Você está na página 1de 23

MINING 301-21:CUT-AND-COVER CONSTRUCTION IN TUNNELING

(Much of this lesson was taken from J.L.Wilson, “Cut-and-Cover Tunnel Structures, Tunnel
Engineering Handbook, 1996)

Introduction

It is not the intent of this lesson to teach you how to design structures that are used in cut and
cover construction of tunnels. This is a Civil Engineering specialty, just as longwall coal mining
is a Mining Engineering specialty, and you wouldn’t expect to learn that in one lesson. However,
I will try to cover the various methods used for cut and cover construction and support and we
will review the design consideration that must be accounted for in designing for cut-and-cover
structures.

Shallow-depth tunnels, such as large sewer tunnels, vehicular tunnels, and rapid transit tunnels,
are frequently designed as structures to be constructed using the cut-and-cover method. Tunnel
construction is characterized as "cut-and-cover" construction when the tunnel structure is
constructed in a braced, trench-type excavation ("cut") and [the resulting structure] is
subsequently backfilled ("covered"). For depths up to 35-45 ft this method is often cheaper and
more practical than underground tunneling, and depths of 60 ft or more are quite common for
rapid transit cuts. This chapter discusses the design and construction of the larger cast-in-place
concrete structures used as sewer tunnels or transportation tunnels for pedestrian, vehicular, or
rapid transit traffic. The tunnel is typically designed as a box-shaped frame, and due to the
limited space available in urban areas, it is usually constructed within a braced excavation.
Where adequate space is available, such as in open areas beyond urban development, it is often
more economical to use open-cut construction. Where the tunnel alignment is beneath a city
street, cut-and-cover construction interferes with traffic and other activities. This disruption is
lessened through the use of decking over the excavation, placed immediately following removal
of the first lift of excavation. The deck is left in place with construction proceeding below it until
the stage is reached for final backfilling and surface restoration.[ Figures 17-1 through 17-4,
TEH] Fig. 21-1 through 21-4 show cross sections of the more common types of cut-and-cover
tunnel structures.
In this chapter, discussions or characterizations of usual practice in the design and construction of
cut-and-cover tunnel structures refer generally to practice in the United States. The design and
construction of these structures in Canada, Mexico, Europe, Asia, and elsewhere abroad is
similar in many respects, but it can differ in many respects as well.

Subway Line Structures

In cut-and-cover construction between stations, the subway tracks are usually enclosed in a
reinforced concrete double box structure with a supporting center wall or beam with columns.
These tunnel structures are commonly referred to as "line" structures. The track centers are
normally located as close together as possible. In a typical double box section, each trainway will
have a clear width of about 14-15 ft, depending upon width of vehicle and clearances to be
provided for equipment and manways [ Figure 17-1, TEH] Fig. 21-1. The configuration of

1
subway line structures can depart from the typical section to accommodate atypical track
alignment or grade. When, for example, system standards mandate that stations be designed with
a center platform [Figure 17-2] Fig. 21-2, the track centers will need to be widened through an
appropriate transition length upon approaching the station. Occasionally, it will be advantageous
or necessary to gradually change the alignment and grade to the "over and under" position, in
which one track lies above and in line with the other. In general, the configuration of the subway
line structure must be subordinate to system requirements for track alignment and grade.

Subway Stations

Station structures include the trainway for trains, boarding and off-boarding platforms, stairs,
escalators, concourse areas for fare collection, and service rooms. If the line structure is two
circular tunnels constructed by tunneling methods, the station may be designed with a single
center platform. If the line structure is a cut-and-cover double box subway line structure, the
station is usually of the side platform type, except at terminals where center platforms are
normally used to comply with system standards.

The usual perception of a cut-and-cover subway station is that of a two- or three-story reinforced
concrete structure constructed in a rectangular excavation 50-65 ft wide, 500-800 ft long, and 50-
65 ft deep. Typically, the station is a two-story structure with two tracks, as well as the center or
side platforms, supported on the invert. A mezzanine floor typically lies between the roof slab
and the invert/platform level. Fig. 21-2 shows a basic cross section at the 7th/Flower Station,
constructed for the Los Angeles County Metropolitan Transportation Authority (MTA).

The complete subway station will be much more complex than is indicated by Fig. 21-2. Internal
configuration will be significantly affected by the need to provide escalators, stairs, ventilation
requirements, rooms for mechanical aud electrical equipment and other maintenance, and safety
and service facilities. Architectural treatment of the station will also affect internal configuration
and may have an important effect on the design of the basic structure as well. The external
configuration of the station will ordinarily be irregular at and above the mezzanine level, where
station entrances must be provided.

Although the two-story station may be considered conventional, more complex station structures
are not unusual, particularly in and near the center of urban areas. At these locations the subway
stations may be constructed at the intersections of principal system trainway lines, for example.
In such cases boarding and off-boarding platforms in two directions are normally required, so
that in plan view the station structure is constructed in the shape of a cross, with, usually, more
than two stories. Whenever there is a requirement that the subway station accommodate more
than two tracks, the station structure will ordinarily be significantly more complex than the
conventional subway station.

2
Subsurface Highway Structures

Subsurface highway structures of significant length, and constructed by the cut-and-cover


method, are not common. Because of the high cost of the construction of these vehicular tunnels,
the initial and operating cost of ventilation, and cost and practical problems of access and egress,
surface and elevated highways are usually much more feasible. There are, however, occasions
when a cut-and-cover vehicular tunnel is the most feasible or desirable alternative for a proposed
section of a highway.

Cut-and-cover vehicular tunnels are probably most commonly constructed at the approaches to
subaqueous vehicular tubes, due to the depth required for the highway structure to serve in this
capacity. In this application they are usually under the groundwater table and, due as well to their
width, they are typically very massive reinforced concrete structures.Fig. 21-3, [Figure 17-3]
shows a typical cross section of the Central Artery (1-90) Tunnel as it approaches the 1-90
Immersed Tube Tunnel in Boston.

Cut-and-cover vehicular tunnels can be appropriate as well where there is a compelling reason to
restore the highway right of way to its original condition or to provide for other productive use of
the surface property. Cut-and-cover vehicular tunnels have been constructed under major urban
airport runways, for example. A significant portion of the 1-90 freeway on Mercer Island in
Washington State was constructed as a cut-and-cover vehicular tunnel. [There is a highway
tunnel under both the old Stapleton Airport in Denver, and the McCarran Airport in Las Vegas
that were constructed by cut-and-cover.]

Sewer Structures

The cut-and-cover sewer tunnels referred to or discussed in this chapter are large reinforced
concrete box structures used for the transport and/or storage of wastewater. These structures are
larger in area than the largest reinforced concrete pipe or precast concrete cylinder pipe
commercially available. The larger of the transport/storage tunnels constructed for the San
Francisco Clean Water Program are internally 20 ft wide and up to 33 ft or more high. Fig. 21-4,
[Figure 17-4])

TUNNEL DESIGN-STRUCTURAL

Loading--General

The cut-and-cover structure must be designed to have structural capacity sufficient to resist safely
all loads and influences that may be expected over the life of the structure. The principal loads to
be resisted are ordinarily the long term development of water and earth pressures, dead load
including the weight of earth cover, surface surcharge load, and live load.

All loads or potential loads should be categorized similarly to the load categories specified by the
American Association of State Highway and Transportation Officials (AASHTO). The load

3
categories should represent the requirements of the particular cut-and-cover structure under
consideration. For example, the Washington Metropolitan Area Transit Authority (WMATA)
specifies that rapid transit structures must be proportioned for the following loads and forces
when they exist:
• Dead Load (DL)
• Live Load (LL)
• Impact (I)
• Centrifugal Force (CF)
• Rolling Force (RF)
• Longitudinal Braking and Tractive Force (LF)
• Horizontal Earth Pressure (E)
• Buoyancy (B)
• Flood (FL)
• Shrinkage Force (S)
• Thermal Force (T)

In locations where there is a potential for significant seismic activity, earthquake forces (EQ)
must be added.

Loads and forces that must be considered in formulating design criteria for vehicular cut-and-
cover tunnels typically are similar. The loads and forces identified and tabulated in AASHTO
Standard Specifications are ordinarily appropriate for this purpose.

Loads and forces applicable to cut-and-cover sewer tunnels should be similarly identified. In
terms of dead load, forces due to horizontal earth pressure and other applicable loads and forces
are normally similar to those of rapid transit tunnels, except that the internal live load consists of
the loads and pressures imparted to the tunnel structure by the contents of the sewer. In some
cases, the hydraulic grade line may be above the roof of the cut-and-cover sewer, and the
corresponding internal pressures must be properly evaluated. In the latter case, transient forces
may also need to be considered.

Dead Load: The dead load to be considered for the design of cut-and-cover structures normally
consists of the weight of the basic structure, the weight of secondary elements permanently
supported by the structure, and the weight of the earth cover supported by the roof of the
structure and acting as a simple gravity load. The design unit weight for the earth cover should
not be taken as less than 120 pcf for dry fill or less than 130 pcf for moist fill. Some authorities
specify a minimum design unit weight of 130 pcf for the earth cover both above and below the
groundwater table.

Shallow structures of the size and width of subways or vehicular tunnels may impose undue
restrictions upon future loadings over the tunnel if the structure is designed only for the actual
earth cover. For example, structures that pass under city streets must be designed to permit
special vehicle loadings in excess of normal axle loads, such as trucks moving transformers,
boilers, vaults, or other large structural loads. Dead load should be considered as applied in

4
stages representative of all conditions likely to be encountered during the life of the structure. For
example, where removal of earth cover is a foreseeable possibility, and where such removal
could affect design requirements, a separate load case addressing this condition should be
analyzed.

Live Load, Impact, and Other Dynamic Forces: Subway vehicle loads are dependent on the
passenger and maintenance vehicles operating within the system. Standard vehicle loadings (LL)
used in San Francisco, Los Angeles, Toronto, and Washington, D.C., are shown in Fig. 21-5
[Figure 17-5]. In addition, loads due to impact, centrifugal force, rolling force, and longitudinal
braking and tractive force must be accounted for. Impact loads may be defined as statically
equivalent dynamic loads resulting from vertical acceleration of live loads. For subway tunnels
the impact allowance (I) is ordinarily formulated by appropriate modification of the expressions
specified by either AASHTO or the Americaii Railway Engineering Association (AREA), or
both. For preliminary design purposes this allowance may be taken as 10% of the static vehicle
loading. Centrifugal force (CF) can be calculated theoretically as a function of static subway
vehicle loading, design speed, and degree of track curvature. A resulting expression for
centrifugal force is given by both AASHTO and AREA and is used for subway tunnel design.
subway design CF is assumed applied horizontally at a distance representative of the location of
the center of gravity above the top of rail (usually about 5 ft). Rolling force (RF) is usually
considered to be 10% of static subway vehicle loading applied downward on one rail and upward
on the other. Longitudinal breaking and tractive force (LF) is usually considered to be a
longitudinal force equal to 15% of static vehicle loading and applied at the center of gravity of
the subway vehicle. For double-track structures, LL, I, CF, RF, and LF forces must be considered
as applicable to either one or both tracks, and all foreseeable combinations of these forces, in
terms of application and direction, must be analyzed.

Live loads may include pedestrian loading as well as subway vehicle loads. However, in a tunnel
one-story high, pedestrian live load, subway vehicle live load, impact load, and other dynamic
loads (CF, RE LF) are ordinarily transmitted through the invert slab directly to the supporting
ground and, as a consequence, have little or no effect on the proportioning of the structural
elements of the tunnel. These loads will normally affect the basic structural design only if the
tunnel is two or more stories in height.

Cut-and-cover subway structures must also be designed to support surface traffic loading or other
live loading. Usual practice is to base roadway live loads on AASHTO HS 2044 loading. For
structures below a depth of 8 ft, a uniform live load of 300 psf is commonly used. For structures
having less than 8 ft of earth cover, common practice is to design the roof of the subway structure
for the more severe of the following two conditions:

1. Actual depth of cover (DL) plus superimposed HS20-44 wheel load distribution
(LL) is accordance with AASHTO requirements;
2. An assumed future cover of 8 ft (DL) plus a uniform live load of 300 psf.
AASHTO requirements for impact due to roadway live loads were interpreted by
WMATA as follows:

5
Up to 1-ft, 0-in. earth cover I = 30% LL
1-ft, 1-in. to 2-ft., 0-in. earth cover I = 20% LL
2-ft, 1-in. to 3-ft., 0-in. earth cover I = 10% LL
Greater than 3-ft, 0-in. earth cover I = 0% LL

For the most part, live load, impact, and other dynamic forces imparted to cut-and-cover
vehicular tunnels are similar to the corresponding loads and forces imparted to cut-and-cover
subway structures. These loads and forces, and their application, typically conform to or exceed
the requirements contained in the AASHTO specifications for HS2044 loading.

The roof, walls, invert or base slab, and other elements of cut-and-cover sewer structures must be
designed for all foreseeable net internal pressures that can be imparted to these elements by the
fluid contents of the structure. Net internal pressure must be considered to be internal pressure at
any particular operating condition less minimum reliable external pressure due to retained or
supported earth over the life of the structure. Maximum internal pressure must include transient
pressure, if any. Net internal pressure is then treated as internal live load. (Internal live load is
often zero on the wall and roof elements of the structure.) In addition to internal live load, live
load and impact due to surface traffic loading must also be considered. Loading criteria for the
combination of earth cover plus live surface traffic loading most be considered less for shallow
cut-and-cover sewer structures than it is for otherwise similar subway or vehicular structures. It is
recommended, however, that the external roof load on shallow-roof cut-and-cover sewer tunnels
be not less than the weight of the earth cover (DL) plus AASHTO HS20-44 live load plus
impact.

Horizontal Earth Pressure: Horizontal earth pressure (E) may be considered in this chapter to be
lateral pressure due to both retained soil and retained water in soil when water is present.
Horizontal earth pressure may include the effect of surcharge loading resulting from adjacent
building foundation loading, surface traffic loading, or other surface live loading. All of these
components of “E” must be evaluated both in terms of present conditions and future conditions.
This admonition applies particularly to groundwater levels. Where future changes could
adversely affect the subsurface structure, needed protective measures to mitigate the adverse
effects might not be foreseen and might be extremely costly to add to an existing structure.
The soil component of “E” depends on the physical properties of the original ground through
which the cut is being taken. For design and construction purposes, these physical properties are
typically determined by an experienced geotechnical consultant retained to perform a
comprehensive geotechnical investigation of subsurface conditions. These data, as well as
recommended diagrams showing long-term and short-term horizontal earth pressures, are
normally contained in a site-specific geotechnical report.

[Table 17-1] Table 21-1 is a guide showing how different soils may affect horizontal earth
pressure on the structure. However, any attempt to categorize the behavior of retained soils must
be viewed with caution. Stiff, over-consolidated fissured clays, for example, can impart high
lateral pressures to the walls of the subsurface structure. Many subsurface structures will be in

6
Table 21-1 Effect of Soil Type on Lateral Pressure (Generalized)

Soil Type N Value blows/foot Characteristics Lateral Pressure


Dense sand Greater than 30 Difficult to drive a 2 x Low
4 stake with a sledge
hammer

Loose-to-medium Less than 30 Moderate


Sand
Very stiff clay or silt Greater than 16 Can be indented by a Moderate
thumb nail
Medium-to-stiff clay Less than 16 Can be indented by a Moderate-to-high
thumb
Soft clay Less than 4 High

mixed or stratified soils, which cannot be represented in a guide such as Table21-1. Also, many
of these structures will be under the groundwater table. Below the groundwater table, lateral
pressure due to retained soil is or may be considered to be a function of vertical effective stress in
the soil. As a result, the soil component of “E” may be small compared with the total horizontal
pressure due to both retained soil and retained water. Finally, in the modern practice of
formulating lateral pressure diagrams for use as criteria for analysis and design, recommended
maximum horizontal pressure is often equal to or approaching full vertical stress (i.e., the vertical
effective stress plus water pressure), even for soils normally regarded as competent.

The short-term and long-term changes in horizontal earth pressure must be considered. During
the life of the tunnel there may be substantial changes to this loading. Immediately following
construction, the actual short-term earth pressure may be considerably less than long-term design
pressure. In the event of a future excavation parallel and adjacent to the tunnel, unbalanced
lateral pressures may occur with a pressure equal to long-term pressure applied to one side and a
lesser pressure applied to the other. For these reasons, cut-and-cover tunnels should be designed
for both short-term and long-term loading. There are differing opinions on whether or not the
structure should be considered restrained against horizontal translation or proportioned for
stresses resulting from side sway caused by unbalanced horizontal pressures. The approach taken
may also depend on local requirements. A common recommendation for subway structures is that
the tunnel be proportioned for side sway if it is a single story structure, and in the case of two or
more stories, side sway should be considered in the upper story only, with the lower stories
assumed to be restrained against horizontal translation. These assumptions should provide a
competent factor of safety against a future mishap resulting from adjacent construction, and it
will normally be found in proportioning the structural elements of a cast-in-place concrete box
tunnel that they are not unduly increased in size through consideration of this loading. Where
construction bulkhead walls are incorporated into the permanent structure, the design of the

7
connections between walls and roof or invert slabs is governed by the unbalanced lateral load
condition, and a more detailed geotechnical evaluation is appropriate.

The assumed magnitude of long-term and short-term horizontal earth pressure on subsurface
tunnel structures has varied considerably with public authorities. These values have depended in
part on the type of tunnel, the location of the tunnel, and other local factors, as well as the
physical properties of the soil. However, maximum horizontal earth pressure (soil component)
should never be taken as less than vertical effective stress multiplied by an appropriate at-rest
(Ko) coefficient.

Similarly, the criteria to be used to account for unbalanced horizontal earth pressure have varied.
In the design of the Toronto Subway, where groundwater lay below the tunnel structures, short-
term reduced loading resulting from future adjacent construction was determined by multiplying
the design value of horizontal earth pressure by a reduction factor of 0.5. For the design of the
WMATA subway, unbalanced horizontal earth pressure was defined as long-term horizontal
earth pressure applied to one side of the tunnel opposed by short-term horizontal earth pressure
applied to the other.

Diagrams from which maximum and minimum earth and water pressures were computed, for use
as loading criteria in the design of cut-and-cover subway structures by WMATA, are presented in
Fig. 21-6 [Figure 17-6]. Both long-term and short-term loadings are shown. These generalized
diagrams were applicable to a wide range of noncohesive and cohesive soil types to be found in
the greater Washington, D.C., area. The soil component of maximum horizontal earth pressure
was taken as equal to effective vertical earth pressure, to account for the effect of continuing
vibration of the subway and the outward deflection of the structure.

Experience in or with the design, construction, and performance of vehicular cut-and-cover


tunnels is not comparable with that for rapid transit systems. However, engineering practice with
respect to horizontal earth pressure loading on vehicular cut-and-cover tunnels should clearly be
appropriately similar to that for rapid transit tunnels.

A somewhat less conservative approach is normally taken in evaluating long-term horizontal


earth pressure applicable in the design of cut-and-cover sewer structures. The reinforced concrete
box structures constructed as part of the San Francisco Clean Water Program, for example, were
designed for a long-term horizontal earth pressure equal to at-rest pressure plus hydrostatic
pressure plus lateral pressure due to surface traffic surcharge. For noncohesive soils and other
strong soils, at-rest pressure was computed with the coefficient, Ko, equal to 1 - sin φ, where φ
equals the effective stress friction angle. For saturated, soft to medium clays and silty clays (bay
mud), Ko was determined by laboratory test and analysis to be on the order of 0.55 (when applied
to effective vertical stress). Additional horizontal pressure due to surface traffic surcharge was
taken as a nominal, uniformly applied pressure of 150 psf.

Buoyancy: When the groundwater table lies above the bottom of the invert or base slab of a
subsurface structure, an upward pressure on the bottom of the base slab, equal to the piezometric

8
head at that level, must be accounted for. For a rectangular box, this upward pressure multiplied
by the width of the base slab is the buoyant force (B) per lineal ft of structure. When the reliable
minimum weight of the structure plus the fill above the structure (DL min.) exceeds B by an
adequate factor of safety (FS), the structure is considered stable against uplift due to B. Opinions
differ somewhat on an appropriate value for FS. For the design of the San Francisco Bay Area
Rapid Transit System (BART), an FS  1.1 was considered satisfactory when side friction was
neglected. Side friction between the walls of the structure and the retained soil offers additional
resistance to B, but this resistance, except for its influence on the selection of FS, is normally
neglected.

When B exceeds DL min., other resisting features must be incorporated into the design. Some of
the features that may be provided for this purpose are the following:
• The weight of the structure may be increased by thickening the walls, roof, or base
slab. Also, the base slab may be widened to increase the weight of earth
resistance.

• Tension piles designed to provide a tensile force on the base slab can be provided.
Both steel piles and prestressed concrete piles have been used in this application.

• Tie-down anchors, resembling permanent tie-back anchors, can be provided.


Drilling for the anchors is accomplished at some convenient time after the base
slab is placed. The anchor heads are located in formed recesses in the base slab.
After completion of the tie-down installation, the recess is filled with concrete.
The type of anchor used will depend in part on whether the anchor can be founded
in bedrock beneath the structure or in competent soil.

Flood (FL). Where there is a potential for river floods or other flooding that could add loads to
subsurface structures, the design of the structures should allow for this loading as required by the
particular type of structure and the conditions affecting each location.

Shrinkage and Thermal Forces. Between transverse joints in cut-and-cover tunnel structures
constructed of reinforced concrete, shrinkage forces (S) and thermal forces (T) are accounted for
by the longitudinal reinforcement in the walls, roof, and invert slab. The stresses produced by
these forces are typically normal to the principal stresses caused by DL, LL, and I and, therefore,
do not enter into frame analysis of the structure.

Earthquake Forces. Major codes that address the seismic design of surface structures in the
United States contain no provisions for underground structures. The general view is that
underground structures are much less affected by seismic motion than are surface structures.
Although this view is substantiated by limited observations of the performance of underground
structures during seismic events, some severe damage has been reported. In areas identified as
subject to significant seismic activity, it is therefore necessary to determine the extent to which
earthquake forces (EQ) should be considered in the design. In making this determination, the
importance of the structure, the consequences of damage due EQ, the type of soil in which the

9
structure is founded and the potential for liquefaction, and the location of potentially active faults
at or in the vicinity of the site should all be considered.

The geotechnical data needed to assess earthquake forces and risk should be furnished by a
geotechnical consultant experienced in this field. In addition to these data, the geotechnical
consultant should, where appropriate, develop design values for the magnitude and distribution
of EQ. Most such evaluations of EQ have resulted in an assumed lateral pressure to be
superimposed on the horizontal earth pressure, E, on one side of the structure. Thus, a
representation of both the increase in lateral pressure due to retained soil and the racking effect
on the structure during earthquakes is provided for design purposes. Where evaluations of EQ
have been made, it has to date been common to find that the resulting loading condition is less
severe than other load cases without EQ. The effect of EQ on the proportioning of the structure
can also often be considered negligible because of the stiffness of the soils in which the structure
is founded, and because of increased allowable stresses in the elements making up the structure
when EQ is added to other loads.

Load Combinations and Unit Stresses. Except for earthquake forces, permitting allowable
stresses to exceed basic unit stresses for particular combinations or groupings of the loadings
identified in this lesson does not apply to the design of cut-and-cover tunnel structures. When EQ
is omitted from consideration, any combination of these loads that can exist simultaneously
should not produce stresses exceeding basic unit stress. Basic unit stress is considered to be
allowable unit stress in each of the structural materials comprising the structure, as specified in
applicable codes.

When earthquake forces are being considered, an increase in allowable stress is permitted. Usual
practice for reinforced concrete design, using service-load methodology, is to permit 133% of
basic unit stress for the load combination of DL + E + B + FL + EQ. Using load factor
methodology, the comparable increase in stress is represented by the reduced design strength
permitted in applicable codes.

Loading Cases

Cut-and-cover tunnel structures restrain earth but are not free to yield significantly. Nevertheless,
there is a need to consider unbalanced and other atypical loading on these structures. The
particular load cases to be analyzed will depend on the type of structure, its location, the type of
ground in which the structure is founded, the location of the groundwater table, and other local
factors. All reasonably foreseeable temporary and permanent loading cases that would affect the
design of the structure should be investigated. The system design criteria developed by WMATA
for reinforced box and station sections specify that, as a minimum, the following three basic
loading cases be investigated:

Case I: Full vertical and long-term horizontal load, as recommended by the General Soils
Consultant.

10
Case II: Full vertical load, long-term horizontal load on one side, and short-term
horizontal load on the other side, as recommended by the General Soils Consultant. In
underground box structures potentially subject to unequal lateral pressures, the structural
analysis shall consider the top slab as both restrained and unrestrained against horizontal
translation in arriving at maximum shears, thrusts, and moments.

Case III: Full vertical load with short-term horizontal load neglecting groundwater
pressure on both sides, as recommended by the General Soils Consultant.

The system design criteria adopted by the Los Angeles MTA is the same as above except
tthat one additional case is added:

Case IV: Only dead vertical load with long-term horizontal load including hydrostatic
pressures.

Both WMATA and MTA specify that for stress analysis, variations in the elastic support of the
subgrade shall be considered for the different loading cases as appropriate.

A similar approach to the establishment of loading cases as criteria for design should be taken in
designing cut-and-cover vehicular tunnels, and, ordinarily to a less severe extent, sewer tunnels.
For the design of large reinforced concrete box structures constructed as part of the San Francisco
Clean Water Program, the following loading cases, with the box empty, were established:

Case I: Vertical DL + LL
Horizontal, both sides E with traffic surcharge
Case II: Vertical DL + LL
Horizontal E without traffic surcharge
Case III: Vertical DL only
Horizontal E with traffic surcharge
Case IV: Vertical DL + LL
Horizontal No load

In the above sewer tunnel case, LL was assumed to be AASHTO HS20-44 traffic loading, DL was
assumed to be the weight of the fill above the structure plus the weight of the structure as
applicable, and horizontal forces were as previously given herein. Also, with the box full of
wastewater, the hydraulic grade line lay below the permanent groundwater level, so that internal
live load essentially (lid not affect the structural design.

Frame Analysis

In the analysis of the structural frame of the tunnel, loads and pressures representing each loading
case are applied, and the shears, thrusts, and bending moments for each element of the frame are
(in most cases) determined through rigid frame analysis using commonly accepted methodology.

11
In modern practice this methodology is contained in specific computer programs for structural
analysis. Except for particularly wide invert spans, usual practice is to assume that vertical
reactions are uniformly distributed over the bottom of the invert slab. This assumption results in
maximum slab bending moments and is therefore conservative. Fig. 21-7 [Figure 17-7] contains
illustrations of the three cited WMATA loading cases applied to a typical WMATA reinforced
concrete subway line structure and the configuration of each resulting bending moment diagram.

Frame analysis of reinforced concrete is complicated by its nonlinear nature, and more importantly
by the fact that the cracked moment of inertia (Ic) is typically much less than the gross moment of
inertia (Ig). If strains are not a concern, it is common to use gross values of El, since only relative
internal reactions (forces and moments) are desired.

Foundations

For almost all cut-and-cover tunnel structures, the vertical pressure imparted to the subgrade
beneath the structure is less than the in situ total vertical stress in the soil at that level before
construction. Further, most of these structures are founded in competent soils with safe bearing
capacity considerably exceeding applied vertical pressure. For these reasons, pile support for cut-
and-cover tunnels is an uncommon requirement.

There are, however, occasions when the geotechnical consultant will recommend that the
subsurface structure be supported on piles. Such cases can occur when the structure is founded in
weak soils such as soft clay or silt and it is determined that the behavior of the supporting soil
cannot be reliably predicted, that predicted differential settlement would adversely affect the
structure, or that there is not adequate bearing capacity. The requirement for pile support under
any of these conditions is more common for cut-and-cover sewer tunnels.

Reinforced Concrete Design

In the modern practice of reinforced concrete design, the predominant code defining design
methodology is the ACI Manual of Concrete Practice, issued by the American Concrete Institute
(ACI). Other major codes or specifications for design are similar in format and often identical to
the ACI code. There are, however, significant differences between ACI requirements and
AASHTO requirements, for example, with respect to the level of stress actually permitted in
reinforced concrete.

The principal design methodology addressed in the ACI code is the Strength Design Method or
Load Factor Design Method. Although the Strength Design Method is the predominant design
method in use today for reinforced concrete structures, this methodology is not universally
employed in the design of cut-and-cover tunnel structures. The much older Service Load Design
Method (commonly referred to as the Working Stress Method) is still employed by many public
authorities.

12
Modem practice in the design of reinforced concrete for cut-and-cover rapid transit structures is
probably best characterized by the system design criteria specified by Los Angeles MTA. MTA
directs that for all underground rapid transit cast-in-place reinforced concrete structures, including
box lines and stations, the design shall be by the Strength Design Method of ACI-318, or Service
Load Design Method of AASHTO Specifications, as applicable. For cut-and-cover vehicular
tunnels, current practice may be characterized similarly, except that both the Strength Design
Method and the Service Load Design Method are carried out according to AASHTO
Specifications.

Most cut-and-cover sewer tunnels are combined sewers carrying both sewage and stormwater, and
they may carry some industrial wastes as well. Such structures are normally designated
Environmental Engineering Concrete Structures. Reinforced concrete for these structures should
conform the requirements of ACI 350 R as well as the applicable requirements of ACI 318. Both
the Strength Design Method and the Service Load Design Method (ACI Alternate Design
Method) are used in the current practice of the design of these structures.

Reinforced concrete design for cut-and-cover tunnel structures must also conform to all local and
other mandated codes, except when particular provisions of such codes are clearly, or can be
shown to be, not applicable. In general, the codes and specifications cited above will contain the
same or more severe provisions than other, mandated codes. Where earthquake forces are a factor,
the structure should be designed for a desired degree of ductility and toughness as well. To
incorporate these provisions into the reinforced concrete design, authoritative and pertinent
literature on the subject of seismic design should be consulted.

The cited codes and specifications may be regarded as definitive for the proportioning of the
subsurface structure for resistance to principal shears, thrusts, and bending moments. Engineering
practice with respect to providing shrinkage and temperature reinforcement, however, is less well
defined. Cut-and-cover structures are normally subjected to a much smaller range of ambient
temperature than are surface structures. Also, these structures are commonly more massive than
are the surface structures for which the cited codes are applicable. In the opposite vein, and
particularly when these structures lie below the water table, it is almost always desirable that they
be designed so that only minimal cracking is to be expected. These factors are typically taken into
account in formulating criteria for shrinkage and temperature reinforcement. For the design of
rapid transit walls and roof slabs, with transverse joints about 50 ft apart, it is common to provide
temperature and shrinkage reinforcement, on both faces of the wall or slab, in the amount of 80-
100% of normal ACI 318 -7.12 requirements, up to a specified maximum. (No. 7 at 12 in. at each
face is a common specified maximum.) Treatment of invert slabs has been similar to that for walls
and roof slabs. In some cases subgrade drag may need to be investigated.

Minimum requirements for shrinkage and temperature reinforcement, as specified by AASHTO,


have not been considered applicable for cut-and-cover highway tunnels. For these structures,
engineering practice in this regard has been similar to that for rapid transit tunnels.

13
In the design of cut-and-cover sewer tunnels, both infiltration and exfiltration (leakage) are
concerns. ACI 350R offers definitive criteria and discussion applicable to the design of shrinkage
and temperature reinforcement for these structures. The text of ACI 350 R includes an evaluation
of the effectiveness of shrinkage-compensating concrete in reducing the required amount of
shrinkage and temperature reinforcement.

Joints

The amount of shrinkage and temperature reinforcement that should be provided is also
considered to be a function of the distance between transverse joints that will dissipate shrinkage
and temperature stresses in the direction of the reinforcement. Engineering practice has differed
with respect to both the spacing and type of transverse joint that may be used successfully Fig. 21-
8 [Figure 17-8]. In the United States, typical spacing of transverse joints has ranged from 30 to 65
ft, but more commonly is in the range of 35-50 ft. The WMATA cut-and-cover structures contain
transverse contraction joints spaced 50 ft (maximum) apart, with occasional expansion joints as
well. For many other cut-and-cover tunnels there has been more recognition of the relatively small
range of ambient temperature to which these structures are subjected by providing transverse
construction joints in the design. (A construction joint is generally defined here as a joint through
which reinforcement passes.) Cut-and-cover tunnels (line structures and stations) on the BART
system were provided which transverse construction joints spaced typically 35 ft apart, with
occasional expansion joints. This or similar practice has been common for more recently
constructed cut-and-cover tunnels. Some examples are the following:
• Los Angeles MTA Transverse construction joint spacing was typically 45 ft.

• Central Artery, Boston. Transverse construction joint spacing was typically 40 ft.

• San Francisco Clean Water Program. Transverse construction joints were specified
to be 30 ft apart. Increased spacing was allowed up to 60 ft with a specified
increase in longitudinal reinforcement. Longitudinal reinforcement specified was
0.3% for 30-ft joint spacing. For each additional 10-ft increment of spacing of
joints, an additional 0.1 % longitudinal reinforcement was specified. Actual
spacing was usually 50 ft.

Horizontal construction joints are normally specified to be located at the underside of roof slabs,
floors, beams, or girders and at the top of invert slabs. Waterstops are usually required at all
transverse joints and sometimes at horizontal joints as well, (See discussions below on waterstops,
joint treatment, and watertightness.)

SHORING SYSTEMS

Neat-line excavation in soils requires that temporary walls (commonly referred to as "sheeting" or
"shoring walls") be in place before significant cut-and-cover excavation commences. As the
excavation progresses, the shoring walls must be supported so that they retain the vertical faces of
the excavation. The function of the supported shoring walls (shoring system) is usually to prevent

14
detrimental settlement of the around, utilities, and buildings at the side of the cut as
well. The design of the support system will depend on many factors, including the following:
• The physical properties of the soil throughout and beneath the cut.
• The position of the groundwater table during construction. The width and depth of
the excavation.
• The configuration of the subsurface structure to be constructed within the cut.
• The size foundation design and proximity of adjacent buildings.
• The number, size, and type of utilities crossing the proposed excavation. Also, the
presence of utilities adjacent to the excavation.
• Requirements for street decking across the excavation.
• Traffic and construction equipment surcharge adjacent to the excavation
• Noise restrictions in urban areas.

The depth of excavation required for the construction of subsurface structures discussed in this
lesson is rarely less than 25 ft. For the sake of clarity and reference in this lesson, it is arbitrarily
assumed that "cut-and-cover excavation" exceeds 25 ft in depth. Additionally, cases where
bedrock lies above the bottom of the excavation are not addressed in this here.

COMMON TYPES OF SHORING WALLS

Cross sections through each of the common types of shoring walls discussed in this chapter are
illustrated on Fig. 21-9, [Figure 17-9]. Soldier pile and lagging walls and steel sheet piling walls
are often classified as "flexible" walls. In the same context, continuous concrete diaphragm walls,
which are ordinarily much stiffer, are classified as "rigid" or "semirigid" walls, depending upon
actual stiffness.

Soldier Piles and Lagging

A very common type of shoring wall is soldier piles and lagging. The soldier piles are usually
steel wide flange (WF) or bearing pile (HP) shapes installed prior to the excavation by driving or
drilling. Spacing of soldier piles can range from 3 ft or less to 10 ft or more; 5-8 ft spacing is most
common. The soil between the soldier piles is retained with horizontal wood lagging placed
between the soldier piles as the excavation progresses Fig. 21-9a,[ Figures 17-9a] and Fig. 21-10,
[Figure 17-10]. The wood lagging is usually placed behind the inside flanges of the soldier piles
but is occasionally placed in front of the soldier piles ("contact lagging") using proprietary
hardware or welded, threaded studs with retaining plates to secure the lagging to the soldier piles.
During excavation, the depth of exposure below the last-placed lagging may be as little as I ft, as
in the case of say saturated silt (see below), or as much as 4 or 5 ft in competent, cohesive or
semicohesive soils. The restriction in depth of unsupported cut is the height of the cut that will
remain stable until the lagging can be placed, but in soils it should not be more than 5 ft.

When soldier piles are placed in drilled holes, the portion of the drilled hole below excavated
subgrade is sometimes backfilled with 3,000 psi (or more) concrete, either to provide for needed

15
bearing capacity when the soldier pile is axially loaded or to improve available passive resistance
below subgrade. More commonly, however, this portion of the drilled hole is backfilled with a
lean concrete. Above excavated subgrade, good construction practice can range from filling the
hole with lean concrete to filling it with a soil-cement mix, depending on the importance of
avoiding loss of ground when excavating the soil face to fit the lagging boards in place.

Soldier pile and lagging walls are particularly advantageous when many underground utilities will
cross over the excavation. The utilities are exposed prior to installation of the soldier piles, and in
most cases the soldier piles can be placed so that they straddle the utilities. After the soldier piles
are installed, and during the first step in the cut-and cover excavation, the utilities are either
independently supported across the excavation, or they are suspended from street decking that
spans the excavation.

When the cut is in saturated, pervious, or semipervious soils; or when there are zones of saturated
soils within the cut, soldier piles and lagging are rarely feasible unless the groundwater table is
lowered. Soldier piles and lagging have been used in some cases where the groundwater table lies
a few feet above the excavated subgrade in semipervious soils, but this practice is not
recommended, except in some otherwise very difficult circumstances.

When the groundwater table lies at or has been lowered to a level below excavated subgrade,
soldier piles and lagging can be used more or less routinely in a wide variety of soils. Soldier piles
and lagging are also used when the cut lies in weak or decomposed rock not capable of supporting
itself or of being supported with modern shotcrete and rock bolt techniques. Soldier piles and
lagging are not ordinarily considered suitable for cuts in very soft to soft clays, saturated silts,
loose silty sands, or in any soil that is potentially unstable during excavation.

Steel Sheet Piling

Continuous steel sheet pile walls are composed of rolled Z-shaped or arch-shaped interlocking
steel sections. Because of their greater stiffness and resistance to bending, Z-shaped sections are
almost exclusively used as steel sheet piling for cut-and-cover construction Fig. 21-9b, [Figure
17-9b].
Interlocking steel sheet piling is typically used in saturated pervious or semipervious soils and
other loose or weak soils that do not permit the easy placing of lagging. When the groundwater
table cannot be lowered, interlocking steel sheet piling is normally adequately effective in cutting
off concentrated flow through pervious ground both above and below the excavated subgrade.
Interlocking steel sheet piling is used as well in competent sandy soils when groundwater is not a
concern, if there are sufficiently few utility crossings and other subsurface obstacles and it is
found economically advantageous to do so. The sheet piles are driven with either impact-type or
vibratory-type hammers, depending in part on the type of soil. Vibratory hammers are more
commonly used in sandy soils. In modern pile driving practice, with the assistance of pre-auguring
techniques, sheet piles can be driven with a vibratory hammer in dense sand to a depth of 65 ft or
more with relatively good driving production. For the San Francisco Clean Water

16
Program, in soft to medium clays, sheet piles up to 90 ft long were driven with a vibratory
hammer.

Diaphragm Walls

In shoring system design and construction, the term diaphragm wall refers to continuous shoring
walls that are reinforced concrete, a combination of concrete and structural steel. or similar
systems. The walls are constructed from the ground surface and are ordinarily designed so that
they are for construction purposes, watertight. The more common diaphragm walls, and their
applications, are discussed briefly below.

Slurry Walls: Although the term slurry wall sometimes has a broader meaning, it usually refers to
a reinforced concrete wall [as used in cut-and-fill construction] placed in a deep trench usually 2-3
ft wide. The wall is constructed in increments, or panels. Panel lengths have ranged from 7 to 20
ft, but in cut-and-cover construction they are usually 10-15 ft long. The panel is excavated with a
special clamshell-type bucket. The sides of the panel excavation are stabilized by filling the panel
with a bentonite slurry and maintaining the level of the slurry at or near the ground surface
throughout the excavation. Upon completion of the panel excavation, a preassembled steel
reinforcing "cage" is lowered into the slurry-filled panel. Concrete is then placed in the panel by
tremie techniques, displacing the slurry.

It is important (almost always) in slurry wall construction that the joints between panels be
watertight. The most common type of joint is formed with a circular end pipe. Fig. 21-9c, [Figure
17-9c] illustrates the joint configuration formed by this method. The end pipe is a steel tube
inserted at one end of the excavated panel as a stop for tremie concrete. Some time after the start
of the tremie concrete pour, the end pipe is rotated to break bond; it is subsequently slowly
extracted to produce a formed, semicircular joint, which can be cleaned when the next panel is
excavated. The procedure shown in Fig.21-9c indicates a continuous operation in which one panel
at a time is constructed, setting the end pipe at the leading edge. Another procedure is to construct
alternate"primary" panels, setting end pipes at both ends. The resulting "secondary" panels
between primary panels are then constructed.

There are many variations to the more common procedures described [above]. Polymer drilling
fluid has been used in lieu of bentonite slurry on several recently constructed slurry wall projects.
More sophisticated types of slurry wall joints have been used to improve watertightness at the
joints.

Slurry walls can be constructed in soil to depths exceeding 180 ft. For cut-and-cover construction,
slurry walls deeper than about 100 ft are not common, however. Occasionally it is necessary to
key the bottom of the slurry wall into rock. Shallow keys in soft to medium-hard rock have been
successfully excavated using percussion tools developed for this purpose. Keys into medium-hard
to hard rock have been difficult to achieve.

17
Soldier Pile and Tremie Concrete Walls. Soldier pile and tremie concrete (SPTC) walls are
composed of soldier piles spaced at relatively close centers with a good-quality concrete placed
between the soldier piles, thus forming a very stiff continuous wall. Soldier piles are typically 24-
36 in. deep, rolled beams or deeper, built-up sections. Fig. 21-9d, [Figure 17-9d] illustrates the
principal steps in the construction of an SPTC wall:
1. Alternate soldier piles are placed in predrilled holes. Bentonite slurry is used for
hole stabilization if required. Each hole is then filled with sand or other weak
backfill.

2. The slot between alternate soldier piles is then excavated with a specially designed
clamshell tool. Bentonite slurry is used to stabilize the sides of the excavated slot
in a similar manner to that for slurry walls.

3. The intermediate soldier pile is lowered into the slot. Unreinforced concrete is then
placed in the slurry-filled slot on both sides of the intermediate pile
simultaneously, using the tremie method.

Soldier pile spacing for SPTC walls typically ranges from 4 to 6 ft, with the upper limit being held
to about twice the nominal wall thickness. Apart from the great strength that can be achieved
using these walls, there is the added advantage that the soldier pile element of the wall can be
extended deeper than the tremie concrete element. Thus. the soldier piles can be extended below
the tremie concrete into very strong soil or into bedrock when there is a structural reason to do so.

Drilled Pile Walls. In this lesson, "drilled pile walls" refers to walls formed by abutting cast-in-
place concrete or reinforced concrete piles, concrete and soldier beams placed in drilled holes, or
combinations of these concepts. Drilled holes range in diameter from about 2 to 4 ft. Depending
on soil and groundwater conditions, the excavation can be made with or without casing, either in
dry or slurry-stabilized holes. Cast-in-place concrete or reinforced concrete piles placed in a single
line or row, tangent, nearly tangent, or slightly overlapping with each other, have been called
"contiguous," "secant," or (sometimes) "tangent" pile walls. However, in recent years the
configuration shown on Fig. 21-9e, [Figure 17-9e] has been utilized on several cut-and-cover
projects and is probably more common. With this configuration, considerably more strength can
be built into the drilled pile wall. Drilled pile walls are especially suitable when the soil-rock
contact lies near or above the cut-and-cover subgrade and a diaphragm wall is either required or
most suitable above the rock line. This type of wall (or a wall of similar configuration) can be
keyed into rock or constructed as a wall to support weak rock, using commercially available
drilling equipment.

Soil-Cement Mixing Walls [A Form of Jet Grouting]. Soil-cement mixing walls are composed of
soldier piles placed in overlapping columns of in situ soil that has been mixed with a measured
amount of cement or cement and bentonite. The soil-cement columns are constructed with special
proprietary equipment, which employs a bank (usually three) of hollow-shaft augers. The augers
penetrate the soil to the depth of the wall and then are withdrawn, mixing the soil with cement
(and bentonite if needed) in the process, thus creating the overlapping circular columns of soil-

18
cement. Before the soil-cement has hardened, soldier piles are pushed into the alternate (usually)
soil-cement columns. The augers are normally about 24-36 in. in diameter. Soldier piles range
from W18 to W30 spaced approximately 3 ft on center (with 24-in. columns) to 4 ft on center
(with 36-in. columns). The unconfined compressive strength of the soil-cement ranges front about
75 psi in soft clays and saturated silts to about 200 psi in stronger sandy soils. Soil-cement mixing
walls may be constructed to any reasonable depth required for cut-and-cover construction. Wall
depths up to 180 ft have been reported in other applications Fig. 21-9f [see Figure 17-9f].

Applications for Diaphragm Walls. Diaphragm walls are used generally where it is required that
the shoring wall be watertight and where, at the same time, more wall stiffness or resistance to
bending is needed than can be provided by heavy steel sheet pile sections. Diaphragm walls have
been constructed in virtually all soil types, but usually in very soft to medium clays, saturated silts,
or saturated, loose silty or clayey sand. They are usually constructed where surface settlement
adjacent to the cut must be minimized. The stiffer of the diaphragm walls have been specified in
many urban cut-and-cover projects to obviate the need for underpinning adjacent buildings.

The construction of diaphragm walls causes much less noise and vibration than does the driving
of sheet piles, and a diaphragm wall may be chosen over a sheet pile wall in some cases on this
account alone. Where a diaphragm wall is to be used, the choice of the type will depend primarily
on the required stiffness and resistance to bending and shear, actual subsurface conditions, and
cost.

Occasionally, slurry walls or SPTC walls are utilized both as shoring walls and as the wall
element of the permanent structure to be constructed within the cut. SPTC walls were utilized in
this capacity in the design of several BART subway stations.

COMMON TYPES OF SHORING WALL SUPPORT

Internal Bracing

Most cut-and-cover excavations are relatively narrow and, as a result, internal bracing composed
of multiple tiers of horizontal, structural steel framing is the most common type of shoring wall
support used. In a typical excavation, the principal components of each internal bracing tier are
longitudinal beams, or “wales," and transverse compression members, or “struts," arranged
generally as shown on Fig. 21-10, [Figures 17- 10] and Fig. 21-11, [Figure 17-11].

The bracing tiers must be positioned so that they support the shoring wall and permit efficient
construction of the permanent structure .Fig. 21-12, [Figure 17-12] shows the general sequence of
construction operations typically employed during the construction of a subway line structure. The
shoring walls are shown as constructed at the structure neat line (except for wall tolerance
allowed), as is usually the case. During the excavation stage, vertical spacing of bracing tiers is
often specified to be a maximum of 15-16 ft, sometimes 12 ft when it is crucial to minimize
adjacent ground settlement. The maximum depth of cut in any excavation step is usually specified

19
to be 3 ft below the centerline of the next bracing tier to be installed (dimension Z, Fig.21-12).
The amount of settlement of the ground adjacent to the cut is generally considered to be largely a
function of vertical spacing of bracing tiers and shoring wall stiffness. It is sometimes better to
increase wall stiffness to permit larger vertical spacing of bracing tiers, when the larger spacing is
needed to avoid interference of the bracing with the reinforced concrete construction.

Occasionally, it is not feasible to avoid such interference, and different removal techniques or
supplementary bracing methods are required. During the bracing removal stage, the shoring wall
does not depend upon the soil below subgrade for support, and larger vertical shoring wall spans
can ordinarily be permitted.

Struts in the internal bracing framing need to be spaced far enough apart so that excavating
equipment can operate efficiently. Strut spacing is usually in the range of 10-15 ft, but larger
spacing (up to 25 ft) is sometimes used to permit more clearance for construction activities.
However, such large spacing is often very costly because of the much heavier wales that result,
and it can be undesirable as well because of the inward wall movement that accompanies the
increased wale deflection. Where there is no axial load in wales it is usually more economical to
make the wales two to three strut spaces long and discontinuous at the wale ends.Fig. 21-11
shows two framing plans representative of common internal bracing framing.

Internal bracing is compatible with any of the shoring wall types already discussed. Horizontal
force from the shoring wall is transferred to the wale at each soldier pile, at each sheet pile web, or
in the case of slurry walls, at heavy steel bearing plates embedded in the slurry wall at its inside
face. (The bearing plates are fastened to the reinforcing cage when the Slurry wall is constructed.)
The shoring, walls cannot be placed with sufficient accuracy to permit the wales to bear directly
on the soldier piles. sheet piles, of slurry wall bearing plates. The gap between these wall elements
and the wale is typically filled with a structural "packing"[blocking] (see Fig. 21-11 ). The wales
are supported by structural steel brackets ("lookouts") mounted on the soldier piles, sheet piles, or
slurry wall bearing plates.

Many framing concepts different from the typical framing shown in Fig. 21-11 are employed.
Irregular framing is usually required for irregularly shaped cut-and-cover excavation. Secondary
framing may be required to brace the weak axis of struts in the wider excavations. When slurry
wall panels can be constructed so that they are installed symmetrically about the longitudinal
centerline of the cut-and-cover excavation, slurry wall panels are sometimes braced directly by
struts, thus eliminating the need for wales.

Tie-Backs

A tie-back is a form of support in which the horizontal earth pressure (E) acting on the shoring
wall is resisted by an anchor assembly, which in turn deposits its load into soil or rock far enough
behind the wall to have no significant effect on the wall. Tie-backs anchored in soil are commonly
referred to as "soil anchors"; tie-backs anchored in rock are "rock anchors." A tie-back consists of

20
three principal elements (see Fig. 21-13, [Figure 17-13]):

1. An anchor zone, which acts as a reaction to horizontal earth pressure (E) on the
shoring wall.

2. A tie element, which transfers the load from the wall to the anchor zone.

3. A wall reaction assembly at the point of wall support.

To locate the anchor zone safely behind the shoring wall, the anchor zone is typically placed
behind an "assumed failure plane." Theoretical establishment of this plane so that it is
representative of a safe installation requires complex analysis. In most cases, the assumed failure
plane is established by an experienced geotechnical consultant, who in turn relies on both
experience from past performance of tieback installations and theoretical considerations. Fig. 21-
13 shows the range in which the assumed failure plane usually falls.

There are many types of soil and rock anchors. The tie element for all of the more commonly used
tie-backs is either a single threadbar 1-1-3/8 in. in diameter (usually high strength, as
manufactured by Dywidag), or multiple high strength strands, each (in most cases) 0.6 in. in
diameter. The tie elements are installed in holes drilled from inside the excavation through the
shoring wall at an inclination usually in the range of 15-30 from horizontal for soil anchors, and
up to (commonly) 45 for rock anchors. The anchor zone is created by filling the drilled hole
throughout the anchor zone with sand-cement grout or neat cement grout, depending upon the
type of anchor. The zone between the anchor zone and the shoring wall is commonly referred to as
the “unbonded zone" or "unbonded length." Throughout this length the tie element is covered
with a plastic tube so that none of the tie-back load is transferred into the ground. (Other
techniques are also used to create the unbonded length.)

Drilled holes for the most common tie-backs range from 8 to 18 in. in diameter; about 12 in. is
typical. Grout is placed in the drilled hole by gravity or at modest pressure (about 50 psi) when
hollow-stem augers are used to drill the hole. The anchor resistance is developed through grout-to-
soil friction. For these "conventional" tie-backs, working capacities in competent soils ranging
from 70 to 140 kips are common.

More sophisticated tie-backs of much higher capacity are also utilized on cut-and-cover projects.
The type commonly referred to as "regroutable" employs equipment, hardware, other materials
and methods that permit the injection of grout within the anchor zone at very high pressure (up to
300 psi or more) and in multiple applications. These regroutable tie-backs require a relatively
small drilled hole (usually about 5 in.). In competent soils, tie-back capacities up to 400 kips or
more have been installed. Regroutable tie-backs have also been installed successfully in medium
clays (Su = 500-1000 psf) with a working capacity up to 150 kips.
Tie-backs are also compatible with any of the shoring wall types discussed. When the shoring wall
is soldier and lagging, sheet piles, or any of the diaphragm walls that use soldier piles, usual
practice is to mount the anchor head assembly on double-channel wales, which in turn are

21
mounted on and react against the soldier piles or sheet piles. When the shoring wall is a slurry
wall, the anchor head assembly typically reacts directly on the slurry wall concrete (see Fig. 21-
14,[Figures 17-14] and Fig. 21-15, [Figure 17-15].

When tie-backs are utilized as support for shoring walls, excavation proceeds in lifts that
correspond to the vertical spacing of the tie-backs. Each succeeding increment of excavation
cannot commence until the tie-back at that lift has been successfully post-tensioned to its design
working load at the wall reaction assembly. Normally, the tie-backs cannot be post-tensioned until
5 days after completion of the, grouting of the anchor zone.

Tie-backs can be considered an alternative to internal bracing when the following conditions
exist:
• There is ample width within the excavation for tie-back installation.

• Permission is granted by the property owner to install tiebacks in the ground


adjacent to the cut.

• There is no significant piezometric head behind the shoring wall at the level of the
tie-back installation. (To date, attempts to install tie-backs when there is
piezometric head exceeding a few feet behind the shoring wall have had limited
success.)

• The soil behind the shoring wall is sufficiently competent to permit successful tie-
back installation.

• There are no subsurface obstacles such as deep basements beneath adjacent


buildings.

Although all these conditions are often present, it is uncommon to find that tie-backs are an
economical alternative when the excavation is less than about 65 ft wide.

Soil Nailing

In some of the shallower cuts, installation of a shoring wall prior to the excavation can be avoided
by utilizing the soil nailing method. Soil nailing is an in situ reinforcing technique that consists of
installing passive inclusions (nails) into the undisturbed natural soil mass to retain excavation.
The inclusions are usually steel reinforcement bars that either are placed in drilled bore holes and
grouted along their entire length or (sometimes) are driven into place, To provide local stability
between the nails, an outside facing is provided. The facing generally consists of 3-6 in. of
reinforced shotcrete. Nailing differs from tie-back support: the nails are passive elements that are
not post-tensioned. In addition. the nail density in the retained soil is much larger and the nails are

shorter than for tie-back systems. Fig. 21-16, [Figure 17-16] illustrates the process of constructing
soil nail support.

22
The soil nailing technique is best suited to dry or moist noncohesive or semicohesive soils that
will stand vertically for short durations during construction. In heterogeneous soils, with cobbles,
boulders, and weathered rock zones, this method offers the advantage of small diameter, shorter
drill holes. In cohesive soils subject to creep, however, soil nailing is usually either not feasible or
is uneconomical, even at relatively low stress levels. The sequence of soil nailing is shown in Fig.
21-17.

In North America, the retained depth of excavation using the system has generally been less than
30 ft. However, excavations as deep as 60 ft have been supported.

The mobilization of soil reinforcement interaction requires a relative displacement of soil and
reinforcement. Therefore, in urban sites, the use of this technique can be limited by requirements
for minimal settlement and movement of the ground adjacent to the excavation.

Several examples of Cut and Cover excavations are shown in Fig. 21-18 through Fig. 21-23.

Reading Assignment:

Excavation Support by Anchorage Systems: Five pages, P1-1 through P1-5.

23

Você também pode gostar