekelom and Forster (1960) correlated the modulus of elasticity to the CBR
values using wave propagation techniques. Black (1962), and Agarwal and Ghaneker
(1970) have developed regression models for the prediction of bearing capacity of
soils based on the CBR value, and the plasticity of soils. A number of studies have
been made to estimate the CBR value of soils based on the soil-classification, other
index values, and physical properties of the soil. Most of these investigations focus on
specific soil types of particular regions.

Chen et al. (2005) reported studies conducted by AASHTO (1993) for the
design of pavements where a correlation model as in Eq. 2.36 was developed for the
determination of the modulus of elasticity of the subgrade (Es) based on CBR
observations for experiments conducted on fine-grained soils with a soaked CBR of
10 or less.

36
Es = 10.342 x CBR Eq.
2.36

Empirical correlations between the elastic modulus (E) the CBR, and the DCPI
have been proposed by a number of researchers. The Transportation Research
Laboratory (TRL) developed a relationship as given in Eq.2.37, connecting the elastic
modulus (E) and the CBR of the subgrade soil based on investigations performed by
Powell et al (1984). This equation has been established primarily based on data
relating modulus measured using the wave propagation technique, to the in-situ CBR
tests performed on remolded and undisturbed subgrade soils (Jones, 1958) as reported
by Edil and Benson (2005).

E = 17.6 (CBR) 0.64 Eq. 2.37

The National Council of Highway Research Program (NCHRP, 2001)


proposed Eq. 2.38 based on the index properties of the soil such as D60, and a
modified form of Eq. 2.37 given in Eq.2.39 for the modulus of resilience measured in
Psi. D60 represents the diameter of the sieve size in mm through which 60% of the soil
fractions pass through. Eq.2.38 holds good for D60 values ranging between 0.01 mm
and 30 mm.

CBR = 28.09 (D60)0.358 Eq.


2.38

E = 2555 CBR0.64 Eq.


2.39

Livneh and Goldberg (2001) carried out comparative German light drop
weight (LDW), and DCP tests. The relationship between the modulus measured by the
LDW (ELDW) in MPa and the in situ CBR values obtained from the DCPI is expressed
as follows for clayey and sandy soils, respectively.

E LDW = 600 ln (300 / (300 - 6.019 CBR1/1.41) Eq. 2.40

E LDW = 600 ln (300 / (300 - 4.035 CBR1/1.41) Eq. 2.41

In an attempt to correlate alternative approaches to subgrade evaluation to the


CBR method, Nazzal (2003) has suggested the use of the expression as given in Eq.
2.42 with an R-square value of 0.84 for predicting the CBR value based on the
modulus of the subgrade (EG) measured in MPa, estimated using the GeoGauge.

37
CBR = 0.00392 (EG) 2 – 5.75 Eq.
2.42

Nazzal (2003) also performed studies on correlating the CBR values to the
modulus of resilience measured using the Prima 100 LFWD (Epfwd), and developed the
expression as given in Eq. 2.43 for Louisiana soils. The relationship was developed
for Epfwd values ranging between 2.5 and 174.5 MPa, and had an R-square value of
0.83.

CBR = - 14.0 + 0.66 (Epfwd) Eq.


2.43

Abu-Farsakh et al. (2004) performed studies on correlating the stiffness


modulus values determined using the Geo-gauge (EG) to the CBR values based on
laboratory investigations. The following correlation with an R2 value of 0.62 was
obtained.

log (EG) = 1.89 + 1.48 log (CBR) Eq.


2.44

They also developed a correlation between the stiffness modulus values


determined using the Geo-gauge (EG) and the CBR values based on field
investigations. The following correlation with an R2 value of 0.84 was obtained.

CBR = 0.00392 (EG) 2- 5.75 Eq.


2.45

In addition to the above, Abu-farsakh et al. (2004) also made efforts to


develop a correlation between the modulus of resilience determined using the LFWD
and the CBR values. It was found that there was no clear correlation between them
since the data was widely scattered. However, a log-log relationship as given below,
with a low R2 value of 0.36, was suggested.

Log (ELFWD) = 1.149 + 0.702 log (CBR) Eq.


2.46

Wu and Sargand (2007) reported the use of a relationship developed by the


U.S. Army Corps of Engineers Research, and the Development Center Waterways
Experiment Station, relating the values of E, the modulus of resilience (measured in
MPa), and the CBR observations as given below:

38
E = 37.3CBR0.711 Eq.
2.47

Wu and Sargand (2007) also reported the use of a similar expression


developed by the Danish Road Laboratory as given below:

E = 10CBR0.73 Eq.
2.48

2.7 STUDIES ON CORRELATIONS BETWEEN MODULUS OF RESILIENCE,


AND DCPI

This section provides information on correlation studies undertaken by various


researchers relating the modulus of resilience of soils and the DCP observations. Abu-
Farsakh et al. (2004) have reported a theoretical relationship given below, developed
by Chua and Lytton (1981) correlating the DCPI observations (measured in
mm/blow) to Es, the modulus of resilience measured in MPa, with an additive
constant (B) the value of which, depends upon the principal stress difference 2τo. The
values of 2τo for plastic clays, clayey soils, silty soils, and sandy soils are 25, 50, 75,
and 150 respectively, while the values of B for these soils are 2.22, 2.44, 2.53, and
2.63 respectively.

Log (Es) = B - 0.4 log (DCPI) Eq.


2.49

De Beer (1990) proposed a simple relationship with an R-square value of 0.76


between the back-calculated modulus of resilience (Es) determined using a Heavy
Vehicle Simulator, and the penetration indices (DCPI) of the DCP measured
expressed in mm per blow, as expressed in Eq. 2.50, based on 86 observations:

Log (Es) = 3.05 − 1.07 Log (DCPI) Eq.


2.50

Experiments conducted by Hassan (1996) succeeded in the development of a


simple regression-model where the penetration index (DCPI) expressed in inches per
blow for the DCP tests were correlated to the modulus of resilience of soils (ΜFWD)
measured in psi as expressed in Eq. 2.51, for fine-grained soils at optimum moisture
content (OMC):

39
ΜFWD =7013.065 −2040.783 ln (DCPI) Eq.
2.51

Based on DCP tests and CBR-DCP relationships developed in Malaysia during


the National Axle Load study conducted in 1987, Chai and Roslie (1998) proposed the
following model where the penetration index (DCPI) expressed in blows per
penetration of 300mm for the DCP tests, were correlated to the modulus of resilience
(ΜFWD) measured in MPa as expressed in Eq. 2.52.

ΜFWD = 17.6(269/ DCPI) 0.64 Eq.


2.52

Chai and Roslie (1998) also developed a relationship between the back
calculated modulus of resilience (Eback) in MPa, and the DCPI value in the following
form:

Eback = 2224 DCPI - 0.996 Eq.


2.53

Jianzhou et al. (1999) found that there was a strong relationship between DCPI
and the moduli back-calculated from FWD in the following form:

Eback = 338 DCPI - 0.39 Eq. 2.54


George and Uddin (2000) developed relationships between the modulus of
resilience (MR) and DCPI as a function of moisture content, liquid limit, and density.
Due to the MDOT requirements for being able to correlate MR in real time, they also
provided simpler one-to-one relationships between DCPI and MR. The following
relationship was developed for fine-grained soils:

MR = 532.1 DCPI-0.492 Eq.


2.55

The relationship for coarse-grained soils is of the following form:

MR = 235.3 DCPI-0.475 Eq. 2.56

Seyman (2003) has reported the use of relationships developed by Pen (1990)
where the back-calculated modulus of resilience (Es) of pavements measured in MPa
can be estimated using the penetration-rate (DCPI) of DCP tests as in Eq. 2.57, and
Eq. 2.58, where the calculation of the modulus of resilience was performed using the

40
Phoenix system, and the Peach system respectively. The R-square values of these
relationships developed were 0.56 and 0.81 respectively.

Log Es = 3.250 – 0.89 log DCPI Eq.


2.57

Log Es = 3.652 – 1.17 log DCPI Eq.


2.58

Srinivasa kumar et al. (2003) reported regression models between DCP values
and back-calculated subgrade moduli (MR) as a function of field moisture content
(FMC) and, the total over burden (pavement) thickness H for the subgrade soils. Eq.
2.59 developed for clayey soils had an R 2 value of 0.14, a standard error of 0.102, and
a t-statistic of -12.4. Eq. 2.60 for silty soils had an R 2 value of 0.787, and a standard
error of 4.2.

MR = 357.87 DCPI-0.6445 Eq.


2.59

MR = 40.71 – 0.0307 H + 25.46 H0.2 – 0.0584 DCPI2 – 4.194 DCPI Eq. 2.60

Gudishala (2004) reported the study conducted by George et al. (2000) where
the DCPI values, the actual moisture content (wc %), the liquid limit (LL %), the
density ratio (γdr,), the coefficient of uniformity (cu), and the moisture ratio (wcr) are
correlated to the field moisture/optimum moisture) the modulus of resilience observed
in the laboratory. The density ratio (γdr,) refers to the ratio between the field density
and the maximum dry density, whereas, the moisture ratio (wcr) refers to the ratio
between the field moisture content and the optimum moisture content. They proposed
two different models based on their investigation, one for coarse grained soils and the
other for fine grained soils with ao, a1, a2 and a3 as regression coefficients. The models
proposed for fine grained soils and coarse grained soils are given below:

MR = a0 (DCPI)a1( γdra2+(LL/wc)a3) Eq.


2.61

MR = a0 (DCPI/log cu) a1(wcra2+γdra3) Eq.


2.62

Chen et al. (2005) combined the results of the studies of AASHTO (1993) and
Powell et al. (1984) to obtain a direct relationship between the penetration rate

41
(DCPI) for DCP tests measured in mm per blow, and the modulus of resilience
(calculated in MPa) as in Eq. 2.63.

Es = 664.67 x DCPI –O.7168 Eq.


2.63

Further studies by Chen et al. (2005) using DCPs and FWDs for roads in
Texas provided correlations connecting the layer-moduli (Es), and the penetration-rate
(DCPI) measured in mm per blow for DCP tests, as expressed in Eq. 2.64. This
relationship had an R-square value of 0.855 and a mean-square error of 0.15.

Es = 537.76 x DCPI -0.6645 Eq.


2.64

Abu-Farsakh et al. (2005) in their investigation analyzed to find the best


correlation between MFWD and the DCPI. The results of this analysis yielded the non-
linear regression model with an R-square value of 0.91 based on the results of
regression analysis between MFWD and DCPI of values ranging from 10 to 60
mm/blow.

ln(MFWD) = 2.35 +(5.21/ln DCPI) Eq.


2.65

Mohammad et al. (2007) presented two models vide Eq. 2.66 and Eq. 2.67, for
prediction of the modulus of resilience (MR) of cohesive subgrade soils of water-
content w (expressed in percentage), and dry unit weight γd expressed in (kN/m3). The
R2 values of these relationships were observed to be 0.91, and 0.92 respectively, while
the RMSE values were 6.1 Mpa and 5.9 Mpa respectively.

MR = 151.8/ (DCPI) 1.096 Eq.


2.66

MR = 165.5(1/DCPI1.147) + 0.0966(γd/w) Eq.


2.67

Nazzal et al, (2007) proposed a relationship between the ELFWD and the DCPI
as follows, with an R-square value of 0. 87 and an RMSE value of 20, respectively.

ELFWD = 5301.54/ (8.31+ DCPI1.44) Eq.


2.68

42
Abu-farsakh et al. (2004) conducted regression analysis between the back-
calculated modulus of resilience obtained from the tests using FWD (MFWD), and the
stiffness modulus (EG) measured using Geo-gauge for field data and developed a
linear correlation with an R2 value of 0.81.

MFWD = -20.07 + 1.17 (EG) Eq.


2.69

2.8 STUDIES ON CORRELATIONS BETWEEN BACK-CALCULATED


MODULUS OF RESILIENCE FROM FWD TESTS AND SSG STIFFNESS

Back-calculated modulus of resilience of subgrades and base-courses obtained


from tests using the falling weight deflectometer (FWD) are now used extensively in
pavement evaluation and design. Wu et al. (1998) developed the relationship between
the modulus of resilience back-calculated from the FWD (EFWD) measured in MPa,
and the SSG stiffness (KSSG) measured in MN/m as in Eq. 2.70 with an R-square value
of 0.66 in the following form:

EFWD = 22.96e 0.12 kSSG Eq.


2.70

Wu et al. (1998) also noted that the difference between the results of the FWD
and the SSG could be ascribed to in-situ variability of material properties. The EFWD
value is obtained based on seven deflection measurements, measured over a distance
of about 2 meters. It is therefore a weighted average value for different layers on the
other hand, the SSG measures only the near-surface soil stiffness right underneath its
ring-foot, with the measurement influencing a depth of less than 0.23 meters.
Moreover, on the basis of the test results, it was concluded that the SSG was much
more sensitive when the materials were soft and that the FWD become more sensitive
while testing stiffer materials. In other studies, Chen et al. (1999) suggested that a
general linear relationship between EFWD measured in MPa, and KSSG measured in
MN/m as in Eq. 2.71 could be developed, with an R-square value of 0.82.

EFWD=37.65kSSG - 261.96 Eq. 2.71

Chen et al. (1999) also indicated that the base moduli obtained from the SSG
studies were smaller than those obtained from the FWD. In addition, they reported
that discrepancies between EFWD, and kSSG could have been due to inaccuracies with the

43
measurement of EFWD and the fact that the SSG could not give reliable results for soils
stiffer than 23 MN/m.

2.9 STUDIES ON CORRELATIONS BETWEEN UCS AND DCP

McElvaney and Djatnika (1991), based on laboratory studies, have concluded


that DCPI values can be correlated to the unconfined compressive strength (UCS)
measured in KPa for soil-lime mixtures. The investigations were made on original soil
samples and samples mixed with lime in various percentages. It was observed that for
lower ranges of strain, the DCPI values correlated well with the un-confined
compressive strength, and was independent of the lime content in the soil. This
indicated that the correlation was mainly a function of strength, and not upon the
manner in which the strength was developed. Based on data analysis of original soil
samples and samples mixed with lime considered together, three relationships were
developed. The first relationship developed as shown in Eq. 2.72 implied that there
was a 50 percent probability that the values using the regression equations would
underestimate the real value. The second relationship developed as in Eq. 2.73 had a
confidence of 95%, where the probability of underestimation was lesser than 15
percent. The third relationship given by Eq. 2.74 was developed for a confidence of
99%, where the probability of under estimation was lesser than 50%.

Log UCS = 3.56 - 0.807 Log (DCPI) Eq.


2.72

Log UCS = 3.29 - 0.809 Log (DCPI) Eq.


2.73

Log UCS = 3.21- 0.809 Log (DCPI) Eq.


2.74

Additionally, the high correlation of DCP values to the CBR observations has
been used to characterize stabilized base-courses, and subgrades in many projects
(Little et al., 1995).

A graphical relationship between DCPI, in terms of mm/blow, the CBR, and


the unconfined compressive strength (qu) is shown in Fig. 2.4. The Mn/DOT
developed this relationship from studies on Minnesota Road Research Project
(Mn/ROAD) using the DCP (Burnham and Johnson, 1993).

44
Fig. 2.4 Relationship between DCPI, CBR and UCS
2.10 STUDIES ON CORRELATIONS BETWEEN DCP AND STANDARD
PENETRATION TEST

Sowers and Hedges (1966) developed a correlation between DCPI and


observations of the standard penetration test. The correlation equation took the form
as given below for SPT values lesser than 10mm/blow (Burnham and Johnson, 1993).

Log (DCPI) = -A + B Log (SPT) Eq.


2.75
The study involved the use of a DCP of slightly different design. The design
specification enabled the classification of the DCP as a "light penetrometer".
Later Livneh and Ishai (1988) developed a relationship between DCP and the
SPT, of the following form as reported by Luo et al. (1998) for SPT blow counts less
than 30.
Log (DCPI) = -A+B log (NSPT) (Valid for NSPT < 30) Eq.
2.76
where DCPI is the penetration index (mm/blow); NSPT is the SPT blow count; and
A, and B are regression coefficients.
2.11 STUDIES ON CORRELATIONS BETWEEN PLATE LOAD TEST (PLT)
VS DCP, GEOGAUGE, AND LFWD

Konard and Lachance (2000) developed a relationship connecting the


observations made using the DCP to the modulus of deformation obtained using the
plate load test (EPLT) for unbound aggregates and gravelly and sandy soils. The
correlation obtained is given in Eq. 2.77.

45
Log (EPLT) = −0.88405 Log (DCPI) + 2.90625 Eq.
2.77

where DCPI is the DCP penetration index in millimeters per blow (mm/blow)
using a 51- mm diameter cone, and a 63.5-kg hammer dropping through a height of
760 mm, with EPLT is measured in MPa.

Seyman (2003) developed two correlations relating the average penetration


rate or penetration index determined using the DCP to the initial loading deformation
modulus (EPLT(i)), and the reloading deformation modulus (EPLT(R2)) of plate load tests
with R-square values of 0.619, and 0.766 respectively as shown below:

EPLT(i) = 7000/(6.1+ DCPI1.5) Eq.

2.78 EPLT(R2) =2460*DCPI-1.285 Eq.

2.79

Abu-Farsakh et al. (2005) developed correlations between the DCPI and the
EPLT(i) and EPLT(R2) for the field data, and the results yielded the following non-linear
models with an R-square values of 0.94 and 0.95 respectively.

E PLT (i) = (17421.2 /(DCPI 2.05 + 62.53)) – 5.71 Eq.


2.80

E PLT (R2) = (5142.61/ (DCPI 1.57 – 14.7)) – 3.49


Eq.2.81

They also developed correlations between the DCPI and the EPLT(i) and EPLT(R2)
for the combined field and laboratory data to find the best correlations, and the results
yielded the following non-linear models with an R-square values of 0.67 and 0.78
respectively.

E PLT (i) = (9770 /(DCPI 1.6 + 36.9)) – 0.75 Eq.


2.82

E PLT (R2) = (4374.5/ (DCPI 1.4 – 14.9)) – 2.16 Eq.


2.83

CNA Consulting Engineers, Minneapolis conducted extensive field tests to


study the relationship between the deformation modulus obtained from Quasi-Static
Plate Load Tests (QSPLT), and the stiffness-modulus determined using the GeoGauge.

46
The results of this study are presented in Eq. 2.84 through Eq. 2.86 for initial loading
deformation modulus E (QPLT)i, unloading deformation modulus E (QPLT)u and reloading
deformation modulus E (QPLT)R which have R-square values of 0.66, 0.27, and 0.23
respectively. These results suggest that the values of the reloading deformation
modulus obtained from the QSPLT, and the magnitude of the stiffness-modulus
obtained using the GeoGauge show almost the same trend in variation. In this
investigation, it was also observed that the values of the stiffness-modulus predicted
by the GeoGauge (EG) were nearly seven times higher than that of the initial-loading
deformation modulus. Moreover, it was found that variations in the stiffness-modulus
obtained using the GeoGauge agreed well with the variations in the initial loading
deformation modulus, when compared to the variations in the unloading deformation
modulus, and the reloading deformation modulus (Petersen et al. 2002).

E (QPLT)i = 0.3388(EG) + 84.7 Eq.


2.84

E (QPLT)u = 0.6158(EG) + 10.3 Eq.


2.85

E (QPLT)R = 0.8962 (EG) + 25.9 Eq. 2.86

Kamiura et al. (2000) proposed the following relationship between the


observations obtained for the LFWD - Prima 100, and that of the PLTs, for various
types of subgrade materials including volcanic soils, silty sands, and mechanically
stabilized crushed-stone. In this relationship, kLFWD and k30 represents the ratio of the
stress on the loading plate of the LFWD, and the ratio of stress on the plate of 300mm
diameter used in the PLT respectively, to the corresponding deflections measured.

Log (kLFWD/k30) = 0.0031 Log (kLFWD) +1.12 Eq.

2.87

Livneh and Goldberg (2001) have reported the approach adopted by the
German Code, where the use of the following equation connecting the reloading
deformation modulus (EPLT(R2)) obtained from the plate load test, and the modulus of
resilience (ELFWD) (measured in MPa) obtained using the German Dynamic Plate test
(GDP), is recommended in the design of flexible pavement structures.

47
EPLT(R2) = 600 –(300/(300 − ELFWD)) Eq.
2.88

Seyman (2003) derived two correlations relating the modulus of resilience


obtained using the LFWD (ELFWD), to the initial deformation modulus (EPLT (i)), and the
reloading deformation modulus obtained from the PLT (EPLT (R2)), with R-square values
of 0.844, and 0.897 respectively as shown below:

EPLT(i) = 0.907*(ELFWD) -1.8 Eq.


2.89

EPLT(R2) = 28.25* e0.006(ELFWD) Eq. 2.90

Abu-Farsakh et al. (2004) derived correlations between the modulus of


resilience obtained using the LFWD (ELFWD), the initial deformation modulus obtained
using the PLT (EPLT (i)), and the reloading deformation modulus obtained using the PLT
(EPLT (R2)) for field data as shown below. The R-square values for these expressions are
0.92, and 0.94 respectively.

EPLT (i) = 22+ 0.7 (ELFWD) Eq.


2.91

EPLT (R2) = 20.9 + 0.69 (ELFWD) Eq. 2.92

They also developed similar equations for combined field and laboratory data
with R-square values of 0.87, and 0.87 respectively as shown below:

EPLT (i) = 0.71 (ELFWD) + 18.63 Eq.


2.93

EPLT (R2) = 0.65 (ELFWD) + 13.8 Eq.


2.94

Abu-Farsakh et al. (2004) developed correlations as given below, relating the


EPLT(i), EPLT(R2), and EG for the laboratory data. The correlations followed non-linear
models with R-square values of 0.83 and 0.69 respectively.

EPLT(i) = 15.5*e 0.013(EG) Eq.


2.95

EPLT(R2) = 15.8*e 0.011(EG) Eq. 2.96

48
They also developed correlations between EPLT(i), EPLT(R2) and EG for the field
data as shown below. The relationships followed non-linear models with R-square
values of 0.87 and 0.90 respectively.

E PLT (i) = -75.58 + 1.62 (EG) Eq.


2.97

E PLT (R2) = -65.37+1.50 (EG) Eq. 2.98

In addition to the above, they also developed correlations between the EPLT(i),
EPLT(R2) and the modulus of stiffness measured using the GeoGauge (EG) for the
combined field and laboratory data. The following correlations with R-square values
of 0.72 and 0.59 respectively were obtained.

EPLT (i) = 1.168 (EG) - 37.42 Eq.


2.99

EPLT (R2) =10(1.2 (log (EG)) -1.39) Eq.


2.100

Nazzal et al. (2007) developed linear regression models relating the modulus

of resilience obtained using the LFWD (ELFWD), to the initial deformation modulus

(EPLT (i)), and the reloading deformation modulus obtained from the plate load test

(EPLT (R2)) with R-square values of 0.92, and 0.97 and RMSE values of 13.82, and 13.4

respectively.

E PLT (i) = 1.041 (ELFWD) Eq.


2.101

E PLT (R2) = 0.875 (ELFWD) Eq.


2.102 As part of the above study, two non-linear multiple regression curves
were formulated to predict the values of the initial deformation modulus (EPLT (i)) and
the reloading deformation modulus obtained from the PLT (EPLT(R2)) using the modulus
of resilience obtained using the LFWD (ELFWD), the voids-ratio (e), and the water-
content (wc). The results yielded the models shown in the Eq. 2.103 and Eq. 2.104
with R-square values of 0.985 and 0.986, and RMSE values of 5.89 and 9.10,
respectively:

49
E PLT (i) = 1.03*(ELFWD)1.031(e)-0.524 (wc)- 0.0887 Eq. 2.103
EPLT(R2)= 1.22*(ELFWD)1.064(wc)-0.195(e)- 0.021 Eq.
2.104

In the above study, it may be noted that the multiple regression models relating
observations of the PLT and the LFWD have slightly higher R-square values, and
smaller RMSE values when compared to simple linear regression models relating
PLT, and LFWD observations. This suggests that water-content, and voids-ratio too
can be used as independent variables.

Kim et al. 2007 developed a correlation for Korean soils between the static
deformation modulus of PLT and the modulus of resilience of LFWD with an R-
square value of 0.77 as given below:

ELFWD = 1.41 EPLT – 7.48 Eq.


2.105

The above studies indicate that the LFWD serves as an alternative to static
plate load test. The correlations developed between the two tests indicate the
applicability of LFWD in place of the PLT.

2.12 STUDIES ON CORRELATIONS BETWEEN SOIL STIFFNESS, AND


ELASTIC PROPERTIES

The soil stiffness measured using the Soil Stiffness Gauge or Geo-gauge
(SSG), can be used to provide the elastic/ stiffness-modulus of the subgrade materials
close to the surface. This equipment uses a rigid ring-shaped foot, resting on a
linearly-elastic, homogeneous, and isotropic infinite half-space. The stiffness (kSSG)
can be used to determine the Young’s modulus (or the modulus of elasticity) of soil
represented as ESSG based on the following equation (Egorov 1965):

KSSG=ESSGR/(1- ν2) ω(n) Eq.


2.106

where, ν is the Poisson’s ratio of the materials; R is the outside radius of the ring
(57.15 mm); and ω(n) is a constant which is a function of the ratio of inner diameter,
and the outer diameter of the ring; and ESSG, and KSSG are in MPa and MN/m,
respectively. Table 2.3 summarizes the values of Poisson’s ratio for various materials
(Huang 1993).

50
For the ring geometry of the SSG, the constant ω(n) is equal to 0.565, and
hence the above equation (Eq. 2.107) can be expressed as,

KSSG=1.77ESSGR/ (1- ν2) Eq.


2.107

Table 2.3 Values of Poisson’s ratio for various materials


Material Range Typical
Hot mix asphalt 0.30 – 0.40 0.35
Portland Cement Concrete 0.15 – 0.20 0.15
Untreated granular materials 0.30 – 0.40 0.35
Cement-treated granular materials 0.10 – 0.20 0.15
Cement-treated fine grained-soils 0.15 – 0.35 0.25
Lime-stabilized materials 0.10 – 0.25 0.20
Lime-fly ash mixtures 0.10 – 0.15 0.15
Loose sand or silty sand 0.20 – 0.40 0.30
Dense sand 0.30 – 0.45 0.35
Fine-grained soils 0.30 – 0.50 0.40
Saturated soft clays 0.40 – 0.50 0.45
Source: Huang,1993

But the Young’s modulus (ESSG) and the shear modulus (GSSG) are related
through Eq. 2.108 based on which Eq. 2.109 is obtained. These expressions are given
as follows:

ESSG= 2 GSSG(1+ν) Eq.


2.108

KSSG=3.54GSSGR/ (1- ν) Eq.


2.109

2.13 STUDIES ON CORRELATION BETWEEN MODULUS FROM SEISMIC


TESTS AND SSG STIFFNESS

Wu et al. (1998) developed the correlation between the results of the SSG
tests, and the modulus obtained from seismic tests, including dirt-seismic pavement
analyzers (D-SPA) and spectral analysis of surface waves (SASW) for soil
characteristics varying from soft to medium-stiff subgrades to very stiff bases. It was
indicated that the elastic modulus obtained from the SSG was about three times
smaller than that obtained from seismic tests. A linear relationship with R-Square
value of 0.62 was obtained between the seismic modulus (ESEIS) measured in MPa and
the SSG stiffness (kSSG) measured in MN/m as in Eq. 2.110

51
ESEIS = 47.53kSSG +79.05 Eq. 2.110

Chen et al. (1999) conducted a similar study, and developed a relationship


with an R2 value of 0.81 as in Eq. 2.111 between E SEIS from dirt seismic pavement
analyzers (D-SPA), and spectral analysis of surface waves (SASW) for soft - to
medium-stiff subgrades and for very stiff bases and kSSG. The seismic modulus (ESEIS)
was measured in MPa, and the SSG stiffness (kSSG) was measured in MN/m.

ESEIS = 55.42kSSG -162.94 Eq. 2.111

The operation of the SSG is very simple can be used for quality control.
However, it yields a reliable stiffness value used for a shallow depth. Seismic tests can
generate a stiffness-depth profile, but its operation is more complicated (Chen et al.
1999).
2.14 A STUDY ON CORRELATIONS AMONG SOIL PROPERTIES

In a study by Jian-Hua Yin, (1999) relationships were developed correlating


soil properties such as the among clay content (C), water content (w), liquid limit
(wL), plasticity index (Ip) and plastic limit (wp), compression index (Cc), unloading
reloading index (or swelling index) (Cr), and the coefficient of secondary
consolidation (Cơ) for Hong Kong Marine soils (see Eq. 2.112 to Eq. 2.117).
w = 1.53C + 13.26 R2 = 0.96 Eq. 2.112
wL =1.70 C + 13.5 R2 = 0.995 Eq. 2.113
Ip = 1.20 C R2 = 0.90 Eq. 2.114
Cc = 0.0138 Ip + 0.00732 Eq. 2.115
Cr = 0.00219 Ip - 0.0104 Eq. 2.116
Cơ = 0.000369 Ip – 0.00055 Eq. 2.117

They reported that the friction angle Φ’ of Hong Kong Marine deposits
(HKMD) decreased with an increase in Ip in a trend similar to that for the soil reported
by Lupini et al.(1981). Also the Young’s modulus E50 increased with an increase in
effective confining pressure σ3, but decreased with an increase in clay content (or Ip).
2.15 STUDIES ON RELIABILITY OF OBSERVATIONS OF PFWD
COMPARED TO FWD

Fleming et al. (2000) conducted field tests correlating the observations of the
moduli of sub-grades obtained using the falling weight deflectometer, to the results
obtained using LFWDs such as the Prima 100 Portable falling weight deflectometer

52
(LFWD-Prima100), the German Dynamic Plate (GDP), and the Transport Research
Laboratory (prototype) Foundation Tester (TFT). The studies showed that the
modulus of resilience of the sub-grade (MFWD) measured using the FWD was closely
correlated to the values estimated by Prima 100 PFWD.

Nazzal (2003) also conducted studies on correlating back-calculated modulus


of resilience values (MFWD) obtained using FWDs to the estimated values of the
modulus of resilience of sub-grade obtained from observations using LFWDs (ELFWD).
Eq. 2.118 provides a very high correlation, with an R-square value of 0.94, at a
significance level of lesser than 99.9%, and a standard error of 33.1. The relationship
was developed for observations of elastic moduli estimated using LFWDs, within the
range of 12.5 to 865 MPa.

MFWD = 0.97 (ELFWD) Eq. 2.118

A non-linear multiple regression analysis conducted by Nazzal et al. (2007)


relating the modulus of resilience measured using the FWD (MFWD), to that measured
using the LFWD (ELFWD), produced the following model with an R-square value of
0.98 and an RMSE value of 5.89. In this relationship, e and wc represent the voids
ratio, and the moisture-content of the soil, respectively.

MFWD= 6.398(ELFWD) 0.732(e)-0.086(wc)-0.334 Eq. 2.119

The above mentioned investigations indicate that the FWDs provide better
estimates of the modulus of resilience of sub-grades. But the use of FWDs mounted
on heavy vehicles calls for deployment of huge financial resources, in addition to the
involvement of more skilled man-power. However, investigations cited above prove
that studies on modulus of resilience of sub-grades can also be performed effectively
by using light falling weight deflectometers (LFWDs).

Apart from the Prima 100 LFWD, cheaper and more efficient portable FWDs
such as the Loadman impulse test device (Gros, 1993), Inspector-2, and Zon ZFG
2000 were also developed later. It was observed that the Loadman portable falling
weight deflectometer provided a composite modulus that was lesser than that provided
by the Prima 100. Also, the moduli estimated using Prima 100 LFWD correlated
better to that estimated using the FWD with an R-square of 0.552, while the
composite moduli estimated using the Loadman PFWD revealed a lesser correlation
to FWD estimates at an R-square of 0.245 (Steinert et al. 2005).

53
Gros (1993), and Whaley (1994) revealed that the Loadman is sensitive to
heterogeneous layers. The depth of influence is also smaller when compared to that of
FWDs. Due to these reasons the value of modulus of resilience of the sub-grade,
estimated by Loadman is slightly higher when compared to that observed that using
the FWD. The influence of stone-particles on Loadman, was also found to be higher.
However, Whaley (1994) concludes that the FWD and the Loadman are best suited
for determining the layer-moduli and deflections.

Furthermore, on investigations in India comparing the results of the Loadman


to that of FWD, it was found that the modulus of resilience values estimated by both
the methods had a reasonably high correlation with an R-Square of 0.49. Comparison
of average-deflections estimated using the Loadman (LWSMEAN), and the Benkelman
beam (BBMEAN), resulted in the development of a regression equation as in Eq. 2.120
with an R-square of 0.86, and a standard error of 0.14 at 95% significance level
(Bennet, 1994). The data for tests conducted at the centre and the edge were recorded
separately in this study. However, it may be noted that the FWD and Loadman
produced lesser deflections when compared to that of the Benkelman beam (Davies,
1997).

BBMEAN = -0.4268 + 2.2302 LWSMEAN Eq. 2.120

The above studies indicate that PFWDs can be used effectively, in estimating
the sub-grade-moduli and pavement deflections reliably. Also, there exists a need to
link the results of traditional sub-grade evaluation techniques to that of simple and
portable equipment such as the PFWDs, for the benefit of road-engineers of
developing nations, and lesser developed countries. The present study focuses on the
use of Inspector-2 PFWD manufactured by Englo, of Estonia for the estimation of the
sub-grade modulus of resilience.

2.16 STUDIES ON INFLUENCE OF SOIL PARAMETERS ON THE


MODULUS OF RESILIENCE

The modulus of resilience of granular materials depends upon various factors


including the stress state, water content, dry density, and gradation (Seed et al. 1967;
Thompson, 1969; Hicks and Monismith, 1971). Uzan (1985) developed a correlation
model that could handle all types of unbound paving materials ranging from highly
plastic clays to clean granular bases.

54
Investigations based on triaxial tests conducted by Hicks and Monismith
(1971) indicated a steady decrease in the modulus of resilience (MR) with a decrease in
dry density, and with an increase in the degree of saturation up to the optimum water
content.

2.17 STUDIES ON INFLUENCE OF GRADATION, AND FINES ON


MODULUS OF RESILIENCE

Hicks and Monismith (1971) also revealed that the influence of gradation was
not clearly defined for various types of aggregate materials. In one category of
aggregate materials, the modulus of resilience was found to increase as the percentage
passing 75 micron (No. 200) sieve increased, while for another category of aggregate
materials, the opposite trend was observed. However, it was found that regardless of
the aggregate type, the Poisson ratio was found to decrease with increasing fines-
content.

2.18 STUDIES ON INFLUENCE OF WATER-CONTENT AND DRY-DENSITY


ON MODULUS OF RESILIENCE

Many researchers have studied the effect of water-content on the modulus of


resilience. Hicks and Monismith (1971) reported an apparent reduction in modulus of
resilience with increase in water content. Barksdale and Itani (1989) observed a
significant decrease in modulus of resilience for four categories of subgrade materials
tested upon soaking. Raad et al. (1992) concluded that the effect of water on the
resilient properties seemed to be most significant in well-graded materials with a high
amount of fines.

2.19 STUDIES ON INFLUENCE OF CONFINING STRESSES AND STRESS


HISTORY ON MODULUS OF RESILIENCE

The response of a material to cyclic loading is very much dependent on the


stress level. Hicks and Monismith (1971) reported that the stress level affected the
modulus of resilience significantly. They observed that the modulus of resilience
increased considerably with increasing confining stress, and slightly, with increasing
axial stress. It was found that in general, the modulus of resilience and the resistance
to permanent deformations increased with an increase in the mean stress, and the
confining-stress. In addition, Davich et al. 2004 also confirmed from their
investigations that modulus of resilience and Young’s modulus were increased
considerably with increasing confining stress.

55
2.20 STUDIES ON THE SHEAR STRENGTH OF COMPACTED SOILS

Rosenqvist (1959) observed that the mechanical or physical component of


shear strength is contributed mainly by the granular particles and that for a soil of a
given composition, the magnitude of this component depends on the effective normal
pressure between the particles.

In investigations on unsaturated clays compacted to moisture-contents below


OMC (optimum moisture content), Michaels (1959) observed a reduction in cohesion
with decreasing moisture content. In the case of silty clay loams of bulk densities
above 1.20 g/cu cm, Panwer et al. (1972) observed a decrease in cohesion for an
increase in the soil moisture content, while the angle of internal friction was found to
be decreased.

Ghazali (1981) observed that cohesion is dependent on the clay content, water
content and the dry-density, while the angle of internal friction is dependent on the
clay-content and the moisture-content only. In the case of unsaturated compacted fine
soils, Fredlund and Rahardjo (1993) observed that the shear strength was further
affected by the suction forces developed internally among the soil particles. Brackley
(1973, 1975) describes a model of unsaturated clay fabric/ structure, where the clay
particles grouped together within ‘packets’, are held together by suction, and where
the packets themselves act as individual particles. In investigations on sandy loams,
Korayem et al. (1996) confirmed that cohesion increased with the increase in the
initial bulk density and decreased with increasing soil moisture-content. In addition to
the above, Escario and Saez (1986), Alonso et al. (1990), Fredlund and Rahardjo
(1993), Maatouk et al. (1995), Wheeler and Sivakumar (1995), Leroueil and Barbosa
(2000), and Kong and Tan (2000) have performed studies on the shear strength of
unsaturated compacted soils.

2.21 STUDIES ON EFFECT OF SOIL STRUCTURE/ FABRIC AND


MOISTURE ON THE STRENGTH OF UNSATURATED SOILS

Lambe and Whitman (1979) observe that for a given compaction effort and
dry density, the soil tends to be more flocculated on the drier side of OMC as
compared to the dispersive structure towards the wetter side of OMC, and that
flocculated soils have higher strengths than dispersed soils of the same voids content.
Also, Zein (cited in Toll, 2000) observes that although compacted materials are highly

56
aggregated on the dry side of optimum moisture content, aggregation does not exist
on the wet side.

Previous studies by Michaels (1959), Diamond (1970), and Brackley (1973,


1975), and investigations by Zein (1985), Delage et al. (1996), Kong and Tan (2000),
and Toll (2000) report that soils compacted to the drier side of optimum have greater
strengths due to the presence of an aggregated fabric/ structure in soil samples. Such
aggregation was not found to exist in the materials compacted to the wetter side of
optimum (Cokca et al. 2004).

Seed et al. (1961) states that the inter-particle forces, the external forces at the
time of formation of the soil, and the stress history, are responsible for the ‘structure’
of compacted cohesive soils. Leroueil and Vaughan (1990) observe that in addition to
the effects of the initial porosity and stress history, the ‘structure’ of soils play a major
role in determining the engineering properties. A number of studies have been
performed on the structure of compacted soils and the effects of moisture-content on
the structure. Delage et al. (1996) observe that a study on clay platelets and their
aggregations at the micro structural level is essential in understanding the higher
structural levels. In clays, the water content controls the ease with which soil
particles can be rearranged under compactive effort (Mitchell 1993). Increasing the
moisture content up to the OMC tends to increase the inter-particle repulsions,
permitting a more orderly arrangement of the soil particles for a given compactive
effort.

2.22 SUMMARY AND REASONS FOR UNDERTAKING PRESENT STUDY

From the previous sections it was observed that the urge to determine and
correlate the strength properties of soil subgrade between the traditional test methods
and among traditional and non destructive test methods is not a recent idea. The main
reason for this urge was the need to simplify the procedures and economically predict
the properties of soil subgrades without sacrificing quality and reliability. The above
sections provide details on various types of correlations developed between the DCP,
CBR, PLT, the UCS observations, the modulus of resilience measured using various
devices such as the FWD, and the LFWD, and other soil parameters. Apart from the
development of correlations, the influence of the field density, moisture content,
gradation, and on the modulus of resilience was also discussed in the relevant sections

57
above. The above studies also indicate that the LFWD can be used effectively as an
alternative to static plate load test. The LFWD is also found to give good correlations
with the DCP, and the CBR. The LFWD, thus has the potential to be an effective tool
for QC/QA procedures although limited research has been performed using the same
(Seyman, 2003).

Also, in the above review of literature, it can be observed that a number of


researchers have contributed towards the development of a number of relationships
between various parameters for predicting the strength and stiffness of various soil
conditions. However, investigations relating observations made using the LFWD or
PFWD, and the DCP to the observations made using traditional pavement evaluation
approaches for laterite soils were found to be lacking, and inconclusive.

In view of the above, it was felt that investigations using various destructive
and non-destructive approaches need to be performed on laterite soils of the District
of Dakshina Kannada of Coastal Karnataka in Peninsular India, which is
predominantly characterized by the presence of laterite soil subgrades. Investigations
were focused on the study of the strength and stiffness characteristics of laterite soils,
and blended laterite soils, using the DCP, PFWD, CBR, PLT, triaxial test, and the test
for unconfined compressive strength (UCS). This study will enable road engineers to
effectively utilize locally available laterite soils for pavement construction.

58

Вам также может понравиться