Correlating Strength and Stiffness Data of the PENCEL Pressuremeter and Triaxial Compression Tests in Florida

by

Jacob William Jansen

Bachelor of Science Florida Institute of Technology 2015

A thesis submitted to the College of Engineering at Florida Institute of Technology in partial fulfillment of the requirements for the degree of

Master of Science in Civil Engineering

Melbourne, Florida April, 2017

©Copyright 2017 Jacob W. Jansen All Rights Reserved

The author grants permission to make single copies ______

We the undersigned committee, having examined the attached thesis; “Correlating Strength and Stiffness Data of the PENCEL Pressuremeter and Triaxial Compression Tests in Florida Sands”

By Jacob William Jansen Hereby indicate its unanimous approval.

______Paul J. Cosentino, Ph.D., P.E Professor, Civil Engineering Thesis Advisor

______Rodrigo Mesa Arango, Ph.D. Professor, Civil Engineering

______Matthew Jensen, Ph.D. Professor, Mechanical and Aerospace Engineering

______Ashok Pandit, Ph.D., P.E Professor, Civil Engineering Department Head

Abstract

Correlating Strength and Stiffness Data of the PENCEL Pressuremeter and Triaxial Compression Tests in Florida Sands

By: Jacob William Jansen

Major Advisor: Paul J. Cosentino, Ph.D., P.E

The poorly graded sands found throughout Florida provide geotechnical engineers with a difficult challenge when performing testing samples in laboratory tests.

These challenges have caused lab tests such as the triaxial compression test to be overlooked. Since geotechnical engineers estimate strength and stiffness parameters from basic field tests, they often produce overly conservative designs.

Understanding how the automated in-situ PENCEL Pressuremeter (PPMT) test correlates with the triaxial compression test can reduce the time and costs associated with laboratory triaxial testing. Results from triaxial tests yield a

Young’s Elastic Modulus, , and internal angle. Results from a

PPMT test yield a pressuremeter modulus, lift off pressure, and a limit pressure.

The different types of outputted data do not allow for direct comparisons to be made between the triaxial compression test and the PPMT test. This research seeks to correlate the outputted data.

This research involved twenty PPMT tests performed in poorly graded sands, with loose to medium dense texture. PPMT results were compared with results from twenty-one triaxial compression tests performed using removed from the test iii sites. The triaxial test density ranged from 20% to 65% of the relative density.

An equation from Baguelin (1978) was proven to correlate triaxial shear strength with PPMT limit pressure.

Correlations indicate that triaxial elastic modulus and triaxial shear strength correlate moderately . The triaxial modulus is 93 times greater than the shear strength. The PPMT modulus correlates well with the limit pressure, the PPMT modulus is 8.3 times greater than the limit pressure in loose to medium dense sands

(R2=0.89). The triaxial elastic modulus and PPMT elastic modulus correlation show PPMT moduli being on average 60% greater than triaxial moduli in similar density and confining conditions. The correlations from this study indicate that data from the triaxial compression test and the PPMT test can be correlated.

iv

Table of Contents Abstract ...... iii Table of Figures ...... vii List of Tables ...... ix List of Symbols ...... x 1 Introduction ...... 1 1.1 Background ...... 1 1.2 Objective ...... 2 1.3 Approach...... 2 1.3.1 Literature Review ...... 3 1.3.2 Site Selections ...... 3 1.3.3 Laboratory Testing ...... 3 1.3.4 Field Testing ...... 4 1.3.5 Results ...... 4 1.3.6 Analysis ...... 4 1.3.7 Conclusions and Recommendations ...... 4 2 Literature Review ...... 5 2.1 The Pressuremeter ...... 5 2.1.1 Pressuremeter Development...... 5 2.1.2 Pressuremeter Insertion ...... 7 2.1.3 Pressuremeter data interpretation ...... 9 2.1.4 Variations of strain-controlled PMT Tests ...... 15 2.1.5 Pressuremeter Theories...... 17 2.2 Triaxial Testing ...... 18 2.2.1 Apparatus ...... 18 2.2.2 Test Description ...... 22 2.2.3 Test Data ...... 26 2.3 Methods of determining the at rest earth pressure ...... 30 2.3.1 Jaky Determination, 1944 ...... 30 2.3.2 Laboratory Methods ...... 31 v

2.3.3 In-Situ tests ...... 33 3 Description of Test Sites ...... 38 3.1 Test Site Locations ...... 38 3.1.1 Florida Tech Overflow lot ...... 39 3.1.2 Southgate Field ...... 40 4 Test Methods ...... 42 4.1 In-Situ tests ...... 42 4.1.1 PPMT ...... 42 4.2 Laboratory Tests ...... 49 5 Results and Correlations ...... 51 5.1 Soil Properties Results ...... 51 5.1.1 Grain Size ...... 51 5.1.2 Optimum Moisture ...... 52 5.1.3 Relative Density ...... 53 5.2 Triaxial Results ...... 55 5.3 Pressuremeter Results ...... 57 5.4 Correlations ...... 59 5.4.1 Triaxial correlations between strength and stiffness ...... 59 5.4.2 Pressuremeter moduli and strength correlation ...... 60 5.4.3 Triaxial and pressuremeter stiffness correlation ...... 64 6 Conclusions and Recommendations ...... 68 6.1 Conclusions ...... 68 6.2 Recommendations ...... 69 References ...... 71 Appendix A ...... 74 Appendix B ...... 85

vi

Table of Figures

Figure 2-1: From right to left; Pencel, Ménard, and Texam Pressuremeter types (From RocTest) ...... 6 Figure 2-2: Typical Pressuremeter results curve (from Shaban 2016) ...... 9 Figure 2-3: Determination of lift off pressure for a self-boring and pre-bored pressuremeter (Mair and Wood 1987)...... 10 Figure 2-4: PMT unload-reload cycle (From Shaban 2016) ...... 12 Figure 2-5 Typical Load/Unload PMT test data ...... 16 Figure 2-6 Inflated probe shapes in an unconfined environment vs in a confined environment (from Murat and Lemoigne 1988) ...... 17 Figure 2-7 Durham Geo Load frame ...... 19 Figure 2-8 Durham Geo triaxial cell ...... 20 Figure 2-9 Humboldt data aquisition unit ...... 21 Figure 2-10 Triaxial control panel ...... 22 Figure 2-11 Idealized relation for dilation angle, Ψ, from triaxial results ...... 24 Figure 2-13a: Typical Mohr's Circle for CD triaxial data (From Holtz and Kovacs 1981) ..... 28 Figure 2-13b: Typical Mohr's Circle for CU triaxial data (From Holtz and Kovacs 1981) ..... 28 Figure 2-14: Soft Oedometer Ring (Kolymbas, 1993) ...... 32 Figure 2-15: Correlation between the SPT N values, normalized effective overburden, and the triaxial compression phi value (DeMello, 1971) ...... 33 Figure 2-16: Correlation between CPT data and the effective phi angle in soils (Robertson and Campanella, 1983) ...... 35 Figure 2-17: Chart developed by Mair and Wood (1987) to determine the phi value using strain slope...... 37 Figure 3-1: General location of testing sites on the FIT campus shown by stars...... 39 Figure 3-2: Arial overview of the overflow test site. The transect on which tests were performed is shown by the yellow line...... 40 Figure 3-3: Southgate field test site. The transect tested is shown by the yellow line...... 41 Figure 4-1 Pressuremeter control unit with added digital instrumentation (From Shaban, 2016) ...... 43 Figure 4-2 Screenshot of the APMT user interface and data reduction ...... 44 Figure 4-3 Typical membrane calibration curve for PPMT tests ...... 46 Figure 4-4 Typical volume calibration curve from a PPMT test ...... 47 Figure 4-5 driving guide, with thin wall driving tube (From Shaban 2016)...... 49 Figure 5-1 Grain size distributions for test sites in FIT campus ...... 52 Figure 5-2 Standard Proctor moisture density data from a mixed sample, optimum moisture content was determined to be 12% ...... 53 vii

Figure 5-3 Typical Triaxial stress-strain plot ...... 56 Figure 5-4 Correlation between the triaxial initial moduli and the shear strength of the soil at 5% strain ...... 59 Figure 5-5 Comparison between the calculated limit pressure, measured initial modulus and shear strength ...... 61 Figure 5-6 Correlation between PMT initial modulus and PMT limit pressure using all data points ...... 62 Figure 5-7 Limit pressure vs pressuremeter modulus for loose to compact sands ...... 63 Figure 5-8 Relationship between strength and stiffness data for both triaxial and PPMT tests ...... 64 Figure 5-9 Predicted PMT modulus from measured Triaxial data ...... 65 Figure 5-10 Prediction of triaxial moduli using field measured PMT moduli ...... 66

viii

List of Tables Table 2-1 Pressuremeter insertion recommendation table (Winter 1986) ...... 8 Table 5-1 Summary of moisture density results from mixed samples ...... 53 Table 5-2 Summary of maximum density tests ...... 54 Table 5-3 Summary of minimum density tests ...... 55 Table 5-4 Summary of triaxial test results ...... 56 Table 5-5 Averages of triaxial data, based off of density ...... 57 Table 5-6 Summary of PPMT test results ...... 58 Table 5-7 Averages of PPMT data based off of site ...... 58 Table 5-8 Relationships and correlations between the strength and stiffness for PPMT, triaxial, and combined data ...... 64 Table 5-9 Correlation summary ...... 67

ix

List of Symbols Symbol Description DPMT Pressuremeter Diameter Po Lift off pressure G Elastic shear modulus Vm Mean volume of the cylindrical cavity ∆V Change in volume over the corresponding change in pressure, ∆P 휈 Possion’s ratio Pl Limit Pressure Ei Initial elastic modulus

Er Reload elasticmodulus σ’

σtotal Total stress σ1 Maximum principle stress σ3 Confining stress/minimum principle stress Ψ Angle of dilation 휀 Measured Strain 훼 Failure angle 훷 Angle of internal friction 휏 Shear stress 퐾표 At rest earth pressure coefficient 푆푢푝 Pressuremeter shear strength 푆푢푡 Triaxial shear strength

x

Acknowledgements

I would firstly like to acknowledge the faculty and staff of the Civil Engineering

Department at Florida Tech for their support in my graduate studies. I would like to thank my committee chair, Dr. Paul Cosentino for his guidance and helping my academic growth. Additionally I would like to Thank Dr. Alaa Shaban for his help performing the

Pressuremeter tests and helping guide my research process. Finally, I would like to thank my parents and family for the support they provided through my time here at Florida

Tech.

xi

1 Introduction

1.1 Background In the state of Florida, the most commonly found soil is a poorly graded sand, referred to here as Florida sands. Geotechnical investigations used to determine the soil properties of these sands can range from simple observational tests with no sampling, to more involved field testing and/or sampling. These more involved tests can provide the strength and stiffness characteristics of the soil at the site can either be performed in a geotechnical laboratory or in-situ.

A commonly used laboratory test to determine the soil strength and stiffness is the triaxial compression test. This test uses samples transported from the site to a laboratory, where the sample is prepared, and then tested in compression with confinement to obtain the axial strength data. This process of sampling, drying, remolding, and testing take between a hour and a day per test. This test has been used by engineers since the late 1930’s, and provides a wide range of engineering data, under various soil drainage conditions. The data from a triaxial compression test can be used to determine the

Young’s elastic modulus (E), soil shear stress (τ), and angle of internal friction (φ) as a function of unit weight. These engineering parameters are integral parts of many design equations used in geotechnical applications. However, instead of attempting to mimic the conditions of an in-situ soil in a laboratory engineers have designed a device to directly measure the soil’s strength and stiffness characteristics in-situ.

1

The pressuremeter (PMT), first successfully developed in 1956 (Ménard, 1956), has allowed geotechnical engineers to examine soil strength and stiffness in-situ. This device allows the radial strength data of the soil to be measured, without having to transport soil samples and attempt to replicate site conditions in a laboratory. Cosentino et al. (2006) automated the PENCEL pressuremeter unit, leading to a significant increase in its use. The data from the automated PMT can be used to determine the PMT modulus

(E), rebound modulus (Er), lift off pressure (Po), and limit pressure (Pl). These data parameters can take between 30 minutes and an hour per test. These engineering parameters are used in some geotechnical design equations; however they are not as commonly used. The time and cost savings to perform an automated PPMT could lead to large savings in engineering design and construction if the relationship between PPMT data and other more common tests can be determined.

1.2 Objective The objective of this research is to correlate in-situ soil strength and stiffness parameters obtained from PMT tests in Florida sands with the engineering properties obtained from the triaxial compression test.

1.3 Approach This research will use the PENCEL Pressuremeter (PPMT) in pre-bored holes to determine the strength and stiffness characteristics of the soil. Previously determined relationships found in literature will be used to relate the PMT data with the soil shear strength. Finally triaxial tests of samples will be performed to determine the soils’ laboratory shear strength and stiffness. Correlations between the PPMT and the triaxial engineering

2 parameters of the Florida sands will be determined. The steps required in this process are outlined below.

1.3.1 Literature Review A full review of the pressuremeter history and test procedure was performed; this provided the background on the aspects of the pressuremeter being used, as well as prior methods used for testing of different soil parameters. Additionally, a review of triaxial testing and the behavior of in shear was conducted to gain an understanding of the mechanical behaviors exhibited by the soil during testing. Finally a review of the development of methods for determining soil properties from both the

PPMT and triaxial shear test was presented.

1.3.2 Site Selections The sites selected for field testing were selected based upon their location, uniformity, and soil type. The two main sites that were selected contain a loose to medium dense, poorly graded fine sandy soils (SP).

1.3.3 Laboratory Testing Laboratory testing was used to measure data in a controlled and reproducible environment. The soil properties and strengths were determined using the Consolidated

Drained Triaxial test (ASTM D7181). The soil gradation and USCS (Unified Soil

Classification System) classification were determined from ASTM D6913 (soil gradation) and ASTM D2487 (USCS ). In order to determine relative density a minimum and maximum density test was performed (ASTM D4253, ASTM D4254). The optimum moisture content was determined from the standard Proctor test (ASTM D698).

3

1.3.4 Field Testing The sites selected in task 1.3.2 were used for field testing. A transect was set up at both sites, with test points every 25ft along the transect. PPMT tests were performed at each point to gather data over a very contained and uniform sample. A total of 20 test points were accumulated in the field testing procedures.

1.3.5 Results Results from the PPMT tests and triaxial tests were compiled into tables to compare the data from each test. These tables included the site, modulus of elasticity, moisture content, densities, as well as parameters specific to each type of test.

1.3.6 Analysis Test data from both laboratory and field tests were reduced to standard and usable engineering parameters. The data were analyzed using a Excel and R-studio, and multiple linear regression analyses were performed with the data. Additionally, correlations to previous methods discussed in literature were examined.

1.3.7 Conclusions and Recommendations Using the correlations and their corresponding statistical significance, as well as the theoretical results conclusions were made. Recommendations for further studies were developed based on the process and findings from this study.

4

2 Literature Review This section seeks to discuss the background, development, uses and interpretation of the pressuremeter tests and the triaxial tests. Additionally, uses and implications of previous studies will be examined in this section.

2.1 The Pressuremeter

2.1.1 Pressuremeter Development Geotechnical engineers have used laboratory test methods to determine the stress-strain relationships of soils for decades. However, many problems arose from trying to extract, transport, and test undisturbed samples in the lab. The problems associated with transporting and testing caused geotechnical researchers to develop in-situ devices in order to determine the stress-stain relationships of soils. By testing these relationships at the site, researches figured the least amount of disturbance would be applied to the soil, with the main disturbance being due to the testing process itself.

The Pressuremeter was initially developed by Kögler in 1933; however, was not successfully deployed until Louis Ménard in 1956. This device is defined as a “cylindrical device designed to apply uniform pressure to the walls of a borehole by means of a flexible membrane” (Mair and Wood, 1987). Ménard’s Pressuremeter used a three cell system, with all cells being inflated to the same pressure. The cells are rubber membranes fixed around a metal core, and bound by two metal endplates (Mair and Wood, 1987).

The middle cell was used to take measurements, while the two cells on each side, known

5 as guard cells, were used to reduce end effects. The guard cells make the probe function as an infinite cylinder, allowing for the assumption of plane strain to be used for analysis

(Baguelin et al, 1978).

Many different types of pressuremeter’s have been developed since Ménard’s initial

Pressuremeter. The tri-celled Pressuremeter was adapted into a mono-cell version. The two main type of mono-celled Pressuremeter’s are the TEXAM Pressuremeter and the

PENCEL Pressuremeter. The TEXAM is approximately 2.75 inches in diameter by 18 inches long, while the PENCEL is approximately 1.3 inches in diameter by 9 inches long. Both

Pressuremeters are shown in Figure 2-1.

Figure 2-1: From right to left; Pencel, Ménard, and Texam Pressuremeter types (From RocTest)

6

2.1.2 Pressuremeter Insertion Results from the Pressuremeter test will vary depending on how the probe is inserted into the soil. This insertion process will affect the stress-strain characteristics of the soil at the site. The soil type and relative density should be considered when selecting an insertion method.

2.1.2.1 Insertion of Pre-bored Pressuremeter Insertion of a pre-bored Pressuremeter (PBPM) entails lowering the Pressuremeter into a hole slightly larger than the diameter of the probe (between 1.03DPMT and 1.2DPMT). This method is the most common for probe insertion. PBPM works best for shallow depth testing, due to the higher probability of borehole collapse in deeper . The two main methods of borehole preparation for the PBPM are either, drilled or pushed thin wall sampler. Drilled methods include rotary drilling, continuous flight auger, and hand auguring. These methods are not recommended in granular strata due to the large soil disturbance associated with drilling. A pushed or driven thin wall sampler is best used in granular soils. In cohesive soils, the interaction between the soil and the thin walled sampler may cause large disturbance in the layer. In deep test sites a prepared drillers mud is recommended to support the borehole wall (Winter 1986).

2.1.2.2 Insertion of the Self-Boring Pressuremeter The self-boring Pressuremeter (SBPM) has an internal rotary bit leading the

Pressuremeter during insertion. The shavings are flushed with drilling mud up an internal flushing tube, where they are collected in a settling tank. This method of insertion requires the most effort and materials and is primarily used when inserting the probe through cemented layers, or weathered (Winter 1986).

7

2.1.2.3 Insertion recommendation chart The method of insertion of the pressuremeter should be carefully considered, as the insertion method can alter the overall test results. The following table from (Winter,

1986) can be used as a guide for deciding the insertion method. However, additional factors, such as the cohesiveness, depth, grain size, aggregate size for base course, water table, and permeability are important factors to consider.

Table 2-1 Pressuremeter insertion recommendation table (Winter 1986)

8

2.1.3 Pressuremeter data interpretation The Pressuremeter test method produces a stress-strain graph, similar to that of many materials tests. There are three distinctive sections that make up this graph, these are listed and explained below.

Initial phase: is a re-establishing curve portion at which the membrane becomes into a full contact with the walls of a borehole (from point A to point B),

Elastic phase: is a straight-line portion during which the change in volumetric-strains of the membrane are assumed to be constant (from point B to point C),

Plastic phase: is a nonlinear curve portion at which the stressed soil cavity increases significantly with a little increase in applied pressure (from point C to Point D).

Figure 2-2: Typical Pressuremeter results curve (from Shaban 2016) 9

2.1.3.1 Lift off pressure

To estimate the lift off pressure (Po) is estimated from the curve where the tangent line from the initial phase intersects the tangent line from the elastic phase on a pre-bored

PMT test, while it is the displacement of the graph above the strain axis on a self-boring

PMT test. The data reduction process is typically done by hand, drawing the two tangent lines on a scale graph and visually determining this stress point. A hand analysis method is highly variable, and may be best represented as a range instead of a single point. The lift off pressure can be approximately related to the in-situ horizontal at rest earth pressure for pre-bored tests. It is however not practical to predict the horizontal earth pressures due to the relaxation and disturbance of the surrounding soil during the boring process.

The following figures show the methods of determining the lift off pressure (Mair and

Wood, 1987).

(a): SBPMT Curve (b): PBPMT Figure 2-3: Determination of lift off pressure for a self-boring and pre-Curvebored pressuremeter (Mair and Wood 1987).

10

2.1.3.2 Initial Elastic Modulus The straight portion of the curve between the lift off pressure and the plastic zone (figure

2-2 section BC) is the soils elastic response region of the sample. Soil is considered elastic in this region due to the straight line nature of this curve. Lamè (1852) proposed that the radial expansion of a cylindrical cavity in an infinite elastic medium is shown with the following equation:

∆P G = V ( ) (2-1) m ∆V

where:

G : Elastic shear modulus Vm : Mean volume of the cylindrical cavity ∆V : Change in volume over the corresponding change in pressure, ∆P

The shear modulus is found in the initial elastic portion of the pressuremeter curve, which makes it prone to influence from soil disturbance (Baguelin, 1978). To account for the influence of soil disturbance, an unload reload cycle is preformed to determine the reload modulus; these cycles are shown in Figure 2-4. The entire portion of the unload-reload cycle could be used to determine the shear modulus; however a single section over the anticipated strain region can be used to get a more precise modulus.

11

Figure 2-4: PMT unload-reload cycle (From Shaban 2016)

The initial elastic modulus can be related to the shear modulus by using Possion’s ratio

(휈). This relationship is as follows:

E G = (2-2) 2(1 + 휈)

Therefore, Young’s elastic modulus can be directly measured by substituting Equation (2-

2) in Equation (2-3) as given below:

∆P E = 2(1 + 휈)V ( ) (2-3) m ∆V

Possion’s ratio is often assumed to be 0.3 in non-saturated soil, 0.4 in mostly saturated soil, and 0.5 in fully saturated soil.

12

In many cases, however, it is not practical to plot PMT data using volumetric expansion

(∆V), instead radial strain should be used. Radial strain is used instead, because of the differences in probe volumes can cause inconsistent results between different apparatuses. Instead, Briaud (1986) suggests that the strain be represented in radial strain units (εr). The conversion between volumetric strain and radial strain can be easily made, when using the assumption that the fluid being used in the pressuremeter is homogenous and incompressible. A change in volume, converted to a change in radius

(∆R) can then be inserted into the following equation to determine the radial strain.

2πRf − 2πRo Rf − Ro ∆R εr = = = (2-4) 2πRo Ro Ro where:

Ro : Initial cavity radius

R푓: Final cavity radius ΔR: Change in radius

Using this radial strain, the equation for the area of a circle, and the assumption the probe expands with a uniform circular cross section, Equation 2-4 can be combined with

Equation 2-3 to determine the incremental Young’s modulus of elasticity in terms of the radial strain. This is shown in Equation 2-5 below.

13

2 2 ΔR2 Δ푅1 [(1 + 푅 ) + (1 + 푅 ) ] E = (1 + 휈)(푃 − 푃 ) 표 표 2 1 ΔR 2 Δ푅 2 (2-5) [(1 + 2) − (1 + 1) ] 푅표 푅표 where:

Ro : Initial cavity radius υ : Poisson’s ratio

P1 : Cavity radial stress at the beginning of the pressure increment

P2 : Cavity radial stress at the end of the pressure increment

ΔR1: Increase in probe radius at the beginning of the pressure increment

ΔR2: Increase in probe radius at the end of the pressure increment

This initial modulus is ideally uniform through the initial elastic phase of the stress vs strain plot. However, there is typically variation between adjacent data points, therefore a large segment of the elastic portion of the curve is usually used in order to decrease any noise in the data.

2.1.3.3 Limit Pressure The limit is defined as the pressure reached for the infinite expansion of the cylinder

(Briaud, 1986). This is the point where no additional pressure is needed to be applied to continue to apply strain to the soil. The limit pressure (Pl) typically cannot be reached in the normal testing procedure, due to the large strain that is applied. Baguelin (1978) defined the limit pressure as the point where the cavity is twice the initial size (Vf=2Vo).

There are many different ways to extrapolate the data to reach the limit pressure, many of these methods were developed by Baguelin (1978), however the most reliable is to

14 extrapolate the data by hand (Briaud 1986). The limit pressure is shown in Figure 2-2 at point D.

Although disturbance caused by the borehole can have a large effect, up to 20%, on Po and E, the limit pressure is typically not influenced from borehole disturbance, due to the large applied strains in the determination of the limit pressure (Briaud 1986). Soil at radial distances up to 1.6Ro can be remolded into a disturbed annulus without affecting the limit pressure (Baguelin, 1978). It should however be noted, that the disturbance should be minimized in order to provide accurate data.

2.1.4 Variations of strain-controlled PMT Tests The Pressuremeter is a device that lends itself to testing sites in multiple ways, ranging from a single load/unload test to an in depth cyclical and creep test. These will be discussed with their uses and interpretations. The tri-celled probes use stress controlled procedures, while mono-cell probes use strain controlled procedures. This study will focus on the mono-cell, strain controlled tests.

2.1.4.1 Load/unload test The load-unload strain controlled sequence test is the quickest strain controlled test in pressuremeter testing. Once the borehole is created, and pressuremeter inserted, the volume is increased using equal incremental volumes until the probe’s maximum volume is reached. Each incremental reading is recorded after the pressure in the probe stabilizes.

Once the maximum volume is reached, the volume is reduced using equal volume increments.

15

The load/unload test will provide a lift off pressure (Po), initial elastic modulus (Ei), limit pressure (Pl), as well as a rough value of a reload modulus (Er). A typical data graph of this test is shown in Figure 2-5.

Figure 2-5 Typical Load/Unload PMT test data

2.1.4.2 Unload-reload loop The pressuremeter test with an intermittent unload reload loop is another type of strain controlled PMT test. Briaud (1986) added the single unload reload cycle to improve the

PMT test for pavement testing. The probe is inflated to a pressure “P” then incrementally deflated to “0.5P”. The probe is then gradually re-inflated and the test continues to its limit pressure, or max volume, whichever occurs first. By initiating the unload/reload loop, the entire curve was able to be graphed, as well as provided more reliable data. This test was shown in Figure 2-4.

16

2.1.5 Pressuremeter Theories

2.1.5.1 Plane Strain Assumption All pressuremeter theories base many of the assumption that the pressuremeter expands in the bounds of plane strain conditions. Holtz and Kovacs( 1981), say that plane strain conditions occur when one dimension is significantly longer than the other relative dimensions. By having one dimension significantly longer, it allows this axis to be called infinite, and analysis can be done in only one direction. The plane strain assumption with the pressuremeter relies on an infinitely long cylinder expanding in only the radial direction. End effects can play a large role in changing the assumption of an infinite cylinder. Figure 2-9 below shows a PPMT probe as slightly egg shaped when inflated in the air. However, Murat and Lemoigne (1988)show that the PPMT probe becomes much more cylindrical when it is inflated in a confined environment.

Figure 2-6 Inflated probe shapes in an unconfined environment vs in a confined environment (from Murat and Lemoigne 1988)

17

2.1.5.2 Shape Effects of PMT Probe The shape of the pressuremeter probe plays a large role in determining the quality and reliability of data. The ratio of length to diameter is known as the slenderness ratio or the

L/D ratio. This ratio is important in monocell probes to be able to use the assumption of an infinite cylinder. This dimension is not as important with tri-cell probes, because the guard cells reduce the end effects that are present on monocell devices.

Hartman (1974) shows that the maximum error in soil modulus will occur when the L/D ratio is equal to 1, or the probe is spherical. This error is shown to be 33% larger than the actual soil modulus. Current probes have an L/D ratio much greater than 1, meaning the error in soil modulus is much less than the 33% found in spherical probes. An L/D ratio of greater than 6.5 will yield an error of less than 5% for soil modulus, and in many cases can be neglected (Briaud, 1986). Even though the modulus has little change with the L/D ratio, the limit pressure is more greatly influenced by a change in slenderness. Laier

(1973) found that a decrease in the L/D ratio by a factor of two increased the Pl by a factor of 1.28. This can lead to errors in the Pl of 10% to 15% between different types of probes, under the same test conditions.

2.2 Triaxial Testing Triaxial testing is to be performed to determine the laboratory properties of the soils being examined in this study.

2.2.1 Apparatus The triaxial apparatus consists of four main components, a motor controlled load frame, a triaxial cell, triaxial panel, and a data acquisition unit. A thorough understanding of each is critical for the operator. 18

2.2.1.1 Load frame

Figure 2-7 Durham Geo Load frame

A variable speed motor is attached to the load frame so a variable load rate from 0.0001 inches per minute to 0.2 inch per minute can be used during testing. This range allowed for a large range in the strain rate to be applied to the samples.

19

2.2.1.2 Triaxial Cell

Figure 2-8 Durham Geo triaxial cell

The triaxial cell, from Durham Geo Slope Indicator Inc., can be used to test samples from

1.3” in diameter to 4.2” in diameter; as well as samples up to 9” in height.

2.2.1.3 Data Acquisition Unit The Humboldt Data acquisition unit ® is calibrated with 1000 lbs, 5000 lbs, and 10,000 lbs load cells. This system allows for tests to be performed with greater accuracy, depending on the total load applied to the sample. In addition, the data acquisition unit includes a displacement transducer with an accuracy of 0.0001 inches.

20

Figure 2-9 Humboldt data aquisition unit

2.2.1.4 Triaxial Panel The triaxial panel is used to measure the cell pressure and volume changes in the sample.

Pressure regulators in the panel allow for pressure control to 0.1 psi, and volume increments of 0.1 ml.

21

Figure 2-10 Triaxial control panel

2.2.2 Test Description The triaxial test can be performed under a variety of load, sample, and drainage conditions. These features allow for a variety of types of tests to be conducted, allowing for a broad spectrum of data to be collected. Tests of undisturbed samples can be performed by extruding the sample from a Shelby tube directly into the sample membrane. However, many times samples are remolded into the sample membrane.

Remolding of samples allows for tests to be performed at a wide variety of densities. The three main types of triaxial tests are; Consolidated Drained (CD), Consolidated Undrained

22

(CU), and Unconsolidated Undrained (UU). These tests are discussed in the following sections.

2.2.2.1 Consolidated Drained (CD) During Consolidated Drained (CD) test, or ‘S’ test, the soil is consolidated prior to by opening the drainage valve to the specimen (figure 2-11), and applying a hydrostatic confining pressure to the soil. This hydrostatic pressure is equal in all directions, producing isotropic consolidation (Holts and Kovacs, 1981). Consolidation occurs until volume change in the sample has stopped. The drainage lines remain open during shear testing to allow for volume change to occur. The load which applies the normal stress during shear must be applied slowly and the loading rate is based on the permeability of the sample, in order to prevent pore pressures from building.

Since drainage is permitted during shearing the effective stress (σ’) is equal to the total stress (σtotal). When effective stress equals total stress the Mohr’s circle analysis is simplified. If these two stress’ are not equal, then the total stress circle is shifted by the value of shown in Figure 2-13b.

2.2.2.1.1 Behavior of Sands during CD tests The CD test allows for volume change to occur during shear, making the initial of the sample important (Holtz and Kovacs, 1981). A sample with a high void ratio is

‘loose’ while a sample with a lower void ratio is ‘dense’. Additionally, the sample must be fully saturated to observe volume change during shear. The volume change is measured using a burette attached to the sample. When a loose sand is sheared, the void ratio will decrease from the initial loose void ratio, down to a constant, characteristic, void ratio

23

(Holtz and Kovacs,1981). This characteristic void ratio is known as the critical void ratio, it is the “ultimate void ratio at which continuous deformation occurs with no change in the principle stress difference (σ1-σ3)” (Casagrande, 1936). Dense sands display the opposite affect. The volume of dense sands will decrease slightly during shear, then the volume will increase as the sample dilates until its critical void ratio is reached (Holtz and Kovacs,

1981). The ultimate values of the deviatoric stresses (σ1-σ3) should be the same for both loose and dense sands (Hirschfeld, 1963). Volume change vs. strain graphs can be used to determine the angle of dilation (Ψ). This angle can be used to determine the liquefaction potential of sands, as well as losses of strength in sands (Holtz and Kovacs, 1981). These trends are shown in figure 2-14.

Figure 2-11 Idealized relation for dilation angle, Ψ, from triaxial results

24

2.2.2.2 Consolidated Undrained The Consolidated Undrained (CU) test, also known as the ‘R’ test, requires the same initial set up as the CD test. Once the sample is prepared and saturated, the drainage valve is opened, and a hydrostatic confining pressure is applied to the sample to consolidate it.

Once volume change stops, the drainage valve is closed. When drainage is not allowed, no volume change can occur during shear. In the CU test the pore water pressure is typically measured, allowing the effective stress to be calculated. This test may be either a total or effective stress test (Holtz and Kovacs, 1981). Because both total and effective stresses are determined, this is the most common triaxial test performed. Testing labs also prefer this test because the load can be applied at a more rapid rate, since drainage is not allowed during shearing.

2.2.2.2.1 Behavior of Sands in CU tests By preventing volume change and the tendency of soil in shearing to change volume, other than when initially set to critical conditions, will cause either a positive or negative pore water pressure to develop (Holtz and Kovacs, 1981). In a Mohr’s circle analysis of a

CU test, the pore pressure will shift the total stress circle along the sigma axis. The pore pressure shift is shown in Figure 2-13b below. Both circles have the same diameter, because total deviator stress is equal to effective deviator stress. Even though the deviator stress remains constant, the shift along the sigma access significantly changes the friction angle (Holtz and Kovacs, 1981).

25

2.2.3 Test Data The triaxial shear test will produce multiple types of data for analysis. These are all determined from data reduction methods. The modulus of elasticity (E), Mohr’s circle and failure envelope, friction and failure angle, and the angle of dilation can be determined from the triaxial test.

2.2.3.1 Modulus of Elasticity The Modulus of Elasticity can be determined as either strain based or volume based.

Strain based moduli, known as deformation modulus, is calculated from load and displacement. This is shown in Equation 2-6.

휎 퐸 = (2-6) 휖

where:

E : Deformation elastic modulus σ: Applied deviatoric stress 휀 : Measured Strain

Additionally, in triaxial tests where the change in volume is measured, the shear modulus

(G) can be determined. The shear modulus relates the applied stress to the volumetric strain of the sample. This is shown in Equation 2-7.

26

∆P G = ∆V (2-7) ( ) 푉0

where:

G : Elastic shear modulus V0 : Initial Volume of sample in test ∆V : Change in volume over the corresponding change in pressure, ∆P

2.2.3.2 Mohr’s Circle The data from the triaxial tests can be used to construct Mohr’s circles. The Mohr’s circle analysis can be used to determine the shear and normal forces at any plane in a sample.

In addition, the friction angle, and failure angle can be determined graphically from this analysis. The circle is constructed using two points. The first point is the confining stress applied to the sample, while the second point is the sum of the confining stress and the deviator stress. These two points are the principle stresses, and represent the points of zero shear in the sample. Examples of Mohr Circle’s for CD, and CU tests are shown in the figures below.

27

Figure 2-13a: Typical Mohr's Circle for CD triaxial data (From Holtz and Kovacs 1981)

Figure 2-13b: Typical Mohr's Circle for CU triaxial data (From Holtz and Kovacs 1981)

28

2.2.3.3 Angle of Internal Friction The angle of internal friction can be found from the Mohr’s circle analysis above. It can be determined by drawing a line from the origin to a tangent point on the Mohr’s circle. All granular soils will have a of nearly zero, as well as normally consolidated clays

(Holtz and Kovacs, 1981). However, over consolidated clays will have a cohesion value due the preconsolidation hump. This angle can be measured directly from the plot, or can be calculated through trigonometric relations. It is however, easiest to measure directly from the plot. The angle of internal friction will intersect with the failure angle at the same tangent point on the circle. The failure angle (α) can be calculated as:

훷 (2-8) 훼 = 45° + 2

The failure relationship can be derived from the obliquity relationships, where the inclination of the Mohr failure envelope is at its maximum (Holtz and Kovacs, 1981).

In addition to using the Mohr’s circle approach from triaxial testing, the friction angle can be determined from a . The direct shear test will yield a shear and normal force at failure. These two values can be used to determine the friction angle geometrically.

휏 훷 = 푡푎푛−1( ) (2-9) 휎

where:

Φ : Friction angle τ: Shear stress at failure σ : Applied stress at failure

The relationship is only applicable to granular soil, where the cohesion of the soil can be assumed to be zero. If there is cohesive material, multiple trials must be conducted, and 29 the cohesion can be determined from the intercept with the shear axis. Additionally, the friction angle can be determined from each test using a trigonometric relationship, using the confining stress and the maximum applied stress of the test. This relationship is shown in Equation 2-10:

휎1 − 휎3 훷 = 푠푖푛−1( ) (2-10) 휎1 + 휎3

where:

Φ : Friction angle

σ1: Applied stress at failure σ3 : Applied confining stress at failure

The benefit of this form of the equation is a direct calculation from triaxial data. This allows for more data points be found for each set of tests, one phi value per confining stress. The use of the Mohr’s Circle and the listed geometric relationships can provide the majority of engineering data for a given soil.

2.3 Methods of determining the at rest earth pressure

2.3.1 Jaky Determination, 1944 In 1944 Dr. Josef Jaky derived a theoretical equation for determining the coefficient of earth pressure at rest. His method used an infinatly long prismatic soil section with side slopes at the natural , which Jaky assumed to be equal to the angle of internal friction (phi). Jaky then used a series of differential equations to determine a mathematical relationship between the angle of internal friction and the at rest earth pressures.

30

Jaky’s initial solution to the differential equation’s simplified into what is shown in

Equation 2-11. However for Φ values between 20 and 45 degrees, the second portion of

Equation 2-11 simplifies to 0.9. Since most soils have Φ values withing 20 to 45 degrees,

Jaky simplified the equation to what is shown in Equation 2-12.

2 1 + 푠푖푛훷 퐾 = (1 − 푠푖푛훷) ( 3 ) (2-11) 표 1 + 푠푖푛훷

where:

Φ : Friction angle Ko: At rest earth pressure coefficient

The final equation was then simplified to:

퐾표 = 0.9(1 − 푠푖푛훷) (2-12)

In a follow up paper in 1948, Jaky drops the 0.9 from the equation leaving the equation that is currently used by many engineering texts.

2.3.2 Laboratory Methods

2.3.2.1 Triaxial

The Ko value can be determined multiple ways with the triaxial test data. The simplest and quickest way to determine the at rest pressure coefficient is to use a Mohr’s circle analysis to determine the phi angle, and then the angle can be plugged into one of the equations derived by Jaky.

A more complex approach to find Ko is to determine the horizontal and vertical pressures during shear with no volume change. By not allowing volume change in the sample the stresses are characteristic of the at rest condition. The confining stress must be adjusted

31 during axial loading so no compression or dilation occurs. When the horizontal pressure is adjusted with the vertical pressure, Terzaghi’s assumptions of σh/σv can be directly measured.

2.3.2.2 Soft Oedometer Ring The Soft Oedometer Ring (SOR) is a laboratory test apparatus used to obtain the engineering parameters of soils (Kolymbas, 1993). The SOR consists of a load frame, strain and load gauges, and an extensible metal ring. By using the SOR, the problems with non- homogeneous deformations found within triaxial testing are greatly reduced, due to the shallow depth of the ring (0.75”-1.5”). By keeping the depth of the ring small, load is transferred equally throughout the soil structure, as well as the skin friction between the ring and soil sample is reduced. The SOR is shown in Figure 2-14 below. The SOR test method can work efficiently to determine the Ko value of a soil, because it measures horizontal and axial strain directly. There is no need to use correlations and equations to determine the at rest condition. However, to avoid failing the soil into an active condition the test must occur far below the stress limits (Kolymbas, 1993).

Figure 2-14: Soft Oedometer Ring (Kolymbas, 1993) 32

2.3.3 In-Situ tests

2.3.3.1 Standard Penetration Test The Standard Penetration Test (SPT) is an in-situ test which requires a 140 lbs hammer to be dropped from a height of 30 inches to drive a sampler tube through a soil stratum. The number of blows required to drive the sampler head 12 inches into a soil layer is reported as the ‘N’ value as blows per foot. There is no strong method for determining the Ko value from the SPT. However, DeMello (1971) produced an empirical correlation between the

Φ’ value of an uncemented sand in triaxial compression and the N value from SPT tests.

This Φ’ can then be used in the Jaky (1941) equation to determine a reasonable Ko value.

The chart from DeMello (1971) is shown in Figure 2-15 below.

Figure 2-15: Correlation between the SPT N values, normalized effective overburden, and the triaxial compression phi value (DeMello, 1971)

33

2.3.3.2 Cone Penetration Testing The (CPT) was developed in the 1950’s and is under constant improvement, from a mechanically driven analogue device to the current hydraulically pushed digital version (Schmertmann, 1978). The latest CPT cone is an electric device that is instrumented to record the cone resistance (qc) and the side friction (fsc) as it is hydraulically pushed through the soil. An empirical correlation similar to the SPT correlation was determined by Robertson and Campanella (1983). Their correlation uses the value of qc, as well as the effective stress in order to make a correlation to a Φ’ value from a triaxial compression test. This chart is shown in the figure below. Once the Φ’ value is determined from the Robertson correlation, this value may be used in the Jaky

(1944) equation to determine the Ko value.

34

Figure 2-16: Correlation between CPT data and the effective phi angle in sand soils (Robertson and Campanella, 1983)

2.3.3.3 Pressuremeter Lift off pressure is the existing or in-situ horizontal pressures. The most common pressuremeter method of determining the at rest earth pressure is to use the lift off pressure. This method however is highly variable to the quality and type of the borehole.

Different boring methods produce different amount of borehole disturbance, causing variability in the lift-off pressure. To reduce the amount of disturbance, a self-boring pressuremeter can be used to determine more accurate lift-off pressures.

35

Another method, called the strain slope method, uses the slope of the log-log stress strain diagram to correlate an angle of internal friction. By using the slopes of the elastic portion of the stress strain diagram, and assuming a critical value phi angle (between 30 and 35 degree’s for sands), the chart shown in Figure 2-17 can be used to determine an angle of internal friction. By using the angle determined from the chart, the at-rest earth pressure coefficient can be determined from the Jaky (1944) method. The strain slope method was developed by (Mair and Wood, 1987); however this method is most reliable with a self- boring pressuremeter. This method can be used with pre-bored and pre-driven pressuremeters, however the larger the disturbance of the soil, the less reliable the results will be. This method still assumes a value needed to determine the angle of internal friction, making this method subject to borehole disturbance as well as assumptions based on the soil type.

36

Figure 2-17: Chart developed by Mair and Wood (1987) to determine the phi value using stress strain slope.

37

3 Description of Test Sites

3.1 Test Site Locations Two test sites were used on the campus of the Florida Institute of Technology (FIT) in

Brevard County, FL. The first site was located east of Country Club Rd in the remaining undeveloped lot, and the second site North of East University Blvd and East of SR 507

(Babcock St.) in the current Southgate intramural fields. The site locations are shown in

Figure 3-1 below. These two sites contained a stratum, greater than 5 feet, of a poorly graded sand material, with similar densities throughout the sites, and varied densities.

These two sites were chosen due to their accessibility for testing, the range of densities found at the site, and thin layer of vegetation.

38

Figure 3-1: General location of testing sites on the FIT campus shown by stars.

3.1.1 Florida Tech Overflow lot The first site tested was at the grass overflow lot on the southwest corner of the Olin

Complex on the FIT campus (Figure 3-2). A 250 ft transect was measured to allow for reproducible tests (Figure 3-2). An overflow lot was selected due to previous compaction from vehicular traffic; slightly increasing densities of the top strata at this site. There is very little variation in the soil gradation across the site, which will cause the moisture

39 density, and minimum and maximum densities to be similar across the soil (Holtz and

Kovacs, 1981). The moisture contents of the soil at this location varied from 11% to 21%, this variation was due to water table flow direction, and grass cover holding moisture in some area’s while exposed sand in other areas.

Figure 3-2: Arial overview of the overflow test site. The transect on which tests were performed is shown by the yellow line.

3.1.2 Southgate Field The second site test was the southgate field, located at the corner of Babcock st. and

University Blvd. Similar to the first testing location, this is a grass covered field located on the FIT campus (Figure 3-3). The major difference is that the southgate field does not

40 experience vehicle traffic on its surface, making the density at this site lower than at the overflow lot site. The soil present at this site is a poorly graded sand soil (SP) with a gradation similar across the length of the test site. The moisture content in this location ranged from 10% to 30%. This large variation was due to rainstorms between testing days.

Figure 3-3: Southgate field test site. The transect tested is shown by the yellow line.

41

4 Test Methods The tests methods were separated between in-situ and laboratory tests. The testing procedures for each method are described here.

4.1 In-Situ tests

4.1.1 PPMT

The PPMT test is comprised of three different parts: saturation, calibration, and testing.

The basic procedure involves calibrating the pressuremeter unit, preparing the test site and borehole, and conducting the test. The test involves water being forced through tubing into a probe using a crank handle and piston housed in box called the control unit.

4.1.1.1 Control Unit The PPMT control unit used to perform the tests, has been instrumented with digital sensors that are an upgrade, developed by Cosentino et al (2006) to improve the accuracy of the results. These digital instruments are used to measure and record the pressure and volume during testing, as shown in Figure 4-1. These items, as well as the standard measurement items, are housed in a plastic case.

42

Figure 4-1 Pressuremeter control unit with added digital instrumentation (From Shaban, 2016)

The linear string potentiometer and pressure transducer, which yield volume change and pressure respectively, were added to digitally and accurately record the applied volume and corresponding pressure. These instruments are connected to a data port, where the outputs can be viewed with a computer. The instrumentation process can be found in

Cosentino et al (2006) (PP. 54-57). In addition to the digital instruments, the following components are contained in the pressuremeter control unit.

 A 138 cm3 screw piston

 A 2500 kPa dial pressure gauge

 A volume indicator

 Several valves and tubing

4.1.1.2 Automated Pressuremeter Software The digital instrumentation added to the pressuremeter needed to be read by a software package. The calibrated digital outputs from the pressure transducer and the 43 potentiometer were inputted into a Labview ® based software package called The

Automated Pressuremeter Software (APMT) (Cosentino et al, 2006). By using a Labview ® based software, an easy to use graphical user interface (GUI), shown in Figure 4-2, was developed.

The APMT software records the data from the calibration processes, test and site parameters, and field test data. These inputs are then reduced in the program to produce a final data curve. The raw data is shown in Figure 4-2 as the red line, the reduced data is shown in blue. Data is then outputted in a comma separated value (CSV), for use in other data management programs.

Figure 4-2 Screenshot of the APMT user interface and data reduction

44

4.1.1.3 PPMT Calibration

The PPMT calibration process consists of three main steps: saturation, membrane calibration, and volume calibration. These steps are necessary to ensure the PMT is properly operating and for proper data reduction.

4.1.1.3.1 Saturation The entire PPMT system must contain no air voids, the saturation process is intended to ensure no air voids are present during testing. The cylinder is filled and emptied multiple times with de-aired water. While the cylinder is still connected to the de-aired water source, the tubing and probe are inflated using the de-aired water. The tubing and probe are lightly agitated to pass any air bubbles to the valve at the end of the probe. De-aired water is pushed through this valve to remove the air bubbles. This inflation and agitation process is repeated until no air bubbles can be detected in the system. The probe is then deflated and the PMT control unit is disconnected from the de-aired water source. A full description of the saturation process is given in ASTM 4719-07.

4.1.1.3.2 Membrane Calibration The membrane calibration is used to measure the pressure exerted by the flexible membrane at any given volume. The membrane calibration is performed using the APMT program. With the probe placed at the same elevation as the control unit, and the membrane able to expand unobstructed, the probe is expanded in 5 cm3 increments.

Pressure readings are recorded from the APMT software at every volume increment.

Readings should be taken for the entire range of the predicted test volume, 95 cm3 is typically a sufficient volume for this calibration. A typical membrane resistance curve is shown in Figure 4-3

45

140

120

100

80

60

Pressure(kPa) 40

20

0 0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80 85 90 95 Volume (cm3)

Figure 4-3 Typical membrane calibration curve for PPMT tests

4.1.1.3.3 Volume Calibration The last step in the calibration process is a volume calibration. The volume calibration yields the change in the system volume during testing. As pressure increases during testing the tubing, piston, and other components of the system can expand; this expansion is accounted for with the volume calibration. The probe is fitted into a rigid metal sleeve, and the volume is incrementally increased at a rate of 5 cm3 until full contact is made with the metal sleeve. Once full contact has been made (see Figure 4-4), pressure is incrementally increased at a rate of 500 kPa until 2500 kPa is reached. For data reduction, the point where full contact is made (usually a sharp increase in pressure), is set to the origin and then represents the volume change in the system.

46

3000

2500

2000

1500 Full contact made

Pressure(kPa) 1000 with calibration sleeve 500

0 0 5 10 15 20 25 30 35 40 45 Volume (cm3)

Figure 4-4 Typical volume calibration curve from a PPMT test

4.1.1.4 PPMT Field Test

The PPMT field test procedure follows, in general, the procedures presented in the ASTM

D4719-07 using the equal volume method. Slight variations were made between the

ASTM method, in particular in the borehole creation, and the volume increments used.

1) Place control unit at desired test location, and attach probe and computer to the

control unit.

2) Secure the borehole driving guide (Figure 4-5) to the location using soil nails, or

equivalent method to prevent shifting of the guide during the boring process.

3) Drive thin walled boring tube (outside diameter 1.3”) through the guide sleeve

using a 5 pound sledge hammer, taking care to remove and empty the sampler

periodically throughout the boring process. By emptying the sampler, the chances

of soil plugging the tube reduces.

47

4) Collect a moisture content sample from the base of the bore hole.

5) Slide the PPMT probe into the hole immediately after the boring is complete to

reduce the chances of the borehole caving in.

6) Using the APMT software, select the appropriate calibration data files to associate

with the test. Input the general test data required (date, location, borehole

number, control unit height, and test depth).

7) Select ‘run Automated PMT test’ from APMT software menu. Then increase the

volume in the probe by 5 cm3, selecting ‘take reading’ on APMT screen after each

increase in volume. Repeat this step until the limit equilibrium (Pl) is reached or

the unit hits its maximum volume.

8) Reduce the volume 2 cm3 per increment, selecting ‘take measurement’ after each

increment until a volume of 0.75 x Vmax is reached for the rebound slope.

9) Select ‘Save test’ from APMT test screen to complete the test.

48

Figure 4-5 Borehole driving guide, with thin wall driving tube (From Shaban 2016) 4.2 Laboratory Tests

Laboratory tests were used to classify and determine soil properties, including USCS soil classification methods, grain size analyses , moisture density relationships, minimum and maximum densities, and consolidated drained triaxial tests. Each test procedure will be briefly discussed, along with deviations from the corresponding ASTM methodology.

4.2.1.1 Unified Soil Classification (ASTM D2487) Soil at each site was classified using the Unified Soil Classification System (USCS). The procedure for the USCS classification follows the flow chart in the ASTM D2487 procedure.

49

4.2.1.2 Grain Size Analysis (ASTM D6913) The grain size analysis was performed to compare the soil gradation at each test site.

Samples were taken from each site, dried, and organic material removed for each. The grain size analysis was performed according to ASTM D6913.

4.2.1.3 Standard Proctor (ASTM D698) A standard proctor test was performed on the soil samples received from the field. This test allowed the optimum compaction moisture to be determined for the relative density testing.

4.2.1.4 Relative Density (ASTM D4253, D4254) The minimum and maximum density was determined to compare data on a percentile scale. Using both dry and optimum moisture conditions, the maximum and minimum density was determined.

4.2.1.5 Consolidated Drained Triaxial (ASTM D7181) Mechanical properties of the soil were determined by using a consolidated drained (CD) triaxial test. The test was run in accordance with ASTM D7181, the method of sample compaction is not specified in ASTM D7181; the method used for compaction is given below.

1) Calculate the total mass of soil needed for a set value of dimensions (1.4” x 2.8”)

2) Define three equal lift heights and separate the total mass of soil into three equal

parts.

3) Using gentle tamping and vibration, compact each lift in the mold until the final

height for each lift is met.

4) After seating the top cap onto the sample, verify the final height and diameter.

50

5 Results and Correlations

5.1 Soil Properties Results Laboratory classification, moisture density, and relative density of the soil were performed to properly classify the soil, and allow for the soil to be grouped into appropriate categories for correlations. Results of these tests are given in the following sections.

5.1.1 Grain Size All soils in this study were found to be poorly graded sand (SP) according to the Unified

Soils Classification System (USCS). The grain size analyses are shown in Figure 5-1. The research and testing were performed on SP type soils, results and correlations were made only on data from SP soils. The average coefficient of uniformity (Cu) is 2.5 and the average coefficient of curvature (Cc) is 1.3. The low Cu value (Cu< 6) show a uniformly graded soil, and a Cc value between 1 and 3 shows the soil is not gap graded. These shape parameters show that uniformity is the reason the sand is poorly graded, not gap graded.

51

Grain size distribution FIT campus

# 4 #40 #100 100%

90% Overflow lot North 80% Overflow lot Center 70% Overflow lot South 60%

Percentfiner 50% Southgate Field

40%

30%

20%

10%

0% 10 1 0.1 0.01 Sieve Size (mm)

Figure 5-1 Grain size distributions for test sites in FIT campus

5.1.2 Optimum Moisture Standard proctor tests (ASTM D698) were performed on soil from the overflow lot test site to determine the optimum compaction moisture of the soil. This moisture content would then be used to determine the compaction for the maximum relative density test. The results of the standard proctor tests are shown in Figure 5-2. The maximum wet proctor density is 126 pcf when compacted at 12% moisture content, producing to a 113 pcf dry density. The shape of the moisture density curve shows a steep drop off in density when the soil is compacted dry of optimum, while a smaller loss of density is experienced when the soil is compacted wet of optimum. Therefore the relative density test should be performed at optimum moisture density, however deviations should err to wet of optimum. 52

Table 5-1 Summary of moisture density results from mixed samples

Wet Dry Trial MC Density density % pcf pcf 1 7% 117 109 2 8% 122 112 3 12% 126 113 4 16% 125 108 5 21% 122 101

Standard Proctor Results 128 126

124 122 120 118 116

WetDensity (pcf) 114 112 110 0.0% 5.0% 10.0% 15.0% 20.0% 25.0% Moisture Content

Figure 5-2 Standard Proctor moisture density data from a mixed sample, optimum moisture content was determined to be 12%

5.1.3 Relative Density The minimum and maximum relative densities were determined in order to determine the consistency of the soil tested. The soil from the test site, as well as in the triaxial tests ranged from loose to medium dense (20% to 65% relative density). The relative density is important to consider, because the behavior of sands change as the relative density increases. Additionally, the PPMT soil strengths corresponded with loose to medium dense, as defined by Briaud (1986).

53

5.1.3.1 Maximum Density To determine the range of densities to perform triaxial tests, minimum and maximum relative density tests (ASTM D4253/4253) were performed. The test was performed at both oven dry conditions and near optimum moisture content. A maximum density of 137 pcf was determined with a moisture content of 14%. The water content is very close to the 12% determined from the standard proctor test. Additional compaction effort will result in greater density at a greater moisture content (Holtz and Kovacs, 1981). A summary of the maximum density tests are shown in table 5-2 below.

Table 5-2 Summary of maximum density tests

Sample W.C Wet Density Dry Density % pcf pcf 1 0% -- 102 2 0% -- 103 3 0% -- 104 4 11% 122 110 5 14% 137 120 6 13% 135 119 7 13% 136 120

5.1.3.2 Minimum Density In order to determine the relative density of the samples the minimum relative density was determined. Minimum relative density tests are performed at oven dry conditions, as per ASTM D4254. A summary of the minimum density tests are shown in Table 5-3 below.

54

Table 5-3 Summary of minimum density tests

Mold Mold Total Soil Density Weight Volume Weight Weight lbs ft3 lbs lbs lbs/ft3 9.63 0.033 12.47 2.84 85.2 9.63 0.033 12.48 2.85 85.5

5.2 Triaxial Results Consolidated drained (CD) triaxial tests were performed on the SP samples from the FIT test sites. Confining stresses of 5 psi, 10 psi, and 15 psi were chosen for testing to allow for better quality data to be collected. Tests at low confining stress (less than 3 psi) often have too much noise in the data to allow for accurate data analysis. The raw data was recorded using the Humbolt Data Acquisition software and reduced using Microsoft Excel.

The initial modulus (from Equation 2-6), and the normal stress at 5% strain were determined. The normal stress at 5% strain was chosen as the deviator stress because this value corresponded best to the maximum applied normal stress. While in a few cases larger applied stress’ at different strains were found, the stress at 5% was always within one psi of maximum normal stress. A typical triaxial stress strain plot is shown Figure 5-3.

To develop the Mohr-Coulomb failure envelope the normal stress at 5% strain was set as the deviator stress (σ1- σ3); confining stress was set as the minor principle stress (σ3). The shear stress was determined by using half the deviator stress at 5% strain. The major principle stress (σ1) was determined by adding the deviator stress to the confining stress.

A typical Mohr’s circle and failure envelope is shown in Figure 2-13. The angle of internal friction was determined by using Equation 2-10. The results of the 21 triaxial tests are summarized in Table 5-4 below.

55

40 Deviator Stress at 5% strain 35

30

25 20

15 Stress(psi) 10 5 5% Strain E 0 i 0% 5% 10% 15% 20% Strain

Figure 5-3 Typical Triaxial stress-strain plot

Table 5-4 Summary of triaxial test results

Sample Density Confining Initial Deviator Friction Shear Pressure Moduli Stress Angle Stress (5% ε) (5% ε) pcf psi psi psi deg psi

1 100 5 1306 18.3 40.3 9.2 2 100 10 1760 35.3 39.7 17.7 3 100 15 2674 45.4 37.0 22.7 4 100 5 1257 15.5 37.4 7.8 5 100 10 2455 32.4 38.2 16.2 6 100 15 2625 50.8 39.0 25.4 7 100 5 1040 16.8 38.8 8.4 8 100 10 1594 31.7 37.8 15.9 9 100 15 2607 40.7 35.1 20.4 10 90 5 1029 19.6 41.5 9.8 11 90 10 2448 37.8 40.8 18.9 12 90 15 2812 55.5 40.5 27.8 13 90 5 691 18.2 40.2 9.1 14 90 10 1253 36.0 40.0 18.0 15 90 15 1746 47.5 37.8 23.8 16 90 5 880 13.7 35.3 6.9 17 90 10 1513 24.1 33.1 12.1 18 90 15 1877 33.7 31.9 16.9 19 105 3 870 15.2 45.8 7.6 20 105 5 990 25.3 45.8 12.7 21 105 7 1340 29.5 42.7 14.8

56

Table 5-5 Averages of triaxial data, based off of density

Average Average Confining Average Average Density Initial Normal Pressure Shear Stress Phi Angle Modulus Stress pcf psi psi psi psi degrees 5 1201.0 16.9 8.4 38.8 100 10 1936.3 33.1 16.6 38.6 15 2635.3 45.6 22.8 37.0 5 866.7 17.2 8.6 39.0 90 10 1738.0 32.6 16.3 38.0 15 2145.0 45.6 22.8 36.7

5.3 Pressuremeter Results Twenty pressuremeter tests were performed using the PENCEL Pressuremeter. Ten PPMT test were performed in the overflow lot (OF) and ten were performed in the Southgate field (SG). These ten tests were performed along the transects shown in Figure 3-2 and

Figure 3-3, spaced twenty-five feet apart. The APMT software was used to determine the three soil parameters, initial modulus, limit pressure, and lift off pressure. The results of the twenty PPMT performed are summarized in Table 5-6.

57

Table 5-6 Summary of PPMT test results

Initial Limit Lift off Briaud’s Site Number Depth Modulus Pressure Pressure Classification1 -- -- ft psi psi psi OF 101 2.48 1695 260 7.98 Dense OF 102 2.44 1779 206 7.25 Compact OF 103 2.48 1913 165 5.80 Compact OF 104 2.48 1612 166 5.51 Compact OF 105 2.41 1651 173 8.70 Compact OF 106 2.41 1534 151 4.06 Compact OF 107 2.54 1280 146 5.80 Compact OF 108 2.34 1299 179 5.08 Compact OF 119 2.48 2910 212 7.25 Compact OF 120 2.48 2425 184 6.53 Compact SG 201 2.48 650 78 5.80 Compact SG 202 2.48 790 73 2.90 Compact SG 203 2.51 841 70 3.63 Loose SG 204 2.48 585 57 5.08 Loose SG 205 2.51 719 63 6.96 Loose SG 206 2.48 543 54 2.61 Loose SG 207 2.54 710 59 3.63 Loose SG 208 2.54 721 60 3.63 Loose SG 209 2.54 727 56 2.90 Loose SG 210 2.54 665 56 3.05 Loose 1 Soil texture classifications from Briaud 1986

Table 5-7 Averages of PPMT data based off of site

Average Average Average Site Initial Limit Lift off Modulus Pressure pressure -- psi psi psi OF 1810 184 6.40 SG 695 63 4.02

58

5.4 Correlations

5.4.1 Triaxial correlations between strength and stiffness The correlation process began by examining the relationship between the initial triaxial modulus (Ei) and the triaxial shear (Su) strength of the soil. The results are plotted in

Figure 5-4. The most promising relationship was produced by power regression was produced as shown in Equation 5-1. The triaxial initial modulus is about 100 times greater than the shear strength. This correlation was developed from 21 tests in SP sands ranging from 20% to 65% of relative density.

0.84 퐸푖 = 167푆푢 (5-1) 푅2 = 0.72 3000

2500

2000

1500

1000

500 Triaxial Triaxial Initial Modulus(psi)

0 0 5 10 15 20 25 30 Shear Strength (psi)

Figure 5-4 Correlation between the triaxial initial moduli and the shear strength of the soil at 5% strain

59

5.4.2 Pressuremeter moduli and strength correlation An equation developed by Baguelin (1978) (Eq 5-2), was used to relate the PPMT limit pressure to the shear strength. This relationship was used to calculate shear strengths from the field measured PPMT limit pressures.

0.75 푆푢 = 0.21 × 푃푙 (5-2)

Baguelin’s equation was used, instead of developing a site specific equation due to limitations in testing capabilities. A separate, in-situ, shear testing apparatus, such as a vane shear test probe, would be needed to develop a site specific shear versus limit pressure equation.

The calculated shear strengths (Eq 5-2), with confidence intervals, were then added to the triaxial modulus versus shear strength plot (Figure 5-4). The confidence intervals of 95%,

99%, and 99.9% were then compared for both limit pressure versus shear and initial modulus versus shear. In Figure 5-5 the 99.9% confidence interval (α=0.001), which includes the 85% of measured data, shows the confidence intervals of the two data sets overlapping at a limit pressure of 200 psi. A confidence interval of 99.9% was used because it considers more data in the analysis than a 95% confidence interval, which only uses 60% of measured data. Since a limited number of data points were measured, reducing the total number of points used by 40% may significantly bias the results.

Measured and predicted values begin to differ significantly for dense and very dense sands (> 50% for Dense sand, >85% for Very Dense sand).

60

3000 300 Triaxial Data 2500 250

Confidence

interval 2000 200 Power ( Limit Pressure) 1500 150

1000 100

Limit Limit pressure(psi) Initial Initial modulus(psi)

500 50

0 0 1 10 100 Shear strength (psi)

Figure 5-5 Comparison between the calculated limit pressure, measured initial modulus and shear strength

The relationship between the pressuremeter initial modulus was then compared to the limit pressures, allowing for a correlation to be made between the PMT limit pressure and the PMT initial modulus. This relationship is shown in Figure 5-6, and the regression equation is.

퐸푝푚푡 = 8.6 × 푃푙 + 180 (5-3) 푅2 = 0.75

61

3500

3000

2500

2000

1500

1000

500 PressuremeterModulus(psi) 0 0 50 100 150 200 250 300 Limit Pressure (psi)

Figure 5-6 Correlation between PMT initial modulus and PMT limit pressure using all data points

When the data for only loose to compact sands (limit pressure <200 psi and Initial modulus < 2000 psi) the regression correlation increases, as well as the relationship between these values reduces from 8.6 times greater to 8.3. The data for the loose to compact sands are shown in Figure 5-7 and the regression equation is shown in Equation

5-4.

퐸푖 = 8.3 × 푃푙 + 180 (5-4) 푅2 = 0.89

62

2500

2000

1500

1000

500 PressuremeterModulus(psi)

0 0 50 100 150 200 Limit Pressure (psi)

Figure 5-7 Limit pressure vs pressuremeter modulus for loose to compact sands

The correlation between the PMT limit pressure and the triaxial initial modulus was compared to the limit pressure vs PMT initial modulus correlations from Cosentino et al.

(2007). Cosentino et al. found the PMT initial modulus was 8 to 16 times greater than the

PMT limit pressure, with lower density soils ranging toward 8 times greater, and very dense soils ranging toward 16 times greater. Figure 5-6 shows the triaxial initial modulus is 8.6 times greater than the PMT limit pressure, while Figure 5-7 shows the modulus is only 8.3 times greater. The ratio determined in testing is within the range Cosentino presented for loose soils, verifying the relationship in Equation 5-4. Since the densities from the two field test sites do not overlap, there is a void in the data, most prominent on the limit pressure axis.

The shear strength and the initial modulus for both the PPMT and triaxial tests were plotted together to show that the strength parameters have similar influence on both the triaxial and PPMT initial modulus. The relationship between the strength and stiffness for 63 both PPMT and triaxial data is a power relationship, with stiffness increasing as strength increases as shown in Figure 5-8.

3500 Triaxial Data 3000 PMT Modulus 2500

2000

1500

Initial Initial Modulus(psi) 1000

500

0 1 10 100 Shear strength (psi)

Figure 5-8 Relationship between strength and stiffness data for both triaxial and PPMT tests

Table 5-8 Relationships and correlations between the strength and stiffness for PPMT, triaxial, and combined data

Data Power equation R2 Linear equation R2

1.14 PPMT 퐸푖 = 121 × 푆푢 0.88 퐸푖 = 185 × 푆푢 − 152 0.76

0.84 Triaxial 퐸푖 = 167 × 푆푢 0.72 퐸푖 = 93 × 푆푢 + 232 0.73

0.77 Combined 퐸푖 = 223 × 푆푢 0.74 퐸푖 = 88 × 푆푢 + 440 0.63

5.4.3 Triaxial and pressuremeter stiffness correlation Equations 5-2 and 5-4 were then used to develop a predicted PMT modulus from each measured triaxial shear strength. Equation 5-2 was used to predict the limit pressure

64 from the measured triaxial shear strength. Using these limit pressures the PMT modulus was then predicted using Equation 5-4. These predicted PMT moduli and corresponding triaxial moduli are shown in Figure 5-9. The resulting predicted PMT modulus is on average 68% larger than the measured triaxial modulus, indicating a triaxial modulus could give a more conservative design value. The results of predicting the PMT modulus from the triaxial modulus are shown in Figure 5-9.

퐸푃푀푇 푝푟푒푑푖푐푡푒푑 = 1.68 × 퐸푡푟푖 − 250 (5-5) 푅2 = 0.72

Predicted PMT modulus using Triaxial data

7000 6000 5000 4000 3000 2000

PredictedPMT modulus 1000 0 0 500 1000 1500 2000 2500 3000 Measured Triaxial Modulus (psi)

Figure 5-9 Predicted PMT modulus from measured Triaxial data

Assuming Equation 5-5 can be used to predict moduli from PMT limit pressures, the correlation was worked the opposite direction. The triaxial initial modulus was predicted using the measured field PMT moduli. Using Equation 5-4, the limit pressure was determined from PMT moduli, Baguelin’s Equation (5-2) was then used with the predicted limit pressure to determine the shear strength, and then finally the corresponding triaxial

65 modulus was determined from the shear strength Equation 5-1. Figure 5-10 is a plot of these data, producing a linear relationship between the triaxial and PMT moduli. The regression equation (Eq 5-6) has a very strong correlation and indicates the triaxial moduli on average 50% less than the corresponding PMT moduli.

퐸푡 푝푟푒푑푖푐푡푒푑 = 0.5 × 퐸푝푚푡 + 330 (5-6) 푅2 = 0.91

1400

1200

1000

800

600

400

200 Predicted Triaxial PredictedTriaxial Modulus(psi) 0 0 500 1000 1500 2000 2500 Measured PMT Modulus (psi)

Figure 5-10 Prediction of triaxial moduli using field measured PMT moduli

66

Table 5-9 Correlation summary

Equation Description Equation R2 number Triaxial modulus vs triaxial shear 퐸푖 = 93 × 푆푢 + 230 0.73 5-1

Limit pressure vs PMT modulus in loose to compact 퐸푖 = 8.3 × 푃푙 + 180 0.89 5-3 sands PMT modulus conversion from triaxial data 퐸푃푀푇 푝푟푒푑푖푐푡푒푑 = 1.68 × 퐸푡푟푖 − 250 0.72 5-4

Triaxial modulus conversion from PMT data 퐸푡 푝푟푒푑푖푐푡푒푑 = 0.5 × 퐸푝푚푡 + 330 0.91 5-5

67

6 Conclusions and Recommendations

The research objectives in this study were to correlate the data from PPMT tests to the data produced from laboratory triaxial testing. The parameters used from the PPMT tests were the limit pressure and the initial elastic modulus, because these parameters were least influenced by the insertion process. The parameters used from the triaxial tests were the shear strength, angle of internal friction, and the initial elastic modulus. These parameters were used because of their common use in many geotechnical design calculations. The following conclusions and recommendations were made based on the test results.

6.1 Conclusions

All conclusions made would be applicable only to loose to medium dense, poorly graded sand soils.

1) From the triaxial testing, the shear strength can be predicted with moderate

accuracy (R2 of 0.73). This relationship is needed to relate shear strength from

PPMT data to initial triaxial modulus.

2) Using both previously developed correlations and correlations developed from

testing, the triaxial initial modulus is on average 60% less than the PPMT initial

modulus, when using PPMT data corresponding to limit pressures less than 200

psi. The triaxial initial modulus can be predicted with more certainty using field

68

PPMT data (R2 of 0.91) than the PPMT initial modulus being predicted by using

triaxial data (R2 of 0.72).

3) The shear strength and limit pressure equation developed by Baguelin (1978) falls

within one standard deviation of the triaxial and shear strength correlation

developed in this study (α=0.001) in the range of 200 psi limit pressure or less.

6.2 Recommendations Although strong correlations were developed from the testing, recommendations are made to ensure accurate use of the correlations and recommendations for further studies of this topic.

1) Due to large deviations, greater than one standard deviation, in the correlations

when the soil initial moduli is greater than 2000 psi or the PPMT limit pressure is

greater than 200 psi, as shown in Figure 5-5, it is recommended that these

correlations only be used when soil limit pressure is below 200 psi and the initial

modulus is less 2000 psi. These values correspond with Briaud’s loose to compact

soil, or the USCS loose to medium dense soils.

2) In order to insure accuracy when using these correlations, it is recommended that

the correlations be used with testing in poorly graded sand soil. The consistency

of the soil tested should be classified as loose to medium dense.

69

3) It is recommended for further studies that densities between 65% and 100% of

relative density be used in order to verify that these correlations are accurate

irrespective of the soil density.

70

References

Baguelin, F., Jezequel, J. F., & Shields, D. H. (1978). In Clausthal (Ed.), The pressuremeter and engineering (1e éd.) (1st ed.). Germany: Trans Tech Publications.

BRIAUD, J. C. (1986). Pressuremeter: Application to Pavement Design,

Casagrande, A. (1936). The determination of the pre-consolidation load and its practical significance. First International Conference on and Foundation Engineering, Cambridge. , 3. pp. 60-64.

Chen, Y. C. (1990). Correlating the pressuremeter with the . Unpublished Master, Texas Tech,

Cosentino, P. J. (1987). Pressuremeter moduli for airport pavement design. Unpublished Doctoral, Texas A&M University,

Cosentino, P. J. (2007). Correlations between PENCEL pressuremeter, cone penetrometer and dilatometer parameters. Transportation Research Board Proceedings,

Demello, V. (1971). The standard penetration test- A state of the art report. 4th PanAmerican Conference on Soil Mechanics and Foundation Engineering, 1(1), 1-86.

Erbland, P. J. (1993). Modifying the pavement pressuremeter to enable low strain level resilient moduli determination. Unpublished Masters, Florida Institute of Technology,

Hartman, J. P. (1974). Finite element parametric study of verticle strain influence factors and the pressuremeter test to estimate the settlement of footings in sand. Unpublished Doctoral, University of Florida,

Hirschfeld, R. C. (1963). Stress-deformation and strength characteristics of soils. Unpublished Doctoral, Harvard,

Holtz, R., & Kovacs, W. (1981). An introduction to . New Jersey: Prentice-Hall, Incorporated.

Jaky, J. (1944). A nyugalmi nyomas tenyezoje (the coefficient of earth pressure at rest). Mernok Es Epitesz-Egylet Kozlonye [in Hungarian], 78(22), 355-358.

Knittel, A. (1993). Comparing resilient moduli for granular bases and materials using standard and modified pavement pressuremeters. Unpublished Masters, Florida Institute of Technology,

71

Kolymbas, D., & Bauer, E. (1993). Soft oedometer--A new testing device and its application for the calibration of hypoplastic constitutive laws. Geotechnical Testing Journal, 16(2), 263-270.

LAIER, J. E. (1973). Effects of Pressuremeter Probe length/diameter Ratio and Borehole Disturbance on Pressuremeter Test Results in Dry Sand.,

Lamè, G. (1852). Leçons sur la théorie mathématique de l'élasticité des corps solides . Paris: Bachelier.

Mair, R., & Wood, D. (1987). In CIRIA (Ed.), Pressuremeter testing methods and interpretations (1st ed.). Boston: ButterWorths.

Method for consolidated drained triaxial compression test for soils(2011). ASTM International.

Michalowski, R. (2005). Coefficient of earth pressure at rest. Journal of Geotechnical and Geoenvironmental Engineering, 131(11), 1429-1433.

Murat, J. R., & Lemoigne, Y. (1988). Improved calibration and correction techniques for pressuremeters. Geotechnical Testing Journal, 11(3)

Riggins, M. (1981). The viscoelastic characterisation of marine sediment in large scale simple shear. Unpublished Doctoral, Texas A&M University,

Robertson, P., & Campanella, R. (1983). Interpretation of cone penetration tests: Part 1 and part 2. Canadian Geotechnical Journal, 20, 718-745.

Schmertmann, J. (1978). Guidlines for cone penetration test: Performance and design. Federal Highway Administration,

Schmertmann, J. (1986). Suggested method for performing the flat dilatometer test. Geotechnical Testing Journal, 9(2), 93-101.

Shaban, A. (2016). Evaluation of unbound pavement layers moduli using the miniaturized pressuremeter test. (Doctoral, Florida Institute of Technology). Proquest Dissertation Publishing,

Standard practice for classification of soils for engineering purposes (unified soil classification system)(2011). ASTM International.

Standard test methods for laboratory compaction characteristics of soil using standard effort (12 400 ft-lbf/ft3 (600 kN-m/m3))(2012). ASTM International.

72

Standard test methods for maximum index density and unit weight of soils using a vibratory table(2016). ASTM International.

Standard test methods for minimum index density and unit weight of soils and calculation of relative density(2016). ASTM International.

Standard test methods for particle-size distribution (gradation) of soils using (2009). ASTM International.

Winter, E. (1986). The pressuremeter in geotechnical practice No. 1089)Transportation Research Board.

73

Appendix A Pressuremeter Data

74

Overflow lot site 101 800 700

600

500 400

300 Stress(kPa) 200 100 0 0% 5% 10% 15% 20% 25% Strain

A 1- Overflow lot 101

Overflow lot site 102 900 800 700 600 500 400

Stress(kPa) 300 200 100 0 0% 5% 10% 15% 20% 25% Strain

A 2- Overflow lot 102

75

Overflow lot site 103 1200

1000

800

600

Stress(kPa) 400

200

0 0% 5% 10% 15% 20% 25% Strain

A 3- Overflow lot 103

Overflow lot site 104 1000 900 800

700 600 500 400 Stress(kPa) 300 200 100 0 0% 5% 10% 15% 20% 25% Strain

A 4- Overflow lot 104

76

Overflow lot site 105 900 800 700 600 500 400

Stress(kPa) 300 200 100 0 0% 5% 10% 15% 20% 25% Strain

A 5- Overflow lot 105

Overflow lot site 106 800 700

600

500 400

300 Stress(kPa) 200 100 0 0% 5% 10% 15% 20% 25% Strain

A 6- Overflow lot 106

77

Overflow lot site 107 800 700

600

500 400

300 Stress(kPa) 200 100 0 0% 5% 10% 15% 20% 25% Strain

A 7- Overflow lot 107

Overflow lot site 108 700 600

500 400 300

Stress(kPa) 200 100 0 0% 5% 10% 15% 20% 25% Strain

A 8- Overflow lot 108

78

Overflow lot site 119 1400 1200

1000 800 600

Stress(kPa) 400 200 0 0% 5% 10% 15% 20% 25% Strain

A 9- Overflow lot 119

Overflow lot site 120 1200

1000

800

600

Stress(kPa) 400

200

0 0% 5% 10% 15% 20% 25% Strain

A 10- Overflow lot 120

79

Southgate site 201 450 400 350 300 250 200

Stress(kPa) 150 100 50 0 0% 5% 10% 15% 20% 25% Strain

A 11- Southgate 201

Southgate site 202 500 450 400

350 300 250 200 Stress(kPa) 150 100 50 0 0% 5% 10% 15% 20% 25% Strain

A 12- Southgate 202

80

Southgate site 203 450 400 350 300 250 200

Stress(kPa) 150 100 50 0 0% 5% 10% 15% 20% 25% Strain

A 13- Southgate 203

Southgate site 204 400 350

300

250 200

150 Stress(kPa) 100 50 0 0% 5% 10% 15% 20% 25% Strain

A 14- Southgate 204

81

Southgate site 205 450 400 350 300 250 200

Stress(kPa) 150 100 50 0 0% 5% 10% 15% 20% 25% Strain

A 15- Southgate 205

Southgate site 206 350 300

250 200 150

Stress(kPa) 100 50 0 0% 5% 10% 15% 20% 25% Strain

A 16- Southgate 206

82

Southgate site 207 400 350

300

250 200

150 Stress(kPa) 100 50 0 0% 5% 10% 15% 20% 25% Strain

A 17- Southgate 207

Southgate site 208 400 350

300

250 200

150 Stress(kPa) 100 50 0 0% 5% 10% 15% 20% 25% Strain

A 18- Southgate 208

83

Southgate site 209 400 350

300

250 200

150 Stress(kPa) 100 50 0 0% 5% 10% 15% 20% 25% Strain

A 19- Southgate 209

Southgate site 210 400 350

300

250 200

150 Stress(kPa) 100 50 0 0% 5% 10% 15% 20% 25% Strain

A 20- Southgate 210

84

Appendix B Triaxial Data

85

Overflow lot center (90 pcf/5 psi) 18 16 14

12 10 8

Stress(psi) 6 4 2 0 0% 2% 4% 6% 8% 10% 12% 14% Strain (in/in)

B 1- Overflow lot center (90 pcf 5psi)

Overflow lot center (90 pcf/10 psi) 30

25

20

15

Stress(psi) 10

5

0 0% 2% 4% 6% 8% 10% 12% 14% 16% Strain

B 2- Overflow lot center (90pcf/10psi)

86

Overflow lot center (90 pcf/15 psi) 45 40 35

30 25 20

Stress(psi) 15 10 5 0 0% 2% 4% 6% 8% 10% 12% 14% 16% 18% Strain

B 3- Overflow lot center (90pcf/15psi)

Overflow lot north (90pcf/5psi) 18 16 14

12 10 8

Stress(psi) 6 4 2 0 0% 2% 4% 6% 8% 10% 12% 14% Strain

B 4- Overflow lot north (90pcf/5psi)

87

Overflow lot north (90pcf/10psi) 35 30

25 20

15 Stress(psi) 10 5 0 0% 2% 4% 6% 8% 10% 12% 14% Strain

B 5- Overflow lot north (90pcf/10psi)

Overflow lot north (90pcf/15psi) 45 40 35

30 25 20

Stress(psi) 15 10 5 0 0% 2% 4% 6% 8% 10% 12% 14% 16% Strain

B 6- Overflow lot north (90pcf/15psi)

88

Overflow lot south (90pcf/5psi) 20 18 16

14 12 10 8 Stress(psi) 6 4 2 0 0% 2% 4% 6% 8% 10% 12% 14% 16% 18% Strain

B 7- Overflow lot south (90pcf/5psi)

Overflow lot south (90pcf/10psi) 35 30

25 20

15 Stress(psi) 10 5 0 0% 2% 4% 6% 8% 10% 12% 14% 16% 18% Strain

B 8- Overflow lot south (90pcf/10psi)

89

Overflow lot south (90pcf/15psi) 70 60

50 40

30 Stress(psi) 20 10 0 0% 2% 4% 6% 8% 10% 12% 14% 16% Strain

B 9- Overflow lot south (90pcf/15psi)

Overflow lot center (100pcf/5psi) 20 18 16

14 12 10 8 Stress(psi) 6 4 2 0 0% 2% 4% 6% 8% 10% 12% 14% 16% Strain

B 10- Overflow lot center (100pcf/5psi)

90

Overflow lot center (100pcf/10psi) 40 35

30

25 20

15 Stress(psi) 10 5 0 0% 2% 4% 6% 8% 10% 12% 14% 16% 18% Strain

B 11- Overflow lot center (100pcf/10psi)

Overflow lot center (100pcf/15psi) 50 45 40

35 30 25 20 Stress(psi) 15 10 5 0 0% 5% 10% 15% 20% 25% Strain

B 12- Overflow lot center (100pcf/15psi)

91

Overflow lot north (100pcf/5psi) 25

20

15

10 Stress(psi)

5

0 0% 2% 4% 6% 8% 10% 12% 14% 16% 18% Strain

B 13- Overflow lot north (100pcf/5psi)

Overflow lot north (100pcf/10psi) 40 35

30

25 20

15 Stress(psi) 10 5 0 0% 5% 10% 15% 20% Strain

B 14- Overflow lot north (100pcf/10psi)

92

Overflow lot north (100pcf/15psi) 50 45 40

35 30 25 20 Stress(psi) 15 10 5 0 0% 2% 4% 6% 8% 10% 12% 14% 16% 18% Strain

B 15- Overflow lot north (100pcf/15psi)

Overflow lot south (100pcf/5psi) 20 18 16

14 12 10 8 Stress(psi) 6 4 2 0 0% 2% 4% 6% 8% 10% 12% 14% 16% 18% Strain

B 16- Overflow lot south (100pcf/5psi)

93

Overflow lot south (100pcf/10psi) 35 30

25 20

15 Stress(psi) 10 5 0 0% 5% 10% 15% 20% Strain

B 17- Overflow lot south (100pcf/10psi)

Overflow lot south (100pcf/15psi) 60

50

40

30

Stress(psi) 20

10

0 0% 5% 10% 15% 20% Strain

B 18- Overflow lot south (100pcf/15psi)

94

Mixed sample (105pcf/3psi) 18 16 14

12 10 8

Stress(psi) 6 4 2 0 0% 2% 4% 6% 8% 10% 12% 14% Strain

B 19- Mixed sample (105pcf/3psi)

Mixed sample (105pcf/3psi) 18 16 14

12 10 8

Stress(psi) 6 4 2 0 0% 2% 4% 6% 8% 10% 12% 14% Strain

B 20- Mixed sample (105pcf/3psi)

95

Mixed sample (105pcf/7psi) 35 30

25 20

15 Stress(psi) 10 5 0 0% 2% 4% 6% 8% 10% 12% 14% 16% Strain

B 21- Mixed sample (105pcf/7psi)

96