You are on page 1of 15

Mechanically Stabilized Earth Pressures against

Unyielding Surfaces Using Inextensible Reinforcement


Wyatt W. Kelch 1; Michael B. Rodgers, Ph.D., P.E. 2; Taylor A. Rawlinson, Ph.D. 3;
Michael C. McVay, Ph.D. 4; Rodrigo A. Herrera, P.E. 5;
and David J. Horhota, Ph.D., P.E. 6
Downloaded from ascelibrary.org by University of Central Florida on 02/07/24. Copyright ASCE. For personal use only; all rights reserved.

Abstract: This article presents experimental data on the earth pressures that develop between mechanically stabilized earth (MSE) wall
panels that are tied together in an unyielding condition. Prior to design and construction, an extensive literature review was conducted to
ensure the MSE wall adhered to the local standard specifications for road and bridge construction, current design codes, and the construction
sequencing specified by the wall panel manufacturer. The Strong Wall at the University of Florida’s Structures and Materials Laboratory was
utilized to construct a model full-scale MSE wall using inextensible reinforcement, with 140 locations continuously monitored. The MSE
wall was divided into two different relative compaction zones, 95% and 103% of the reinforced soil’s optimum dry density, to investigate
a range of friction angles that can develop during conventional construction. To achieve the necessary aspect ratio of 0.3 specified by practi-
tioners for use in real-world applications, a reaction frame was incrementally loaded on top of the reinforced backfill to simulate additional
wall height and overburden stress. Earth pressure coefficients were found for each compacted soil lift and incremental load from the reaction
frame for both states of soil density. A force equilibrium analysis was conducted to ensure all forces within the experimental setup were
accounted for to validate the results. From the derived earth pressure coefficients, it was observed that earth pressures moved from a passive
condition to an active or at-rest condition as the soil-height increased above each reinforcement level. It was concluded that increased lateral
stress develops from the compaction effort in an unyielding condition, which is not accounted for using conventional MSE design methods.
Consequently, an equation was developed that incorporates a variable friction angle (φ) based on the compaction effort for an unyielding
condition that closely followed the trends of the measured results. DOI: 10.1061/JGGEFK.GTENG-11799. © 2024 American Society of
Civil Engineers.

Introduction precast concrete, are mechanically attached to the embedded rein-


forcements. The frictional resistance that develops between the soil
Mechanically stabilized earth (MSE) walls are a type of reinforced and reinforcement surface opposes the lateral earth pressure exerted
earth retaining structure commonly used for bridge abutments, on the wall facings, holding the panels in place, and retaining the
highway separations, and when the construction area is limited. earthen backfill.
MSE structures use metallic or geosynthetic reinforcing strips or MSE wall reinforcements are either extensible or inextensible.
grids compacted between layers of soil. Wall facings, typically Extensible reinforcements such as geosynthetics allow sufficient
internal deformation to occur, which engages full active pressure
1
Graduate Research Assistant, Engineering School of Sustainable Infra- within the reinforced mass with a linear active zone defined by
structure and Environment, Univ. of Florida, 365 Weil Hall, Gainesville, the Rankine failure plane. MSE structures supported by extensible
FL 32611 (corresponding author). Email: wkelch@ufl.edu
2
Research Assistant Scientist, Engineering School of Sustainable Infra-
reinforcements do not behave as rigid bodies and tension does not
structure and Environment, Univ. of Florida, 365 Weil Hall, Gainesville, develop over the full reinforcement length. Conversely, inexten-
FL 32611. ORCID: https://orcid.org/0000-0001-9975-5849. Email: michael sible reinforcements such as metallic strips prevent internal defor-
.rodgers@essie.ufl.edu mation and tension develops over the full reinforcement length.
3
Research Assistant Scientist, Engineering School of Sustainable Infra- The reinforced mass supported by inextensible reinforcement tends
structure and Environment, Univ. of Florida, 365 Weil Hall, Gainesville, to behave as a rigid body with a bilinear failure plane (Anderson
FL 32611. ORCID: https://orcid.org/0000-0002-9522-2911. Email: taylor et al. 2010).
.rawlinson@essie.ufl.edu
4 In most cases of MSE wall construction, the far end of the soil-
Professor, Engineering School of Sustainable Infrastructure and Envir-
onment, Univ. of Florida, 365 Weil Hall, Gainesville, FL 32611. ORCID:
embedded reinforcement is not attached to a structure. This allows
https://orcid.org/0000-0002-5731-5711. Email: mcm@ce.ufl.edu minor lateral deformation of the wall panels to occur, as the lateral
5
Senior Geotechnical Engineer, Florida Dept. of Transportation, Central earth pressure imposed on the wall panels by the backfill pushes
Office, 605 Suwannee St., Tallahassee, FL 32699. Email: rodrigo.herrera@ them slightly outward. The lateral earth pressure that develops
dot.state.fl.us within the reinforced zone in this case is approximately equal to
6
State Geotechnical Materials Engineer, Florida Dept. of Transportation, the active lateral pressure developed in conventional earth pres-
State Materials Office, 5007 Northeast 39th Ave., Gainesville, FL 32609. sure theory. The tension that develops in the reinforcements from
Email: david.horhota@dot.state.fl.us
the induced active lateral pressure dictates the internal stability
Note. This manuscript was submitted on March 24, 2023; approved on
October 17, 2023; published online on January 25, 2024. Discussion period and controls the reinforcement density required by the AASHTO
open until June 25, 2024; separate discussions must be submitted for in- (2007) design code.
dividual papers. This paper is part of the Journal of Geotechnical and In practice, cases such as road widening may require an MSE
Geoenvironmental Engineering, © ASCE, ISSN 1090-0241. structure to be constructed in front of an existing MSE wall or other

© ASCE 04024009-1 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2024, 150(4): 04024009


solid facing. The ratio of the distance between the existing wall face design code (AASHTO 2007), and the construction sequencing
and new wall face (base, B) to the new wall height (H) is known as specified by the wall panel manufacturer (RECo 2011). As depicted
the aspect ratio. A range of aspect ratios was examined in Kniss in Fig. 1 and detailed in Table 1, the model MSE wall was con-
et al. (2007), which concluded that traditional MSE walls should structed using the Strong Wall in the UF Structures Laboratory. The
not be constructed if the aspect ratio is less than 0.3 due to suscep- right side of the Strong Wall is removed in Figs. 1(a and b) for
tibility of sliding, overturning, and plunging failure. At aspect ratios viewing purposes. Fig. 1(b) also indicates the undercompacted side
less than 0.3, MSE wall reinforcements are often directly tied to an and overcompacted side of the construction area, and a portion of
existing MSE wall or other structure using an inextensible rein- the reinforced steel plate on the overcompacted side is removed to
forcement type. This direct connection creates an “unyielding con- show the MatJack airbags and soil-bearing plates. A construction
dition” that prohibits stress relief through minor deformation of time-lapse of the model MSE wall is provided in Video S1.
the retaining structure that typically occurs in conventional MSE
wall design. Unfortunately, little information is available regard-
ing the earth pressures that are generated for this design scenario. Design Methodology
Downloaded from ascelibrary.org by University of Central Florida on 02/07/24. Copyright ASCE. For personal use only; all rights reserved.

Current design recommendations for MSE walls in Federal High- From the findings of Kniss et al. (2007) and discussions among UF
way Administration Geotechnical Engineering Circular, GEC, researchers, the FDOT, and MSE wall panel manufacturers, it was
number 11 (FHWA et al. 2009a, b) indicate that “much higher” determined that, in practice, MSE panel reinforcements are only
reinforcement tension develops when wall panels are tied together, tied to an unyielding structure when the aspect ratio is 0.3 or less.
but the actual soil pressure and reinforcement tensions developed Because the proposed MSE wall and base soil layer needed to fit
in the unyielding case are not well understood/defined. This arti- within the UF Strong Wall area (Fig. 1), and the base of the MSE
cle presents experimental data collected from an instrumented, wall needed to be approximately 3 m to use standard compaction
unyielding MSE wall constructed using the Strong Wall in the equipment, it was determined that a reaction frame would be neces-
Structures and Materials Laboratory at the University of Florida sary to simulate additional wall height and overburden stress on top
(UF), which is described in the following section. The MSE wall
of the constructed MSE wall to achieve the necessary aspect ratio.
was constructed with inextensible reinforcement and was divided
The dimensions of the nominal 1.52 × 1.52 m square panels
into overcompacted and undercompacted zones, with earth pres-
used for construction were 1.46-m wide by 1.50-m tall. This pro-
sures measured using wall-mounted and soil-embedded earth pres-
vided 75 cm of vertical spacing between four rows of reinforcement
sure cells (EPCs), and reinforcement tension measured using full
for a total constructed wall height of 3 m. To satisfy the unyielding
bridge strain gauges attached to the steel reinforcement strips.
condition, the constructed area was designed to use a second set of
Based on the experimental data, a better understanding of the earth
identical wall panels placed up against the back of the UF Strong
pressures developed during this unyielding scenario was identified,
Wall that would tie directly to the front-facing panels. This ensured
and a new equation that incorporates a variable friction angle (φ)
that the anchor points both lined up from front to back and that the
based on the compaction effort is suggested.
reinforcement connections, which often control internal stability,
were representative of field standards. Since the as-built distance
Model MSE Wall between the front and rear MSE wall panels was 2.97 m, the re-
action frame on the top of the MSE structure was designed to sim-
The full-scale model MSE wall constructed during this study was ulate an additional 6.9 m of overburden pressure for a total wall
designed to comply with current design standards and to closely height of 9.9 m and an aspect ratio of 0.3. Flat galvanized steel
match real-world field conditions. This included following the strips measuring 50-mm wide and 4-mm thick were used as the
Florida Department of Transportation’s (FDOT) standard specifica- inextensible reinforcement, as specified by the wall panel manufac-
tions for road and bridge construction (FDOT 2017), AASHTO’s turers for this design scenario.

Fig. 1. (a) Model MSE wall without reinforced fill; and (b) model MSE wall with reaction frame.

© ASCE 04024009-2 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2024, 150(4): 04024009


Table 1. Model MSE wall experimental components and functions
ID Experimental component Experimental function
A UF strong wall Supports model MSE wall layout
B High strength anchor points Counteracts surcharge loads induced by the reaction frame
C Base soil retention blocks Retains base layer soil and routes free draining water
D Base layer of soil Supports constructed area and provides a gravel drainage layer
E Concrete leveling pads Supports MSE wall panels
F Front MSE wall panels Simulates new MSE wall in an unyielding condition
G Rear MSE wall panels Simulates existing MSE wall in an unyielding condition
H Separation boundary Minimizes influence from adjacent compaction effort
I Steel reinforcement strips Connects front and rear wall panels for an unyielding condition
J Soil-embedded EPCs Measures vertical earth pressure
K Wall-mounted EPCs Measures lateral earth pressure
Downloaded from ascelibrary.org by University of Central Florida on 02/07/24. Copyright ASCE. For personal use only; all rights reserved.

L Threaded rods Connects the reaction frame to reinforced anchor points


M Reinforced backfill Simulates two controlled compaction efforts for investigation
N Soil-bearing plates Distributes uniform surcharge loads to reinforced soil
O MatJack airbags Induces controlled surcharge loads to reinforced soil
P Reinforced reaction plates Provides uniform reaction surface to counteract MatJack loads
Q Reaction girders Provides reaction frame resistance to counteract MatJack loads
R Potentiometer locations Measures wall displacement

Conventional MSE design theories for panels constructed in a that decreases to ka (active geostatic stress) at a depth of 6 m below
yielding condition were reviewed for both internal and external sta- the backfill surface, as shown in Fig. 2(a).
bility calculations and later compared to the measured results found Next, the simplified coherent gravity method, or simplified
from the Model MSE wall constructed in an unyielding condition method (Allen et al. 2001), combines favorable elements of several
(Chalermyanont and Benson 2004, 2005; Lawson and Yee 2005; design methods allowed by the AASHTO standard specifica-
FHWA et al. 2006a, b, 2009a, b; WSDOT 2010; Anderson et al. tions. Instead of using ka and k0 directly as in the coherent gravity
2012; Kim and Salgado 2012). First, the coherent gravity method, method, a kr =ka curve was developed to account for differences in
which was developed to estimate steel strip reinforcement stresses reinforcement type, as shown in Fig. 2(b). The simplified method
for precast panel faced MSE walls (Anderson et al. 2010), was ex- suggests a lateral pressure varying from a multiple of ka at the top
amined. The method treats the reinforced soil as a rigid body with of the wall, which is a function of the reinforcement type, to a near-
lateral forces from the retained soil acting upon the rigid body sim- active earth pressure state at a depth of 6 m below the backfill
ilar to those found in conventional gravity-type retaining systems. surface.
The soil friction angle is used to calculate a lateral earth pressure Differing from the coherent gravity and simplified methods,
coefficient that is applied to the vertical stress to derive lateral the Spangler and Handy method (Spangler and Handy 1984) con-
stress. It is assumed that the potential stress carried by a reinforce- siders the silo effect, or soil arching (Handy 1985; Paik and Salgado
ment is equal to the lateral soil stress acting on a defined tributary 2003), for narrow conventional MSE walls using reinforcements
wall area, which will be discussed later. The top of the wall is as- tied in a yielding condition in front of a stabilized face (Kniss
sumed to have a lateral earth pressure of k0 (at-rest geostatic stress) et al. 2007). The method proposes that friction between the wall

Fig. 2. (a) Coherent gravity (reprinted from Anderson et al. 2012, © ASCE); and (b) simplified methods (reprinted with permission from AASHTO
LRFD Bridge Design Specifications, 2007, published by the American Association of State Highway and Transportation Officials, Washington, DC).

© ASCE 04024009-3 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2024, 150(4): 04024009


Downloaded from ascelibrary.org by University of Central Florida on 02/07/24. Copyright ASCE. For personal use only; all rights reserved.

Fig. 3. (a) MSE wall front profile and tributary wall areas; and (b) MSE wall in plan view.

face and soil provides a vertical shear load that resists soil settle- controlled the design (RECo 2003). Although factored loads for
ment, reducing vertical overburden pressure and, consequently, certain earth pressure coefficients exceeded this controlling factored
lateral earth pressure. This method suggests that the reduction resistance, strain gauges mounted on reinforcement strips allowed for
in lateral pressure results in an overly conservative reinforcement the safe monitoring of reinforcement tensions as the MSE structure
design while the vertical shear load is redistributed as increased was constructed and the reaction frame was incrementally loaded.
bearing pressure directly below MSE wall panels.
The tributary wall area used for internal stability design with the
Instrumentation and Data Acquisition
coherent gravity and simplified methods (AASHTO 11.10.6.2.1) is
defined as a half-panel tall by two-panels wide, as indicated by the Vibrating wire EPCs were used to measure both vertical and hori-
diagonal hatch marks in Fig. 3(a). The figure presents the front pro- zontal stresses in the reinforced backfill. Sixteen Geokon 4810
file of the model MSE wall with the front wall panels and base soil (Lebanon, New Hampshire) vibrating wire “fatback” EPCs were
retention blocks removed to display the reinforcement and wall- installed on the back face of the front MSE wall panels to measure
mounted EPC locations within each tributary area. Additionally, horizontal stress at each row of reinforcement, as depicted in Fig. 3(a).
the naming convention for the four rows of reinforcement is shown These EPCs are specifically designed to measure earth pressures
in Fig. 3(a), while the arrows identify the four columns of reinforce- against a solid surface. Per manufacturer recommendation for in-
ment strips that were instrumented with strain gauges. Fig. 3 shows stallation, a layer of mortar was first applied to the gauge location
that the wall was divided into two halves by a separation barrier on the wall panel prior to placement. EPCs were pressed into the
composed of a wood frame and high-strength Styrofoam board. mortar layer prior to curing to ensure a completely smooth surface
Each row of reinforcement included four steel reinforcement strips free of point loads.
tying the rear and front MSE wall panels together on both sides of Vertical stress was measured by 36 Geokon 4800 vibrating wire
the separation barrier. The instrumented strips were placed in the EPCs at each reinforcement level as well as 15.24 cm below the
two center columns of reinforcement per compaction side to reduce center of the four concrete leveling pads bearing the MSE wall
any frictional effects generated by the side Strong Walls and sep- panels. The layout of the embedded EPCs measuring vertical stress
aration barrier, both of which were lined with plastic sheeting to at each reinforcement level is depicted in Fig. 3(b). The embedded
reduce friction at the boundaries (Fang et al. 2004). The letters in EPCs were placed in line with the wall-mounted EPCs at each
Fig. 3 refer to Table 1. reinforcement level to best capture the vertical and horizontal stress
The modified proctor test (AASHTO T-180) covers the determi- relationship. Prior to construction, each EPC was tested and cali-
nation of the moisture-density relations of soils when compacted brated inside a small soil box by applying incremental loads with
in a mold and identifies a maximum dry density and optimum mois- a calibrated instrumented loading frame. For installation, soil was
ture content for a given soil. A relative compaction effort target placed and compacted 15.24 cm above the target EPC depth. Each
of 95% of the maximum dry density from T-180 was applied on soil-embedded EPC location was then carefully excavated and lev-
the left (undercompacted) side, and a target of 103% of T-180 was eled at the base. The soil-embedded EPCs were then placed and
applied on the right (overcompacted) side. This was intended to checked with a digital level. Per manufacturer recommendation,
provide a high and low internal friction angle (φ) for the reinforced a powdered sand layer was placed at the bottom of each excavated
soil and represent the compaction effort range commonly experi- area prior to gauge placement, to reduce the development of point
enced in the field. loads. Once the gauges were confirmed to be level, the excavations
The AASHTO (2007) code for load and resistance factor design were backfilled with soil in 2.54-cm hand-compacted layers.
(LRFD) was used in compliance with the FDOT’s standard spec- Five strain gauge locations were installed on 16 of the 32 steel
ifications for road and bridge construction (FDOT 2017) for all reinforcement strips identified in Fig. 3(a). Each strain gauge loca-
other stability design checks and soil testing to closely follow con- tion comprised four individual VPG C2A-06-062LW-350 strain
ventional design practices. From the stability analyses, it was found gauges arranged in a full bridge formation to eliminate the influ-
that the reinforcement strip bolt hole bearing resistance limit state ence of bending and temperature, so that only axial forces were

© ASCE 04024009-4 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2024, 150(4): 04024009


measured within the reinforcements. Waterproof-shielded direct Table 2. Soil properties of base soil and MSE reinforced fill from direct
burial cable was used, and multiple protective coatings were ap- shear testing following AASHTO T-236 (AASHTO 2013)
plied at each gauge location. This was done to ensure the gauges Low High
and wires could survive the high stresses induced by compaction compaction compaction
and overburden (Chen and Fang 2008; Duncan and Seed 1986; Engineering property effort effort
Duncan et al. 1991), as well as exposure to water introduced during Relative compaction, RC (%) 95.7 103.5
compaction. Each instrumented reinforcement strip was incre- Dry density, γ d (kN=m3 ) 15.9 17.2
mentally load-tested before and after each protective coating was Density index/unit weight, I d (%) 64.3 102.8
applied, and just prior to installation. Individual calibration curves Moisture content, w (%) 12.8 12.8
for the 80 full bridge locations were developed from a total of 240 Internal friction angle, φ (degrees) 31.0 40.5
load tests. The percent difference of each calibration curve before
and after the protective coatings were applied was less than 0.1%.
Full bridge strain gauges were also installed on the last sections 12.7% obtained from AASHTO T-180. Table 2 provides the soil
Downloaded from ascelibrary.org by University of Central Florida on 02/07/24. Copyright ASCE. For personal use only; all rights reserved.

of the six threaded rod assemblies to measure the load applied by engineering properties obtained from tests performed at the
the reaction frame. The instrumented threaded bar sections were FDOT’s State Materials Office.
also load-tested and calibrated up to the maximum expected load The coarse aggregate used in the drainage layer met the
for each bar. Calibration curves were found for each bar by com- requirements of FDOT section 901 (FDOT 2017) with a size dis-
paring applied loads to monitored strain, as was done for the gauges tribution of No.89. Each base soil layer was compacted to 103.5%
mounted on the steel strip reinforcements. of T-180 using a vibratory plate compactor to mitigate a potential
To monitor wall movement, Micro-Epsilon WDS-750-P60-CR-P bearing capacity failure of the wall footing based on factored loads.
draw-wire displacement sensors (Vishay Measurements Group, Nuclear density testing in accordance with FM 1-T 238 (FDOT
Raleigh, North Carolina), also known as potentiometers, were 2017), ASTM D4643 (ASTM 2017), and AASHTO (2007) was
installed at the base of each MSE wall compaction zone. The performed at four locations for each compacted lift to ensure the
locations of the two displacement sensor draw-wires, attached to target compaction was achieved. After the third soil lift was com-
the wall panels, are indicated by the letter R in Fig. 3(a). The base pacted, the four MSE wall panel leveling pads were installed at an
of each displacement sensor was mechanically fastened to the soil embedment depth of 15.2 cm from the top of the base soil layer.
retaining blocks and housed in a protective enclosure. An EPC was installed 15.2-cm beneath each leveling pad to record
Two separate data acquisition modules were required to handle vertical stress beneath the wall panels. Starting at the surface of the
the 140 monitored locations. A Campbell Scientific CR6 (Logan, base soil, a reinforced high-strength Styrofoam barrier was con-
Utah) was used to collect data from the EPCs at each reinforcement structed, which divided the area into two equal halves as depicted
level, as well as the strain gauges mounted on the reinforcements. in Figs. 1 and 3. The barrier was constructed in stages as the wall
A Campbell Scientific CR10X was used to monitor the four EPCs height increased and provided separation between the 95% and
beneath the MSE wall leveling pads, the displacement sensors, and 103% compaction effort zones. This allowed the investigation to
the six full bridge strain gauge locations on the reaction frame quantify the effects of an unyielding condition for a wide range
threaded rods. The DAQ modules were programmed to continu- of friction angles developed by undercompacting and overcompact-
ously collect synchronized data every 3 min. ing the reinforced fill, without influence from the adjacent compac-
For each reinforcement level, average measurements were de- tion effort.
rived for the vertical earth pressure, lateral earth pressure, and strip A construction and quality control manual for reinforced earth
tension using the four soil-embedded EPCs, two wall-mounted square panels (RECo 2011) provided by the MSE wall panel manu-
EPCs, and the average strip tensions from both reinforced strips, facturer was followed for proper construction sequencing and
respectively. The average values were then used during construc- technique. Panels were maneuvered and placed using the labora-
tion earth pressure coefficient analyses. Averages were used to ac- tory overhead crane. The rear Strong Wall face was layered with
count for the possible research-induced effects of the separation 12.7 cm of high-strength Styrofoam to provide clearance between
barrier and side Strong Wall boundary conditions; therefore, the the Strong Wall and back MSE wall. The rear unyielding MSE
intent of the averaging was to balance the influence of different panels were centered on the rear leveling pads flush against the
interfaces on each side of the reinforced zone. 12.7-cm Styrofoam layer. Strong Wall anchor points were used
to brace the rear panels in place before fill and reinforcements
were added. The front-yielding MSE panels were placed, leveled,
MSE Wall Construction and braced using conventional MSE wall construction techniques
To best represent real-world conditions, a soil base layer was (Fig. 4) with the exception of wall brace anchorage, which was
first constructed for the unyielding MSE wall to bear upon within attached to the retention block threaded rods as opposed to soil-
the bin-shaped Strong Wall area. To achieve this, a 76.2-cm tall, embedded wood posts used in the field. The first set of panels
61.0-cm thick retaining wall was constructed across the open side consisted of alternating half- and full-height panels at 1.90-cm
of the Strong Wall using three reinforced concrete blocks to retain horizontal spacing for the front and rear walls. Once in place,
the base layer soil. Threaded rods tied to the floor anchors ran EPCs were installed on the inside wall face of the front MSE wall
through holes cast into the retaining blocks, and bearing plates and panels at the first reinforcement level to measure horizontal stresses
nuts were tensioned down on the blocks using a jacking system to [Fig. 3(a)]. Filter cloth was then glued over all panel and side wall
ensure the blocks did not yield from the lateral stresses. joints to avoid soil loss at the wall seems.
The base soil consisted of three 15.2-cm (6-in.) compacted Per construction specifications, the initial soil layers beneath the
layers of soil overlying a 15.2-cm nonwoven geotextile encapsu- first row of reinforcement were prevented from interacting with the
lated gravel drainage layer. The soil used throughout the project wall panels until the first row of reinforcements were connected to
was an AASHTO classified A-3 poorly graded sand typical for the MSE wall panels as depicted in Fig. 4(a); this procedure was
MSE wall construction in Florida, with a maximum dry density to prevent lateral earth pressures from pushing the unsecured panels
(γ d-max ) of 16.6 kN=m3 and optimum moisture content (wopt ) of outward at the base during compaction. After the first row of

© ASCE 04024009-5 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2024, 150(4): 04024009


Downloaded from ascelibrary.org by University of Central Florida on 02/07/24. Copyright ASCE. For personal use only; all rights reserved.

Fig. 4. (a) First course of backfill; and (b) second course of backfill. (Reprinted from RECo 2011, with permission from The Reinforced Earth
Company.)

reinforcements was connected, fill and compaction efforts were MatJacks for the 95% of T-180 compaction effort. Bearing soil
extended to the wall face for all following lifts [Fig. 4(b)]. plates were arranged to maximize coverage of the reinforced soil’s
Once soil layers for both compaction zones were level with footprint to apply a simulated surcharge over the entire reinforced
MSE wall panel reinforcement connections, the steel reinforcement zone. MatJacks were centered on the soil-bearing plates as depicted
strips were installed, tying the front wall panels to the rear unyield- in Fig. 5(b) to avoid application of asymmetric loading on the soil.
ing wall panels. Next, a 15.2-cm compacted lift was constructed The incremental MatJack pressures applied during loading were
over the reinforcements for both compaction zones, and soil- controlled by a manifold using air regulators and monitored using
embedded earth pressure cells were installed 15.2-cm deep, level digital pressure gauges.
with the reinforcement, following the installation process previ- A reinforced steel plate was placed symmetrically on top of
ously outlined. Each compaction zone had four embedded EPCs the MatJack arrangements in both reinforced fill areas with I-beam
for measuring vertical stress at each reinforcement row, as shown reinforcements welded perpendicular to the MSE wall faces
in Fig. 3(b). This process was repeated for the remainder of [Figs. 6(a and b)]. Two custom-made girders were positioned
construction with wall panels, reinforcements, threaded rod seg- perpendicular to the I-beam reinforcements and aligned with the
ments, and reinforced foam boards being installed when necessary. row of three threaded bar assemblies at the front and rear of the
Nuclear density testing in accordance with FM 1-T 238 (FDOT MSE structure, as shown in Fig. 6(a). Each custom girder com-
2017) was conducted in two locations for both compaction zones prised two 5.94-m long C-channels welded back-to-back with
for each 15.2-cm compacted lift to ensure the target compaction
15.2-cm (6-in.) clear spacing using steel plates. The final threaded
was reached within each compaction zone. Data collection began
bar sections, instrumented with strain gauges, were run through
for each gauge as it was installed allowing for the live monitoring
the clear space between the welded C-channels with strain gauges
of reinforcement tensions, horizontal stress, vertical stress, and wall
positioned at the midspan. Steel bearing plates and nuts were then
displacement during construction.
installed on each thread bar assembly to lock the two custom gird-
ers and reaction frame in place. Strain gauge readings from the six
Simulated Wall Height Reaction Frame Design instrumented bar sections were used to ensure that each threaded
Fig. 5(a) shows the model MSE structure after the final layer bar assembly was under equal tension before MatJack loading
of reinforced soil was compacted, and wall panel bracings were began. Fig. 6(b) depicts the fully constructed MSE wall with the
removed. Fig. 5(b) depicts the layout of soil-bearing plates and reaction frame in place.

Fig. 5. (a) Completed MSE wall before reaction frame installation; and (b) MatJacks and soil-bearing plates on the top of the 95% of T-180
compaction effort side.

© ASCE 04024009-6 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2024, 150(4): 04024009


Downloaded from ascelibrary.org by University of Central Florida on 02/07/24. Copyright ASCE. For personal use only; all rights reserved.

Fig. 6. (a) Welded C-channels on top of the reinforced steel plates; and (b) completed MSE wall after reaction frame installation.

Test Results and Discussion then divided by the average dry unit weight measured from nuclear
density testing during construction to estimate simulated soil heights.
Nuclear Density and Loading Frame Results The surcharge was applied in eight load phases shown in Table 4.
A blowout occurred in a MatJack on the 103% compaction
Two nuclear density tests were performed in each compaction zone effort side twice at two different locations during Load Phases 2
for each 15.2-cm soil lift. Based on the results of all nuclear den- and 3. Both times, the reaction frame was removed, and faulty
sity tests, the average relative compaction (RC) from the under- MatJacks were replaced. Load Phases 4–8 were completed without
compacted side was 95.7% of T-180 and the average relative issue. The applied surcharge at Load Phase 8 equated to 6.89 m
compaction from the overcompacted side was 102.8% of T-180. of overburden for the 95% of T-180 side and 6.91 m of overbur-
A summary of the average soil relative compaction and density den for the 103% of T-180 side. This achieved the necessary aspect
index/unit weight [Id , ASTM D4253 (ASTM 2019) and ASTM ratio of 0.3 when added to the constructed wall height of 3 m.
D4254 (ASTM 2016)] between MSE wall reinforcement levels Figs. 7(a and b) combine the depth profiles of the vertical earth
for both compaction zones is shown in Table 3. pressures measured by nuclear density testing (ND) during con-
Table 3 shows that the compaction effort from Row 1 to Row 2 struction with the results of the applied surcharge loading derived
is noticeably lower than the other levels for the 103% of T-180 from threaded bar (DD) measurements to produce a ND/DD pro-
compaction effort. It is suspected that this is due to the construction file, and compare them with the vertical earth pressures measured
method illustrated in Fig. 4(a), where the backfill is prevented from by EPCs at Row 1 (R1 EPC).
contacting MSE wall panels until the backfill height is 15.2 cm Good agreement is observed in Fig. 7(a) between the applied
above the first row of reinforcements. Consequently, the vibratory and measured vertical stress at nearly every soil height above
compaction effort applied to the lifts directly above the first row of Row 1 for the 95% compaction effort besides the first two data
reinforcements was distributed into the relatively loose backfill di- points after the end of construction (EOC), where MatJack failures
rectly behind the wall panels. Compacting the soil to approximately caused variability in the EPC measured vertical earth pressure. This
103% of T-180 was easily achieved after the second row of rein- indicates that there was a minimal loss of vertical stress due to soil
forcements was secured. arching in Row 1 on the 95% compaction effort side. Conversely,
Each incremental surcharge load applied by the reaction frame Fig. 7(b) shows that the measured vertical stress in Row 1 for the
from pressurizing the MatJacks was sustained until the measured 103% compaction effort was less than the applied pressure at every
earth pressures, strip tensions, and reaction bar tensions stabilized. soil height above Row 1 beyond a height of 0.76 m when the sec-
The incremental surcharge pressures for each compaction effort ond row of reinforcement strips was added. This indicates that the
were derived based on the average loads measured from the reac- higher compaction effort resulted in some degree of soil arching
tion frame instrumented rods divided by the surface area of the steel and shear transfer to the wall panels that decreased the overburden
plates bearing on the soil. Each incremental surcharge pressure was pressure within the reinforced zone.

Table 3. Soil properties of base soil and MSE reinforced fill


Soil height (m) 95% of T-180 103% of T-180
Layers Bottom Top γ d (kN=m3 ) RC (%) I d (%) γ d (kN=m3 ) RC (%) I d (%)
Base layer 0.00 0.46 17.2 103.3 101.9 17.2 103.3 101.9
Base layer to Row 1 0.46 0.76 16.3 98.3 77.1 17.1 102.7 99.1
Row 1 to Row 2 0.76 1.52 15.9 96.0 65.9 17.0 102.3 96.7
Row 2 to Row 3 1.52 2.29 15.8 95.3 62.6 17.1 102.9 100.0
Row 3 to Row 4 2.29 3.05 16.0 96.1 66.4 17.1 102.9 100.0
Above Row 4 3.05 3.41 15.9 95.7 64.5 17.3 103.9 105.1

© ASCE 04024009-7 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2024, 150(4): 04024009


Table 4. Simulated surcharge loading results These alternating directional shifts in wall displacement are corre-
95% of T-180 103% of T-180 lated to construction processes such as adding wall panels, attach-
(γ d ¼ 15.91 kN=m3 ) (γ d ¼ 17.05 kN=m3 ) ing subsequent levels of reinforcement, and removing wall panel
bracings. This is also supported by the measured displacements
Load Surcharge Soil Surcharge Soil
phase (kPa) height (m) (kPa) height (m)
that occurred during surcharge loading where alternating direc-
tional shifts were not observed, and the panels only shifted away
1 15.03 0.94 18.10 1.06 from the reinforced backfill. The wall displacements presented in
2 27.24 1.71 30.07 1.76 Figs. 8–11 were all measured from the same potentiometers at the
3 35.67 2.24 44.62 2.62
base of each compaction zone. The bottom wall displacement was
4 53.58 3.37 69.71 4.09
5 63.06 3.96 73.21 4.29 adjusted to the soil height above each row of measurements in the
6 73.93 4.65 83.26 4.89 following figures.
7 93.27 5.86 98.63 5.79
8 109.69 6.89 117.74 6.91
Downloaded from ascelibrary.org by University of Central Florida on 02/07/24. Copyright ASCE. For personal use only; all rights reserved.

Reinforcement Row 2 Results Analysis


Row 2 showed similar results to Row 1. For the 95% compaction
effort, the kh for Row 2 decreased until a soil height of approx-
Reinforcement Row 1 Results imately 5.85 m above Row 2, where kh stabilized in an active
condition, in Fig. 9(a). The active condition is supported by the
The ratio of lateral to vertical stress (kh ) decreased at Row 1 for continuous wall movement away from the soil mass measured dur-
both compaction efforts during construction. For the 95% of ing surcharge loading shown in Fig. 9(c).
T-180 compaction effort, shown in Fig. 8(a), the kh moved from For the 103% compaction effort, the Row 2 kh decreased
a passive condition during early compaction to an at-rest condi- throughout construction and surcharge loading to below an active
tion at the end of construction (EOC, i.e., horizontal black dashed condition, in Fig. 9(b). As previously discussed, minimal wall
line). The kh further decreased during the surcharge loading until movement was observed for the higher compaction effort. Also
an approximate height of 6 m above Row 1, where kh is seen to observed for both Rows 1 and 2 were higher tension loads in the
stabilize in the active condition. The active condition is confirmed reinforcements for the lower compaction effort compared to the
by the continuous wall displacement (δ wall ) seen in Fig. 8(c). higher compaction effort. This provides evidence that panel dis-
The kh for the 103% of T-180 compaction effort decreased placement is a key determining factor in reinforcement tensions.
until a soil height of approximately 1.98 m above Row 1, where
it stabilized above an at-rest condition until 3.69 m, where the kh
began a decline to below an active condition by the end of Reinforcement Row 3 Result Analysis
loading [Fig. 8(b)]. Unlike the 95% of T-180 compaction effort The third row indicated a change in behavior for both compaction
side, limited wall displacement was measured for the 103% side. zones. The 95% compaction zone produced a kh that decreased
The lack of wall movement further suggests that shear stress until a soil height of approximately 3.38 m above Row 3, then
at the soil–wall interface was increased for the higher compac- stabilized in an at-rest condition until approximately 5.76 m of
tion effort. soil height, shown in Fig. 10(a). Past this, kh continued to decrease
Figs. 8(c and d) indicate that the wall panels actually moved slightly below the at-rest condition. Additionally, reinforcement
toward the reinforced fill at a soil height of 2.62 m above Row 1, strip tensions were much higher in the top two rows of reinforce-
which is when the temporary braces were removed at EOC. This ment when compared to the bottom two rows. As suggested by
followed the trends seen in Figs. 8(c and d) in relation to wall the coherent gravity method, higher locked-in compaction forces
displacement that is observed to move both toward and away in the top half of the wall fully engaged tension forces in the
from the backfill during construction for both compaction efforts. reinforcement, while compaction in the lower half of the wall was

(a) (b)

Fig. 7. Depth profiles of applied vertical stress (ND/DD) compared to measured vertical stress at Row 1 (R1 EPC) for (a) 95% compaction effort; and
(b) 103% compaction effort.

© ASCE 04024009-8 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2024, 150(4): 04024009


Downloaded from ascelibrary.org by University of Central Florida on 02/07/24. Copyright ASCE. For personal use only; all rights reserved.

(a) (b)

(c) (d)

Fig. 8. Depth profiles of Row 1 earth pressures and coefficients for (a) 95%; and (b) 103% compaction efforts; and strip tensions and wall
displacements for (c) 95%; and (d) 103% compaction efforts.

insufficient to fully engage tension forces in the reinforcement Reinforcement Row 4 Results Analysis
strips. During the loading phase, the top panels were prevented
from moving outward until bottom panels fully engaged their An issue arose at the fourth level of reinforcement when the divider
reinforcement strips, resulting in the at-rest condition seen in separating the two compaction zones was observed to be bowing
Fig. 10(a). toward the undercompacted zone at the top where the divider was
Similar to the 95% compaction zone, Row 3 in the 103% weakest (Fig. 11). This portion of the divider had the least amount
compaction zone indicated a change in soil behavior in the upper of soil on each side to counteract the large surcharge load induced
half of the wall, likely due to locked-in compaction forces as sug- 0.37 m above the reinforcement level. Due to the 103% side be-
gested by the coherent gravity method. The kh at Row 3 began ing loaded to a higher pressure (i.e., qs103% ¼ 117.7 kPa versus
above the passive state, then slowly declined to the at-rest con- qs95% ¼ 109.7 kPa) to achieve the necessary surcharge height, it is
dition during the last load phase, Fig. 10(b). Also, similar to the believed that the soil in the 103% side at Row 4 began moving
95% side, strip tensions in Rows 3 and 4 of the 103% side were toward the 95% zone, which resulted in an increase in lateral stress.
much higher than Rows 1 and 2, further indicating that the upper This was expressed through an increasing kh value with depth on
half of the wall was exposed to higher locked-in compaction the 95% side and a near-active condition being reached on the
forces, as suggested by the coherent gravity method. It is believed 103% side, shown in Figs. 11(a and b). Although the analysis at
that these locked-in compaction forces in the upper half of the Row 4 did match the observed behavior, these results were not used
wall account for the inward displacement seen at the base of the to derive earth pressure coefficients because they were deemed to
wall as a slight rotational component was induced by the locked- be unrepresentative of the true soil behavior had the divider not
in compaction forces when bracings were removed. As discussed, failed in the uppermost segment.
the displacement measurements presented in Fig. 10 were taken Overall, strip tensions were higher at each row on the 103% side
from the potentiometers at the base of the wall for each compac- compared to the 95% side after construction but were then lower
tion effort. If wall displacements had been measured at the Row 3 than the 95% side at maximum surcharge loading. This is a strong
reinforcement level, a displacement shift away from the backfill indication that strip tension is largely influenced by wall move-
would likely have been observed at the EOC when bracings were ment, because the 95% side experienced less surcharge load than
removed. the 103% side but did experience five times the wall displacement.

© ASCE 04024009-9 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2024, 150(4): 04024009


Downloaded from ascelibrary.org by University of Central Florida on 02/07/24. Copyright ASCE. For personal use only; all rights reserved.

(a) (b)

(c) (d)

Fig. 9. Depth profiles of Row 2 earth pressures and coefficients for (a) 95%; and (b) 103% compaction efforts; and strip tensions and wall
displacements for (c) 95%; and (d) 103% compaction efforts.

Force Equilibrium Analysis the applied soil height and wall weight, indicating a small degree
of soil arching. An overburden reduction of 1.45 kPa was found
An analysis of the forces and stresses acting on the wall confirmed
that force equilibrium had been achieved for both compaction between the applied and the measured stress at Row 1 for the maxi-
efforts. A diagram of the forces and pressures acting on the wall mum surcharge load. This overburden reduction of 1.45 kPa was
at row one is shown in Fig. 12. These forces and pressures included then converted to a vertical force by multiplying the stress by the
the average vertical stress (σvRow1 ), average lateral stress (σhRow1 ), reinforced zone area. Dividing this force between the two pads and
simulated surcharge (qs ), shear stress at the wall (τ w ), weight of the their areas resulted in a vertical stress of 6.77 kPa on each leveling
wall panels (W w ), weight of the soil column acting on the leveling pad due to the arching effect. This was in good agreement with the
pad (W s ), and the force at the leveling pad (FLP ). measured additional average leveling pad pressure, beyond the
It was found that frictional resistance generated at the soil–wall weight of wall panels and soil column, of 6.87 kPa. The minimal
interface led to an arching effect that reduced overburden pressure 1.5% difference between these values indicates force equilibrium
within the reinforced soil zone and, consequently, lateral stress at was achieved.
the wall. For force equilibrium, the shear stress distributed to the The estimated shear stress available at the soil–wall interface
wall face due to frictional resistance should be equal to the reduc- (τ w ) was then calculated as 26.04 kPa for the 95% side using a
tion in overburden pressure, which is then translated into additional soil–wall friction angle of 30° (Gomez et al. 2003). The additional
leveling pad pressure/force beyond the weight of the wall and soil average leveling pad pressure of 6.87 kPa was then converted into a
column acting on the leveling pad. The depth profiles of the com- vertical force by multiplying the value by the leveling pad area.
bined weight of the wall and soil column (SC), leveling pad pres- This vertical force was then divided by the area of the wall facing
sure under the front wall (LPF), leveling pad pressure under the above Row 1 to find the equivalent shear stress of 0.83 kPa. This
back wall (LPB), and the average leveling pad pressure (LPAVG) indicates that the measured equivalent shear stress at the soil–wall
are depicted for both compaction effort zones in Fig. 13. interface did not exceed the available shear stress, and the limited
For the 95% side, it is shown in Fig. 13(a) that the average lev- wall height to surcharge height had no effect on the shear transfer
eling pad pressure recorded at EOC was consistent with the com- from soil arching.
bined weight of the soil and the wall. After surcharge loading was Fig. 13(b) shows that the average pressure on the 103% leveling
applied, the average leveling pad pressure slightly increased above pads exceeded the combined weight of the leveling pad soil column

© ASCE 04024009-10 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2024, 150(4): 04024009


Downloaded from ascelibrary.org by University of Central Florida on 02/07/24. Copyright ASCE. For personal use only; all rights reserved.

(a) (b)

(c) (d)

Fig. 10. Depth profiles of Row 3 earth pressures and coefficients for (a) 95%; and (b) 103% compaction efforts; and strip tensions and wall
displacements for (c) 95%; and (d) 103% compaction efforts.

and wall panels by 38.99 kPa. The average additional leveling pad earth pressure coefficients in both compaction zones moved from
pressure on the 103% side was five times larger than on the 95% a passive condition to an active condition for the first two rows.
side, indicating a much higher degree of soil arching. The overbur- Meanwhile, the third row in both compaction efforts moved toward
den reduction from arching was calculated to be 8.31 kPa, which an at-rest condition, providing more evidence that the wall behaved
translated to a leveling pad pressure of 38.79 kPa. This resulted in a differently in the upper and lower halves, as suggested by the co-
negligible 0.52% difference from the measured additional average herent gravity method.
leveling pad pressure, and indicates that force equilibrium was
achieved.
The equivalent shear stress calculated from average leveling pad Design Recommendations for Soil Pressure against
pressure on the 103% side of 4.73 kPa did not exceed the estimated Unyielding Surfaces
shear stress available at the soil–wall interface of 14.95 kPa. This After comparing the earth pressure coefficients derived from both
further confirmed that the limited wall height to surcharge height compaction efforts to the earth pressure coefficients estimated from
had no effect on the shear transfer from soil arching. the coherent gravity method, the simplified method, and Spangler
and Handy method (Fig. 15), it is clear that the conventional meth-
ods inadequately quantified the additional lateral stress developed
Derived Earth Pressure Coefficients
from compaction efforts in an unyielding condition using inexten-
Figs. 14(a and b) show the derived earth pressure coefficients for sible reinforcement.
the first three reinforcement rows along with active, at-rest, and A new equation was subsequently developed based on the re-
passive earth pressure coefficients for the 95% side and 103% side search results that considered the movement of earth pressure co-
compaction efforts, respectively. Row 4 was removed from the efficients from the passive condition to an active or at-rest condition
analysis due to bowing of the separation barrier at the top of the as previously described. Additionally, the equation also considered
reinforced fill, as previously discussed. The 95% compaction effort a variable range of φ based on the values found for the under and
resulted in a φ ¼ 31.2° while the 103% compaction effort produced over compacted zones which correlates well to the ranges most
a φ ¼ 39.5°, which provided a large range of φ based on the com- likely found in practice. This allows the expected compaction ef-
paction effort for the investigation. As seen in Figs. 14(a and b), fort to be considered directly in design when using the equation.

© ASCE 04024009-11 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2024, 150(4): 04024009


Downloaded from ascelibrary.org by University of Central Florida on 02/07/24. Copyright ASCE. For personal use only; all rights reserved.

(a) (b)

(c) (d)

Fig. 11. Depth profiles of Row 4 earth pressures and coefficients for (a) 95%; and (b) 103% compaction efforts; and strip tensions and wall
displacements for (c) 95%; and (d) 103% compaction efforts.

assumed that concrete facing panels and flat steel reinforcement


strips are used in design

kh ¼ kp@OC × zb ð1Þ

where kh = design earth pressure coefficient; kp@OC = passive earth


pressure coefficient for an over-compact state of soil density; z =
depth below the top of the wall (m); and b = exponent parameter
that quantifies a variable φ based on the compaction effort and geo-
static state of stress in the soil (active, at-rest, passive).
Laboratory testing must be performed to fully develop the equa-
tion. Modified proctor tests must be performed at 95% of T-180 and
above 100% of T-180 to satisfy compaction requirements in addi-
tion to direct shear testing to determine the internal friction angle
at each state of soil density. For demonstration, the 103% of T-180
compaction effort measured within the reinforced soil zone from
nuclear density testing is used for equation development. The fol-
lowing trial first sets kh to an active condition for 95% of T-180 and
Fig. 12. Force equilibrium diagram. z to 6 m based on the observed data trends and coherent gravity
design methodology. After solving for b, Eq. (2) is produced

b ¼ logðka@95 =kp@103 Þ= logðzÞ ð2Þ


Eq. (1) was developed to estimate the expected design earth pres-
sure coefficients of the unique MSE wall configuration where the The b parameter found in Eq. (2) is then used in Eq. (1) for
aspect ratio is equal to or less than 0.3, and inextensible reinforce- design. Following the design methodology of the coherent gravity
ment is mechanically fastened between stiff wall facings. It is and simplified methods, a value of ka@95 is used rather than the kh

© ASCE 04024009-12 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2024, 150(4): 04024009


Downloaded from ascelibrary.org by University of Central Florida on 02/07/24. Copyright ASCE. For personal use only; all rights reserved.

(a) (b)

Fig. 13. Depth profiles of vertical earth pressure for the soil column (SC), front leveling pad (LPF), back leveling pad (LPB), and leveling pad average
for (a) 95% compaction; and (b) 103% compaction efforts.

(a) (b)

Fig. 14. Depth profiles of earth pressure coefficients for (a) 95% compaction effort; and (b) 103% compaction effort.

Fig. 15. Conventional design methods coherent gravity, simplified, and Spangler and Handy compared to earth pressure coefficients measured in an
unyielding condition.

© ASCE 04024009-13 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2024, 150(4): 04024009


Downloaded from ascelibrary.org by University of Central Florida on 02/07/24. Copyright ASCE. For personal use only; all rights reserved.

(a) (b)

Fig. 16. Developed design equation: (a) kp − ka ; and (b) kp − k0 with a 6-m cutoff compared to earth pressure coefficients measured in an unyielding
condition.

derived from Eq. (1) beyond a depth of 6 m. As seen in Fig. 16(a), • Higher compaction efforts increase soil arching and shear trans-
the design equation using ka@95 fits well for Rows 1 and 2, which fer to wall panels in an unyielding condition. The shear trans-
moved from a passive condition to an active condition, but it under- ferred at the soil–wall interface is then translated into increased
estimated the earth pressure coefficients for Row 3 from both com- bearing pressure on the MSE wall leveling pads.
paction zones. • Increased shear transfer at the soil–wall interface due to soil
Because the upper half of the wall was observed to move from arching leads to a reduction in vertical and lateral stress as
a passive condition to an at-rest condition for both compaction described by the Spangler and Handy method.
efforts, the at-rest earth pressure coefficient from 95% of T-180 • As described by both the coherent gravity and simplified meth-
was also used in the equation development to quantify behavior ods, higher locked-in compaction forces were generated in the
observed in Row 3 for both compaction efforts. The b parameter upper half of the constructed wall compared to the bottom half,
considering the at-rest earth pressure is shown in Eq. (3) regardless of compaction effort. The distinction between the
upper and lower region of a wall should be treated with cau-
b ¼ logðk0@95 =kp@103 Þ= logðzÞ ð3Þ tion due to the limited wall height of 3 m for the experimental
MSE wall.
The results provided in Fig. 16(b) show that the updated equa- • In an unyielding condition, earth pressure coefficients move
tion can quantify the behavior of Row 3 in the upper half of the wall from a passive condition to an active or at-rest condition as soil
for both compaction zones well. The only two outlying data points height is increased above the reinforcement level. Earth pres-
occurred during the first two load phases when MatJacks failed sure coefficients developed in the unyielding condition gener-
and can be disregarded. In summary, two equations considering a ally stabilized in an active or at-rest condition at an approximate
variable φ based on the compaction effort were developed that fol- depth of 6 m as suggested by the coherent gravity and simplified
lowed the data trends well for the upper and lower halves of the methods.
wall. It is advised that, due to the limited constructed wall height • An equation that incorporates a variable friction angle based on
of 3 m, caution be taken when distinguishing between the upper the compaction effort for an unyielding condition and FDOT
and lower halves of a wall. It is entirely possible that behavior ob- design and construction requirements was developed that fol-
served in Row 3 of the experimental MSE wall could occur in the lowed the trends of the measured data well.
lower half of a much taller wall. One possible explanation of the
different behavior observed between the upper and lower halves of
the experimental MSE wall could be the construction procedures Data Availability Statement
implemented, where soil is not placed or compacted against the
wall facing until after the first row of reinforcements are connected. Some or all data, models, or code that support the findings of this
study are available from the corresponding author upon reasonable
request.
Conclusions
Based on the results of this study, the following conclusions can be Acknowledgments
drawn for an MSE wall with an aspect ratio less than or equal to 0.3
in an unyielding condition using inextensible reinforcement tied The assistance of the FDOT’s State Materials Office as well as
between stiff wall facings: the district and central office Geotechnical Engineers is greatly

© ASCE 04024009-14 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2024, 150(4): 04024009


appreciated. The authors thank the FDOT’s Geotechnical Field Chen, T. J., and Y. S. Fang. 2008. “Earth pressure due to vibratory com-
Specialists: Bruce Swidarski, Todd Britton, Kyle Sheppard, and paction.” J. Geotech. Geoenviron. Eng. 134 (4): 437–444. https://doi
Travis “Dalton” Stevens, Geotechnical Laboratory Manager Bill .org/10.1061/(ASCE)1090-0241(2008)134:4(437).
Greenwood, and Geotechnical Laboratory Specialists Mike Risher. Duncan, J. M., and R. B. Seed. 1986. “Compaction-induced earth pressures
The authors also thank Keith Brabant and the Reinforced Earth under K0-conditions.” J. Geotech. Eng. 112 (1): 1–22. https://doi.org
Company for their guidance on MSE wall construction, and Mark /10.1061/(ASCE)0733-9410(1986)112:1(1).
Duncan, J. M., G. W. Williams, A. L. Sehn, and R. B. Seed. 1991.
Wallace at Campbell Scientific for his programming expertise.
“Estimation earth pressures due to compaction.” J. Geotech. Eng.
Without their assistance, this research would not have been pos-
117 (12): 1833–1847. https://doi.org/10.1061/(ASCE)0733-9410(1991)
sible. This work was supported by the Florida Department of
117:12(1833).
Transportation’s State Materials Office through research contract Fang, Y. S., T. J. Chen, R. D. Holtz, and W. F. Lee. 2004. “Reduction of
No. BDV31-977-89. The opinions, findings, and conclusions ex- boundary friction in model tests.” Geotech. Test. J. 27 (1): 3–12. https://
pressed in this publication are those of the author(s) and not neces- doi.org/10.1520/GTJ10812.
sarily those of the Florida Department of Transportation or the
Downloaded from ascelibrary.org by University of Central Florida on 02/07/24. Copyright ASCE. For personal use only; all rights reserved.

FDOT (Florida Department of Transportation). 2017. Standard specifica-


United States Department of Transportation. tions for road and bridge construction. Tallahassee, FL: FDOT.
FHWA (Federal Highway Administration), R. R. Berg, B. R. Christopher,
and N. C. Samtani. 2009a. Design of mechanically stabilized earth
Supplemental Materials walls and reinforced soil slopes—Volume I. Publication No. FHWA-
NHI-10-024, FHWA GEC 011-Vol I. Washington, DC: FHWA.
Video S1 is available online in the ASCE Library (www.ascelibrary FHWA (Federal Highway Administration), R. R. Berg, B. R. Christopher,
.org). and N. C. Samtani. 2009b. Design of mechanically stabilized earth
walls and reinforced soil slopes—Volume II. Publication No. FHWA-
NHI-10-025, FHWA GEC 011-Vol II. Washington, DC: FHWA.
References FHWA (Federal Highway Administration), N. C. Samtani, and E. A.
AASHTO. 2007. LRFD bridger design specifications. 4th ed. Washington, Nowatzki. 2006a. Soils and foundation reference manual—Volume I.
DC: AASHTO. Publication No. FHWA-NHI-06-088. Washington, DC: FHWA.
AASHTO. 2013. Direct shear test of soils under consolidated drained FHWA (Federal Highway Administration), N. C. Samtani, and E. A.
conditions. AASHTO T-236. Washington, DC: AASHTO. Nowatzki. 2006b. Soils and foundation reference manual—Volume II.
Allen, T., B. Christopher, V. Elias, and J. DiMaggio. 2001. Development of Publication No. FHWA-NHI-06-089. Washington, DC: FHWA.
the simplified method for internal stability design of mechanically sta- Gomez, J. E., G. M. Filz, and R. M. Ebeling. 2003. “Extended hyperbolic
bilized earth walls. Rep. No. WA-RD 513.1. Olympia, WA: Washington model for sand-to-concrete interfaces.” J. Geotech. Geoenviron. Eng.
State DOT. 129 (11): 993–1000. https://doi.org/10.1061/(ASCE)1090-0241(2003)
Anderson, P. L., R. A. Gladstone, and J. E. Sankey. 2012. “State of the 129:11(993).
practice of MSE wall design for highway structures.” In Geotechnical Handy, R. L. 1985. “The arch in soil arching.” J. Geotech. Eng. 111 (3):
Engineering State of the Art and Practice: Keynote Lectures from Geo- 302–318. https://doi.org/10.1061/(ASCE)0733-9410(1985)111:3(302).
Congress 2012, GSP 226. Reston, VA: ASCE. https://doi.org/10.1061 Kim, D., and R. Salgado. 2012. “Load and resistance factors for internal
/9780784412138.0018. checks of mechanically stabilized earth walls.” J. Geotech. Geoenviron.
Anderson, P. L., R. A. Gladstone, and J. L. Withiam. 2010. “Coherent Eng. 138 (8): 910–921. https://doi.org/10.1061/(ASCE)GT.1943-5606
gravity: The correct design method for steel-reinforced MSE walls.” .0000664.
In Proc., Earth Retention Conf., 3GSP 208. Reston, VA: ASCE. Kniss, K. T., S. G. Wright, J. Zornberg, and K. Yang. 2007. Design con-
https://doi.org/10.1061/41128(384)51. siderations for MSE retaining walls constructed in confined spaces.
ASTM. 2016. Standard test methods for minimum index density and unit Rep. No. FHWA/TX-08/0-5506-1. Austin, TX: Texas DOT.
weight of soils and calculation of relative density. ASTM D4254. West Lawson, C. R., and T. W. Yee. 2005. “Reinforced soil retaining walls with
Conshohocken, PA: ASTM.
constrained reinforced fill zones.” In Proc., Slopes and retaining struc-
ASTM. 2017. Standard test method for determination of water content
tures under seismic and static conditions, GSP 140. Reston, VA: ASCE.
of soil and rock by microwave oven heating. ASTM D4643. West
Paik, K. H., and R. Salgado. 2003. “Estimation of active earth pressure
Conshohocken, PA: ASTM.
against rigid retaining walls considering arching effects.” Géotechnique
ASTM. 2019. Standard test methods for maximum index density and
unit weight of soils using a vibratory table. ASTM D4253. West 53 (7): 643–653. https://doi.org/10.1680/geot.2003.53.7.643.
Conshohocken, PA: ASTM. RECo (Reinforced Earth Company). 2003. Documentation and test results
Chalermyanont, T., and C. H. Benson. 2004. “Reliability-based design for for clip angle connections used in reinforced earth structures. Sterling,
internal stability of mechanically stabilized earth walls.” J. Geotech. VA: RECo.
Geoenviron. Eng. 130 (2): 163–173. https://doi.org/10.1061/(ASCE) RECo (Reinforced Earth Company). 2011. Design manual for reinforced
1090-0241(2004)130:2(163). earth walls. Sterling, VA: RECo.
Chalermyanont, T., and C. H. Benson. 2005. “Reliability-based design for Spangler, M. G., and R. L. Handy. 1984. Soil engineering. New York:
external stability of mechanically stabilized earth walls.” J. Geotech. Intext Educational Publishers.
Geoenviron. Eng. 5 (3): 196–205. https://doi.org/10.1061/(ASCE)1532 WSDOT (Washington State Department of Transportation). 2010. Geotech-
-3641(2005)5:3(196). nical design manual. M-46-03.01. Olympia, WA: WSDOT.

© ASCE 04024009-15 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2024, 150(4): 04024009

You might also like