juran1991

Upload: leo-schloss

Post on 08-Jul-2018

216 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/19/2019 juran1991

    1/38

    26

    GROUND ANCHORS AND SOIL NAILS

    IN

    RETAINING STRUCTURES

    ILAN

    JURAN,

    D.Sc.

    Professor and Head

    VICTOR ELIAS, P.E.

    Department

    of

    Civil and Environmental Engineering

    Brooklyn Polytechnic University

    V. Elias

    &

    Associates, P.A.

    Consulting Engineers

    26.1 INTRODUCTION

    Ground anchor and soil nail retaining systems are designed to

    stabilize and support natural

    and

    engineered structures

    and

    to

    restrain their movement using tension-resisting elements. The

    basic design concept consists of transferring the resisting tensile

    forces generated in the inclusions into the ground through the

    friction

    (or

    adhesion) mobilized

    at

    the interfaces. These systems

    allow the engineer to efficiently use the in-situ ground in

    providing vertical or lateral structural support. They present

    significant technical advantages over conventional rigid gravity

    retaining walls or external bracing systems

    that

    result in

    substantial cost savings and reduced construction period.

    Therefore, during the past few decades, ground anchors, and

    more recently soil nails, have been increasingly used in civil

    engineering projects.

    The use

    of

    these systems in permanent structures requires

    careful evaluation of the durability of the structural elements

    and

    assessment

    of

    the long-term system performance. A variety

    of inclusions, corrosion-protection systems, and installation

    techniques have been progressively developed by specialty

    contractors. This chapter briefly describes the construction

    process and the main structural elements. I t presents the main

    aspects

    of ground-inclusion interaction, illustrates the observed

    behavior of instrumented str\lctures, and outlines durability

    considerations, performance criteria, and design approaches

    that have been developed to ensure the internal and external

    stability of these composite retaining systems.

    26.2 PRINCIPLES, HISTORICAL DEVELOPMENT,

    AND

    FIELDS OF

    APPLICATION

    26.2.1 Permanent

    Ground Anchors

    Permanent ground anchors are prestressed cement-grouted

    tendons used in soils or rock to restrain

    and

    control the

    displacements

    of

    structural elements such as walls or slabs.

    They have been developed mainly by specialty contractors

    involved in temporary excavation

    support

    systems and in some

    cases are p roprieta ry. The anchors a re installed in drilled holes

    and prestressed

    to

    the design load in order to mobilize and

    transfer the required resisting force from the ground to the

    868

    structural element. Temporary ground anchors are used for a

    specified construction period

    and

    their service life is generally

    less than 2 years. Permanent ground anchors are corrosion

    protected to insure their long-term performance throughout the

    design service life of the structure.

    Figure 26 .1 shows a schematic diagram of a permanent

    ground a nchor. The basic components of the ground

    anchor

    are:

    • The tendon is made of prestressing steel wires, strands, or

    bars and includes:

    a. The anchor bond length-where the tendon

    is

    fixed in the

    primary grout bulb

    and

    transfers the tension force

    to

    the

    surrounding ground. The

    anchor

    bond length is designed

    to

    provide the required load pull-out capacity of the

    anchor.

    b.

    The unbonded length-where the tendon

    is

    free to elongate

    elastically

    tran

    sferring the resisting force from the anchor

    bond length to the structural element (i.e., wall face, slab,

    etc.). I t s designed

    to

    reach the underlying substratum or, in

    Fig. 26.1 Permanent ground anchors.

    H.-Y. Fang (ed.), Foundation Engineering Handbook 

    © Springer Science+Business Media New York 1991

  • 8/19/2019 juran1991

    2/38

    homogeneous soils, to locate the anchor bond length

    beyond the potentially unstable soil mass adjacent to the

    structural element.

    • The anchor grout, also called primary grout, is generally a

    portland cement-based mixture or a polymer resin and is

    used to transfer the anchor force to the ground. Secondary

    grout can be injected into the drilled hole after stressing to

    provide corrosion protection for unsheeted tendons.

    • The anchorage is

    a device attached to the tendon that consists

    of a plate and an anchor head (or threaded nut) and permits

    stressing and lock -off of the prestressing steel.

    During the past 50 years, permanent ground anchors have

    been extensively used by contractors to provide vertical and

    lateral support for natural and engineered structures. Typical

    applications of ground anchors are illustrated in Figure 26.2.

    They have found widespread acceptance in a variety of civil

    engineering projects including cut slope retaining systems,

    tied-back diaphragm or soldier pile walls, bridge abutments,

    stabilization of natural slopes and

    cliffs,

    tunnel portals, under

    pinning, repair or reconstruction of quay walls, dam spillways,

    loading ramps, hangars, etc. They have also been frequently

    used as tied own supports for dams, transmission towers, and

    waterfront structures, primarily to resist uplift water pressures

    and rotational loadings.

    Tiebacks were first used to anchor structures in rock. The

    earliest permanent rock tied owns were installed by the French

    engineer Coyne for anchoring the Jument lighthouse (1930)

    and raising the Cheurfas Dam, Algeria (1934).

    By

    the late

    1950s,

    use of permanent rock tiedowns had become common practice

    in

    renovation and construction of dams (Evans, 1955; Morris,

    1956; Middleton, 1961) and towers (Weatherby, 1982). In the

    1950s contractors began to use tiebacks for temporary supports

    Fractured

    sandstone

    ...........

    /

    Permanent

    tiebacks

    (a)

    .D

    Wall

    (c)

    Ground Anchors and Soil Nails in Retaining Structures 869

    of deep excavations. The first permanent soil tiebacks in the

    United States were installed in

    1961

    in a very stiff silty clay for

    the construction of retaining walls for the Michigan expressway

    (Jones and Kerkhoff,

    1961).

    However, in spite of long-term

    European experience, permanent ground anchors had not been

    in common use in the United States until the late

    1970s,

    mainly

    because of engineering concerns with regard to long-term

    performance, potential time-dependent (creep) movement, cor

    rosion protection of the tendon, and the need to establish

    reliable quality control testing procedures to verify the short

    and long-term holding capacity. Technological efforts have been

    continuously invested by specialty contractors to overcome

    these limitations, develop efficient corrosion-protection systems,

    improve grouting methods and installation procedures, and

    increase the tension capacity of the prestressed tendons.

    The rapid acceptance and growing use of ground anchors

    can be attributed mainly to significant technical advantages

    resulting in

    substantial cost savings and reduced construction

    period. Specifically, in urban areas the use of ground anchors

    often allows significant reduCtion in right-of-way acquisition

    and permits the elimination of temporary support systems,

    external bracings, or the need for underpinning existing structures

    near to excavation sites. The increasing confidence in ground

    anchor use for permanent structures is primarily due to reliable

    quality control procedures that involve routine performance

    and proof testing of all production anchors under loads

    exceeding the design load. Performance specifications and codes

    of practice, based on experience and long-term observations of

    permanent anchor installations, have been developed in Euro

    pean countries (French Recommendations, Bureau Securitas,

    1977;

    FIP

    Rules, 1974; German Standards, DIN, 1972, 1976;

    PTI Recommendations, 1980) and more recently in the United

    States (FHWA; see Cheney, 1984) to specify design, construction,

    and monitoring procedures

    (b)

    (d)

    Permanent

    tiebacks

    Existing

    dam

    • ' , . "

    '-

    ... . F. ..

    Permanent

    tiedown

    Fig.26.2 Typical applications of permanent ground anchors. (a) Concrete wall.

    (b)

    Landslide and tunnel portal.

    (c)

    Permanent tower

    tiedown.

    (d)

    Dams.

    (After Weatherby,

    1982.)

  • 8/19/2019 juran1991

    3/38

    870 Foundation Engineering Handbook

    26.2.2 Soil Nailing

    Soil nailing is an in-situ soil reinforcement technique that has

    been used during the past two decades, mainly in France and

    Germany, in cut slope retaining systems and slope stabilization.

    The fundamental concept of soil nailing consists of reinforcing

    the ground by passive inclusions, closely spaced, to create in

    situ a coherent gravity structure and thereby to increase the

    overall shear strength of the in-situ soil and restrain its

    displacements. This technique has emerged essentially as an

    extension ofthe "New Austrian Tunneling Method" (Rabcewicz,

    1964-65), which combines reinforced shotcrete and rockbolting

    to provide a flexible support system for the construction

    of

    underground excavations.

    Although soil nailing technology is relatively new, it has

    been used in a variety of civil engineering projects including

    stabilization of railroad and highway cut slopes (Rabejac and

    Toudic,

    1974; Hovart and Rami, 1975; Stocker et

    aI.,

    1979;

    Cartier and Gigan, 1983; Schlosser, 1983); construction of

    excavation retaining structures in urban areas, for high-rise

    buildings and underground facilities (Louis, 1981; GassIer and

    Gudehus, 1981; Shen et aI., 1981); landslide stabilization

    (Guilloux et aI., 1983; Blondeau et

    aI.,

    1984); tunnel portals in

    steep and unstable stratified slopes (Louis, 1981); and other

    civil and industrial projects. Typical applications of soil nailing

    are illustrated in Figure 26.3. Several nailed soil-retaining

    structures have been instrumented to establish a data base for

    evaluation of structure performance and development of reliable

    design methods (Stocker et

    aI.,

    1979; GassIer and Gudehus,

    1981; Schlosser, 1983; Plumelle, 1986). In North America, the

    system was initially used in Vancouver, B.C., in the late 1960s

    in

    temporary excavation supports for industrial and residential

    buildings (Shen et

    aI.,

    1981). Presently, soil nailing systems can

    (a)

    (i) Conventional

    method ---

    I'"

    ///t t

    /

    I

    Reinforced ,

    concrete

    (c)

    (ii) Austrian tunneling

    method

    Anchor pin

    be considered for any temporary

    or

    permanent application

    where conventional cut retaining systems, such as cast-in-place

    reinforced-concrete walls or tied-back walls, are applicable. As

    demonstrated by GassIer and Gudehus (1981), soil-nailed

    retaining structures can withstand both static and dynamic

    vertical loads at their upper surface without undergoing excessive

    displacements. Therefore, they can be effectively used in the

    construction of bridge abutments. Soil nailing also appears

    to provide an efficient and economical technique for repair and

    reconstruction of existing structures, particularly tie-back walls

    and reinforced soil retaining structures.

    In soil-nailed retaining structures, the inclusions are generally

    steel bars or other metallic elements that can resist tensile

    stresses, shear stresses, and bending moments. They are either

    placed in drilled boreholes and grouted along thei r total length

    or driven into the ground. The nails are not prestressed but

    are closely spaced (e.g., one driven nail per 2.5 ft2, one grouted

    nail per 10-

    50

    ft

    2

    ) to provide an anisotropic apparent cohesion

    to the native ground. The facing of the soil-nailed structure is

    not a major structural load-carrying element but rather ensures

    local stability of the soil between reinforcement layers and

    protects the ground from surface erosion and weathering effects.

    It generally consists of a thin layer of reinforced shotcrete (4- to

    6-in thick), constructed incrementally from the top down. The

    facing and the nails are placed immediately after each excavation

    stage to restrain ground decompression and therefore to prevent

    deterioration of the original mechanical properties and shear

    strength characteristics of the native ground. Prefabricated or

    cast-in-place concrete panels have increasingly been used

    in

    the

    construction of permanent structures to satisfy specific aesthetic

    and durability design criteria.

    As with ground anchors, soil nailing has been primarily used

    for temporary retaining structures. This is mainly due to the

    ----

    (b)

    2f}[

    TESTB

    ~ t

    .. .. 1

    20

    ft

    J

    (d)

    Fig. 26.3

    Typical applications

    of

    soil nailing.

    (a)

    Landslide.

    (b)

    Retaining structures.

    (c)

    Tunnel portal.

    (d)

    Abutments.

  • 8/19/2019 juran1991

    4/38

  • 8/19/2019 juran1991

    5/38

    872 Foundation Engineering Handbook

    valves located along the bond zone. Typical examples of

    regroutable anchors are the Soletanche

    IRP

    anchor (Jorge,

    1969), the TMD-Bachy anchor (Clement and Navarro, 1972),

    and the TPT anchor (Mastrantuono and Tomiolo,

    1977).

    The

    Soletanche IRP anchor and the tube-a-manchette packer are

    schematically illustrated

    in

    Figure

    26.5.

    Single and multi-underreamed anchors are used in stiff cohesive

    soils and weak rock. Their installation process involves the use

    of an underreaming device that is basically a cutting tool. First,

    the cylindrical anchor shaft is drilled, usually using a continuous

    flight auger. Then, the cutting tool is mechanically expanded

    at a controlled rate to the design size of the underream. The

    soil is removed by water flushing; neat cement grout is

    tremie-grouted into the drill hole; and the tendon is inserted.

    Depending on the underreaming device, several underreams

    I Steel tendon

    Sleeve-pipe

    2 Plastic sheath

    ~ ~ ~ S I

    IT - ' - - 3

    Annular space (epoxy-

    or

    cement-filled)

    Anchor

    Strand

    Rubber

    manchette

    Grouting

    pressure distends rubber

    manchelle and forces grout through

    sealing grout

    sleeve grout to secure tube

    it manchette in hole

    (b)

    Fig.26.5 (a) Schematic section of an IRP anchor. (b) Detail of

    tube-a-manchette for pressure grouting control. (After Pfister et al.,

    1982.)

    can be cut simultaneously. The spacing between the underreams

    is

    selected to induce a shear failure along a cylinder passing

    through the tips of the underreams.

    The main structural element of each ground anchor

    is

    the

    steel tendon, which may consist of bars, wires, or strands.

    Strands and wires have advantages with respect to tensile

    strength (ultimate tensile strength: 270ksi for strands and

    240 ksi for wires), and ease of transportation, storage, and

    fabrication. However, bars (ultimate tensile strength of 150 to

    160

    ksi) are more readily protected against corrosion and, in

    the case of shallow, low-capacity anchors, are usually easier

    and cheaper to install. Often, availability and cost

    will

    be the

    determining factors. In the United States, bars and seven-wire

    strands are the most commonly used tendons. High-capacity

    tendons made of 18 strands with a diameter of 0.50 or 0.60

    inch are also available for high-capacity tied own applications.

    B.

    Soil Nailing

    The steel reinforcing elements used for soil nailing can be

    classified

    as

    (a) driven nails, (b) grouted nails, (c) jet-grouted

    nails, and (d) corrosion-protected nails.

    Driven nails,

    commonly used in France and Germany, are

    small-diameter (15 to 46 mm) rods or bars, or metallic sections,

    made of mild steel with a yield strength of 350

    MPa

    (50 ksi).

    They are closely spaced

    (2

    to 4 bars per square meter) and

    create a rather homogeneous composite reinforced soil mass.

    The nails are driven into the ground at the designed

    inclination using a vibropercussion pneumatic or hydraulic

    hammer with no preliminary drilling. Special nails with an axial

    channel can be used to allow for grout sealing of the nail to

    the surrounding soil after its complete penetration. This installa

    tion technique

    is

    rapid and economical (4 to 6 per hour).

    However, it is limited by the length of the bars (maximum

    length about

    20

    m) and by the heterogeneity of the ground (e.g.,

    presence of boulders).

    Grouted nails are generally steel bars (15 to 46 mm in

    diameter) with a yield strength of 60 ksi. They are placed in

    boreholes (10 to 15 cm in diameter) with a vertical and

    horizontal spacing varying typically from 1 to 3 m depending

    on the type of the in-situ soil. The nails are usually cement

    grouted by gravity or under low pressure. Ribbed bars can be

    used to improve the nail-grout adherence, and special perforated

    tubes have been developed to allow injection of the grout

    through the inclusion.

    let-grou ted nails are composite inclusions made of a grouted

    soil with a central steel rod, which can

    be

    as thick as

    30

    to

    40 cm.

    A technique that combines the vibropercussion driving and

    high-pressure (greater than

    20

    MPa) jet grouting has been

    developed recently by Louis (1986). The nails are installed (Fig.

    26.6) using a high-frequency (up to 70 Hz) vibropercussion

    hammer, and cement grouting

    is

    performed during installation.

    The grout is injected through a small-diameter (few millimeters)

    longitudinal channel in the reinforcing rod under a pressure

    that

    is

    sufficiently high to cause hydraulic fracturing of the

    surrounding ground. However, nailing with a significantly lower

    grouting pressure (about 4 MPa) has been used successfully,

    particularly in granular soils. The jet-grouting installation

    technique provides recompaction and improvement of the

    surrounding ground and increases significantly the pull-out

    resistance of the composite inclusion.

    Corrosion-protected nails generally use double protection

    schemes similar to those commonly used in ground anchor

    practice. Proprietary nails have recently been developed by

    specialty contractors (Intrafor-Cofor; Solrenfor) to be used in

    permanent structures. These corrosion-protection schemes are

    described in a later section of this chapter.

  • 8/19/2019 juran1991

    6/38

    Fig. 26.6a Construction process of a soil-nailed wall illustrating

    excavation, shotcreting, and nailing.

    Fig. 26.6b Jet bolting: installation of reinforcing elements. 1.

    Vibropercussion hammer.

    2.

    Sliding support.

    3.

    Reinforcement to

    be inserted. 4. Sliding guide. 5. Fixed guide. 6. Soil to be treated.

    26.3.2

    Facing and

    Structural

    Retaining Elements

    In an anchored retaining system the wall has a major structural

    role. It has to resist the tensile forces transferred by the anchors,

    the lateral pressure applied by the retained soil, and bending

    moments. The wall has to be stiff enough to restrain the ground

    displacement induced by the excavation process. The facing in

    a nailed soil-retaining system has only a minor mechanical role.

    The maximum tensile forces generated in the nails are signifi

    cantly greater than those transferred to the facing. The main

    function of the facing is to ensure local stability of the ground

    between the nails and to limit its decompression. Hence, the

    facing has to

    be

    continuous, fit the irregularities of the cut slope

    surface, and be flexible enough to withstand ground displacement

    during excavation. The structural elements used to build the

    anchored wall are therefore basically different from those used

    to construct the facing of a nailed soil-retaining structure.

    A. Structural Elements

    of an

    Anchored Wall

    An anchored wall can be constructed with a wide variety

    of structural elements, using different installation techniques.

    Selection of the structural element for a specific application will

    generally depend on the subsurface soil (or rock) type, ground

    water conditions, local construction practice, availability of

    material and equipment, and performance requirements. The

    Ground Anchors and Soil Nails in Retaining Structures

    873

    structural elements can be evaluated in terms of their stiffness,

    ease of handling and installation, durability, water-tightness or

    continuity, and ease of removal. The elements commonly used

    can be broadly classified into four major categories: driven

    sheet piles, soldier piles and lagging walls, cylinder walls, and

    concrete diaphragm or slurry walls. Typical properties of each

    system are indicated in Table

    26.1.

    Sheet-pile walls usually consist of interlocking steel sheets

    driven into the ground prior to excavation. They are fairly

    impervious and easy to handle and install in soft clays,

    cohesionless silts, or loose sands. However, they are difficult to

    use in compact granular soils containing cobbles or boulders.

    As compared with other elements, they are relatively flexible

    and the wall displacement

    will

    in general be larger. They

    are commonly used for marine bulkhead construction (see

    Chapter 12).

    Soldier piles

    and lagging (Figure 26.7a) usually consist of

    steel H-beams that are either driven into the ground or placed

    in predrilled boreholes prior to excavation. Concrete bored piles

    with reinforcement or permanent casing have also been used.

    As

    excavation proceeds, the ground between these piles is

    retained by lagging of wood planks, cast-in-place, or precast

    concrete elements. H-beam soldier piles and lagging walls are

    probably the excavation support system most widely used in

    the United States for temporary supports. They are easy to

    install in most types of soils, and present a significant advantage

    specifically in compact or irregular strata that would obstruct

    sheet piling. They can be readily adapted to different site

    conditions and irregular wall

    alignments. The main disadvantage

    of this retaining system is that the wall is rather pervious and

    subsurface water

    flow

    may cause local instabilities. A properly

    lagged wall should permit drainage, draw down, and fluctuation

    of water level without flow of the retained soil.

    Cylinder walls consist of an array of cylindrical caissons that

    are usually constructed of reinforced concrete or mixed-in-place

    soil-cement and are closely spaced to form a continuous wall.

    They can

    be

    cast-in-place and installed using several techniques

    such as hollow-stem augers, rotary drilling equipment, deep

    mixing methods, or jet-grouting. Depending on the stiffness of

    the individual cylinders, such a wall may be rigid enough to

    support lateral loads with limited deflection. To achieve water

    tightness and properly retain the soil, shotcrete or lagging in

    the space between the cylinders may be required. Alternatively,

    the cylinders can overlap to produce a continuous, impervious

    wall. In addition to their rigidity, cylinder walls offer the

    advantage of adaptability to irregular site alignments and can

    be used in a variety of ground conditions.

    Slurry walls

    or concrete diaphragm walls are generally

    formed in a trench supported by viscous mud slurry (see Chapter

    20). Concrete is tremied into the trench, displacing the mud

    slurry upward. Reinforcement of the wall is made by vertical

    steel sections, precast reinforced-concrete members, or cages of

    reinforcing steel. Recent developments include the use of precast

    concrete panels. These walls can be designed to achieve a

    specified degree of stiffness and water-tightness, and can be

    integrated in the permanent structure. They are often used

    where lowering of the water table would adversely affect

    adjoining structures. Their main disadvantage is the relatively

    high cost and the need for specialized construction equipment

    and experienced contractors. They also may present environ

    mental problems pertaining to slurry disposal.

    Cast-in-place reinforced-concrete panels

    have been used in

    the construction of multi tied-back walls (Kerisel et aI., 1981).

    Figure 26.7b shows a schematic cross section of a 30-m deep

    open excavation retained by ten layers of prestressed anchors.

    This anchored wall was constructed from the top down with

    successive stages of (1) excavation, (2) in-place casting of the

    reinforced-concrete panel (2-ft thick, 9-ft high), and (3) anchoring.

  • 8/19/2019 juran1991

    7/38

    874 Foundation Engineering Handbook

    TABLE 26.1 PROPERTIES OF

    STRUCTURAL WALL

    ELEMENTS.

    EI-KSF/F

    System

    x

    10

    3

    3 to 50

    VERTICAL

    STEEL SHEETING

    SWF

    to 14WF

    al

    6ft to 8ft c-

  • 8/19/2019 juran1991

    8/38

    0

     

    2

    6

    m

     

    C

     

    =

     

    0

     

    <

    \>

     

    =

     

    3

     °

     

    2

    6

    m

     

    2

     

    2

    4

    m

     

    C

    '

    =

    l

    O

    k

    P

    a

     

    3

     

    2

    8

    m

     

    =

     

    3

    2

    °

     

    E

     

    4

     

    2

    8

    m

    ·

     

    a

     

    S

     

    2

    8

    m

     

    N

     

    6

     

    2

    8

    m

     

    7

     

    3

    0

    m

     

    C

    =

    2

    0

    k

    P

    a

     

    '

     =

     

    8

    °

     

    -

    -

    -

    -

    S

    c

    h

    e

    m

    a

    t

    i

    c

     

    e

    r

    o

    S

     

    e

    e

    t

    i

    o

    n

     

    f

    i

    g

    .

     2

     

    .

     

    7

     

    (

    ,

    )

     

    S

    o

    l

    d

    i

     

    p

    H

     

    d

     

    l

    o

    g

    g

    i

    n

    g

     

    ,

    n

    ,

    h

    o

     

    .,

    ,

    1

    1

    .

     

    (

    C

    o

    u

     

    .

     

    N

    i

    ,

    h

    o

    /

    '

    o

    n

     

    c

    o

    n

    ,

    ,

     

    u

    w

    o

    n

     

    C

    o

    .)

     

    (

    b

    )

     

    L

     

    .

    ,

    I

    m

     

    o

    p

    e

    n

     

    i

    o

    n

     

    n

     

    o

    n

    t

    e

     

    C

    a

    r

    l

    o

    .

     

    K

    e

    r

    i

    s

    e

    l

     e

    t

     

    8

    1

    .

     

    1

    9

    8

    1

    .

    )

     

    (

    b

    )

     

    b

    )

     

    c

    o

    n

    t

    .

     

    87 

  • 8/19/2019 juran1991

    9/38

    6

     

    S

    H

    O

    i

    C

    R

    E

    i

    E

     

    f

    /

    l

    .

    C

    I

     

    -

    l

    G

     

    G

    O

     

    G

     

    .

     

    ?

    \

    .

    .

    A

    1

    E

     

    .

    I

    '

    O

    '

    J

    .

    '

    I

    '

     

    C

    O

    /

    )

    .

    1

    £

    .

    0

     

    c

    l

    J

    i

     

    i

     

    f

    l

    i

     

    f

    l

    E

    I

    J

    )

     

    p

    l

    I

    .

    I

    N

    i

     

    P

    I

    I

    .

    t

    J

    \

     

    f

    i

    g

    .

     

    2

    6

    .

    1

     

    _

    G

    R

    O

    U

    i

     

    f

    '

    /

    l

    -

    O

     

    /

    l

    -

    $

     

    R

    E

    Q

    \

    .

    \

    I

    R

    E

    '

    .

    O

     

    ,

     

    O

    R

    ,

    /

    l

    -

    R

     

    f

    '

    t

    >

    c

    l

    (

     

    ,

    0

     

    C

    O

    l

    '

    f

    (

    '

    /

    l

    -

    I

    N

     

    ,

    N

    O

     

    S

    1

    /

    l

    -

    C

    ;

    E

     

    G

    R

    O

    I

    f

    t

    .

     

    N

    /

    l

    -

    I

    \

    .

    .

     

    l

    S

    i

     

    S

    i

    A

    G

    E

     

    G

    R

    O

    \

    .

    )

    "

    r

     

    ,

    0

     

    @

     

    >

    '

     

    <

    I

    e

     

    ,

    O

    R

     

    .

     

    .

    '

     

    s

    o

    O

    '

     

    o

    v

    <

    '

     

    .

    "

    e

     

    .

    o

    r

     

    ,

     

    <

    o

     

    r

    .

    <

    r

     

    ,

    e

     

    d

     

    ·

    1

     

    s

    h

    o

    t

    c

    r

    e

    t

    e

     

    t

    a

    c

    i

    n

    g

    ·

     

    l

     

    c

    )

     

    c

    o

    n

    s

    t

    r

    U

    c

    t

    i

    o

    n

     

    e

    t

    a

    l

     

    1

     

    a

     

    ,

     

    (

    b

    )

     

    .

     

    .

    .

     

    '

    ,

    ,

    .

     

    ,

    ,

    0

    '

     

    '

    '

     

    P

    '

    o

    '

     

    ,od 

    0' 

    0

    """"

    '

    -

    \

     

    e

    ,

    u

    ,

     

    S

    O

    /

    ,

    ,

    ,

    '

    O

    ,

    -

    )

     

    2

    6

     

    8

     

    I

     

    a

    )

     

    P

    l

    e

    t

    a

    b

    l

    icated steel 

    p

    a

    n

    e

     

    s

    .

     

    fig· 

    \

     

  • 8/19/2019 juran1991

    10/38

    26.3.3

    Drainage

    Groundwater (see Chapters 1 and 7)

    is

    a major engineering

    concern relative to construction of anchored and nailed soil

    retaining structures.

    An

    appropriate drainage system must be

    provided to (a) prevent generation of excessive hydrostatic

    pressures on the facing (or the structural wall element), (b)

    protect the facing element and particularly shotcrete facing from

    deterioration induced by water contact, (c) prevent saturation

    of the nailed ground, which can significantly affect the structure

    displacement and may cause instability during and after excava

    tion. In anchored walls, prefabricated vertical drains, porous

    engineering fabrics, or subhorizontal drains can be used for

    drainage ofthe subsurface

    flow.

    In soil nailing, shallow drainage

    (plastic pipes, 10 cm in diameter,

    30

    to 40 cm long)

    is

    usually

    used to protect the facing, while subhorizontal slotted plastic

    tubes are used for deep drainage of the nailed ground. In the

    case of permanent structures with prefabricated panels, a

    continuous drain such

    as

    a geotextile can be placed behind the

    facing.

    26.4

    SOIL-INCLUSION

    INTERACTION:

    PULL-OUT

    CAPACITY ESTIMATES

    The load-transfer mechanisms between a grouted anchor (or

    nail) and the subsurface soil (or rock) as well as the ultimate

    pull-out capacity depend upon several parameters, including

    installation technique, drilling and grouting method, grouting

    pressure,

    size and shape of the grouted inclusion, engineering

    properties of the in-situ soil and specifically its relative density

    (or overconsolidation ratio), permeability, and shear strength

    characteristics (see Chapter 3).

    The grain size and porosity of the in-situ soil govern the

    grout conductivity. In sands, gravels and weathered rock, with

    hydraulic conductivities of 10-

    1

    to

    1O-

    2

    cm/sec, grout will

    permeate through the pores or natural fractures of the ground.

    In fine-grained cohesionless soils (silts and fine sands), with

    hydraulic conductivity smaller than 10-

    3

    cm/sec, the grout

    cannot penetrate the small pores but rather compacts locally,

    under pressure, the surrounding ground. Increasing the grout

    pressure

    will

    induce a greater grout permeation into the ground

    and/or a more effective ground densification. Consequently,

    under high-pressure grouting, high radial stresses are locked

    into the soil surrounding the anchor, increasing its pull-out

    capacity.

    26.4.1 Load Transfer in Ground Anchors

    Figure 26.4 illustrates the basic soil-inclusion interaction

    mechanisms for the main types of ground anchors.

    Tremie-grouted straight-shafted anchors,

    which are more

    commonly used in rock and very stiff to hard cohesive soils,

    generate their pull-out resistance through the lateral shear

    mobilized at the grout-ground interface. The pull-out capacity

    of these anchors is often estimated by

    P

    =

    n;'D'L"ult (26.1)

    where ul t

    is

    the ultimate lateral shear stress at the ground-grout

    interface (also called shaft friction), D and L are, respectively,

    the effective diameter and bond length of the grouted anchor.

    It should be noted that the effective anchor diameter

    D

    is

    difficult to estimate, since it is highly dependent upon ground

    porosity and grout permeability.

    I t is

    commonly assumed that

    in

    competent rocks (Littlejohn and Bruce, 1975)

    'ul t

    =

    lO%'Sa

    for

    Sa

    <

    600 psi (26.la)

    Ground Anchors and Soil Nails in Retaining Structures 877

    where Sa

    is

    the uniaxial compressive strength. In cohesive soils,

    'ul t =

    IX'

    S.

    (26.1b)

    where

    IX is

    an adhesion factor, and S. is the average undrained

    shear strength of the soil.

    The adhesion factor (IX) generally varies (Tomlinson, 1957;

    Peck, 1958; Woodward et

    aI.,

    1961) within the range of

    0.3

    to

    0.75, with the lower values obtained for stiffer and harder clays.

    Low-pressure grouted anchors

    are installed under an effective

    grout pressure lower than 150 psi (or, in cohesive soils, under

    pressure that would not fracture the ground) most commonly

    using hollow-stem augers or tremie grouting, an open hole in

    cohesive soils or cored rotary-drilled holes in cohesion1ess soils.

    The grouting pressure

    will

    induce an increase of the effective

    diameter of the grout bulb by permeation or local compaction

    ofthe ground. Therefore, the pull-out resistance ofthese anchors

    is highly dependent upon the grout pressure. It

    is

    primarily

    derived from the ultimate interface shear stress but an end

    bearing resistance can be mobilized owing to an effective

    increase of the grout bulb diameter. The pull-out resistance

    is

    commonly estimated using Equation 26.1.

    For cohesionless soils,

    'ul t

    =

    p'

    A .

    tan

    cfJ

    (26.1c)

    where

    p

    is

    the effective grout pressure, cfJ

    is

    the internal friction

    angle of the soil, and A

    is

    a dimensionless empirical coefficient

    smaller than 1 (Hanna, 1982). For practical applications

    p is

    generally limited to less than

    50

    psi or 2 psi per foot of

    overburden (Littlejohn, 1970).

    High-pressure grouted anchors

    are installed under effective

    grout pressures exceeding 150 psi often using postgrouting

    techniques or pressure injection (feasible only in cohesionless

    soils). Figure

    26.9,

    by Jorge (1969), illustrates the significant

    effect of grouting pressure on the ultimate load-transfer rate

    (or ultimate lateral shear stress) of multiphase postgrouted

    anchors in different types of soils. The high-pressure grouting

    25

    20

    .::

    "

  • 8/19/2019 juran1991

    11/38

    878 Foundation Engineering Handbook

    Z

    .><

    ~

    ~

    'u

    '

    .

    '

    .>

    01)

    c

    '>,

    ....

    ...

    '

    -0

    '

    .Q

    u

    ;;;

    'E

    5

    2000

    1800

    1600

    1400

    1200

    1000

    800

    600

    400

    200

    o

    Bond-to-ground length L , m

    dense

    Typeo/soil

    Density

    SPT

    blows/30 em

    Gravelly

    Very dense

    120

    -

    sand

    Dense 60

    -

    Medium

    dense

    43

    U

    = dr:Jd

    lO

    I- -

    ...

    = 1.6/0.16

    Loose

    I I

    x

    Sandy gravel

    U=

    15/0.3

    Very den e

    >

    130

    ... 11.3 = Diameter

    of

    grouted body D = 11.3

    em

    Sandy

    gravel

    U=5to10

    Gravelly and

    = = = U=8to lO

    ~ ~ ~ ~ a n d

    ~ ~ ~

    Medium to coarse

    sand

    =-=.,.,,;,.,

    (with gravel)

    U = 3.5to 4.5

    Diameter of grouted bodies

    D

    = 10to IS em

    Fig.26.10 Ultimate load holding capacity

    of

    anchors in sandy-gravel and gravelly sand.

    (After Ostermayer and Scheele, 1977.)

    results in a grout root (or fissure) system that mechanically

    interlocks with the surrounding ground, increasing substantially

    the pull-out capacity of the anchor. Particularly, in dense

    granular soils this interlocking phenomena generates high

    tendency for the soil to dilate, which in turn results in a normal

    stress concentration at the

    grout-ground

    interface. The effect

    of pressure injection on the soil-anchor interaction is difficult

    to evaluate. Empirical relationships were provided by Oster

    mayer (1974) for estimating the ultimate lateral shear stress for

    high-pres sure-grouted anchors, with and without postgrouting,

    in

    fine-grained soils (sandy silts to highly plastic clays). Oster

    mayer and Sheele (1977) developed empirical curves, reproduced

    in Figure 26.10, to estimate the ultimate pull-out capacity of

    pressure-injected anchors in granular soils as a function of

    anchor length, soil type, density, and uniformity. These curves,

    derived from 30 pull-out tests on anchors installed under grout

    pressures of about 70 psi, illustrate that the ultimate capacity

    of the anchor is not proportional to its length.

    Underreamed anchors,

    which are mainly used

    in

    stiff to hard

    cohesive soils derive their pull-out capacity from adhesion along

    their shaft above the underreams, end bearing of the first

    underream, and lateral shear along a cylinder established by

    the tips of the underreams. For the cylinder to

    be

    effectively

    established, the spacing between the underreams should not

    exceed

    1.5

    times their diameter (Bassett, 1977). Estimate of the

    pull-out capacity of these anchors (Littlejohn, 1970) is based

    on empirical formulas that are conventionally used for pile

    design in

    cohesive soils.

    The load transfer along pressure-injected and high pressure

    postgrouted anchors has been investigated by several authors

    (Bustamente,

    1975,

    1976; Ostermayer and Sheele, 1977; Shields

    et

    aI.,

    1978; Bustamente, 1980; Davis and Plumelle,

    1982).

    Figure

    26.11 illustrates, for ultimate pull-out loads, the distributions

    of the lateral interface shear stress along pressure-injected

    anchors in gravelly sands of different densities (Ostermayer and

    Sheele, 1977). Similar results were reported for postgrouted

    anchors (Bustamente, 1972) in river sands (Fig. 26.12a) and for

    straight-shafted anchors (Feddersen, 1974) in highly over

    consolidated, stiff, plastic clays. Figure 26.12b shows the results

    Soil density

    Bond/engtll

    L, .,m

    0

    Very dense

    2.0m

    4.5m

    0

    Dense

    3.0m

    0

    Medium

    2.0m

    dense

    4.5m

    '

    Loose

    2.0m

    ...

    4.5m

    300

    1200

    \Loo

    1000

    ME

    900

    Z

    800

    ><

    3

    700..

    c

    .2

    600

    u

    :E

    SOO

    c

    :.;;;

    til

    Very

    dense

    • . ~ a x l

    ,

    • 0 0

    -: \ Medium

    " • •

    den

    e

    __

    _

    O - - = =

    -

    ,;;;.:;.;-

    - , -

      • max

    l ,

    =

    mean

    "

    ~ D , ~ . . . Loose

    ~ - - -   - -- -- - - -- -- -_ ...,

    - - - - - ' _ ~ max

    t , "" mean "

    ...

    100

    Length. m

    L,.

    = 2.0 m

    L,.

    = 3 .0 m

    L,

    = 4.5 m

    Fig. 26.11 Distribution

    of

    the lateral interface shear stress along

    pressure-injected anchors at the ultimate load.

    (After Ostermayer

    and

    Scheele, 1977.)

  • 8/19/2019 juran1991

    12/38

    o

    2

    3

    E

    E. 5

    o

    6

    7

    8

    9

    10

    train

    200

    400

    (a)

    Ground Anchors and Soil Nails in Retaining Structures 879

    o

    E

    Pull

    -o

    ut load distribution

    Displace

    ment V, mm

    Mobilization of lateral friction

    (b)

    Fig.26.12 (a)

    Distribution

    of

    deformation along the length

    of an

    IRP anchor.

    (After Bustamante,

    1972). (b) Mobilization of the lateral

    friction along an anchor in a plastic clay. (Winnezeele, Bustamente, 1980.)

    of a pull-out test on an instrumented anchor

    in

    a plastic clay

    (Bustamente, 1980). The slope of the tension force distribution

    along the anchor corresponds to the lateral interface shear stress

    mobilized at a specific depth under the applied pull-out force.

    As shown in Figure 26.12, the shear stress-upward anchor

    displacement curves obtained for different depths indicate

    overconsolidation of the subsurface soil layer and illustrate that

    the anchor displacement required to

    fully

    mobilize the ultimate

    shear stress is about 5 to 10 mm. The results of these studies

    demonstrate that in dense granular soils and highly overcon

    solidated clays the load-transfer rate along the anchor is not

    constant and the pull-out capacity is therefore not proportional

    to the anchor length.

    The variation of the load-transfer rate along the anchor is

    mainly the result of its extensibility during pull-out testing.

    It

    is primarily dependent upon the relative rigidity (or elastic

    modulus ratio) of the anchor and the grout-soil interface and

    is

    particularly pronounced in highly dilatant stiff soils. Wernick

    (1977), Schlosser and Elias (1978), and Plumelle and Gasnier

    (1984) have shown that the restrained tendency of the soil to

    dilate during shearing results in a normal stress concentration

    at soil-inclusion interfaces that affects significantly the load

    transfer rate. As shown in Figure

    26.13,

    vertical stresses as high

    as four times the overburden pressure were measured in a

    medium dense river sand at the anchor interface. The higher

    the (anchor-to-soil) elastic modulus ratio the more uniform is

    the load-transfer rate. Figure 26.11 shows that in loose to

    medium dense granular soils the interface lateral shear stress

    is

    approximately constant along the anchor.

    It is

    of interest to

    indicate that these results are consistent with those obtained

    for ribbed metallic strips in Reinforced Earth structures

    (Schlosser and Elias,

    1978).

    Several authors have attempted to analyze the load transfer

    along the anchor using the "t-z" method (Coyle and Reese,

    1966; Davis and Plumelle, 1982), which is commonly applied

    in design of friction piles,

    or

    more complex interface soil models

    (Zaman et al., 1984; Frank et al., 1982). However, rational

    anlysis of the

    ground-anchor

    interaction requires appropriate

    interface properties that are difficult to estimate.

    3

    100

    200

    Distance, d, cm

    Fig. 26.13

    Restrained dilatancy effect around a ground anchor.

    (After

    Plumelle

    and

    Gasnier, 1984.)

  • 8/19/2019 juran1991

    13/38

    88 Foundation Engineering Handbook

    26.4.2 Soil Nail Interaction

    In soil nailing, similarly to ground anchors, the load transfer

    mechanism and the ultimate pull-out resistance of the nails

    depend primarily upon soil type and installation technique.

    The pull-out resistance of

    driven nails

    in a dense granular

    soil was correlated by Cartier and Gigan (1983) with design

    recommendations for Reinforced Earth walls (Schlosser and

    Segrestin, 1979). These recommendations use the concept of an

    "apparent friction coefficient" that

    is

    derived from Equation

    26.1

    assuming

    (26.1d)

    where

    y

    is

    the unit weight of the soil

    h

    is

    the overburden height above the nail

    /1*

    is

    the apparent friction coefficient

    As

    shown in Figure 26.14, the apparent friction coefficient

    (/1*)

    obtained from pull-out tests in a nailed granular wall corresponds

    to the design value generally used for the ribbed metallic strips

    in Reinforced Earth walls. At relatively low depth, owing to

    the restrained dilatancy

    effect, the value of /1* is significantly

    greater than 1 and it decreases with depth to tan

    (4)

    ).

    Laboratory-scale pull-out tests were conducted

    by

    Elias and

    Juran (1988) in a medium dense sand to evaluate the effect of

    the nail installation process on the apparent friction coefficient.

    Figure

    26.15

    shows that the construction process for Reinforced

    Earth (i.e., placing the nail during the construction and com

    pacting the sand around the nail) produces a substantially

    higher apparent friction coefficient than nailing by driving the

    nail into the compacted sand embankment. In the latter case,

    nail driving

    will

    significantly reduce the restrained dilatancy

    effect on the pull-out resistance. Therefore, design guidelines

    for Reinforced Earth walls cannot be extrapolated to soil-nailed

    structures.

    Grouted nails

    are generally gravity-grouted. Their pull-out

    resistance is therefore expected to be approximately the same

    as tha t of an equivalent straight-shafted anchor installed under

    low (or no) grout pressure. The drilling of the borehole for the

    grouted nail produces an unloading of the disturbed surrounding

    soil that can significantly affect its mechanical properties. The

    soil-nail interaction is primarily dependent upon soil recom

    paction due to grouting. In cohesionless soils, grouting pressures

    of

    50

    to

    100

    psi are commonly used to prevent caving as the

    casing

    is

    withdrawn. This grouting pressure

    will

    induce ground

    5

    10

    o

    Z,m

    t/(kN/m)

    10 20

    - , -

    t/AVERAGE

    17

    kN/ml

    _ _ 5

    10

    SOIL: Sand = 33° c = 10 kPa

    NAILS: Driven profile

    o

    2

    F

      Specifications for

    _ reinforced

    earth

    _

    tan

    -

    ,m

    Fig.26.14

    Soil-reinforcement

    friction between a driven nail and

    a granular soil.

    (After

    Cartier

    and

    Gigan, 1983.)

    5

    ;:

    ;;

    c-

    .2

    U

    4

    :E

    '0

    3

    C

    Q)

    'u

    it:

    2

    .,

    2

    0

    u

    C

    ..,

    :;;

    Q.

    Q.

    <

    o

    Depth of nail, cm

    Fig. 26.15

    Laboratory pull-out test results.

    (1)

    Nails placed

    during backfilling.

    (2)

    Nails inserted during excavation. (After Elias

    and Juran, 1988.)

    recompaction associated with grout penetration into permeable

    gravelly seams, thereby increasing substantially the pull-out

    resistance of the nail. Apparent friction coefficient values as

    high as 3 to 6 have been reported (Elias and Juran, 1988).

    Figure 26.16a shows a cross-sectional

    view

    of an excavated

    (a)

    (b)

    Fig.26.16

    (a)

    Cross-sectional

    view of

    an excavated grouted nail

    in a granular soil, illustrating effect of

    grout

    permeation.

    (b)

    Sectional

    view of

    an excavated nail in

    silty

    clay soil.

  • 8/19/2019 juran1991

    14/38

    200

    150

    z

    -'"

    ~ 100

    50

    o

    ~ L - I

    mum:

    D = 2in

    100

    kN/rn Lille Ch ack

    L , rn

    Fig.26.17 Variation of pull-out resistance of grouted nails

    with

    embedment length.

    (After Louis, 1986.)

    nailed soil, illustrating grout permeation into an alluvial soil

    grouted under pressure ofless than

    70

    psi. Figure 26.16b shows

    that in a fine-grained cohesive soil the tremie grouting results

    in a rather smooth soil-inclusion interface. The presence of

    water at the interfaces, specifically in plastic soils,

    will

    generate

    a lubrification effect, decreasing substantially the pull-out

    resistance of the nail. Figure

    26.17

    (Louis, 1986) shows a

    Ground Anchors and Soil Nails in Retaining Structures 881

    summary of pull-out test results obtained with low-pressure

    grouted nails in different types of soils.

    Jet-grouted nails are installed under a grout pressure that

    can exceed

    20 MPa

    and

    is

    sufficiently high to cause hydraulic

    fracturing of the surrounding ground (Louis, 1981). Similarly

    to high-pressure grouting of anchors, the jet-grouting installa

    tion technique produces a mechanical interlocking between the

    penetrating grout and the surrounding ground that results in

    a substantial increase of the effective nail diameter.

    I t

    also

    provides recompaction of the surrounding ground that signifi

    cantly improves the pull-out resistance of the composite nailed

    soil inclusion. Field pull-out tests on jet-grouted nails (Louis,

    1986) yielded ultimate lateral shear stress values as high as

    400 kPa in sands and 1000 kPa in sandy gravels.

    26.4.3 Estimates

    of

    Pull Out Cpacity from In-situ

    Tests

    To date, estimates of the pull-out resistance of anchors and

    nails are mainly based upon empirical formulas (or ultimate

    lateral interface shear stress values) derived from field experience.

    These formulas are useful for feasibility evaluation and prelimi

    nary design. Table

    26.2

    provides a summary of estimated

    ultimate interface lateral shear stress (or ultimate load-transfer

    rate) values for soil nails and ground anchors as a function of

    soil (or rock) type and installation technique. Recently, increasing

    attempts have been made to develop field correlations between

    the ultimate lateral shear stress ('t

    ul t

    ) and the engineering

    properties of soils obtained from commonly used in-situ tests

    such

    as

    the Standard Penetration Test (Fujita et

    aI.,

    1977) or

    the self-boring pressuremeter test (Bustamente,

    1975, 1976).

    Recognizing apparent similitude between the soil response to

    high-pressure anchor grouting and to the expansion of a

    TABLE 26.2 ESTIMATED ULTIMATE INTERFACE LATERAL SHEAR

    STRESS

    VALUES

    FOR

    GROUND ANCHORS AND SOIL

    NAILS.

    Grouted Nails

    Construction

    Method

    Rotary drilled

    Driven casing

    Jet

    grouted

    Augered

    Soil Type

    Silty sand

    Silt

    Piedmont residual

    Sand

    Dense sand/gravel

    Dense moraine

    Sandy colluvium

    Clayey colluvium

    Fine sand

    (medium dense)

    Sand

    Sand / gravel

    Soft clay

    Stiff to hard clay

    Clayey silt

    Calcareous sandy clay

    Silty sand fill

    Ultimate Lateral Shear Stress,

    kips/

    ft

    Soil

    Nailing

    (Elias

    and

    Juran, 1988)

    2

    to

    4

    1.2

    to

    1.6

    1.5 to 2.5

    6

    8

    8 to

    12

    2

    to

    4

    1

    to

    2

    8

    20

    0.4 to 0.6

    0.8

    to 1.2

    1

    to

    2

    4

    to

    6

    0.4 to 0.6

    Permanent Ground

    Anchors

    (Cheney, 1984

    r

    5 to 9

    7

    to 13

    10 to

    20

    3.5

    to

    4.5

    b

    4.5 to

    8.5

    b

    8.5

    to

    11 .5

    b

    2 to 4

    1.5

    c

    a

    Cheney recommends a safety factor of 2.5 With respect to the ultimate lateral shear stress values indicated

    in

    thiS

    table.

    b Values obtamed for pressure-injected anchors by Jorge

    (1969).

    C Design value proposed by Weatherby (1982) for hollow-stem augered anchor (assuming a diameter of 6

    mches)

    in

    both sandy and clayey soils.

  • 8/19/2019 juran1991

    15/38

    882

    Foundation Engineering Handbook

    (a)

    0.8

    0.7

    0.6

    o

    00

    Q

    Q

    v

    o

    v

    v

    ~

    0.5

    :2

    o

    . :

    ~ y :

    o

    Q

    j 0.4

    v Q

    ~ 8 , Q ~ Q

    v

    ..

    \-t9 .. Y, ' /o /

    .... v

    ••

    . . . . lOXar

    o

    o

    0.3

    o

    • • 00 . . 0

    0.2

    0.1

    o.

    y>y...... ;;.-"000 0

    od.,,< 0 Type IRS: • Bustamante

    et

    al.

    S:>'6"/ , . Fujita et al.

    o Ostermayer and Scheele

    K Koreck

    ~ - . . .

    Type

    IGU:

    ... Bustamante

    et

    al. v Ostermayer

    Loose

    Dense

    Very dense

    20

    40

    60 80 100 120

    SPT(N)

    (b)

    0.3

    IRS

    «l

    c..

    • 0 0 0

    ~ - . '

    o .

    .7.

    ~ ..

    • •

    • •

    0 1 ; ~ ~ 0 ' . . ·v

    2 0.2

    IGU

    '1 -

    0

    '

    ' , v v-.-

    ....... ' 't:>.-

    -

    ~

    _

    ~ v : r Type IRS: • Bustamante et al. oOstermayer

    .1

    v. v "''''

    ,. .... _ _ _ _ V ' Type IGU: ... Bustamante

    et

    al. v Ostermayer

    ?-:..

    .. ... v V '

    Jones, Turner, Spencer

    o

    0.5 1.5 2

    2.5

    5 10

    15 20 25 30

    SPT (N)

    Fig.26.18

    Empirical relationships

    for

    the determination

    of

    the lateral interface shear stress.

    (a)

    Lateral interface shear stress

    for

    sand and

    gravel. (b) Lateral interface shear stress for silty

    clay

    soils. (After Bustamante

    and

    Doix, 1985.)

    pressuremeter cell, the French Central Laboratory of Bridges

    and Roads (L.c.P.c.) has conducted an extensive research

    program including 94 pull-out tests in 34 sites to provide a

    data base for field correlations.

    Figure 26.18 shows the empirical relationships derived by

    Bustamente and Doix (1985) to estimate the ultimate lateral

    shear stress values

    (Tul

    l

    ) in different types of soils and rocks as

    a function of the limit pressure

    PI

    obtained from the pressuremeter

    test or the SPT N value. These guidelines take into account

    the improvement of the soil surrounding the anchors by different

    modes of injection, considering single-stage pressure-grouted

    (lGU)

    anchors (grout pressure of O.5PI < P < PI) and multi

    stage postgrouted (IRS) anchors (grout pressure> PI). Also

    shown in Figure

    26.18

    is the wide scatter of the

    field

    data

    obtained by the L.c.P.c. and other investigators (Ostermayer,

    1974; Fujita et

    aI.,

    1977; Ostermayer and Sheele, 1977; Koreck,

    1978;

    Jones and Turner, 1980; Jones and Spencer, 1984) that

    have been compiled by Bustamente and Doix to establish these

    empirical relationships. The pull-out capacity of the anchor

    is

    estimated using Equation

    26.1.

    The effective anchor diameter

    is

    estimated using a correction factor

    (a)

    to allow for diametral

    expansion due to high·pressure grouting. The

    a

    values for

    IGU

    type anchors range from 1.1 in weathered rocks, silty clays, and

    fine

    sands, to

    1.4 in

    highly dilatant granular soils, while the

    a

    values for IRS type anchors range from 1.4 in granular soils

    and weathered rock, to 1.8

    in

    stiff clays and marls.

    The available field data pertaining to the pull-out capacity

    of nails

    is

    presently still too limited to substantiate development

    of reliable correlations.

    An

    attempt has been made by Guilloux

    and Schlosser (1984) and Louis (1986) to correlate the measured

    pull-out capacity of both driven and grouted nails with the

    French recommendations (L.c.P.c. and S.E.T.R.A., 1985) for

    the determination of lateral shaft friction on bored and driven

    concrete piles from pressuremeter test results. Figure

    26.19

    shows that in fine-grained soils (i.e., fine sands, silts, non plastic

    clays) predicted Tult values correlate reasonably well with

    pull-out test results, while in dilatant gravelly soils, compacted

    moraine, or fissured rocks they may significantly underestimate

    the measured ultimate lateral shear stress.

    It

    appears that further research and

    field

    testing could

    significantly improve the database for estimating the pull-out

    capacity of ground anchors and soil nails. The pressuremeter

    test appears also to provide valuable data for grouting proced

    ure, such as the maximum injection pressure that can be used

  • 8/19/2019 juran1991

    16/38

  • 8/19/2019 juran1991

    17/38

    884

    Foundation Engineering Handbook

    5.

    Durability considerations may impose severe limitations on

    the use of metallic inclusions in aggressive environments.

    . In addition, it should be pointed out that in a n a i ~ e d - s ~ i l

    retaining structure a certain soil-to-reinforcement relatIve dIs

    placement is required to mobilize their interaction and generate

    the required resisting nail forces. Therefore, in urban areas the

    potential use ofthis technique can be limited by the requirement

    to prevent movement of structures in the i m m ~ d i a t e vicinity of

    excavation sites. Monitoring of the structure dIsplacements can

    be implemented to verify that they are compatible with the

    required performance.

    26.6 FEASIBILITY EVALUATION

    To evaluate the feasibility and engineering use of permanent

    ground anchors or soil-nailing systems, soi.l conditions and

    existing physical constraints have to be consIdered.

    The presence of utilities such as subways or other under

    ground facilities and the need to obtain

    n d ~ r g r o u n d e a s e ~ e ~ t s

    may preclude installation of anchors (or naIls) and can slgmfi

    cantly affect the project cost.

    Durability considerations require an evaluation of the aggres

    siveness of the ground and the pore water, particularly when

    field observations indicate corrosion of existing structures.

    Ground anchors and nailing should not

    be

    used for permanent

    structures in corrosive soils (e.g., soils with high contents of

    cinder, ash or slag

    fills,

    rubble

    fills,

    industrial or acid mine

    wastes, etc.). The soil tests most commonly used to evaluate

    ground aggressiveness are electrical resistivity, pH, and .sulfate

    concentration. The critical values for ground aggressIveness

    commonly associated with ASTM standards are outlined in

    FHW A DP-68-IR, Permanent Ground Anchors, and are sum

    marized in Table

    26.3.

    Assessment of the suitability of the subsurface soil (or rock)

    to provide short- and long-term pull-out capacity of the

    a n c ~ o r

    (or nail) requires a determination of its engineering propertIes,

    specifically, shear strength and creep characteristics. ..

    In rock, the overall strength

    is

    controlled by the eXlstmg

    joints or discontinuity system. Highly fractured rocks with open

    joints or cavernous limestone are difficult to grout and therefore

    potential

    use

    of ground anchors or soil nails should be carefully

    assessed.

    Permanent ground anchors and soil nails generally cannot

    be cost-effectively installed in loose granular soils with SPT

    blow count number

    (N)

    lower than

    10

    or relative density lower

    than

    0.30.

    Nailing becomes practically unfeasible in poorly

    graded cohesionless soils with a uniformity coefficient of less

    than 2. In such soils, nailing would require stabilization of the

    cut face prior to excavation by grouting or slurry wall con

    struction.

    In fine-grained cohesive soils, long-term pull-out performance

    of the anchors (or nails)

    is

    a critical design criterion. Permanent

    ground anchors and soil nails are, generally, not feasible in soft

    TABLE 26.3 FEASIBILITY CRITERIA WITH REGARD

    TO GROUND AGGRESSIVENESS.

    Test

    Resistivity

    pH

    Sulfate

    Chlorides

    ASTM

    Standard

    G-57-78

    (ASTM)

    G-51-77 (ASTM)

    California DOT test 407

    California DOT test 422

    Critical Values

    Below 2000ohm/cm

    Below 4.5

    Above 500 ppm

    Above 100 ppm

    cohesive soils with undrained shear strength smaller than

    0.5

    tsf,

    or soils susceptible to creep. A number of national codes

    (German Standards and French Recommendations) index the

    creep susceptibility to the Atterberg limits and natural moisture

    content of the soil. They preclude the use of permanent ground

    anchors in organic soils, and plastic clayey soils with liquid

    limit

    (LL)

    greater than 50 and liquidity index (LJ) greater than

    0.2

    (or consistency index

    ( I )

    less than

    0.9).

    Soils with a plasticity

    index

    (PI)

    greater than 20 must also be carefully assessed for

    creep. In light of the limited experience with soil nailing in

    clayey soils, the applicability criteria developed for ground

    anchors are recommended for feasibility evaluation of soil

    nailed structures.

    26.7 SHORT- AND LONG-TERM PERFORMANCE

    OF ANCHORS AND NAILS

    The effective load transfer from the anchor to the surrounding

    ground requires a relative displacement between these tW?

    components of the retaining system.

    For

    ground anchors, thIs

    relative displacement

    is

    generated by prestressing the anchor

    immediately after installation. In the passive soil nails, resisting

    forces are generated owing to ground displacement during the

    construction. Evaluation of the short- and long-term performance

    of ground anchors and nails requires determination of their

    load-displacement-time behavior for the specific application

    and site conditions. Short-term performance

    is

    defined by a

    time-independent load-displacement relationship, while an

    assessment of the long-term performance should account for

    the effect of time-dependent phenomena such as creep and

    relaxation.

    26.7.1 Short-term Performance

    A static loading of anchors or nails can cause several "short

    term" failure mechanisms:

    a. Failure of the steel tendon or nails.

    b. Shear failure of the soil mass owing to insufficient depth of

    anchor embedment.

    c. Failure of the grout-tendon or nail bond.

    d.

    Failure of the soil-grout bond.

    The engineering design of the anchored (or nailed) retaining

    system for specific application and site conditions should

    provide a proper selection of the inclusion (i.e., mechanical

    properties, length, inclination, spacings, and corrosion protec

    tion) to prevent any of these failure modes.

    (a) Selection

    of

    tendon or nail section should insure that the

    working stress in the inclusion does not exceed its ultimate

    tensile strength with an acceptable factor of safety. The Post

    Tensioning Institute (PTI) recommends limiting the working

    tensile stress in prestressed steel to 60 percent of the ultimate

    tensile strength for permanent structures and 80 percent for

    temporary applications.

    (b) To prevent a shear failure of the shallow soil mass

    overlying the upper anchors, the bond zone should be located

    at a minimum depth of embedment that

    is

    generally of the

    order of 15 ft (4.6 m). This embedment length should also permit

    high-pressure grouting without damage to existing facilities.

    (c) To insure that the strength of the ground is

    fully

    mobilized

    the grout-tendon (or nail) bond should not be exceeded. The

    mechanism of grout- tendon bond involves three components:

    adhesion, friction, and mechanical interlock. The neat cement

  • 8/19/2019 juran1991

    18/38

    grout generally used provides steel-cement bonding values of

    the order of 1 to 2 MPa, mainly owing to the mechanical grout

    interlocking against irregularities in the tendon surface (i.e.,

    ribbing of the bar, threading of rebars, or stranding of wires).

    (d) Failure of the ground-grout bond will result in sliding

    of the anchor or nail. The bonded anchor or nail length should

    be designed to ensure that the force mobilized in the inclusion

    does not exceed its pull-out capacity with an acceptable factor

    of safety. The empirical relations currently employed to estimate

    the pull-out capacity (or the ultimate lateral shear stress) of

    anchors or nails can only be used for a preliminary design.

    Pull-out tests on soil nails are required to provide reliable data

    for final design, and for anchored walls each production anchor

    should be tested to ensure its load-carrying capacity throughout

    the anticipated service

    life

    of the structure. For practical reasons,

    a minimum bond length of 15

    ft is

    generally required for ground

    anchors in soils. Experience has shown that bond lengths

    exceeding

    40 ft

    do not efficiently increase the anchor capacity.

    Anchor inclination

    should be as small as possible. However,

    steep inclinations may be dictated by practical considerations,

    such as right-of way constraints, buried utilities, and soil profile.

    A minimum inclination of about 10° to

    15° is

    generally required

    to facilitate and insure effective grouting, particularly under

    low pressure. Higher inclination of tieback anchors

    will

    result

    in a transfer of significant vertical forces to the structural element

    (i.e., concrete wall or soldier pile) and is generally used only to

    reach deep bearing strata or to avoid existing structures.

    26.7.2 long Term Performance

    Long-term performance of anchors or nails depends primarily

    upon the potential of the ground-inclusion system to creep.

    Theoretically, creep can develop in the three basic components

    of the system: the ground surrounding the bond zone, the grout,

    and the steel (i.e., tendon and/or connections). However, in

    practice, creep deformations of the cement grout and the steel

    are found to be insignificant, while fine-grained clayey soils may

    undergo large creep deformation that

    will

    result in time

    dependent anchor displacement. Large creep displacements

    have been reported for multi-underreamed anchors (Ostermayer,

    1974) and pressure-injected anchors (Bustamente et

    aI.,

    1978;

    Bustamente, 1980) in plastic clayey soils. Relaxation of the steel

    tendon (i.e., stress decrease under constant strain) can also affect

    the long-term performance. However, for a stress level lower

    than the elastic limit of the steel the stress loss

    will

    generally

    not exceed 5 percent of the lock-off stress and its effect on the

    displacement will be negligible.

    Creep

    is

    a time-dependent deformation of the soil structure

    under a sustained loading owing to a continuous fabric reorien

    tation. The creep potential of a clayey soil

    is highly dependent

    upon the composition and structure of its minerals, its de

    positional (preconsolidation) history, and its natural moisture

    content (or consistency index). Several investigators (M urayama

    and Shibata, 1958; Bishop, 1966; Singh and Mitchell, 1968;

    Edgers et aI., 1973) have shown that, as illustrated in Figure

    26.20a, for most soils, under a sustained deviatoric stress, the

    log of strain rate

    is

    linearly decreasing with the log of time.

    Singh and Mitchell (1968) reported that the slope m of this

    linear relationship appears to be a soil property and is indepen

    dent of the deviatoric stress

    level.

    The m parameter, which can

    be obtained from laboratory creep tests, can be used to assess

    the creep potential of the soil. Values ofm smaller than 1 indicate

    a relatively high potential for accelerated creep associated with

    a strength loss that will induce a creep rupture. Bustamente

    (1980) showed that Singh and Mitchell's creep theory appears

    Ground Anchors and Soil Nails in Retaining Structures

    885

    to consistently describe the observed time-dependent anchor

    displacement under a constant load. He therefore suggested

    that the creep displacement under a sustained load can be

    esitmated using Singh and Mitchell's type equation

    Ae«T

    L ll

    = 1l

    o

    +

    - - (t

    1

    -

    m

    - 1)

    (26.2)

    1-

    m

    where

    T

    and L llo are respectively the applied pull-out force and

    the initial displacement prior to creep; A,

    IX,

    and m are interface

    creep parameters that are obtained from the experimental

    log

    l - l o g

    t and log

    f,.l

    - T curves, and

    f,./ is

    the displacement

    rate.

    Figures 26.20b and c (Bustamente, 1980) illustrate the creep

    behavior of an anchor in a plastic clay and the determination

    of the relevant interface creep parameters. The test results

    indicate a steady increase of the creep displacement almost up

    to failure, which

    is

    consistent with the m

    =

    1 value derived from

    the experimental log f,.l-t curves.

    In spite of the apparent similarity between the laboratory

    creep test results and the soil-anchor interface creep behavior

    observed in situ, more fundamental studies are required in order

    to develop a rational creep model for anchors in plastic

    fine-grained soils.

    In practice, the critical creep load of an anchor or nail

    is

    obtained from a load-controlled pull-out test following a

    standard testing and interpretation procedure (DIN 4125,1972,

    1974; Bureau Securitas, 1977; Cheney, 1984). The French

    standard testing procedure

    is

    schematically illustrated in Figure

    26.21a. Figure 26.21b shows actual results of a load-controlled

    pull-out test on an anchor in a plastic clay (Bustamente,

    1980).

    It consists of I-hour sustained load increments of O.lF (where

    Fg

    is

    the elastic limit strength of the steel tendon at which

    permanent elongation is

    0.1

    percent). For each load increment

    the anchor displacement (s) is plotted versus log time (T). An

    upward concavity of the creep curve indicates an accelerated

    creep inducing failure. The slope of the s vs.log T line

    is

    plotted

    against the applied pull-out load to determine the critical creep

    load

    Fc.

    The allowable anchor working load F

    w

    is the smaller

    of either 0.9Fc or

    0.6F

    g

    • The loading increment period can

    significantly affect the test result. Therefore, a second test

    is

    conducted that includes a 72-hour sustained loading stage at

    0.9F

    w

    to verify the long-term anchor performance.

    26.7.3 Repetitive

    loading

    Anchored structures are often subjected to repetitive (or fluctua

    ting) live loads such as tidal variations, wind or sea wave

    loadings, etc. Permanent ground anchors must be designed to

    withstand such repetitive loadings throughout the service period

    of the structure, which may include millions of cycles. Documen

    ted technical data on the long-term performance of anchors

    under repetitive loadings are still very limited. Repetitive

    loading tests on anchors for a seawall in France showed (Pfister

    et a\., 1982) that for peak cyclic load levels smaller than

    63

    percent of Po (where

    Po is

    the ultimate static pull-out capacity)

    anchor displacement became negligible after

    five cycles.

    However,

    for larger cyclic loads anchor displacement continued to increase

    at a constant or increasing rate. Begemann (1973) reported that

    repetitive uplift loads on steel H-piles in sand under cyclic load

    amplitude as low as

    35 percent of Po generated progressive

    pull-out of the piles. Laboratory model studies of repetitive

    loading on plate anchors and friction piles have been conducted

    by several investigators (Hanna et a\.,

    1978;

    Andreadis et a\.,

    1978; Hanna, 1982) and suggest some trends in the anticipated

    anchor response to cyclic loading. Specifically, Al-Mosawe

    (1979) and Hanna (1982) showed that displacement rate (per

  • 8/19/2019 juran1991

    19/38

  • 8/19/2019 juran1991

    20/38

    Ground Anchors and Soil Nails in Retaining Structures 887

    /; f..

    /;

    f..

    /; f.. log

    T

    I ~ .

    10

    6()

    71] ~ i l } ~ i l ]

    (i)

    Creep

    curves

    0.1

    (ii) Critical creep load

    (a)

    2

    - - - -e- - - -

    lOOkN 200kN

    E

    -_e-e-e-e

    300

    ekN""e_-e-e-e___

    E

    - - e - e_ - e - e_e - e - e - . _e_e_e

    -e -e ~ k N

    "

    . e e · _ ____e _ e

    E

    -e-e_e. 500kN- e - e

    __

    u

    -.-e-.-e_e 00 kN -e-e- ._e

    6

    ....

    e_ 7iXJ

    e

    kN'".-e-e-__

    .

  • 8/19/2019 juran1991

    21/38

    888

    Foundation Engineering Handbook

    1.0

    O.OOOOOll:-__

    +

    _ _

    __

    ~ - - - - , ; ; ; - l ; : = - - - : ; ~ .

    1

    Number of cycles, N

    (a)

    Number

    of cycles, N

    (b)

    Fig. 26.22

    (a)

    Effect

    of

    number

    of

    cycles on the

    rate

    of anchor displacement.

    (b)

    Effect of number

    of

    load cycles on anchor displacement.

    P

    u

    = ultimate pull-out load.

    (After AI-Mosawe. 1979.)

    load

    is

    attained. In order to determine long-term creep potential,

    each load increment

    is

    maintained until measured deflection

    is

    negligible (Le., displacement rate

    is

    smaller than a specified

    displacement increment per log cycle of time) and a I-hour

    creep test is conducted under the reference test load. Cheney

    (1984) recommends that the reference test load should be

    1.5

    times the design working load in cohesionless soils and 1.25

    times the design working load in cohesive soils. The performance

    tests are conducted on the first anchors (minimum of two

    anchors) to verify the selected installation procedure and

    provide reference data for the proof tests.

    The proof est consists of a single cycle of ncremental loading

    to the reference test loads specified above followed by unloading.

    Each load increment

    is

    maintained until measured deflection

    is negligible. The test results are compared with performance

    test results on an adjacent anchor.

  • 8/19/2019 juran1991

    22/38

    Three acceptance criteria have been established (Cheney,

    1984):

    1. To ensure that the load transfer reaches the anchor bond

    length, the deflection of the anchor head should exceed 80

    percent of the calculated elastic elongation of the unbonded

    tendon length.

    2. The total anchor deflection measured at the maximum test

    load should not exceed the calculated elastic elongation of

    the tendon length measured from the anchor head to the

    center of the bond length. This criterion (not valid for anchors

    in layered soils or for underreamed anchors) ensures that

    the center of gravity of the bond stress· distribution has not

    been transferred beyond the midpoint of the bond length.

    3. Creep displacement should not exceed 0.08 inches during

    the final log cycle of time.

    26.8 DURABILITY CONSIDERATIONS

    The long-term resistance of permanent ground anchors (or

    nails) depends primarily upon their resistivity to corrosion,

    ground aggressiveness, and groundwater compositions. Under

    ground steel corrosion is induced when a difference in potential

    exists between two points that are electrically connected and

    immersed in an electrolyte. Electrochemical cells may develop

    between the steel tendon and an external metal element or in

    local regions of inhomogeneities within the metal surface of the

    same tendon. In either case the chemical reaction between the

    groundwater and anodic regions in the exposed steel tendon

    results in time-dependent metal loss.

    For

    the corrosion process

    to occur, oxygen has to

    be

    supplied to the metal and therefore

    air-moisture solutions, specifically in industrial areas, and soil

    layers containing high oxygen content are highly corrosive. The

    major variables that affect the corrosion rate are:

    1.

    Ground aggressiveness: organic soils, and acidic or highly

    alkaline soils that contain large concentrations of soluble

    salts such as sulfates, chlorides

    or

    bicarbonates, are highly

    corrosive.

    2.

    Groundwater composition: acidic, alkaline, or salt solutions

    have high electrical conductivity, inducing high corrosion

    rate.

    3. Differential aeration: high oxygen concentration (e.g., in fill

    or near the ground surface) results in a cathodic environment,

    its local variation in the ground generates electrochemical

    cells and thereby accelerates the corrosion rate.

    4. High stresses or cyclic stresses in the steel tendon accelerate

    corrosion and may generate, in anodic environments, brittle

    stress-corrosion cracking.

    5. Environmental hazards including bimetallic action, large

    temperature changes, anerobic bacteria, and stray currents

    in the ground (i.e., currents caused by a mass transit facility,

    electrical transmission, or transport systems) will generate

    a highly corrosive environment.

    The corrosion process can develop through different mecha

    nisms, such as uniform surface corrosion,localized pit corrosion,

    stress corrosion, corrosion fatigue, and hydrogen embrittlement.

    The type of corrosion

    will

    significantly affect the degradation

    rate of the steel tendon and the efficiency of potential protection

    systems.

    Documented technical data on the long-term corrosion

    performance of ground anchors are very limited since only few

    permanent installations have been in service more than 25 years

    (Weatherby, 1982). However, performance trends can be antici

    pated on the basis of extensive research that has been conducted

    by the National Bureau of Standards (Romanoff, 1957) on

    underground corrosion.

    Ground Anchors

    and Soil

    Nails in Retaining Structures

    889

    A detailed review of corrosion-induced anchor failures by

    Weatherby (1982) yielded pertinent conclusions, specifically in

    demonstrating that:

    1.

    Quenched and tempered prestressing steels have been in

    volved in a significant number of tieback failures.

    2. The unprotected portion of the tendon just behind the anchor

    head

    is

    highly susceptible to corrosion.

    3.

    All reported failures occurred in the unbonded length of the

    tendon, mostly near the anchor head, where poor corrosion

    protection (or none) was provided.

    Based on these conclusions, FHWA recommendations for

    permanent ground anchors (Cheney, 1984) require that all

    anchors used for permanent applications be corrosion-protected

    in the unbonded length and at the anchor head. For routine

    applications, only a single degree of corrosion protection is

    required, which may consist of a grease-filled sheath along the

    free

    stressing length and grout cover (minimum 0.5-inch thick)

    in the bond zone.

    A variety of corrosion-protection systems have been devel

    oped. They mostly rely on the following basic principles

    (Weatherby, 1982; Hanna, 1982).

    Simple corrosion protection relies upon cement grout to

    generate a noncorrosive high-pH environment and protect the

    tendon in the bond zone. Plastic sheaths filled with anticorrosion

    grease, special epoxy pitch, or cement mix, and heat-shrinkable

    sleeves are commonly used for corrosion protection along the

    free stressing length.

    Coating

    with electrostatically applied resin-bonded epoxy

    can be applied to increase the corrosion protection in the bond

    zone. Intact resin-bonded coatings, being dielectric,

    will

    preclude

    the formation of galvanic cells in areas affected by microcracking.

    Complete encapsulation of the steel tendon

    is

    accomplished

    by grouting it into a uniformly corrugated plastic or steel tube

    to provide double protection. The annular space between the

    tube and the tendon

    is

    usually filled with neat cement grout

    containing admixtures to control bleeding of water from the

    grout.

    Compression steel tubes are used by European contractors

    to protect the tendon in the bond zone. The tube, which is

    high-pressure grouted into the ground, maintains the pressure

    injected grout under compression, preventing microcracking.

    The unbonded length

    is

    generally protected using a grease-filled

    PVC tube.

    Secondary grouting

    is

    commonly used to protect the unbonded

    length of the tendon. First, the anchor (primary) grout

    is

    tremied

    into the bond zone and the tendon is tested and locked-off.

    Then, the secondary (antibleed) grout is tremied around the

    unbonded length of the anchor, bonding it to the surrounding

    ground. Cheney (1984) recommends that secondary grouting

    be used only for semipermanent or low-risk applications.

    Postgrouting technique can be effectively used to provide

    repeated high-pressure grouting in the bond zone with corrosion

    protection of the tendon.

    For permanent applications of soil nailing, based on current

    experience, it

    is

    recommended (Elias and Juran, 1988) that a

    minimum grout cover of 1.5 inches be achieved along the total

    length of the nail. Secondary protection should be provided

    by

    electrostatically applied resin-bonded epoxy on the bars with

    a minimum thickness of about

    14

    mils. In aggressive environ

    ments,

    full

    encapsulation

    is

    recommended. It may be achieved,

    as for anchors, by encapsulating the nail in corrugated plastic

    or steel tube grouted into the ground. For driven nails, a

    preassembled encapsulated nail, shown in Figure

    26.23,

    has

    been developed by the French contractor Solrenfor (Louis,

    1986).

  • 8/19/2019 juran1991

    23/38

    890 Foundation Engineering Handbook

    Protective

    tube

    with injection

    holes

    Welding

    Fig. 26.23

    TBHA nail patented and developed

    by

    Solrenfor for

    permanent structu res.

    Anti-corrosion

    grease

    Anchorage

    Anchor

    head

    Bearing plate