Você está na página 1de 36

S0300-A8-HBK-010

APPENDIX APPENDIX G G
ANCHORING ANCHORING SYSTEMS SYSTEMS
G-1 INTRODUCTION
This appendix discusses anchoring systems by type, performance, holding capacity, and design, as commonly used in salvage operations. An
anchoring system consists of the anchor itself, the mooring line that transmits forces from moored vessels or pulling systems to the anchor, and
an attachment point or tensioning system on the moored vessel or work platform. The anchor line normally consists of chain, wire rope, or some
combination of the two, but may include ber line or rigid elements. The anchor provides the majority of the resistance to motion, or holding
capacity, but the contribution of portions of the anchor line buried in or lying on the seaoor may be signicant, especially with chain. The
portion of anchoring system outboard of the attachment point or tensioning system, including the anchor, anchor line, and other ancillary devices,
is sometimes called the ground leg.
Figure G-1. Simplified Anchor Types.
(c) GRAPPLING
(d) DIRECT-EMBEDMENT (e) PILE (b) DEADWEIGHT
(a) DRAG-EMBEDMENT
Anchors can be roughly divided into ve types, as shown in Figure G-1:
Drag-embedment anchors,
Deadweight anchors or clumps,
Grappling devices,
Direct-embedment anchors, and
Pile anchors.
G-1 G-1
S0300-A8-HBK-010
Suitability of the ve anchor types for various
Table G-1. Comparison of Anchor Types.
Item Deadweight Pile Direct-
embedment
Drag-
embedment
Grappling
Seafloor Material
Soft clay, mud ++ + ++ ++ o
Soft clay layer (0-20 ft) over hard layer ++ ++ o + o
Stiff clay ++ ++ ++ ++ o
Sand ++ ++ ++ ++ o
Hard glacial till ++ ++ ++ +
Boulders ++ o o o +
Soft rock or coral ++ ++ ++ + ++
Hard, massive rock ++ + + o ++
Seafloor Topography
Slope < 10 degrees ++ ++ ++ ++
Slope > 10 degrees o ++ ++ o
Loading Direction
Omnidirectional ++ ++ ++ o o
Unidirectional ++ ++ ++ ++ ++
Large uplift ++ ++ ++ o ++
Lateral Load Range
To 100,000 lbs ++ + ++ ++ ++
100,000 - 1,000,000 lbs + ++ + ++ o
Over 1,000,000 lbs o ++ o o o
++ Functions well
+ Functions well, but not normally the best choice
o Does not function well
conditions is given in Table G-1. Advantages and
disadvantages of various anchor types are given in
Table G-2. Detailed information on anchor
performance, applicability, and use can be
obtained from the Naval Civil Engineering
Laboratory (Code L 42) Port Hueneme, California,
telephone (805) 982-9419 or AUTOVON 360-
5419. Specialty anchors, such as propellant-
embedment or large drag-embedment anchors, can
be procured through the Ocean Engineering and
Construction Project Office, Chesapeake Division,
Naval Facilities Command (Code FPO-1), tele-
phone (202) 433-3881 or AUTOVON 288-3881.
Determination of holding power of deadweight,
direct-embedment, and pile anchors require an
understanding of basic soil mechanics. The
discussions of these types of anchors (Paragraphs
G-3, G-5, and G-6) draws heavily on the
information presented in Paragraph 3-7.
The information on direct-embedment and pile
anchors is presented to enable the salvage engineer
to conduct preliminary evaluations of the
feasibility of these anchor types for use in various
salvage situations. Whenever possible, the
assistance of a marine geotechnical engineer
should be sought if it appears that direct-
embedment or pile anchors may be particularly
suited to a salvage requirement.
Table G-2. Advantages and Disadvantages of Anchor Types.
Drag-embedment Anchors Deadweight Anchors Pile Anchors Direct-embedment Anchors
Advantages Advantages Advantages Advantages
High capacity (> 100,000 lbs) possible.
Broad range of types and sizes
available.
Standard, off-the-shelf equipment.
Broad use experience.
Continuous resistance can be provided
even if maximum capacity is exceeded.
Anchor is recoverable.
Resists uplift, allowing short mooring line
scope.
No setting distance is required.
Anchor is reliable because most holding
force is due to anchor mass.
Simple, on-site construction is feasible.
Size is limited only by load-handling
equipment.
Economical if material is readily available.
Reliable on thin sediment cover over rock.
Mooring line connection is easy to inspect
and service.
High capacity (>100,000 lbs) possible.
Resists uplift, allowing short mooring line
scopes.
Anchor setting is not required.
Dragging is eliminated.
Drilled and grouted piles are especially
suited for hard coral or rock seaoors.
Simple, on-site construction is feasible.
Anchor does not protrude above seaoor.
Driven piles are cost-competitive with other
high-capacity anchors when driving
equipment is available.
Comes in a wide range of sizes and
shapes, such as pipe and structural shapes.
Field modications permit piles to be
tailored to suit particular requirements.
Accurate anchor placement is possible.
Can be driven into layered seaoor.
High capacity (> 100,000 lbs) is possible.
Resists uplift, allowing short mooring line
scopes.
Dragging is eliminated.
Has higher holding capacity-to-weight ratio
than any other type.
Easier handling due to relatively light weight.
Can function on moderate slopes and hard
seaoors.
1
Easier installation due to possible instant
embedment on seaoor contact.
1
Accurate placement is possible.
Anchor does not protrude above seaoor.
Can accommodate layered seaoors or
seaoors with variable resistance.
Disadvantages Disadvantages Disadvantages Disadvantages
Anchor cannot resist uplift; large line
scopes are needed to cause near
horizontal loading at seaoor.
Does not function in hard seaoors.
Behavior is erratic in layered seaoors.
Penetrating/dragging anchor can
damage pipelines, cables, etc.
Lateral load resistance is low compared to
other anchor types.
Usable water depth is reduced;
deadweight can be an undesirable
obstruction.
Requires large-capacity load-handling
equipment for placement.
Taut moorings may aggravate ship
response to waves (low resilience).
1
Drilled and grouted installation is expensive
and requires special skills and equipment.
Costs increase rapidly in deep water or
exposed locations where special installation
vessels are required.
Special equipment (pile extractor) is
required to retrieve or refurbish the
mooring.
More extensive site data are required than
for other anchor types.
Pile-driving equipment must maintain
position during installation.
1
True for any taut mooring
Susceptible to cyclic load-strength reduction
when used in taut moorings in loose sand or
coarse silt seaoors.
For critical moorings, knowledge of soil
engineering properties is required.
Anchor typically is not recoverable.
Special consideration is needed for
ordnance.
1
Anchor cable is susceptible to abrasion and
fatigue.
1
Gun system is not generally recoverable in
deep water (> 1,000 ft)
Surface vessel must maintain position during
installation.
1
Propellent-embedded anchor
From NCEL Handbook for Marine Geotechnical Engineering, 1985
G-2 G-2
S0300-A8-HBK-010
G-2 DRAG-EMBEDMENT ANCHORS
Drag-embedment anchors (or drag anchors) are used for temporary moorings and as anchor points for beach gear, parbuckling rigs, or other
pulling systems. Drag anchors are generally known by manufacturers tradenames.
Drag-embedment anchor performance is discussed in detail in Appendix G and Paragraph 6-3.4 of the U.S. Navy Ship Salvage Manual, Volume
1 (S0300-A6-MAN-010); Paragraph 7-2.2.1 of that manual describes and illustrates the six anchorsNAVMOOR, Stato, Eells, LWT, Danforth,
and Stocklesscommonly used for Navy moorings and beach gear. The following paragraphs supplement the Salvage Manual discussion, and
provide performance data for some commercial anchors not described in the manual.
The underwater weight of an anchor is less than dry weight because of the difference in buoyancy, but performance criteria are usually based
on dry weight. Advertised or tabulated anchor weights are nominal and may differ from actual weight by as much as 15 percent; for example,
a 6,000-pound NAVMOOR anchor actually weighs about 7,200 pounds.
G-2.1 Function. A properly functioning drag-embedment anchor is lowered or dropped to the seaoor and pulled along the bottom until it
tripsrotates to a position where the uke or ukes are forced into the seabedpenetrates the seaoor, and embeds itself to the depth required
to develop its maximum holding capacity. Successful deployment depends on several factors, including anchor geometry, anchor line length,
and soil conditions. Seemingly minor features can affect anchor performance dramatically. The following paragraphs discuss the effects of
various aspects of anchor geometry on anchor performance. See Paragraph 3-7 for a discussion of soil properties.
G-2.1.1 Tripping. In general, anchors with heavy crowns, small or nonexistent tripping palms, or those with the shank/uke hinge far back
on the uke exhibit tripping problems. Examples of anchors with these inherent features and attendant lack of tripping reliability include
Stockless, LWT, and Stevx (see Figure G-8). Tripping problems occur most often in soft soils and are overcome by proper anchor selection
and deployment. Two platforms are required to deploy a ground leg so that there is a high probability of trippingone platform to pay out
the ground leg, and one to handle, lower, and position the anchor for digging in when the anchor line is tensioned. Paragraph G-2.1.1. of the
U.S Navy Ship Salvage Manual, Volume 1 (S0300-A6-MAN-010) describes this process in greater detail.
G-2.1.2 Penetration. During penetration,
Figure G-2. Forces Affecting Anchor Penetration.
W
W = ANCHOR WEIGHT
F = SOIL REACTION ON FLUKES
T = GROUND LEG TENSION
F
T

the ukes dig into the soil under the action


of the anchors weight, horizontal traction
exerted by the anchor line on the shank,
and soil reaction on the ukes and other
surfaces, as shown in Figure G-2.
Penetration is complete when the when the
upper level of the ukes is ush with the
seaoor surface. The angle formed by the
ukes and shank when fully opened (uke
angle, ) is one of the most important
factors governing anchor penetration. For
a given anchor geometry and soil
combination, there is a critical or optimum uke angle,
c
. For uke angles less than the critical angle, penetration is possible, but is hindered
by the formation of a rigid wedge of soil that adheres to the ukes and increases resistance to penetration. The soil is sheared along a line at
the outer edge of the soil wedge on the ukes; the soil failure line forms an angle with the uke. When the uke angle varies, the sum
+ remains fairly constant. The practical consequence of this process is the formation of a ball of soil that retards penetration as it is pushed
along ahead of the anchor. When the uke angle is approximately equal to the critical angle, penetration is accomplished by simple shearing
of the soil along a surface close to and approximately parallel to the uke surface. The angle is reduced to 0, and the soil wedge to a thin
boundary layer. If the uke angle is greater than the critical uke angle, the ukes rotate sharply and the rear of the anchor rises above the
seaoor, and the anchor tends to break out and may overturn and drag on one side. Without stabilizers, the anchor will slide on the side of the
ukes and not reset or penetrate.
For most articulated, reversible anchors, critical uke angle is approximately:
30 to 35 degrees in granular soils (sand, gravel),
50 degrees in soft soils (mud, silt, soft clay), and
25 to 30 degrees in stiff clays.
G-3 G-3
S0300-A8-HBK-010
Fluke roughness determines the inclination
Figure G-3. Anchor Penetration in Coarse-Grained Soils.
0
0.01 0.1 1.0 10 100
P
R
E
S
E
N
C
E
O
F
E
L
E
M
E
N
T
S
L
A
R
G
E
R
T
H
A
N
(
I
N
)
:
ANCHOR WEIGHT, TONS
ADAPTED FROM THE USE OF ANCHORS IN OFFSHORE PETROLEUM OPERATIONS,
ALAIN PUECH, GULF PUBLISHING CO., 1984
2
4
6
8
ANCHORS DO
NOT PENETRATE
ANCHORS
PENETRATE
P
E
N
E
T
R
A
T
I
O
N
H
A
M
P
E
R
E
D
B
Y
P
R
E
S
E
N
C
E
O
F
L
A
R
G
E
E
L
E
M
E
N
T
S
(B) of the soil reaction on the uke. For
rough ukes (cast steel, ribbed ukes, etc.),
the stress inclination is essentially the same
as the soils internal friction angle. For
smooth ukes (rust-free plate), the stress
inclination is much smaller than the friction
angle and seldom greater than 25 degrees.
The smaller stress inclination makes the
uke more resistant to the formation of the
rigid soil wedge; critical uke angle is
greater, broadening the range of soils for
which the anchor is suited.
Anchor penetration is also inhibited by the
presence of large soil elements and the
relative stiffness of cohesive soils. Figures
G-3 and G-4 illustrate the probability of
anchor penetration as functions of anchor
weight and soil features.
G-2.1.3 Burial. After initial penetration
of a properly functioning anchor, the
tractive force exerted by the anchor line
mobilizes soil forces (drag) on various
surfaces of the anchor:
Because of geometry and
angle of attack, the vertical
Figure G-4. Anchor Penetration in Stiff Soils.
0
0.01 0.1 1.0 10 100
S
O
I
L
C
O
H
E
S
I
O
N
(
L
B
/
F
T
2
)
ANCHOR WEIGHT (TONS)
1,000
2,000
3,000
4,000
ANCHORS DO
NOT PENETRATE
ANCHORS
PENETRATE
P
E
N
E
T
R
A
T
I
O
N
P
O
S
S
S
I
B
L
E
B
U
T
N
O
T
G
U
A
R
A
N
T
E
E
D
ADAPTED FROM THE USE OF ANCHORS IN OFFSHORE PETROLEUM OPERATIONS,
ALAIN PUECH, GULF PUBLISHING CO., 1984
component of the soil
resistance to horizontal
movement on the ukes is
directed downwards.
The vertical component of
drag forces on the shank,
anchor line, stocks/stabilizers,
palms, etc., is directed
upwards.
The anchor will bury itself until it reaches
a depth where the upward forces balance
the downward forces. Resistance forces
increase rapidly with depth, as soil strength
increases and length of embedded anchor
line increases. As burial depth increases,
the anchor rotates because the shank tends
to align itself with the anchor line and the
downward component of the soil forces on
the ukes decreases. Most reversible
anchors cannot be buried in dense sands or
stiff clays because soil shear strength is too
high to permit penetration of the shank,
anchor line, and stabilizers. Specialized
anchors have been developed that are
capable of burial in stiff soils.
Fluke angle affects burial in much the same way that it affects penetration. Optimum burial angle is generally slightly smaller than critical
penetration angle.
Overall streamlining enhances anchor burial. The crown, palms, shank, and stock or stabilizers especially inhibit burial, especially if they are
located forward on the anchor so as to interfere with the plastic ow of soil over the ukes. Since these features are necessary to the functioning
of the anchor, they cannot be eliminated. In some anchorssuch as the Hookcrown, palms, and stabilizers are placed to the rear of the
anchor, under the uke, to minimize resistance to burial. Anchors of this type are unilateral, and must be lowered to the seaoor to ensure
tripping and penetration. The bearing area of the underside of the shank is the major surface resisting burial, and the rst to come into play
in most anchors. The greater the bearing area of the shank, the greater the resistance to burial. In general, if the ratio of shank bearing area
to uke area is greater than 0.13, burial is not possible. Minimum shank cross-sectional area is set by design holding capacity; bearing area
is minimized in various anchor designs by using short shanks and/or beveling the edges of the shanks at about 45 degrees.
G-4 G-4
S0300-A8-HBK-010
Anchor line characteristics affect burial
Figure G-5. Holding Capacity Plots for Different Anchor Behavior.
RELATIVE DRAG DISTANCE, L/F
RATED
EFFICIENCY
E
F
F
I
C
I
E
N
C
Y
e
=
T
/
W
2
1
2
3
4
5
4 6 8
SOFT CLAY
TRIPPING
DEFECT
INSTABILITY
PENETRATION WITHOUT
BURIAL
DEEP BURIAL
W = ANCHOR WEIGHT
F = FLUKE LENGTH
T = ANCHOR LINE TENSION
L = DISTANCE
ANCHOR DRAGGING AT
CONSTANT DEPTH
AND TENSION
ADAPTED FROM THE USE OF ANCHORS IN OFFSHORE PETROLEUM OPERATIONS,
ALAIN PUECH, GULF PUBLISHING CO., 1984
depth signicantly. For equal breaking
strengths, the maximum cross section of
chain is several times more than that of
wire rope; chain will create more drag and
resist burial to a greater degree than wire
rope. Anchors on wire rope will achieve
greater burial depth, and therefore greater
holding capacity than the same anchor on
chain. The increase in holding capacity of
the anchor may be offset by the decrease in
the holding capacity of the buried portion
of the anchor line.
Anchors and/or the anchor line can be tted
with special burial devices that increase the
area producing downward components of
soil resisting forces. Typical devices
include auxiliary plates tted to the forward
part of the shank (making the anchor
unilateral), uke extenders, and anchor line
depressors. Anchor line depressors are
streamlined steel castings that enhance
burial of the anchor line, with an active
area of about 30 percent of the anchors
uke area, and weight of about 10 percent
of anchor weight. One manufacturer claims
that three depressors in series can double
the holding power of a Bruce anchor.
Burial and development of holding capacity
can be monitored with a tensiometer. As the ground leg is tensioned, tension is plotted against drag distance as shown in Figure G-5. If tension
increases steadily (exponentially), the anchor has tripped and is penetrating. If the tension indicator uctuates or remains static, tripping and/or
burial is incomplete. Dragging should be halted and the anchor reset.
G-2.1.4 Stability. A stable anchor can be pulled over long distances (several times its dimensions) without lying on its site, overturning, or
breaking out of the seabed. After penetration and burial, traction force is essentially constant. There are two types of instability:
Lateral instability rotation about the axis of the shank/anchor line, occurring most often during penetration, but also during
dragging, causing the anchor to move upwards and possibly break out of the seabed.
Vertical instability the anchor moves upward in a vertical plane dened by the shank/anchor line at the end of penetration or
while dragging, either by translation or rotation about a horizontal axis through the shank attachment point.
Lateral instability is usually related to anchor geometry. Anchors with very long and narrow ukes, such as the LWT and Danforth, are basically
unstable, and are tted with long stabilizers to prevent rotation. Wide uke spacing gives good stability in homogeneous soils, but anchors with
tapered, closely spaced ukes are more stable in heterogeneous soils because there is greater probability that both ukes will operate in soil
of the same strength and consistency. Closely spaced ukes are a disadvantage if the soil contains coarse elements large enough to jam between
the ukes. Fouling by wire rope or other obstructions, especially over only one uke, will cause serious instability. Heavy crowns cause
overturning during penetration, especially in heterogeneous soils. Manufacturing defects or damage, especially those causing asymmetry between
ukes, cause rotational moments as the anchor is dragged, due to the difference in soil forces generated on the ukes. Twisted or bent shanks
or stabilizers and asymmetrically attached pendants can also cause instability.
Lateral instability is avoided or overcome by structural arrangement of the anchor, including the addition of specic components. Stocks at the
forward end of the shank, or stabilizers at the rear of the ukes prevent lateral rotation, and are effective if properly dimensioned. Their
effectiveness is reduced in very soft or uid soils. Stocks or stabilizers add to the bulkiness of the anchor, hamper handling and stowage, and
inhibit burial. Bent, broken, or fouled stabilizers can induce instability by creating laterally unsymmetrical soil forces. Some anchors, such as
the Flipper Delta, Eells, and AC-14 are constructed with plate surfaces in planes different from that of the ukes to provide stability. Certain
nonarticulating anchors, such as the Bruce, are designed to be self-stabilizingwhen the anchor begins to rotate, the change in attitude increases
soil force on the rising side and decreases soil forces on sinking side, returning the anchor to its upright position.
G-5 G-5
S0300-A8-HBK-010
Vertical instability is related to the point of
Figure G-6. Burial Behavior Versus Articulation.
DRAG DISTANCE
B
U
R
I
A
L
D
E
P
T
H
ATTACHMENT POINT/ARTICULATION AXIS:
BEHIND BARYCENTER
NEAR BARYCENTER
FORWARD OF BARYCENTER
1
2
3
1
2
3
ADAPTED FROM THE USE OF ANCHORS IN OFFSHORE PETROLEUM OPERATIONS,
ALAIN PUECH, GULF PUBLISHING CO., 1984
Table G-3. Power Law Coefficients.
Anchor type
a
Soft soils, soft clays and silts Hard soils, sands and stiff clays
H
R
(lbs 1,000) b H
R
(lbs 1,000) b
BOSS 210 0.94 270 0.94
BRUCE Cast 32 0.92 250 0.80
BRUCE Twin Shank (TS) 189 0.92 210 0.94
BRUCE Flat-fluke Twin Shank (FFTS) 250 0.92 -
b
-
b
Danforth 87 0.92 126 0.80
Flipper Delta 139 0.92 -
b
-
b
G.S. (AC-14) 87 0.92 126 0.80
Hook 189 0.92 100 0.80
Lightweight (LWT) 87 0.92 126 0.80
Moorfast 87 0.92
60
100
c
0.80
0.80
NAVMOOR 210 0.94 270 0.94
Offdrill II 87 0.92
60
100
c
0.80
0.80
STATO 210 0.94
250
d
190
g 0.94
STEVDIG 139 0.92 290 0.80
STEVFIX 189 0.92 290 0.80
STEVIN 139 0.92 165 0.80
STEVMUD 250 0.92 -
e
-
e
STEVPRIS (straight shank) 189 0.92 210 0.94
Stockless (fixed fluke) 46 0.92
70
44
f 0.80
Stockless (movable fluke) 24 0.92
70
44
f 0.80
See Figures G-6 and G-7 for plotted results.
a
Fluke angles set for 50 degrees in soft soils and according to manu-
facturers specifications in hard soils, except when otherwise noted
b
No data available
c
For 28-degree fluke angle
d
For 30-degree foot angle (STATO)
e
Anchor not used in this seafloor condition
f
For 48-degree fluke angle
g
For dense sand conditions (near shore)
From NCEL Handbook for Marine Geotechnical Engineering, 1985
application of the tractive force relative to
the barycenter (center of soil pressure on
the uke surfaces). In a freely articulating
anchor (uke-to-traction-line angle not
constrained) the burial path of the anchor is
determined by the relative position of the
articulation axis and the barycenter, as
shown in Figure G-6. Free articulation
achieves maximum burial depth, but is
seldom used because anchor stability is
very sensitive to the location of the
articulation axis. Instead, tractive force is
applied through a rigid shank so that
articulation occurs above the barycenter and
the shank can be blocked to prevent uke
angle from exceeding the critical
penetration/burial angle. As the shank is
shortened, the articulation axis (anchor line
to shank attachment) moves closer to the
barycenter, and the critical angle increases.
Burial depth increases, but the anchor
becomes more sensitive to the soil type and
local soil heterogeneities.
G-2.2 Holding Capacity. Resistance to
horizontal movement, or anchor holding
capacity, increases with burial depth
because deeper soil is generally denser and
stronger, and provides increased resistance
to the anchor moving through it. Holding
power is primarily determined by the mass
of the displaced soil. Therefore, deeply
embedded anchors with a large uke area
perform better because they cause the
largest displacement of strongest soil.
Stable anchors provide constant holding
capacity once they reach maximum burial
depth, even if dragged. For an anchor to
develop its maximum holding capacity, it
must penetrate to its optimum depth. If the
anchors burial is halted by a hard layer, it
will drag along the layer, providing
constant resistance (holding capacity) less
than maximum capacity. Until an anchor
reaches its equilibrium depth, burial depth
is a function of drag distance. If drag
distance is to be limited (to avoid fouling
submarine cables, for example), the anchor
cannot be loaded to full capacity.
The Power Law Method is the best
technique to predict holding capacity.
Holding capacity, H
m
, is determined by:
where:
H
m
= H
r

,
W
a
10,000
b
H
r
= holding capacity of a 10,000-pound (dry weight) version of the reference anchor, from Table G-3
W
a
= anchor weight for which H
m
is to be determined, lbs
b = an exponent depending on the anchor and soil type, from Table G-3
G-6 G-6
S0300-A8-HBK-010
Figure G-7. Anchor Holding Capacity.
ANCHOR AIR WEIGHT (x 1,000 POUNDS)
CAPACITY IN SAND CAPACITY IN MUD
A
N
C
H
O
R
H
O
L
D
I
N
G
C
A
P
A
C
I
T
Y
(
x
1
,
0
0
0
P
O
U
N
D
S
)
1,000
400
500
600
700
800
900
300
200
100
90
80
70
60
50
40
30
20
10
1 2 3 4 5 6 7 8 9 10 20 30 40
D
A
N
F
O
R
T
H
,
L
W
T
S
T
E
V
P
R
I
S
(
S
T
R
A
I
G
H
T
S
H
A
N
K
)
,
B
R
U
C
E
T
S
E
E
L
L
S
,
S
T
O
C
K
L
E
S
S
-
3
5

F
L
U
K
E
A
N
G
L
E
M
O
O
R
F
A
S
T
,
O
F
F
D
R
I
L
L
I
I
-

2
8

F
L
U
K
E
A
N
G
L
E
,
H
O
O
K
S
T
O
C
K
L
E
S
S
-
4
8

F
L
U
K
E
A
N
G
L
E
S
T
A
T
O
-
3
0

F
L
U
K
E
A
N
G
L
E
B
O
S
S
,
N
A
V
M
O
O
R
B
R
U
C
E
C
A
S
T
S
T
E
V
F
I
X
,
S
T
E
V
D
I
G
FLUKE ANGLES SET FOR SAND AS
PER MANUFACTURERS SPECIFICATION
ADAPTED FROM NCEL HANDBOOK FOR MARINE GEOTECHNICAL ENGINEERING, KARL ROCKER JR, 1985
ANCHOR AIR WEIGHT (x 1,000 POUNDS)
1 2 3 4 5 6 7 8 9 10 20 30 40
FLUKE ANGLES SET FOR MUD AS
PER MANUFACTURERS SPECIFICATION
N
A
V
M
O
O
R
,
S
T
A
T
O
,
B
O
S
S
B
R
U
C
E
F
F
T
S
,
S
T
E
V
M
U
D
D
A
N
F
O
R
T
H
,
A
C
1
4
,
L
W
T
,
M
O
O
R
F
A
S
T
,
O
F
F
D
R
I
L
L
I
I
F
L
I
P
P
E
R
D
E
L
T
A
,
S
T
E
V
I
N
,
S
T
E
V
D
I
G
B
R
U
C
E
T
S
,
H
O
O
K
,
S
T
E
V
F
I
X
,
S
T
E
V
P
R
I
S
E
E
L
L
S
,
S
T
O
C
K
L
E
S
S
(
F
I
X
E
D
-
F
L
U
K
E
)
S
T
O
C
K
L
E
S
S
(
M
O
V
A
B
L
E
-
F
L
U
K
E
)
B
R
U
C
E
C
A
S
T
This method produces a straight-line relationship between anchor holding capacity and anchor weight on a log-log plot, as shown in Figure G-7
for various commercial and Navy anchors.
Holding power for a given anchor can also be calculated by multiplying the anchors weight by its efficiency (also called holding power factor).
H = We
where:
H = holding power, lbs
W = anchor dry weight, lbs
e = anchor efficiency, dimensionless
Caution must be applied when determining holding power by anchor efficiency, as holding power is not a linear function of anchor weight.
Anchor efficiencies are valid only for the specied weight. If the efficiency for a given anchor weight is used to predict performance of a larger
anchor of the same type, holding power will be overestimated.
When a single anchor will not develop the required holding capacity, it is common practice to install a second anchor on the same mooring leg
in a piggyback rig. Anchors can be rigged in tandem or doubled. Paragraph 6-3.4.3 of the U.S Navy Salvage Manual, Volume 1 (S0300-A6-
MAN-010) discusses double and tandem anchor rigs in detail. When properly rigged and deployed, tandem anchors can develop more than twice
the capacity of a single anchor in the same soil. Holding capacity can also be increased by use of large weights, or sinkers, attached to anchor
lines to absorb energy and ensure horizontal loading on the anchor. If the sinker-to-anchor-line connection fails, the sinker will be lost and the
entire mooring may fail. Connections must allow free movement of chain links in order to avoid distortion or failure of links. Sinker shackles,
shown in Figure 7-32, or appropriately dimensioned plate shackles should be used to allow free movement of chain links.
G-7 G-7
S0300-A8-HBK-010
Table G-4. Anchor Selection.
Anchor
Soil Type
Anchor
Soil Type
Sands Muds Stiff Clays Heterogeneous Sands Muds Stiff Clays Heterogeneous
Multipurpose Anchors Specific Anchors
Improved Stockless + o ++ ++ Delta Triple ++ o + +
High-performance Stockless ++ + + + Doris mud o ++ o o
Stock ++ o + ++ Hook ++ ++ o o
Stevin ++ + o o Stevshark + o ++ ++
Stockless + o + + Bruce ++ o + +
Flipper Delta ++ ++ o o
++ Functions well
+ Functions, but not the best choice
o Does not function well
From The Use of Anchors in Offshore Petroleum Operations, A. Puech, 1984
G-2.3 Selection. Anchor selection is a two-step process, in which:
One or more anchor types are chosen for use based on overall performance in the expected soil, availability, and cost.
The selected anchor type is sized to develop the required holding capacity.
To aid the selection process, Table G-4 lists anchor applicability by soil type. Table G-5 rates anchor types subjectively, based on eld
experience and test ndings with small anchors.
Table G-5. Rating of Drag-embedment Anchor Types.
Anchor Type
Reliability
Cohesive Soils (clays and plastic silts) Cohesionless Soils (sands)
a
Tripping/Dig-in Stability Holding Capacity Tripping/Dig-in Stability Holding Capacity
Stockless
b
(movable uke) Low Medium Low High Medium Low
Stockless
b
(xed uke) High Medium Low High High Low
G.S. (AC 14)
c

c
Medium High Medium Medium
Danforth Medium Low Medium High Medium Medium
Lightweight (LWT) Low Low Low High Medium Medium
STATO/NAVMOOR
d
High Medium High High High High
Moorfast Medium Medium Medium Medium Medium Medium
Offdrill II Medium Medium Medium Medium Medium Medium
Flipper Delta
c

c
Medium
c

c
Medium
STEVIN
c

c
Medium
c

c
Medium
STEVFIX Low Low High High Medium High
STEVPRIS
c

c
High High High
STEVDIG
c

c
High Medium High
STEVMUD High
c
High
e

e
BOSS High Medium High High
c
High
Hook High High Medium Medium High Medium
BRUCE Cast High High Low High High High
Bruce Twin-shank High High High
c
High High
a
Fluke angle set at manufacturers recommendation for sand
b
With stabilizers (ratings not as high without stabilizers)
c
Insufficient data available for rating
d
Fluke angle set at 30 degrees for sand
e
Anchor not normally used in this seaoor
From NCEL Handbook for Marine Geotechnical Engineering, 1985
G-8 G-8
S0300-A8-HBK-010
G-2.4 Drag Anchor Types. Physical
Figure G-8. Drag-Embedment Anchors.
DEEP PENETRATION ANCHORS
ELBOWED SHANK ANCHORS
STEVIN ANCHORS
HIGH PERFORMANCE STOCKLESS ANCHORS
IMPROVED STOCKLESS ANCHORS
STANDARD STOCKLESS ANCHORS
STOCKED ANCHORS
STEVSHARK
BRUCE
STEVDIG/
STEVIN
BRUCE T.S.
STEVFIX
DANFORTH
ADMIRALTY
AC-14
NAVY
STOCKLESS
ADMIRALTY
AM-7
SINGLE-FLUKE
STOCK DREDGER
MOORING
ANCHOR
BEIJERS HALLS SPECK
STOKES SNUGSTOW WELDHOLD
L.W.T. BOSS
MOORFAST/
STATO/
NAVMOOR
HOOK
STEVMUD
ADMIRALTY
AC-12
DELTA
FLIPPER
DELTA
characteristics of an anchor affect its
tripping, penetration, stability, and overall
performance. Anchors can be classed by
general characteristics such as uke/shank
angle (typical anchors for each class are
shown in Figure G-8); the presence and
geometry of stabilizers, crowns, tripping
palms; and hinge function:
Deeply penetrating anchors of
unique geometry with holding
power roughly proportional to
the third power of penetration
such as Stevshark and Delta
(see Paragraphs G-2.5.1 and
G-2.5.2).
Anchors with elbowed shanks
for deep penetration, such as
Bruce, Hook, and AC-12 (see
Paragraphs G-2.5.3 and G-
2.5.4).
Anchors with large, hollow
ukes, hinges near the center
of gravity, and relatively
short shanks and stabilizers,
known generically as Stevin
anchors (see Paragraph G-
2.5.5).
High-performance stockless
anchors with hinge and
stabilizers at the rear and
relatively long shanks and
stabilizers such as Moorfast,
O f f d r i l l , S T A T O ,
NAVMOOR, LWT, etc. (see
Paragraph G-2.5.6 and the
U.S. Naval Ship Salvage
Manual, Volume 1 (S0300-
A6-MAN-010)).
Improved stockless anchors
with short, thick stabilizers,
hinges at the rear and
relatively short, more or less
square shanks, such as the
AC-14, Stokes, and similar
anchors (see Paragraph G-
2.5.7).
Standard stockless anchors,
such as the Navy Stockless
and similar commercial pat-
terns (see Paragraph G-2.5.8).
Stocked anchors with small
uke area and stabilizers at
the front of the shank such as
the Stock (old-fashioned or
Admiralty), Single Fluke
Stock, Dredger, etc. (see
Paragraph G-2.5.9).
G-9 G-9
S0300-A8-HBK-010
G-2.5 Drag Anchor Notes. The following notes describe features specic to some commonly used drag anchors, such as tripping, penetration,
burial, and effectiveness in various soil types. The NAVMOOR, STATO, LWT, Danforth, Eells, and Navy Stockless anchors, described in
Paragraph 7-2.2.1 of the U.S. Navy Salvage Manual, Volume 1 (S0300-A6-MAN-010), are not addressed here.
G-2.5.1 Stevshark. The Stevshark is a ballastable anchor with xed twin-shank tted with teeth. It is used primarily for temporary and
permanent moorings in stiff clays and heterogeneous sediments, but in practice tripping is generally incomplete in very stiff clays due to teeth
hooking. Burial is impossible in very stiff soils. The anchor breaks out easily.
G-2.5.2 Delta. The Delta is a compact, unilateral, pointed-uke anchor which exhibits excellent tripping and stability in all soil types. Delta
anchors embed deeply in soft soils and are recommended for temporary moorings in mud and sand seaoors. A variation known as the Delta Triple
consists of three Delta anchors welded together. The Delta Triple penetrates rapidly and deeply, but is very bulky and difficult to handle. Delta
Triple anchors are generally used for permanent moorings in sandy soils. The Flipper Delta is a bilateral anchor with cut-out, pointed ukes in the
shape of right triangles with the right angles against the shank. The crown is open and cage-like with the outboard perimeters formed by angled
plates that act as stabilizing ns. Stability and tripping reliability are excellent in all types of soil; burial is deep in soft soils, but hampered by the
crown in stiff or heterogeneous soils. The anchor is bulky on deck, but resists fouling as there are no projecting parts other than the shank.
G-2.5.3 Bruce. The Bruce is a rugged, nonarticulating, self-stabilizing anchor, usually constructed of cast steel. Penetration is satisfactory
in clays, and good for temporary moorings in sand; stability is very good. A welded steel version, with two cross-stiffened shanks is known
as the Bruce TS (twin-shank).
G-2.5.4 Hook. The Hook anchor is a nonreversible, large-area plate anchor with a curved shank, used primarily for permanent moorings in
sand and mud. The Hook anchor trips satisfactorily, but penetration is difficult, and stability is poor in stiff soils.
G-2.5.5 Stevin. Stevin anchors are distinguished by several features:
Strong streamlined shanks, hinged near the anchors barycenter, to enhance penetration.
Hollow ukes to give maximum uke area for anchor weight.
Stabilizers protruding from the ukes to overcome instability inherent in hinging the shank near the barycenter.
Fluke-to-shank angle adjustable for different soil conditions.
Crowns specially designed to enhance tripping.
The Stevin anchor family includes the basic Stevin, Stevdig, Stevx, and Stevmud anchors (the Stevshark is produced by the same manufacturer
as the Stevin anchors, but is a distinct anchor type, as discussed in Paragraph G-2.5.1). Tripping is satisfactory in all soils. Penetration is very
good in sandy and soft seaoors, but difficult in stiff soils. Stability is generally good. Stevin anchors are characterized by reliable holding
power, minimum weight, and easy handling.
The Stevx is based on the basic Stevin design, with a considerably enlarged uke area (about 64 percent). Stock stabilizers combined with
the uke points prevent the anchor from sliding over the bottom on its side. Holding power in mud is about 30 times anchor weight. For very
soft mud, an adapter can be tted to the anchor to increase uke area by 80 percent.
The Stevmud has enlarged ukes to increase holding capacity in very soft mud. The uke area is twice that of the basic Stevin (see Figure
G-8); holding power in mud is about 35 times anchor weight.
G-2.5.6 Moorfast. The Moorfast is a cast version of the Stato anchor, but of heavier construction. The uke area is half that of the Stato,
and holding power is about 14 times weight. Moorfast anchors are provided with wedges to set uke angle for 32 degrees in sand and 50
degrees in mud. The Moorfast crown is suitable for use in mud without modication. T-ATF-166 Class tugs carry a large Moorfast anchor
hawsed in the stem. The Offdrill anchor is similar to the Moorfast, but with slightly smaller ukes.
G-2.5.7 Improved Stockless Anchors. Most of the anchors in this class have holding capacities of about 8 times anchor weight in soft soils,
and from 10 to 15 times anchor weight in sand or rm clay. The AC-14 is used as a ships anchor by the Royal Navy and is approved by
Lloyds and ABS. Fluke angle is 34 degrees. Stabilizer ns ensure continuous penetration. The Danforth-Jackson Stokes anchor is similar.
G-10 G-10
S0300-A8-HBK-010
G-2.5.8 Standard Stockless. Stockless anchors were designed for use as ships anchors. Consequently, they are more easily recovered but
less efficient than higher performance mooring anchors. Fluke angle is set at 45 to 48 degrees for most stockless types. Holding efficiency
is about 4.5 times the anchor weight for U.S Navy Stockless, and 2 to 3 times the anchor weight for most commercial Stockless anchors.
Performance is enhanced by tting stabilizers, and by welding or blocking ukes open at the critical angle for the soil.
G-2.5.9 Stock Anchors. With the stabilizing stock forward of, and at right angles to the uke(s), burial is impossible with stock anchors; the
relative orientation of uke and stock does ensure reliable tripping and penetration, however. They are capable of signicant holding capacity,
often as high as 15 times anchor weight in rm clay or sand. They can resist moderate uplift because the embedded uke digs in like a pickaxe
when the end of the shank is raised. Because of its grappling ability, the old-fashioned Stock and similar anchors can hold on coral or rocky
ground. The single blade salvage anchor, shown in Figure G-9, is commonly used by commercial salvors and is a modication of the admiralty
pattern AM-7 single uke anchor. The
Figure G-9. Specialized Anchors for Soft (Mud) Seafloors.
SINGLE-BLADE
SALVAGE ANCHOR DORIS MUD ANCHOR
enlarged uke area increases holding
capacity in soft soils.
G-2.5.10 Doris Mud. The Doris Mud
anchor shown in Figure G-9 is a unilateral,
nonarticulating anchor especially designed
for soft soils, and does not readily t into
any of the anchor classes dened in
Paragraph G-2.4. Stability is good, but
penetration and burial are possible only in
very soft soils where the large area of the
bulldozer-like blade resists forward motion.
The anchor is bulky and difficult to handle
and break out.
G-3 DEADWEIGHT ANCHORS
Any heavy object that can be placed on the seaoor can be used as a deadweight anchor. Steel, concrete, and ferro-cement clumps are
commonly used. Factors to consider in selection of deadweight anchors and their installation include:
Water depth.
Seaoor slope.
Presence and rate of soil erosion.
Degree of sediment consolidation.
Signicant characteristics of common deadweight anchors are shown in Figure G-10 (Page G-12).
G-3.1 Holding Capacity. Holding power of a deadweight anchor is the force required to lift or drag the large weight over the sea bottom.
Resistance to uplift or vertical force is simply the submerged weight of the anchor, plus suction effects in soft bottoms. Resistance to dragging
results from friction between the seaoor and the anchor. See Paragraph 3-7 for a discussion of basic soil properties and calculations.
Lateral loads result from several causes:
Mooring line tension,
Down-slope force of gravity on a sloping seaoor,
Current drag, and
Storm-wave or earthquake loading.
G-11 G-11
S0300-A8-HBK-010
Figure G-10. Deadweight Anchors.
(a) SINKER
EFFICIENT UPLIFT
EASY TO HANDLE
(b) SQUAT CLUMP
LOW OVERTURNING
MORE AREA CONTACTING SOIL
(c) RAILROAD RAILS
OR SCRAP IRON
LOW BULK, HIGH WEIGHT
LOW COST
(d) CONCRETE SLAB
WITH SHEAR KEYS
HIGH LATERAL CAPACITY
SCOUR CONTROL
(f) MUSHROOM
SHALLOW BURIAL
(g) WEDGE
SHALLOW BURIAL
LOW OVERTURNING
UNI-DIRECTIONAL
(h) SLANTED SKIRT
DEEPER BURIAL
UNI-DIRECTIONAL
(i) HIGH LATERAL
CAPACITY, FREE FALL
FREE-FALL
INSTALLATION
HIGH LATERAL
CAPACITY
(e) OPEN FRAME WITH
WEIGHTED CORNERS
HIGH LATERAL CAPACITY
REDUCED LOWERING
LINE DYNAMIC TENSIONS
SHALLOW BURIAL
(j) FREE FALL (DELCO)
(k) DEAD WEIGHT GEOMETRY
FREE FALL
INSTALLATION
EFFICIENT UPLIFT
ADAPTED FROM NCEL HANDBOOK FOR MARINE GEOTECHNICAL ENGINEERING, KARL ROCKER JR, 1985
D
f
z
s
H
B
SHEAR
KEYS
G-3.1.1 Static Short-term and Cyclic Loading in Cohesive Soils. Static short-term loading and cyclic lateral loading on cohesive soils are
treated as undrained failure problems. The maximum lateral load capacity (parallel to the seaoor), Q
ul
, for an anchor on cohesive soil is:
where:
Q
ul
= s
uz
A + 2s
ua
D
f
B
s
uz
= undrained shear strength of the soil at depth D
f
, [force/length
2
]
s
ua
= average undrained shear strength between the seaoor and depth D
f,
[force/length
2
]
A = foundation or anchor base area, [length
2
]
D
f
= embedment depth of foundation or anchor (depth of shear key tip below the seaoor), [length]
B = minimum foundation or anchor base dimension (usually called the foundation width), [length]
G-12 G-12
S0300-A8-HBK-010
G-3.1.2 Static Short- and Long-term Loading in Cohesionless Soils. For cohesionless soils, lateral load failure is a drained soil failure, and
the maximum lateral load capacity in sliding is:
where:
Q
ul
= W
b
+
b
AD
f
F
ve
cos F
h
sin + R
p
= coefficient of friction between anchor base and soil or between soil and soil when shear keys cause deep failure (shear keys
are described in Paragraph G-3.1.3)
W
b
= underwater (buoyant) weight of anchor, [force]
F
ve
= design environmental loading and mooring line loading in the vertical direction (upward is positive), [force]
F
h
= design environmental loading and mooring line loading in the horizontal direction (down-slope is assumed positive), [force]

b
AD
f
= buoyant weight of soil trapped in shear keys, [force]

b
= buoyant unit weight of soil, [force/length
3
] (see Table 3-3, Page 3-18)
A = base area of the anchor, [length
2
]
D
f
= embedment depth of shear keys, below the seaoor surface, [length]
= seaoor slope angle
R
p
= passive soil resistance on leading edge of base, [force]
The coefficient of friction depends on soil type and anchor
Table G-6. Coefficient of Friction for Deadweight Anchors.
Soil
Internal
Friction
Coefficient
Smooth
Steel
Rough
Steel
Smooth
Concrete
Rough
Concrete
Smooth
PVC
Quartz Sand 0.67 0.27 0.60 0.60 0.69 0.33
Coralline sand 0.67 0.20 0.63 0.63 0.66 0.20
Oolitic Sand 0.79 0.23 0.56 0.58 0.74 0.26
Foraminiferous Sand-Silt 0.64 0.40 0.66 0.67 --- 0.40
From NCEL Handbook for Marine Geotechnical Engineering, 1985
roughness. Table G-6 gives coefficients of friction for typical
materials and marine cohesionless soils. In the absence of
better information, the internal coefficient of friction can be
estimated from the soil friction angle :
= tan ( - 5) degrees for a rough steel or concrete
base without shear keys
= tan for a base with shear keys
Friction angle is given for various soils in Table 3-3 on
Page 3-18.
When the anchor is embedded deeply or tted with shear keys, a wedge of soil in passive failure develops in front of the leading foundation
edge and provides resistance to sliding. In some cases, this passive wedge can contribute about 10 percent of the total lateral resistance.
Because the sediment comprising the passive wedge may be removed by current scour or by animal burrowing, the contribution of the passive
wedge to sliding resistance is usually neglected.
To maintain stability against sliding, a factor of safety, F
s
, can be applied to the lateral load capacity, Q
ul,
to account for uncertainties in soil
data or failure mechanism:
For most applications, a factor of safety of 1.5 to 2 is adequate. The minimum underwater (buoyant) weight of the anchor is derived from the
F
s
=
Q
ul
(W
b

b
AD
f
F
ve
) sin + F
h
cos
maximum lateral load capacity equation, assuming R
p
= 0:
For level seaoors, is 0 and anchor buoyant weight is:
W
b
=
(F
s
+ tan) F
h
F
s
tan
+ F
ve

b
AD
f
Anchors with skirts but without shear keys are more likely to slide along the foundation base rather than at the depth of the skirtthe buoyant
W
b
=

,
F
s
F
h

+ F
ve

b
AD
f
weight of soil in the skirts/keys (
b
AD
f
) is neglected.
G-13 G-13
S0300-A8-HBK-010
G-3.1.3 Shear Keys. Shear keys are
Figure G-11. Soil Failure Modes for Sliding Deadweights.
(a) SLIDING BASE FAILURE
POTENTIAL
FAILURE
PLANES
SEAFLOOR
POTENTIAL
FAILURE
PLANES
POTENTIAL
FAILURE
PLANES
(b) DEEP PASSIVE FAILURE (c) PASSIVE WEDGE FAILURE
F
h
z
s
B
ADAPTED FROM NCEL HANDBOOK FOR MARINE GEOTECHNICAL ENGINEERING,
KARL ROCKER JR, 1985
vertical plates added to the base of
deadweight anchors to increase lateral load
capacity by forcing the failure surface (the
surface on which the anchor slides), deeper
into the seaoor, where stronger soils resist
higher lateral loads. Three possible failure
modes for shallow foundations tted with
shear keys are shown in Figure G-11.
Shear keys should be placed close enough
to each other to force sliding failure to
occur at the base of the shear keys, as
shown in Figure G-11(a).
The number of shear keys, n, required in
each direction is computed by comparing
the design load parallel to the seaoor to the passive resistance developed per key:
where:
n
F
s
F
hp
W
b
sin
R
p
1
F
hp
= resultant of applied loads in the down-slope direction, [force]
F
s
= safety factor
W
b
= buoyant weight of the anchor, [length]
= seaoor slope
Minimum shear key spacing should equal the shear key depth for cohesive soils and twice the shear key depth for cohesionless soils. The
passive resistance developed by one shear key, R
p
is:
where:
R
p
=

1
1
1
]

b
z
2
s
2
+ 2s
ua
z
s
B (cohesive soils)
R
p
=
K
p

b
z
2
s
B
2
(cohesionless soil)
z
s
= shear key depth below base of the anchor, [length]
s
ua
= average undrained soil shear strength between the foundation base and the tip of the key, [force/length
2
]

b
= soil buoyant unit weight, [force/length
3
]
B = shear key width, [length]
K
p
= coefficient of passive lateral earth pressure
= tan
2
(45
o
+ 0.5)
= soil internal friction angle (see Table 3-3)
In cohesionless soils, a shear key depth of 0.05B is appropriate for internal shear keys. The shear key around the edge of the foundation, or
the perimeter skirt, prevents undermining of the anchor by scouring and is normally deeper; a depth of 0.1B is recommended. The depth of
shear keys or perimeter skirts is usually limited by the net downward force available to drive the keys. Shear keys should be designed to
penetrate fully under only the submerged weight of the anchor. Penetration is assisted by installing vent holes in the base to allow water and
soft surcial soils trapped by the keys to escape. Sharpening the leading edge of keys will also aid penetration.
G-3.1.4 Overturning Resistance. If subjected to excessive lateral and uplift loading, a deadweight anchor on a horizontal surface may rotate
about a point near the leading edge of its base. On sloping seaoors, the forces are resolved into components parallel and normal to the seaoor
surface. Stability against overturning is achieved by insuring the resisting or stabilizing moment, M
s
, is greater than the overturning moment,
M
o
. To insure full contact between the base and the supporting soil, the anchor should be designed so that the resultant normal soil reaction,
R
s
, acts within the middle one-third of the base. The point where R
s
crosses the shear key line is then the assumed point of anchor rotation.
The maximum stabilizing moment is:
The soil within the shear keys is assumed to separate from the base during overturning, and does not contribute to the stabilizing moment.
M
s
=
(W
b
F
ve
) B
6
G-14 G-14
S0300-A8-HBK-010
The overturning moment is:
where:
M
o
= F
h
H
1
+ z
s
H
1
= vertical distance from F
h
to the base of the shear key
z
s
= depth of the shear key tip below the foundation base
To maintain stability, M
s
must be greater than M
o
:
For preliminary sizing, the minimum width of the anchor can be calculated from:
W
b
F
ve
B
6
F
h
H
1
+ z
s
To minimize the potential for overturning, the moment arm of the lateral load component (the distance H
1
+ z
s
) should be kept as small as
B =
6F
h
H
1
+ z
s
W
b
F
ve
possible. This is most easily done by minimizing height of the deadweight; H
1
should be limited to 0.25B, if possible.
G-3.2 Shallow Foundations. Lateral load capacity, uplift resistance, and overturning resistance for foundations are calculated by the same
equations used for deadweight anchors, with two modications:
Buoyant weight of the anchor, W
b
is replaced by buoyant weight of the foundation and supported structure, W
bf
+ W
bst
. Buoyant
weight of surface-piercing structures will vary with tide, swell, and other water level uctuations.
Signicant lateral loads may result from current and/or wind loads on the supported structure.
When evaluating a foundations stability against overturning, using the highest values possible for W
bf
and W
bst
may not give a realistic estimate
of the stabilizing moment, M
s
. If lower values for W
b
and W
bst
are possible at the same time the maximum values for F
ve
and F
h
occur, the lower
values should be used. Foundation placement should be smooth and continuous to minimize disturbance to the seaoor soil and creation of an
eccentric foundation orientation. Bearing capacity (downward load), Q
u
is determined by:
where:
Q
u
= A s
u
N
c
K
c
+
b
D
f
K
q
A = effective base area of foundation, [length
2
]
s
u
= undrained shear strength of cohesive soilaveraged over the distance B below the foundation base, [force/length
2
]
N
c
= bearing capacity factor; for undrained failure N
c
= 5.14

b
= buoyant unit weight of soil above the foundation base, [force/length
3
]
D
f
= depth of embedment of foundation
K
c
,K
q
= correction factors which account for load inclination, foundation shape, embedment depth, and inclination of ground. For a
nearly level surface, a nearly square or round shape, and a vertical load, K
c
= 1.2 and K
q
= 1.0.
If the load is eccentric, or includes a moment, bearing capacity is based on a reduced foundation base to soil contact area. Effective dimensions
are calculated from the eccentricity, e. For a rectangular footing, effective length (L), breadth (B), and area (A), are:
L = L - 2e
l
B = B - 2e
b
A = B L
where:
e
l
= distance from center of footing to center of load, measured parallel to the long axis of footing
e
b
= distance measured parallel to short axis of footing
Foundation settlement due to elastic deformations and soil consolidation may pose a signicant problem, even in the absence of a bearing
capacity failure, because such settlements are rarely uniform. The occurrence of differential settlement is greatly enhanced by eccentric loading.
G-15 G-15
S0300-A8-HBK-010
G-4 GRAPPLING DEVICES
Grappling devices are used to engage and hold against solid massive seaoor features, such as coral heads, rock outcrops, and crevices or ledges
in rock and coral bottoms. Holding power depends on the strength of the grappling device and the seaoor features. Correct pulling angle and
constant tension must be maintained to prevent the anchor from losing its grip and sliding.
G-5 DIRECT-EMBEDMENT ANCHORS
Direct-embedment anchors are installed in such a manner that the anchor is buried before the anchor line is loaded, in contrast to drag-
embedment anchors which bury themselves as they are loaded. Deadmen, clumps, or drag anchors placed in excavated pits and buried are crude
but effective direct-embedment anchors, as are drag anchors settled into the seaoor by diver-jetting or explosive-induced vibration. Purpose-
built direct-embedment anchors are primarily plate-type anchors which are inserted vertically into the seaoor and expanded or re-oriented to
increase pullout resistance. There are ve major types of direct embedment anchors:
Propellant-driven,
Vibratory-driven,
Impact-driven,
Jetted-in, and
Augured-in.
Direct-embedment anchors offer signicant advantages over other types of anchors, including a very high holding capacity/weight ratio, resistance
to uplift, and the ability to support short ground leg scopes and tight moorings. Advantages and drawbacks of direct-embedment anchors are
given in Table G-7.
Table G-7. Advantages and Disadvantages of Direct-embedment Anchors.
Anchor Type Advantages Disadvantages
Vibro-driven Deep burial.
High values of holding capacity/weight ratio.
Accept vertical/multidirectional forces.
Emplacement cost increases rapidly with
anchor size and water depth.
Propellant-
embedded
Same as above.
Penetration probable in relatively hard soils (consolidated clays, coral, and
glacial tills).
Rapid installation.
Possibility of misre.
Shock wave, personnel safety.
Frequent damage to anchor chain.
Jetted-in Same advantages as for vibro-driven anchors.
Used in sands which are easily liquied.
Sand returns to denser condition, increasing holding power.
Need for dual system (water injection and
sediment extraction of sand by air lift.
Application limited to thick, sandy beds.
Not good in clay.
Slow and uneconomical.
Augured-in Used for anchoring pipelines to seaoor.
Anchoring in tandem provides torque reaction on each anchor.
500 FSW limit due to difficulty in
supplying hydraulic power to greater
depths.
G-16 G-16
S0300-A8-HBK-010
G-5.1 Propellant-embedded Anchors.
Figure G-12. Installation Sequence for a Propellant-Embedment Anchor.
1. TOUCHDOWN
(FIRING)
2. ANCHOR
PENETRATION
3. ANCHOR
KEYING
4. ANCHORAGE
ESTABLISHED
FROM NCEL HANDBOOK FOR MARINE GEOTECHNICAL ENGINEERING, KARL ROCKER JR, 1985
Propellant-embedded anchors are red into
the seaoor by a gun barrel to achieve high
holding capacities. They have been
developed for both deep and shallow water
use, and have the signicant advantage of
near-instantaneous embedment on seaoor
contact.
Propellant-embedded anchors do not require
a support stand during installation, can be
installed without difficulty on moderate
slope, and do not require deployment
vessels to remain on station for an extended
period, as do vibratory or impact-driven
systems.
After ring on touchdown, the anchor
penetrates and comes to rest deep within
the seaoor substrate. When a load is
applied to the mooring line, the anchor
keys, or rotates, into a position of maximum
resistance. The installation sequence is
shown in Figure G-12.
Because of the limited selection of
propellant-embedment anchors, design is a
matter of selecting an adequately sized
anchor and uke pattern appropriate to the
seabed. The NCEL propellant-embedment
anchor, with various uke patterns, is
shown in Figure G-13 (Page G-19).
Propellant-embedment anchor performance
and physical characteristics are given in
Tables G-8 and G-9 (Page G-18).
Table G-8. Propellant-embedded Anchors (PEA).
Anchor Type
Nominal Capacity (lbs)
x 1,000
Design Operational
Water Depth, ft
Measured or Estimated Holding Capacity (kips)
Sand Clay
a
Coral
NCEL, Port Hueneme, CA
Navy 10K 10 25 - 20,000 30 15 35
b
Navy 20K 20 50
c
- 20,000 60 35 35
d
Navy 100K 100 35 - 20,000 250 150 110
e
SUPSALV 100K 100 35
c
- 700
c
250 150 105
f
NAVY 300K
g
300 50 - 20,000 600 450
U.S. Army Mobility Equipment R & D Center,
Fort Belvoir, VA
XM-50 50 9
c
- 150 70
h
A/S/ Raufoss, Ammunisjonsfabrikker, Norway
REA 250TD 100 180 70
a
Average for soft clays found in ocean basins
b
Average of 10 tests in coral with an unconned strength of 4,200 psi
c
Has been used at this water depth
d
Average of 10 tests in coral with an unconned strength of 3,400 psi
e
Average of 22 tests in coral with an unknown strength
f
Average of 6 tests in coral with an unconned strength of 1,500 - 2,500
psi
g
Anchor is under developmentcapacities are estimated
h
Average of 2 tests in coral of unknown unconned strength
From NCEL Handbook for Marine Geotechnicial Engineering, 1985
G-17 G-17
S0300-A8-HBK-010
Table G-9. PEA Characteristics and Performance.
Anchor Type
10K 20K 100K SUPSALV 100K 300K
a
Anchor System Characteristics
Length, gun with uke, ft 6.5 9.0 12.0 13.0 15.0
Diameter or width, ft 2.0 3.5 6.0 8.0 8.0
Nominal weight, in air, lbs 650 2,000 7,000 14,000 18,000
Operating water depth, ft 25-20,000 50-20,000 35-20,000 25-500 50-20,000
Nominal holding capacity, lbs 10,000 20,000 100,000 100,000 300,000
Fluke velocity, fps 370-390 360-460 380-500 380-500 380-520
Fluke Dimensions
Sand:
length x width, ft 2 x 1 3 x 2 5 x 2.5 5.5 x 2.8 7 x 4
bearing area, ft
2
1.9 5.5 11.0 13.0 24.0
weight in air, lbs 160 290 1,300 4,000
Clay:
length x width, ft 2 x 2 3 x 3 6 x 4 6.7 x 3.3 8 x 7
bearing area, ft
2
3.7 8.5 28.0 22.0 56.0
weight in air, lbs 185 420 2,100 1,900 6,800
Coral:
length x width, ft 2 x 1 3 x 1.5 5 x 2 or 6 x 3.2 6.7 x 3.3
Estimated Penetration, ft
Clay Flukes
Soft basin soil (silty clay) 25.0 35.0 52.0
b
64.0
Distal turbidite (low s
u
) 19.0 27.0 43.0
b
57.0
Distal turbidite, (high s
u
) 19.0 26.0 39.0
b
49.0
Proximal turbidite 17.0 23.0 33.0
b
41.0
Calcareous ooze (deep water) 27.0 39.0 60.0
b
72.0
Coarse calcareous ooze (low s
u
) 25.0 35.0 54.0
b
63.0
Coarse calcareous ooze (high s
u
) 19.0 27.0 42.0
b
50.0
Siliceous ooze 30.0 43.0 65.0
b
79.0
Pelagic clay (low s
u
) 33.0 47.0 68.0
b
81.0
Pelagic clay (high s
u
) 27.0 37.0 52.0
b
63.0
Sand Flukes
Loose sand ( = 30 deg,
t
= 110 pcf)
c
12.0 17.0 25.0
b
30.0
Medium dense sand ( = 35 deg,
t
= 120 pcf)
c
11.0 16.0 23.0
b
27.0
Dense sand ( = 40 deg,
t
= 130 pcf)
c
10.0 15.0 21.0
b
25.0
Coral
a
2-7 3-12 10-37 8-35
a
Estimated parameters for anchor under development
b
No experience
c

t
= Total unit weight
From NCEL Handbook for Marine Geotechnical Engineering, 1985
G-18 G-18
S0300-A8-HBK-010
G-5.1.1 Holding Capacity in Coral.
Figure G-13. Navy Propellant-Embedded Anchors.
10K CORAL FLUKE USED AT
BARBERS POINT, OAHU, HAWAII
EXPERIMENTAL ROCK FLUKE
FOR 20K SYSTEM
10K ANCHOR SHOWING SAND
AND CLAY FLUKES
SAFE-AND-ARM
DEVICE
GUN
BARREL
REACTION
VESSEL
CLAY
FLUKE
SAND
FLUKE
TOUCHDOWN
PROBE
TOUCHDOWN
PROBE (READY
POSITION)
3
.
3
F
T
2
.
0
F
T
G
U
N
A
S
S
E
M
B
L
Y
F
L
U
K
E
A
S
S
E
M
B
L
Y
3
F
T
2
F
T
1
F
T
1 FT
FROM NCEL HANDBOOK FOR MARINE GEOTECHNICAL ENGINEERING, KARL ROCKER JR, 1985
Holding capacitythe load required to pull
the anchor uke out of the seaoor in a
few minutesis thought to result from the
plate-like uke keying or partially keying
into the coral formation under increasing
anchor line load. Because the failure mode
is not well understood, mechanistic models
to predict holding capacity have not been
developed. Holding capacity has not been
related to coral strength; as coral strength
increases, penetration decreases, but holding
capacity remains approximately constant.
Average holding capacity in coral is given
in Table G-8 for various sizes of
propellant-embedded anchors. Holding
capacity of anchor ukes embedded by the
NCEL 10K and 20K systems were
evaluated by a simple regression analysis,
yielding an equation for predicting holding
capacity as a function of kinetic energy:
where:
F
c

,
m v
2
2
0.684
F
c
= holding capacity, [kips]
m = anchor uke and piston
mass, [slugs]
v = initial uke velocity, [fps]
This equation is completely empirical and is
not dimensionally stable. Values used for m
and v must be in the units described. The
use of this equation is therefore limited to
the range of input parameters covered by the
data from which it was developed; it is valid
only for anchor uke shapes similar to the
NCEL plate-like coral uke and for coral
with unconned compressive strengths
ranging from 1,500 to 4,200 psi.
G-5.1.2 Holding Capacity in Rock.
Holding capacity is thought to result from
Table G-10. NCEL 20K Propellent-Embedded Anchor Tests in Rock.
Rock Type Approximate
Compressive Strength
psi
Penetration
ft
Peak Load
(Vertical)
lbs
Comments
Sandstone 2,000 3 45,000 Could not extract, wire failed
Sandstone 2,000 3.6 42,000 Could not extract
Basalt 3,000 3 --- Not tested
Basalt 3,000 3 65,000 Could not extract, wire failed
Basalt 3,000 3 65,000 Could not extract, wire failed
Basalt 3,000 3 75,000 Could not extract, wire failed
high compressive stresses between the rock
and the conical uke and from the bonding
of comminuted rock to the uke surface by
the heat generated during penetration.
Model tests indicate that holding capacity
decreases with rock type in the following
order: granite, basalt, limestone, shale, and
sandstone. The results of six test rings of
an NCEl 20K anchor are shown in Table
G-10. In other tests by NCEL, propellant-
embedded rock ukes have carried vertical
loads greater than 100,000 pounds and
lateral loads approaching 200,000 pounds
without failure. At present, these and similar test results are the only guide to expected performance of the NCEL rock uke. It is not known
how these results may extrapolate to other rock types or to other size anchors.
To date, efforts to develop a reliable holding-capacity equation for the propellant-embedded anchors in hard rock have not been successful. Local
rock strength variations within the rock types tested are believed to be largely responsible for the lack of consistent holding capacity performance.
Work is continuing in this area to better understand embedment anchor behavior and to allow development of a predictive method. NCEL should
be contacted for the best estimates of holding capacity in rock, and anchors should be proof-loaded to verify adequate capacity.
G-19 G-19
S0300-A8-HBK-010
G-5.1.3 Sediment Overburden. The
Figure G-14. Impact-Driven Anchors.
MENARD ROTATING PLATE ANCHOR MARK IV UMBRELLA PILE-ANCHOR
DRIVING
POSITION
IN-SERVICE
POSITION
FINAL
EMPLACEMENT
POSITION
POSITION
AFTER PULL-
OUT TEST
DRIVING
MANDREL
MUD LINE
ENLARGED
PERSPECTIVE
FROM NCEL HANDBOOK FOR MARINE GEOTECHNICAL ENGINEERING, KARL ROCKER JR, 1985
Figure G-15. Jetted-In Anchors.
HYDROPIN ANCHOR
ROYAL DUTCH SHELL JETTED ANCHOR FOR SAND AND CLAY SEAFLOORS
WATER INLET
SETTLED SAND
ANCHOR LINE
PULLING LINE
SEABED
SEABED CEMENT GROUT
WATER
INLET
PLATE
BOLTED TO
ANCHOR
WATER JETS
WATER
JETS
BRIDLE LUGS
BRIDLE
LUGS
PRESSURIZED
WATER
SKIRT
PERIPHERAL
JETS
NOZZLE
AIR
INJECTION
RISER
UNDISTURBED
SAND
UNDISTURBED
CLAY
energy imparted to the anchor uke by the
ring system is attenuated during
penetration through overlying soil layers,
resulting in shallower penetration into the
rock or coral layer and presumably lower
holding capacity. There is insufficient data
to estimate the inuence of different types
or depths of sediment overburden; most
tests have been conducted on bare
formations. NCEL 100K anchors with
coral ukes have been installed through up
10 feet of coralline sands and oozes without
capacity degradation, as judged by proof-
loading. Smaller anchors that penetrate
about half as far as the 100K anchor pre-
sumably would not be affected by similar
sediment layers up to ve feet deep. The
effect of sediment overburden on conical
rock ukes is unknown; tentative guidance
is to limit their use to sediment depths of
less than 10 feet of clay or 5 feet of sand.
G-5.1.4 Topography. Seaoor
topography does not affect holding capacity
directly, but may prevent proper
installation. Flukes may ricochet off
sloping or oblique surfaces. Flukes striking
on top of an outcrop or near the edge of a
ledge may spall the rock or coral without
penetrating deeply. Areas with surfaces
sloping more than 20 degrees or with
vertical or near vertical faces taller than one
uke length should be avoided.
G-5.2 Impact-driven Anchors. Impact- or
hammer-driven anchors include the Navy
umbrella pile and the Menard rotating plate
anchor, shown in Figure G-14. Water depth
is limited by the available pile-driving and
follower equipment for installation, although
impact-driven anchors have been successfully
installed at depths greater than 1,000 feet.
Impact-driven anchors may be an attractive
anchoring solution for salvage if adequate
pile-driving equipment is available.
G-5.3 Jetted-in Anchors. Jetted-in anchors
are buried in the seaoor through water-jet
disturbance of the sediment. The anchor
consists of a cylindrical drum equipped with
one or more injection nozzles along its per-
iphery. The system is embedded by the dual
action of pressurized water injection and
pumping of uidized sediment. Figure G-15
shows typical jetted-in anchors.
G-20 G-20
S0300-A8-HBK-010
Jetted-in anchors function best in sands or granular soils that are easily liquied by the jetting process. In hard clays, shell, and cobble soils,
penetration by jetting is slow and uneconomical. Advantages include deep burial, high holding-capacity-to-weight ratio, and the ability to resist
vertical and multi-directional pullout forces. Disadvantages include the size and complexity of the installation apparatus, and the limited number
of seaoor types in which the system can be used effectively. Jetted-in anchors can sometimes be built on site and installed with available high-
pressure pumps and rehoses.
G-5.4 Vibro-driven Anchors. Vibro-driven anchors are driven vertically into the substrate by cable traction vibro-driving units. Cable traction
is exerted in an off-center direction on the plate to cause it to pivot, mobilizing the passive pressure of the soil.
Hydraulic vibro-driving units are typically limited to 1,000 feet because of the problems associated with loss of hydraulic pressure at depths.
NCEL has developed and successfully tested an experimental, battery powered variation of a vibro-driven anchor for water depths to 6,000 feet.
Further use of this device was discontinued as propellant-driven anchors were developed. The main disadvantage of the vibro-driven anchoring
system is that the installing platform must be maintained in a position directly over the anchor during the period required for proper uke
deployment, typically 15 to 20 minutes. Costs increase proportionately with anchor size and water depth. Vibro-driven anchors are seldom used
in salvage if adequately sized propellant-embedded or impact-driven anchors are available.
G-5.5 Augured-in Anchors. Auger anchors are screw-
Table G-11. Conditions Complicating Direct-embedment Anchor Use.
Seafloor Condition Potential Complication
Hard Strata
Thin soil layer over rock, or thin
soft clay layer over sand.
Glacial erratics or residual
surficial gravel and cobbles.
Nodule or pavement formations
(usually manganese) over soil.
Submarine lava flows.
Soil thickness not sufficient to develop fluke
capacity, but sufficient to consume most of fluke
kinetic energy before it reaches stronger layer.
Damages flukes and limits penetration into
underlying sands and hard clays.
Same as above.
Extremely irregular and complex. Anchors must
be proof-tested to full load to ensure reliability.
Sloping Seafloor
Soil slopes over 10 degrees.
Hard soil or rock scarps or cliffs.
Anchor may affect slope stability, but direct-
embedment anchors have less effect than most
other anchor types.
Improper embedment from deflection or ricochet
of fluke.
Scour Sand waves can be large and move rapidly,
removing significant overburden from shallow
embedded anchors.
Sensitive, Soft Soils
Cohesive soil with sensitivity of 6
or greater.
Deep ocean oozes.
Weak, porous clays (shear
strength to overburden pressure
ratio, s
u
/ p
0
, 0.1 - 0.15).
Fluke installation can remold and weaken soil,
limiting developed holding capacity.
Same as above.
Long-term capacity may be lower than short-
term.
From NCEL Handbook for Marine Geotechnical Engineering, 1985
shaped shafts installed under high torque and some
vertical load. The depth limit is approximately 500 feet
and is imposed primarily by difficulties in supplying
power through hydraulic hoses to the seaoor. Augured-
in anchors are used primarily for anchoring pipelines to
the seaoor and are usually installed in pairs opposite
each other. They may be useful for anchoring temporary
pipelines or other equipment to the seaoor during
certain salvage operations.
G-5.6 Site Data. In nonhomogeneous soils, sediment
type and thickness must be determined. The seaoor
type and approximate consistency must be known in
order to select the appropriate anchor uke type (i.e.,
clay, sand, coral, or rock uke). Depth of sediment strata
must be known to ensure that the anchor uke has suf-
cient sediment thickness to develop the design capacity.
These data are best obtained over a wide area through
acoustic sub-bottom proling and coring. In areas of
large relief, such as areas of outcropping rock, erosion, or
slumping features, a deep tow proling system may be
necessary to obtain an accurate picture of seaoor
topography and distribution of sediment in-ll between
the relief features. Geotechnical properties may be
estimated from soil property proles to make a rough
estimate of capacity in lieu of accurate site-specic data.
For direct-embedment anchors in critical moorings,
where the consequences of a single mooring failure are
severe, data from in-situ tests and good quality soil
cores are required. Where dynamic loads are
signicant, specialized tests using core samples may be
necessary. If sediment consistency or type varies
across the mooring site, cores should be obtained at
each anchor location. Soil cores should be obtained
over the full estimated penetration depth of the anchor.
Small corers, which achieve penetrations of 10 feet in
sands and 30 feet in clays, are often used to obtain the
sediment and dene the upper portion of the geotechni-
cal property prole. With the guidance of geophysical
data, sediments below this sampled depth may be as-
sumed to be similar, and the soil property prole would
be extended to the necessary depth. Expendable
penetrometers can provide additional data where longer
coring is not possible.
G-5.6.1 Complicating or Hazardous Conditions. Direct-embedment anchor systems function well in a wide range of seaoor conditions.
They can be adapted to function well where drag anchors and pile anchors are inefficient or nonfunctional. Extreme soil conditions, such as
very hard or very soft seaoors, complicate the use of direct-embedment anchors, making special efforts necessary during site survey, positioning,
design, installation, and proof-loading. Table G-11 lists complicating or hazardous conditions, and describes their impact on direct-embedment
anchor performance. The approach to most of these complications is to avoid them by relocating the anchor or selecting an anchor system less
sensitive to the problem (e.g., using deadweight, pile, or grappling anchors on rock).
G-21 G-21
S0300-A8-HBK-010
G-5.6.2 Specialized Survey Tools. Two specialized site survey tools have been developed that can support the siting, design, and installation
of the Navys propellant-embedded anchors. The expendable Doppler penetrometer measures the undrained shear strength of the soil indirectly.
Although strength is not determined to the level of accuracy as it is from in-situ measurement or coring, the device is simply employed and
requires little on-site time. The Pinger probe is a 3.5-kHz battery-powered sound source that can be used with a propellant-embedded anchor
system to assist positioning the anchor in complex seaoor conditions. The probe is attached to the anchor systems lowering line about 100
feet above the anchor, and provides a high quality, real-time image of sub-bottom seabed stratication. The installation vessel then maneuvers
the anchor system over a seabed prole until a location is found that maximizes chances for successful installation. Doppler penetrometers and
Pinger probes can be obtained through the Naval Civil Engineering Laboratory (NCEL).
G-5.7 Fluke penetration and Keying. Estimates of propellant-embedded uke penetration are taken from Table G-9 and can be rened from exist-
ing proles of undrained soil strength. The penetration depth of jetted-in anchors is limited primarily by hard layers in the soil prole, which stop
or impede the jet erosion process. Penetration depth of driven anchors depends on the capacity of the available pile-driving equipment.
As the anchor uke moves upward it keys into a horizontal orientationthe position of maximum holding capacity. The keying distance, z
k
, is a
function of uke geometry, soil type, soil sensitivity, and duration of time between penetration and keying. Experience has shown that Navy
propellant-embedded anchor ukes key in about 2.0 uke lengths in cohesive soil and in about 1.5 uke lengths in cohesionless soils.
Although no recommendation is made for
Figure G-16. Soil Failure Modes for Direct-Embedment Anchors.
SOIL BEARING
PRESSURE
FAILURE
SURFACE
FLUKE
WIDTH
FLUKE
WIDTH
SHALLOW ANCHOR FAILURE DEEP ANCHOR FAILURE
FROM NCEL HANDBOOK FOR MARINE GEOTECHNICAL ENGINEERING, KARL ROCKER JR, 1985
altering the above estimate for z
k
, it is
believed that keying distance may be great-
er in highly sensitive soils. Keying dis-
tance in sensitive soils can be shortened by
allowing the anchor to "soak" for 24 hours
or more before tensioning the anchor line.
G-5.8 Static Holding Capacity. Loads on
seaoor anchors are rarely completely static
but often have impulse or repetitive
components. Dynamic holding capacity is
addressed in Paragraph G-6.4.
Holding capacity depends on the soil failure
mode, which in turn is dependent on
relative embedment depth (the ratio of
embedment depth to anchor minimum
dimension, z/B) and on the soil type and
strength. Shallow failure occurs when the
seaoor surface is displaced by the upward
motion of the anchor plate and the soil
failure surface continues up to the seaoor.
Deep failure occurs when the anchor plate
is sufficiently deep within the seabed that
the soil failure surface does not reach the
seaoor. Transition from shallow-to-deep
behavior occurs over a range of relative
Table G-12. Soil Strength Reduction Factor.
Soil Type h
Very soft, moderately sensitive, clayey silt, s
u
1 psi, S
t
3 0.8-0.9
Soft, normally consolidated, silty clay, s
u
2 psi, S
t
3 0.8
Pelagic clay, s
u
1.2 psi, S
t
3 0.7
Foraminiferal sand-silt, 77-86% carbonate, s
u
2.2 psi, S
t
10 0.25
From NCEL Handbook for Marine Geotechnical Engineering, 1985
embedment depths from 2 to 5 in cohesive
soil and 2 to 10 in cohesionless soil.
Failure modes for direct-embedment
anchors are shown in Figure G-16.
G-5.8.1 Short-term Capacity in Cohesive
Soils. Short-term loading conditions exist
when the anchor-caused soil failure is
governed by a soils undrained shear
strength. Failure occurs immediately after,
or within a few minutes of load application,
before signicant drainage of pore water can
take place. Short-term static load capacity in
cohesive soils, F
st
, is:
F
st
A s
u
h N
cs

1
1
]
0.84 0.16

,
B
L
where:
A = projected maximum uke area perpendicular to B = plate minimum dimension, usually width, [length]
direction of pullout, [length] L = plate maximum dimension, usually length, [length]
s
u
= soil undrained shear strength, [force/length
2
] h = soil strength reduction (disturbance correction) factor
(see Paragraph 3-7.1.4)
N
cs
= short-term holding capacity factor in cohesive soil, from Figure G-17.
G-22 G-22
S0300-A8-HBK-010
The disturbance correction factor, h,
Figure G-17. Short-term Holding Capacity Factors for Cohesive Soil.
RELATIVE EMBEDMENT DEPTH, z/B
0 2 4 6 8 10
0
5
10
15
20
N
cs
(a) SHORT-TERM
HOLDING
CAPACITY
FACTOR
N
cs
N
q
(b) LONG-TERM
HOLDING
CAPACITY
FACTOR
(c) DRAINED SOIL
HOLDING
CAPACITY
RELATIVE EMBEDMENT DEPTH, z/B
0 2 4 6 8 10
0
5
10
15
s
u
< 0.75 PSI
s
u
< 0.75 PSI
s
u
< 1 PSI
s
u
< 1 PSI
s
u
< 1.5 PSI
s
u
< 1.5 PSI
s
u
< 4 PSI
s
u
< 4 PSI
RELATIVE EMBEDMENT DEPTH, z/B
0
0
2
20
4
40
6
60
8
80
10
100
2 4 6 8 10 12 14
= 40 *
= 35
= 30
= 25
= 20
FROM NCEL HANDBOOK FOR MARINE GEOTECHNICAL ENGINEERING, KARL ROCKER JR, 1985
NOTES: Z = EMBEDMENT DEPTH
B = PLATE MINIMUM DIMENSION (WIDTH)
* = VALUES FOR CAN BE OBTAINED
FROM TABLE 3-3
accounts for soil remolding during uke
penetration and keying. Values for the
factor, h, were determined for the four soil
types listed in Table G-12 by anchor tests.
Sensitivity, S
t
the ratio of undisturbed to
remolded shear strengthis an important
indicator of the strength reduction resulting
from soil disturbance. For soils with S
t
values considerably different from those in
Table G-12, an estimate must be made for
the h-value. More sensitive soils will
display greater strength reductions.
The holding capacity factor, N
cs
, from
Figure G-17(a), is a function of the soils
undrained shear strength and relative
embedment depth. For the deep failure
mode, N
cs
= 16. If drainage vents allow
water to ow rapidly to the underside of
the plate anchor, the suction formed on the
underside of the plate will be relieved; N
cs
should be reduced to the long-term holding
capacity value, N
c
, from Figure G-17(b).
G-5.8.2 Long-term Capacity in Cohesive
Soils. Long-term loading exists when a
static load is applied to the anchor over a
time long enough to allow virtually
complete dissipation of excess pore water
pressures. The time duration ranges may
be a day for silts, a week for silty clays, or
considerably longer for clays. In cohesive
soils, the long-term holding capacity is
governed by the effective soil drained
strength parametersthe drained cohesion
intercept, c, and the drained friction angle,
. The long-term static holding capacity,
F
lt
, is:
where:
F
lt
= A cN
c
+
b
zN
q

1
1
]
0.84+0.16

,
B
L
A = projected maximum uke area
perpendicular to direction of
pullout, [length]
c = drained soil cohesion,
[force/length
2
](fromTable 3-3)
N
c
= long-term holding capacity
factor in cohesive soil (from
Figure G-17(b))

b
= buoyant unit weight of the
soil, [force/length
3
] (from
Table 3-3)
N
q
= holding capacity factor for a
drained soil condition (from
Figure G-17(c))
F
lt
= long-term static holding
capacity, [force]
F
st
= short-term static holding
capacity, [force]
z = embedment depth [length]
Long-term static holding capacity, F
lt
, must be less than short-term capacity, F
st
the applied long-term load cannot exceed the short-term load
without initiating failure.
G-23 G-23
S0300-A8-HBK-010
For very soft underconsolidated sediments, such as delta mud, the shear failure mode may be different than for normally consolidated sediments.
The reduced holding capacities in these very soft soils can be conservatively predicted by reducing the cohesion, c, and the drained friction angle,
, values before obtaining the holding capacity factors and calculating holding capacity. The reduced values, c and , are determined by:
G-5.8.3 Short- and Long-term Capacity in Cohesionless Soils. In sands and gravel, static-loading- induced excess pore water pressure
c =
2
3
c
= arctan

,
2
3
tan
dissipates rapidly as the load is applied. Soil failure is assumed to be drained for both static short- and long-term loading. Static holding
capacity in cohesionless soils, F, is:
where the holding capacity factor N
q
is obtained from Figure G-17(b) using the relative embedment depth and the soil friction angle.
F A
b
z N
q

1
1
]
0.84 0.16

,
B
L
When dealing with very loose sands (i.e., relative density less than 40 percent), the soil friction angle, , should be reduced in the same manner
as for the drained cohesive soil case before entering the plots Figure G-17.
G-5.8.4 Factors of Safety. For applications where little is known about the soil conditions at the site, a safety factor of 3 is recommended
for holding capacity estimates, i.e., limit applied loads to F/3, where F is calculated holding capacity. A safety factor of 2 can be used when
good site data leads to a high level of design condence.
G-5.9 Dynamic Holding Capacity. Dynamic loads are applied quickly, but for short periods. Dynamic loads are considered to be applied
quickly when the load development takes less than 10 minutes for clays or less than 10 seconds for sand. Dynamic loads are divided into two
categories: (1) cyclic or repetitive loadings and (2) impulse loading (basically a single event). Both types can alter plate anchor holding capacity
by changing the conditions in the soil surrounding the anchor. Cyclic loads typically result from wave forces on moored vessels and attendant
vessel motions. Impulse loads can result from sudden changes in loading of a vessel moored on short scope (heavy lift), vessel collisions, and
similar events. Wind loading is normally quasi-static, but strong gusts may impulse or on near-cyclic loads.
G-5.9.1 Cyclic Loading. For design purposes, cyclic loading is separated into three categories:
Cyclic line loading of the anchor, leading to soil strength loss in the vicinity of the anchor and subsequent failure.
Cyclic line loading that may cause upward anchor movement (creep), possibly moving the anchor into shallower soil; thereby
lowering short-term static holding capacity.
Earthquake-caused cyclic loading of the soil mass resulting in near-complete loss of strength in the entire soil mass and sudden
anchor failure.
Cyclic loads are characterized by a pure cyclic double-amplitude loading component, P
c
, superimposed on a static loading component, P
s
. Cyclic
and static load magnitudes are expressed as a percentage of static short-term anchor holding capacity. Cyclic loads with a double amplitude
less than 5 percent of the static short-term holding capacity are disregarded. Two additional parameters are required to describe a cyclic loading
condition:
n
t
= Total number of load cycles expected in the anchors lifetime to evaluate the potential for anchor creep.
n
c
= The number of cycles that occur in a limited time period required for dissipation of excess pore pressure, t
cd
, to evaluate soil
strength loss and potential for liquefaction.
G-5.9.2 Strength Loss. Virtually all soils are subject to some strength loss from extended cyclic loading. The amount of strength loss,
however, varies considerably depending on soil type, state, and the nature of the cyclic loading. The following factors reduce soil susceptibility
to strength loss:
Denser soil,
More plastic soil,
Lower cyclic loading magnitude,
Smaller number of load cycles, and
Longer time period over which the cycles occur.
G-24 G-24
S0300-A8-HBK-010
Some low-density cohesionless soils are
Figure G-18. Time for Dissipation of Excess Pore Pressure.
SOIL PERMEABILITY, (k) ft/sec
10
-3
10
0
10
0
10
-1
10
-2
10
-3
10
-4
10
-5
10
1
10
1
10
2
10
2
10
3
10
3
10
4
10
5
10
6
10
7
10
8
10
-4
10
-5
10
-6
10
-7
10
-8
10
-9
10
-10
10
-11
T
I
M
E
R
E
Q
U
I
R
E
D
F
O
R
P
O
R
E
P
R
E
S
S
U
R
E
D
I
S
S
I
P
A
T
I
O
N
,
t
c
d
(
s
e
c
)
T
I
M
E
R
E
Q
U
I
R
E
D
F
O
R
P
O
R
E
P
R
E
S
S
U
R
E
D
I
S
S
I
P
A
T
I
O
N
,
t
c
d
(
d
a
y
s
)
B
=
7
F
T
B
=
2
F
T
FROM NEEL HANDBOOK FOR MARINE GEOTECHNICAL ENGINEERING, KARL ROCKER JR, 1985
susceptible to complete liquefaction.
Sediments of this type (uniform ne sands,
coarse silts, and some clean deep sea oozes)
can experience a near-total strength loss un-
der cyclic loading. Plate anchors are not
recommended for these soils if signicant
cyclic loading is expected.
For other soils, the susceptibility of a given
plate anchor to cyclic-load-caused strength
reductions can be evaluated by estimating
the maximum cyclic load that can be
sustained by the anchor without pore
pressure dissipation. The plots in Figure
G-18 give the time required to dissipate
excess pore pressure, t
cd
, as a function of
uke length, B, and soil permeability, k.
Table 3-6 gives permeabilities for typical
soils, but permeability should be de-
termined by testing undisturbed soil from
the vicinity of the anchors soil mass, if
possible. The maximum number of double-
amplitude cyclic loadings that can occur
within the time period t
cd
is estimated from
the known or expected loading conditions.
Figure G-19 is then entered to determine
Figure G-19. Direct-Embedment Anchor Cyclic Load Capacity (Without Soil Strength Loss).
NUMBER OF LOADING CYCLES, n
c
10
0
0
10
20
30
40
50
60
70
80
90
100
10
1
10
2
10
3
10
4
10
5
10
6
10
7
D
O
U
B
L
E
-
A
M
P
L
I
T
U
D
E
C
Y
C
L
I
C
L
O
A
D
(
%
O
F
S
T
A
T
I
C
S
H
O
R
T
-
T
E
R
M
C
A
P
A
C
I
T
Y
)
H
IG
H
L
Y
P
L
A
S
T
IC
C
L
A
Y
S
LOW
PLASTICITY AND MEDIUM
DENSITY SOILS
DATA APPLY TO ANCHORS WITH AN
AVERAGE STATIC LOAD LESS THAN
33% OF STATIC SHORT-TERM CAPACITY
NOTE:
FROM NEEL HANDBOOK FOR MARINE GEOTECHNICAL ENGINEERING, KARL ROCKER JR, 1985
the maximum cyclic load, P
c
, that can be
sustained without signicant loss of soil
strength.
G-5.9.3 Factors of Safety. Because the
design approaches for cyclic loading are
quite conservative, lower safety factors can
be used for the cyclic loading aspects of
anchor design. Safety factors of 1.75 for
calculations based on limited soil data, or
1.25 for those based on detailed soil data,
are appropriate.
G-5.9.4 Impulse Loading. An impulse
load, P
I
, is a single load applied quickly but
for a short periodless than 10 minutes for
clays and less than 10 seconds for sands.
These loads are considered single events
only if enough time elapses between similar
events to allow the soil to return to its
normal state without residual effect. In the
absence of other dynamic loads, an anchor
can resist an impulse load higher than its short-term static capacity. The procedures for predicting holding capacity under impulse loading
presented below are appropriate for use with circular, square, or rectangular (L/B 2) anchor ukes only.
G-25 G-25
S0300-A8-HBK-010
G-5.9.5 Cohesive Soil. For cohesive soils,
LOAD DURATION (sec)
NOTE: WHEN s
u
IS ESTIMATED, USE NORMALLY
CONSOLIDATED, MODERATE SENSITIVITY CURVE
S
O
I
L
S
T
R
E
N
G
T
H
I
N
F
L
U
E
N
C
E
F
A
C
T
O
R
,
I
0.01
1.0
1.2
1.4
1.6
1.8
2.0
2.2
2.4
2.6
2.8
3.0
3.2
3.4
0.1 1.0 10 100 1000
Figure G-20. Strain-Rate Factor (I) for Cohesive Soil.
FROM NCEL HANDBOOK FOR MARINE GEOTECHNICAL ENGINEERING, KARL ROCKER JR, 1985
OVERCONSOLIDATED OR
SENSITIVE CLAYS (S
t
> 5)
NORMALLLY CONSOLIDATED, MODERATELY
SENSITIVE CLAYS (S
t
= 2 TO 5)
NORMALLY CONSOLIDATED, NONSENSITIVE
CLAYS (S
t
< 2)
ALL CLAYS WITH LIQUIDITY INDEX > 1
the maximum impulse load that can be
applied to an anchor, F
I
, is determined by:
where:
F
I
= I R
c
R
I
I
f
F
st
F
st
= static short-term anchor
holding capacity, [force]
I = inuence factor for adjusting
the soil strength for strain
rate from Figure G-20
R
c
= reduction factor for cyclic
loading
R
I
= reduction factor for repeated
impulse loading
I
f
= inertial factor for capacity
increase under very rapid and
short-duration loading (i.e.,
for impulse duration less than
0.01 second)
The factor I is a strain rate used to adjust
for an increase in the soil undrained shear
strength during impulse loading. The factor
R
c
depends on the loading history prior to
LOAD DURATION (sec)
I
N
E
R
T
I
A
L
F
A
C
T
O
R
,
I
f
0.001
1.0
1.2
1.4
1.6
1.8
2.0
2.2
2.4
2.6
2.8
3.0
3.2
3.4
0.01 0.1 1.0 10.0 100.0
Figure G-21. Inertial Factor (I
f
).
B = 7 FT
B = 2 FT
FROM NCEL HANDBOOK FOR MARINE GEOTECHNICAL ENGINEERING, KARL ROCKER JR, 1985
the impulse loading and adjusts F
I
for the
inuence of other nonstatic loads that are
occurring at about the same time. If the
impulse load is the rst event, then R
c
=
1.0. If cyclic loads immediately precede
the impulse event, then:
where P
c
is the design vertical compressive
R
c
= P
c
F
st
load at the foundation pile.
The factor R
I
adjusts F
I
for repeated
impulse loadings. If there is only one
impulse load in a four-hour period, R
I
= 1.
If there is more than one impulse load in a
four-hour period,
where f
c
is the average frequency, in
R
I
1.33 e
1.15f
c
impulses per hour, over a four-hour period.
The factor I
f
increases F
I
for the inertia of the soil mass at very short duration loadings (i.e., where the loading is known to be applied for less
than 0.1 second), and is determined from Figure G-21.
G-26 G-26
S0300-A8-HBK-010
G-5.9.6 Cohesionless Soils. Impulse holding capacity under impulse loading in cohesionless soils is also derived by applying a series of
inuence factors to the calculated short-term static holding capacity:
where:
F
I

,
N
qI
N
q
R
c
R
I
I
f
F
st
F
st
= static short-term anchor holding capacity
N
qI
= cohesionless soil holding capacity factor adjusted for impulse loading
N
q
= cohesionless soil holding capacity factor (from Figure G-17).
R
c
= reduction factor for cyclic loading
R
I
= reduction factor for repeated impulse loading
I
f
= inertial factor for capacity increase under very rapid and short-duration loading (i.e., for impulse duration less than 0.01 second)
The adjusted holding capacity factor (N
qI
) ac-
LOAD DURATION (sec)
S
O
I
L
S
T
R
E
N
G
T
H
I
N
F
L
U
E
N
C
E
F
A
C
T
O
R
,
I
0.01
1.0
1.2
1.4
1.6
1.8
2.0
2.2
2.4
2.6
2.8
3.0
3.2
3.4
0.1 1.0 10.0 100.0 1000.0
Figure G-22. Strain-Rate Factor (I) for Cohesionless Soil.
FINE-SILTY SANDS
MEDIUM-COARSE SANDS

I
= sin
-1
I sin
1 + (I - 1)sin
(UNTYPICAL, OVERCONSOLIDATED CONDITION)
(TYPICAL CONDITION)
FROM NCEL HANDBOOK FOR MARINE GEOTECHNICAL ENGINEERING, KARL ROCKER JR, 1985
counts for the effect of the impulse loading
on the soil friction angle (). N
q
from Figure
G-17(b) can be used for N
qI
by substituting
the adjusted friction angle
I
for , where:
where I is the inuence factor for adjusting

I
sin
1

1
1
]
I sin
1 + (I = 1) sin
the soil strength from Figure G-22. The
factor R
c
is determined in the same manner
as for cohesive soil.
For repeated impulse loads, R
I
is dependent
on the frequency of those impulse loads (f
s
).
If f
s
is less than or equal to one impulse per
10 minutes, R
I
is 1.0. If f
s
is greater than one
impulse per 10 minutes, then R
I
is obtained
from:
where f
s
is the average number of impulses
R
I
2 e
0.116 f
s
per 10 minutes. The inertial factor I
f
is taken
from Figure G-23.
G-5.10 Holding Capacity on Slopes.
Holding capacity of embedded plate anchors
on slopes is affected by the stability of the
slope under the additional inuence of the
embedment anchor, and the inuence of the inclined seaoor on the soil resistance mobilized by the loaded anchor. The inuence of an anchor
on slope stability is extremely complex, involving the effects of anchor installation and anchor loading on the slope. Table G-13 lists factors that
inuence submarine slope stability. All factors lead to a lower resistance to soil mass down-slope sliding and, therefore, greater slope instability.
The inuence of plate anchors on slope stability depends to a high degree on the type and sensitivity of the sediment. Slope angle itself is not a
clear indicator of potential problems.
On inclined seaoors, a signicant portion of the soil shear
Table G-13. Direct-embedment Anchor Effects on
Submarine Slope Stability.
Factor Effects
Impact Loading During Embedment
Effect similar to earthquake loading but with
greater local influence; more critical problem in
loose soils.
Remolding or Disturbance of Soils
During Installation
Effect varies significantly from one soil to another.
Cyclic Loading by Anchor Effect varies significantly from one soil to another.
Local Instability After Anchor Pullout Can progress to major slope failure.
Direct Application of Anchor Load to
Slope
Probably not more significant than a local instabil-
ity problem but can progress into large slide.
From NCEL Handbook for Marine Geotechnical Engineering, 1985
strength is mobilized to support the soil slope against gravi-
tational forces. For down-slope loading, most of the soil shear
stresses developed to resist anchor pullout will be in addition
to those resisting slope failure. The result is that a smaller
amount of the soils shear strength is available to resist anchor
pullout than in a horizontal seaoor. This is less important
when the anchor is loaded in a vertical or up-slope direction.
Holding capacity of a direct-embedment anchor on a specic
slope can be estimated by multiplying the holding capacity
calculated for a horizontal seaoor by a reduction factor, R
s
:
where F
s
is the factor of safety against a slope failure without the anchor. R
s
represents the amount of soil strength remaining, or the amount
R
s
F
s
1
F
s
not mobilized to maintain slope stability. In computing the anchor holding capacity, anchor depth and the holding capacity factors N
c
and N
q
are based on a depth of embedment measured perpendicular to the seaoor.
G-27 G-27
S0300-A8-HBK-010
G-6 PILE FOUNDATIONS AND ANCHORS
Piles are secure, deeply embedded
Table G-14. Pile Anchors.
Characteristics Applications Approx. Maximum
Capacity
Installation Methods Applicable Soil Type
Pipe and H-piles Foundations and anchors Axial: 20,000 kips
Lateral: 1,500 kips
Driven or drilled and
grouted
Soil and rock
Umbrella Piles
a
Anchors 300 kips in sand
100 kips in mud
Driven Soils without boulders
and other obstructions
Chain-in-hole
a
Anchors 550 kips Drilled and grouted Rock with overlying
soil strata
Rock Bolts
a
Anchors 30 kips Drilled and grouted or
mechanically wedged
Rock
Characteristics Advantages Disadvantages Remarks
Pipe and H-piles Easy to splice, high
capacity, can penetrate
through light obstructions
Very high cost Pipe piles resist bending in any direction
Umbrella Piles High capacity in uplift Maximum depth limited
by hammer; soils must be
homogenous
Resistance developed similar to plate-
embedment anchor
Chain-in-hole High capacity Installation may be
difficult

Rock Bolt Very low cost, no heavy


mechanical equipment
necessary
Rock must be competent,
nonfractured (shallow
water only); low capacity
Diver/hand-installed; much smaller than
normal piles
a
Special anchor pile; may be considered a direct-embedment anchor
Adapted from NCEL Handbook for Geotechnical Engineering, 1985
anchoring devices installed by driving or by
drilling and grouting (pile installation
methods are described in Paragraph G-6.6).
High installation costs usually preclude
their use in conditions for which drag-
embedment, deadweight, or direct-
embedment anchors are suitable. Pile
anchors are particularly suited for use when
short-scope moorings are desired, and on
very hard seaoors. Applications and
characteristics of several types of pile
anchor are listed in Table G-14.
Steel pipe sections and H-piles tted with
mooring line connections are the most
commonly used pile types, although
specially designed piles have been
developed to increase lateral or uplift
capacity. Typical anchor piles are shown
in Figure G-23. For simple pipe and H-
piles, axial forces are resisted by soil
friction developed along the pile shaft and
by bearing on the pile tip (for downward
loads). Lateral forces and moments are
resisted by the pile shaft bearing on the
Figure G-23. Anchor Piles.
PIPE PILE
WIDE-FLANGE (WF) SECTION
BUILT-UP SECTION (COMPOSED OF T-SECTIONS)
PADEYE WELD
WELD WELD
BUILT-UP SECTION FOR
MULTIDIRECTIONAL LOAD
FINS TO IMPROVE
LATERAL LOAD
CAPACITY
PADEYE
WEB
PIPE PILE
FOR MULTI-
DIRECTIONAL
LOAD
WIDE-FLANGE
SECTION FOR
UNIDIRECTIONAL
LOAD
MOORING-LINE
CONNECTION
FLANGE
LOAD
L
O
A
D
near-surface soils. Lateral earth pressure
and skin friction effects on a pile anchor
are shown in Figure G-24.
G-6.1 Holding Capacity. Pile anchors or
foundations may be subjected to one or
more of the following loads:
Axial uplift loads,
Lateral loads,
Bending moments, or
Axial downward loads
(compression).
A simplied procedure for the design of
uniform cross-section piles in a nonlayered
seaoor consisting of sand, clay, or
calcareous soils is presented in the following
paragraphs. Design is a trial-and-error
procedure; a pile is selected and then
evaluated for its ability to resist applied loads
without excessive movement and without
exceeding the allowable stresses for the pile
material. The pile is assumed to be a beam
on an elastic foundation with an elastic
modulus that increases linearly with depth.
Design calculations for pile anchors and
foundations are identical; lateral and uplift
loads are governing for anchor piles, while
downward loads are usually governing for
foundation piles. Both anchor and
foundation piles may be subjected to
signicant moments, depending on the point
of application of lateral loads; foundation
piles may also be subjected to uplift loads.
G-28 G-28
S0300-A8-HBK-010
Soil properties for the installation site should be determined by in-situ or laboratory testing, or both. If site-specic soil data are not available,
it may be possible to extrapolate soil properties from geologic and geophysical data from similar areas. Where soil properties vary signicantly
with depth, average properties in the uppermost four pile diameters are used for lateral load analysis, and average properties over the pile length
for axial load analysis.
Figure G-24. Lateral Soil Pressure and Skin Friction on Pile Anchor.
PILE HEAD
PILE
MOORING LINE
LATERAL EARTH
PRESSURE
SKIN
FRICTION
L
O
A
D
G-6.1.1 Lateral Load Capacity. The lateral load capacity, P
h
, is given by:
where:
P
h
=
y
max
(EI)
A
y
T
3
+ aB
y
T
2
Pile length, L
p
, is assumed. A length of L
p
= 3T is suggested as a minimum. For anchor piles, the ratio of maximum lateral deection to pile
P
n
= lateral load capacity, lbs
y
max
= pile head lateral deection, [in.]
EI = pile stiffness, [lb in
2
]
E = modulus of elasticity of pile material, [lb in
2
]
I = moment of inertia of pile cross section, [in
4
]
A
y
, B
y
= deection coefficients, functions of the depth coefficient
L
p
T
, Figure G 25 (Page G 30)
a = height of the pile load attachment point above the seaoor surface, [in]
L
p
= pile length, [in]
T = pile-soil relative stiffness, [in]
=
EI
N
h
0.2
n
h
= coefficient of subgrade reaction, from Figure G 25 (Page G 30), [lb in
3
]
diameter or width (y
max
/D) should be less than 0.1D. Smaller deections may be required for foundation piles. If lateral load capacity, P
h
, is
equal to or slightly higher than anticipated loads, the trial pile is adequate. If P
h
is much greater than anticipated loads, the pile is over-designed.
If cost reduction or material optimization are desired, pile stiffness should be decreased by reducing pile diameter or wall thickness, and/or the
pile length shortened. If P
h
is less than anticipated loads, the pile is under-designed. Holding capacity is increased by increasing:
Pile stiffness by increasing diameter and/or thickness,
Pile length, unless already very long, or
Design depth of pile head.
G-29 G-29
S0300-A8-HBK-010
DEPTH COEFFICIENT
(
z
max
T
=
Lp
T
)
(a) DEFLECTION COEFFICIENT
AT SOIL SURFACE
FROM NCEL HANDBOOK FOR MARINE
GEOTECHNICAL ENGINEERING,
KARL ROCKER, 1985
AFTER EVALUATION OF COEFFICIENTS OF
SUBGRADE REACTION, K. TERZAGHI,
GEOTECHNIQUE, 1955
FROM NCEL HANDBOOK FOR MARINE
GEOTECHNICAL ENGINEERING,
KARL ROCKER, 1985
(
A
y
a
n
d
B
y
)
2
0
1
2
3
4
5
3 4 5 6
A
y
B
y
0
200
150
100
50
0
5 10
(c) n
h
FOR COHESIVE SOILS (b) n
h
FOR COHESIONLESS SOILS
15 20
y
max
D
(%)
y
max
D
(%)
SOFT CLAY
STIFF CLAY
k
l
n
h
=
s
u
k
l
D
0 2 4 6 8 10
120
100
80
60
40
20
0
n
h
(
l
b
/
i
n
3
)
Dr = 35%
Dr = 50%
D
r
=
6
5
%
D
r
=
8
5
%
LOOSE
MEDIUM DENSE
DENSE
Dr = SOIL RELATIVE
DENSITY
D = PILE DIAMETER
OR WIDTH
Figure G-25. Coefficients A
y
, B
y
, n
h
.
G-6.1.2 Uplift Capacity. Pile uplift
Table G-15. Recommended Limiting Values for Unit Skin Friction
and End Bearing for Cohesionless Soils.
Soil Type (deg) N
q
f
s
(max)
(ksf)
q
p
(max)
(ksf)
Noncalcareous Soils
Sand 35 40 2.0 200
Silty sand 30 20 1.7 100
Sandy silt 25 12 1.4 60
Silt 20 8 1.0 40
Calcareous Soils
Uncemented calcareous sand (easily crushed) 30 20 0.3
a
60
Partially cemented calcareous sands with
carbonate content:
0 to 30% 2.0 100
30 to 45% 0.64
a
160
above 45% 0.56
a
140
Highly cemented calcareous soils, i.e., chalk 1.1 140
a
For drilled and grouted piles, the value may approach 2,000 psf, the value for quartz sand: actual value depends
upon installation technique
From NCEL Handbook for marine Geotechnical Engineering, 1985
resistance results from skin friction between
the pile and the soil mass. Pile frictional
resistance is:
where:
Q
s
A
s
f
s
A
s
= the surface area of the pile
below the seaoor [L
2
].
f
s
= average unit skin friction
resistance
Axial load capacity is increased most
effectively by increasing pile length,
although increasing diameter may be
effective.
For cohesionless soils, average unit skin
friction resistance, f
s
, is calculated from:
where:
f
s
= k p
vo
tan ( 5)
k = 0.7 for compression
= 0.5 for uplift.
p
vo
= effective overburden vertical stress in soil, [force/length
2
]
= drained (effective) friction angle from Table G-15 or Figure 3-3
Table G-15 gives limiting values for unit skin friction for cohesionless soils. For piles driven into calcareous soils, the tables limiting values
should be used unless higher values are justied by on-site testing.
For cohesive soils, unit skin friction resistance is calculated differently for normally consolidated and over-consolidated soils. The soil
consolidation state is indicated by the ratio of undrained shear strength and effective overburden pressure, s
u
/p
vo
. If s
u
/p
vo
0.4, the soil is
overconsolidated.
G-30 G-30
S0300-A8-HBK-010
For normally consolidated soils, the average unit skin resistance, f
s
, is equal to:
where L
p
is in feet. The limiting value for skin frictional resistance is the undrained shear strength of the soil, i.e., f
s
s
u
. For overconsolidated
f
s
p
vo

1
1
1
]
0.468 - 0.052 ln

,
L
p
2.0
soils:
For 2.0 < S
u
/p
vo
< 4.0, f
s
= 0.351s
u
.
f
s
= s
u

1
1
1
]
0.468 0.155 ln

,
s
u
p
vo
Average effective overburden soil pressure, p
vo
, at the pile midpoint is:
where:
p
vo
=

b
L
p
2

b
= soil buoyant unit weight, [force/length
3
]
L
p
= pile length, [length]
If the pile is not fully buried, L
p
is the buried length. If the soil unit weight varies with depth, unit weights along the buried length are averaged.
G-6.1.3 Compressive Load Capacity. For foundation piles, resistance to compressive loading comes from frictional resistance along the pile
and from resistance to tip or end penetration. For closed-ended piles, the soil bearing capacity for the pile tip, Q
p
, is:
where:
Q
p
= A
p
q
p
= p
vo, tip
N
q
for cohesionless soils
= 9s
u, tip
for cohesive soils
Q
p
= soil bearing capacity, lbs
A
p
= gross end area of the closed pile, ft
q
p
= unit soil bearing capacity at the pile tip, lb-ft
limiting values from Table G-15
p
vo, tip
= effective vertical stress at pile tip, lb-ft
N
q
= bearing capacity factor from Table G-15
s
u, tip
= soil undrained shear strength at pile tip, lb-ft
Open-ended piles will develop a soil plug inside the open end when installed. The soil plug limits the value of Q
p
to the force required to push
a soil plug up into the pile (a thin-walled pipe). This limiting value is approximately equal to the frictional capacity of the pile, Q
s
. Total pile
capacity in compression, Q
c
from Paragraph G-6.1.2, is thus:
Q
c
= Q
s
Q
p
If pile capacity is less than the design compressive load, capacity can be increased by increasing pile diameter or, preferably, length. Q
p
of an
open-ended pile is signicantly limited by the value of Q
s
. Q
c
may be increased by closing the pile end with a concrete plug or steel plate.
G-31 G-31
S0300-A8-HBK-010
G-6.1.4 Steel Stress Analysis. Maximum stress in the pile under tension (f
maxt
) and compression (f
maxc
) is calculated by:
where:
f
maxt
=
P
t
A
ps
M
max
S
f
maxc
=
P
c
A
ps
+
M
max
S
P
t
= vertical uplift at pile head, lbs
P
c
= horizontal load at pile head, lbs
A
ps
= cross-sectional area of the pile, ft
S = section modulus of the pile, ft
3
M
max
= maximum moment, ft-lbs
The terms A
ps
and S and allowable maximum stress in tension and compression are available from steel design manuals or manufacturers
literature. The values for f
maxt
and f
maxc
are compared with the allowable steel stress in tension and compression for the pile being used. For
most common structural shapes, the allowable maximum stress in tension and in bending is about 60 percent of yield, or about 22,000 psi.
It is also possible to reinforce the pile over the length where high moments exist. While this is a cost-effective alternative, calculation of loads in
piles of variable cross section is beyond the scope of this handbook.
Total maximum moment, M
max
, in the pile is the sum of any applied (design) bending moments, M
a
, and moments created by horizontal loads.
Total moment, M
t
, at any point along the pile is:
where:
M
t
A
m
P
h
T + M
a
B
m
A
m
= nondimensional moment coefficient a function of the depth coefficient z/T, from Figure G-26
P
h
= design horizontal load at the foundation pile, lbs
T = pile-soil relative stiffness = (EI/n
h
)
0.2
, in. (see paragraph G-6.1.1)
M
a
= applied bending moment, in-lb
B
m
= nondimensional moment coefficient, from Figure G-26
It may be necessary to determine M
t
at several locations along the pile in order to nd the maximum moment (M
max
).
Figure G-26. Moment Coefficients A
m
and B
m
.
D
E
P
T
H
C
O
E
F
F
I
C
I
E
N
T
z
/
T
D
E
P
T
H
C
O
E
F
F
I
C
I
E
N
T
z
/
T
0.2 0.2 0.1 0.1
4 4
3 3
2 2
1 1
0 0
0 0 0.4 0.4
A
m
B
m
0.8 0.8 0.6 0.6 1.0 1.0
MOMENT
COEFFICIENT (B
m
)
FOR APPLIED
MOMENT (M)
MOMENT COEFFICIENT
(A
m
) FOR APPLIED
LATERAL FORCE (P)
L
P
/T = 2
L
P
/ T = 2
L
P
/T = 3
L
P
/T = 4
L
P
/T = 10
L
P
/T = 5
L
P
/T = 3
L
P
/T = 10
L
P
/T = 4
L
P
/T = 5
AFTER NONDIMENSIONAL SOLUTIONS FOR LATERALLY LOADED PILES..., H. MATLOCK AND L.C. REESE,
PROCEEDINGS OF THE 8TH TEXAS CONFERENCE ON SOIL MECHANICS AND FOUNDATION ENGINEERING,
BUREAU OF ENGINEERING RESEARCH, UNIVERSITY OF TEXAS AT AUSTIN, 1956
G-32 G-32
S0300-A8-HBK-010
G-6.2 Submerged Anchor Pile Head. For anchor piles driven below the seaoor surface, the mooring line angle at the pile is not the same
as the angle at the seaoor, due to soil bearing resistance against the mooring line. The actual angle at the pile becomes higher and the force
exerted on the pile becomes more of an axial uplift. The seaoor horizontal and vertical (uplift) load components, P
h
and P
u
, are corrected to
reect the actual loads on the pile head. The force corrections are based on several simplifying assumptions, including the assumption that the
changes in both vertical and horizontal force components are due to the soil resistance to horizontal anchor line movement.
The correction to the horizontal force P
h
(horizontal soil force) is:
where:
f
cb
= z
2
c
d
b

b
N
q
for cohesionless soils
f
cb
= 11s
u
d
b
z
c
for cohesive soils
f
cb
= horizontal force exerted on the mooring line by the soil, lbs
Table G-16. Bearing Capacity Factors for Chain
Lateral Force in Sand.
Soil Friction Angle, degrees N
q
20
25
30
35
40
45
3
5
8
12
22
36
After G.G. Meyerhoff and J. F. Adams, The Ultimate Uplift Capacity of Foundations,
Canadian Geotechnical Journal, Nov 1968
z
c
= depth of pile connection below seaoor, ft
d
b
= characteristic mooring line size, ft
= wire rope diameter or 3 (chain diameter)

b
= soil buoyant weight, lb-ft
3
N
q
= bearing capacity factor, from Table G-16
s
su
= undrained shear strength, lb-ft
2
The corrected horizontal and uplift load components, P
h
and P
u
, are:
P
h
= P
h
f
cb
P
u
= P
2
u
+ 2P
h
f
cb
f
2
cb
G-6.3 Pile Anchors in Rock Seaoors.
ROCK
SEAFLOOR
CRUSHED
ROCK
FRACTURED
ROCK
(a) LATERAL BEARING
FAILURE OF ROCK
AND PILE
(b) UPLIFT FAILURE OF
GROUT-TO-ROCK
BONDING
(c) UPLIFT BLOCK FAILURE
OF PILE AND ROCK-MASS
Figure G-27. Pile Anchor Failure Modes in Rock.
There is no rigorous design procedure for
pile anchors in rock because of the
difficulty in characterizing the material
failure mode. Three principal rock failure
modes are illustrated in Figure G-27:
Failure in lateral bearing due
to rock crushing under lateral
load.
Failure in uplift due to failure
of grout-to-rock bonding or
because of a rock-mass
failure in fractured material.
Failure in uplift due to the
piles loosening and loss of
resistance from repeated
lateral loading.
The mode of failure is difficult to establish or predict for a specic location. The strength of a cored sample may be misleading when applied
to the prediction of pile anchor holding capacity in jointed, bedded, faulted, or weathered rock masses.
G-6.3.1 Lateral Capacity. In a rock or hard cemented soil seaoor, a soil cover may be present above the rock, or the rock/cemented zone
may be underlain by soil. For layered soil-rock sites, available computer programs should be used to account for the complexities introduced
by these nonuniform conditions.
G-6.3.2 Soil Overlying Rock. For soil overlying rock, pile capacity design approach depends on the relative depth to the rock, z
s
/T, where
z
s
is the thickness of the soil layer and T is the pile relative stiffness. If z
s/
T is greater than 3.0, the pile can be designed to develop all support
from the soil layer.
G-33 G-33
S0300-A8-HBK-010
If z
s
/T is less than 3.0, a conservative
Table G17. Compressive Strength of Rock.
Rock Type Compressive Strength, s
c
, ksi Buoyant Unit Weight lb/ft
3
Shear Strength, S
s
, ksi
Dolerite 28.4 49.8 123 3.5 8.5
Gabbro 25.6 42.7 123
Gneiss 7.1 28.4 117 2.1 7.0*
Basalt 21.3 42.7 111 2.8 8.5
Quartzite 21.3 42.7 101 2.8 8.5
Granite 14.2 35.6 98 2.0 6.5
Marble 14.2 35.6 98 3.0 6.1
Slate 14.2 28.4 98 2.0 10.2*
Dolomite 11.4 35.6 92 2.5 7.1
Limestone 4.3 35.6 73 1.4 7.1
Sandstone 2.8 24.1 61 1.1 5.7
Shale 1.4 14.2 61 0.4 4.3
Coal 0.7 7.1 5
After I. W. Farmer, Engineering Properties of Rocks, 1968, and Engineering Behavior of Rocks,
1983; and Richard E. Goodman, Introduction to Rock Mechanics, 1980*
approach is to design the pile as if the soil
were not present. That is, the pile is
considered to be cantilevered out of the
rock surface and to resist all forces without
assistance from the soil layer. Stresses
from the applied lateral load must be less
than the rocks compressive strength:
where:
P
h
DL
e
s
c
P
h
= lateral force applied to the
pile head, lbs
D = width or diameter of the pile,
in.
L
e
= effective length of the pile
bearing on the rock strata
equal to the smaller of the
rock layer thickness (z
r
) or
the pile diameter (D), in.
s
c
= rock compressive strength
from Table G-17, lb-in
2
G-6.3.3 Rock Layer Overlying Soil.
When or rock or other hard layer overlies
softer soil, the inuence of the rock layer
may be ignored and the pile designed to
develop all support from the underlying
soil, if the rock layer thickness, z
r
, is less
than 0.2T, where T = (EI/n
h
)
0.2
and n
h
is the
coefficient of soil reaction of the underlying
soil. For rock layer thickness greater than 0.5T, the inuence of the underlying soil may be ignored and the pile designed for the rock
compressive strength. For intermediate values of z
r
, the designer must judge whether to use the soil or rock procedure; the decision depends
upon pile diameter, knowledge of rock layer strength and fracturing, and the layer thickness.
G-6.3.4 Uplift Capacity. Failure in uplift may occur at the grout-to-pile interface, the grout-to-rock interface, or along a rock fracture zone
outside the grouted area. For failure in a fracture zone, a block of rock containing the pile is assumed to be lifted free of the surrounding rock.
In massive, competent rock, uplift capacity is governed by the strength of the grout bond to the pile, by the grout shear strength, or, more rarely,
by the strength of the grout bond to the drilled shaft wall. Uplift capacity, R
a
, of the anchor is given by:
where:
R
a
= s
b
L
r
C
p
s
b
= the lesser of the grout-pile bond strength, grout-rock bond strength, or grout shear strength, [force/length]
L
r
= length of pile embedded in rock, [length]
C
p
= minimum perimeter transmitting the uplift load, [length]
Unless higher bond strengths are veried by testing, the grout-to-steel bonding strength should be limited to 27 psi. The grout-to-rock bond
strength may vary from 0.3 to 1.0 times the rock shear strength, depending on cleanliness of the drilled hole, type of rock, and grouting
procedure.
In fractured rock, anchor uplift capacity is determined by the weighs of the blocks of rock which move with the anchor and by the frictional
force developed between the attached blocks and adjacent blocks. Because of the difficulty in estimating the normal forces acting on vertical
joints and cracks, this frictional force is normally ignored, and uplift resistance taken as the weight of the rock that would be lifted with the pile.
G-34 G-34
S0300-A8-HBK-010
G-6.4 Increasing Lateral Load Capacity.
Figure G-28. Improving Pile Anchor Lateral Capacity.
DISADVANTAGES
ILLUSTRATION
UNIDIRECTIONAL
LOADING.
SOIL IN FRONT OF
PILE MAY BE
WEAKENED.
MORE COSTLY
FABRICATION.
COMPLEX
INSTALLATION.
MORE COSTLY
FABRICATION.
LIMITED
EXPERIENCE
WITH SYSTEM.
ADVANTAGES LATERAL LOAD IS
REDUCED.
RESISTANCE IS
HIGHER
INCREASES
LATERAL
RESISTANCE.
LIMITS PILE HEAD
DEFLECTION AND
BENDING
MOMENT.
INCREASES
LATERAL AND
UPLIFT
RESISTANCE.
LATERAL LOAD
REDUCED.
PROVIDES FOR
SCOUR.
TECHNIQUE LOWERED
ATTACHMENT
POINT
ATTACHING FINS SHEAR COLLARS
WITH ANCHOR
PLATES
BURIED PILE HEAD
LOWERED
ATTACHMENT
POINT
BURIED
PILE HEAD
ATTACHMENT
FINS
CROSS
SECTION
SHEAR
COLLAR
ANCHOR
SHAFT
ANCHOR
PLATE
SHEAR COLLARS
WITH ANCHORS
PLATES
FROM NCEL HANDBOOK FOR MARINE GEOTECHNICAL ENGINEERING, KARL ROCKER JR, 1985
As shown in Figure G-28, there are four
principal means to increase lateral load
capacity of a pile anchor:
Lower the anchor line
attachment point along the
pile length.
Lower the pile head beneath
the soil surface into stronger
soils.
Attach ns or shear collars
near the pile head to increase
lateral bearing area.
Increase pile diameter near
the surface.
G-6.4.1 Load Applied Below the Pile
Head. When the anchor line is connected
to the side of an anchor pile at a distance
of more than ve pile diameters from the
head, the lateral load analysis becomes very
complex and may require use of a computer
program. When calculations must be made
without access to geotechnical engineering
services, the mooring line attachment should be kept less than ve pile diameters. As the connection point is lowered from the pile top to a
point midway down the pile and with the same lateral loads, pile bending moments and deection are reduced signicantly. The mooring line
will approach the pile at a decreasing angle as the connection point is lowered, greatly lowering lateral load and increasing uplift force.
G-6.4.2 Piles with Variable Cross Sections. Increased pile size near the seaoor will increase resistance to lateral loads and bending moments.
Analysis of the response of piles with variable cross sections to lateral loads is complex and, again, best accomplished with the aid of computer
programs. In general, a large increase in pile diameter over a lesser depth is more efficient in reducing deections at the seaoor than is a small
increase in diameter over a greater depth. When enlarged pile sections at the pile head are used, the length of the enlarged section should be
limited to three times the larger diameter.
G-6.5 Special Seaoor Conditions. Steeply sloping seaoors, rock, cobbles, or cemented zones can make installation of driven piles difficult,
although other installation methods may be successful. Drilled and grouted piles may be the method of choice in these environments. Scour
of sediments from around the pile/seaoor interface, typically in areas with swift bottom currents, weakens resistance to lateral force. Soil type
and size and conguration of pile groups inuences scour pattern and rate. Liquefaction of loose granular or sandy sediments by cyclic loading
can effectively remove soil support and cause pile failure. Removal of the surface layers prior to pile installation can signicantly improve pile
performance in granular soils. Slump or sub-sea landslide on slopes can subject piles to high lateral force causing failure or breakout of pilings.
G-6.6 Pile Installation. Piles are installed by one or more of the following methods:
Driving,
Drilling and grouting,
Jacking, or
Jetting.
Small piles can be installed with diver-operated equipment. Pile installation equipment and operators can be provided by Navy Underwater
Construction Teams, the Army Corps of Engineers, or contractors retained through the Supervisor of Salvage. The following information is
provided to give the salvage engineer an idea of the relative complexity, expense, and applicability of pile installation by various methods, but
is in general insufficient for planning major pile installations.
G-35 G-35
S0300-A8-HBK-010
G-6.6.1 Driven Piles. Piles may be driven by impact hammers operated above the water surface, by underwater impact hammers, or by
vibratory hammers.
Piles for piers, harbor structures, bridges, and many offshore structures in shallow water are driven from above the water surface with
conventional hammers like those used to drive piles on land. The pile is made long enough to extend above the water surface when driven to
its design penetration depth or a pile follower is used. The piles are commonly guided by a template that rests on the seaoor, although oating
templates may be used for small, shallow water installations. The pile-driving operation is conducted from a carefully moored work barge that
supports the necessary cranes and auxiliary equipment.
Submarine pile hammers are scaled-up versions of terrestrial hammers, and may be operated by single-acting steam, compressed air, diesel, or
hydraulic power. The rated energy of these hammers varies from less than 100,000 ft-lb per blow to over 1,500,000 ft-lb per blow. Surface-
operated pile drivers have been used in water depths in excess of 1,000 feet.
The success of the surface-driven method of pile emplacement in deep water is dependent upon the presence of the template to act as a guide
for the piles. Without the restraint offered by the template, most of the driving energy would be dissipated by lateral deection of the pile.
For anchor piles driven from the surface without lateral restraint, a reasonable maximum water depth is about 250 feet.
Terrestrial pile hammers may be modied for operation underwater. One manufacturer makes a total of 12 types of steam/compressed air
hammers, with rated energies in air of 8,750 ft-lb to 60,000 ft-lb. These may be operated while submerged with little loss of efficiency. The
modications consist primarily of providing exhaust hoses that extend to the water surface. Because steam cools too much when the hoses are
underwater, compressed air is usually used to operate the hammers.
Vibratory pile drivers are becoming more common in American practice as experience is gained with their use and as more powerful machines
are developed. The machines usually use counter-rotating eccentric weights powered by electric or hydraulic motors to produce the vibratory
forces. The major depth-limiting factors on present systems are the difficulty in handling long lengths of large-diameter, high-pressure hydraulic
lines and the large friction losses in the line. These factors limit the maximum practical water depth of a surface-powered, hydraulic vibratory
drive to about 1,000 feet.
G-6.6.2 Drilling and Grouting. Drilling and grouting is essentially identical to the method used to set a casing for an oil well. A hole of
somewhat larger diameter than the pile is drilled to the proper depth using rotary drilling tools and is cleaned out by pumping seawater through
the drill string. The pile is placed over the drill string and lowered into the hole. Portland cement grout is pumped down the drill string and
forced up outside of the pile to ll the annular void and bond the pile to the soil. The interior of the pile is lled with grout as the drill string
is withdrawn. Piles up to 8 feet in diameter have been placed in water depths in excess of 600 feet by drilling and grouting.
For small piles set in rock, either cement or epoxy grout can be used. Diver-operated hydraulic tools capable of drilling 3-inch diameter holes
to a depth of 20 feet, along with diver-operated grout dispensers, are used by Navy Underwater Construction Teams (UCT) and NCEL. Piles
tting such holes may have capacities in the 10- to 60-ton range, depending on rock strength.
G-6.6.3 Jack-in Piles. Piles may be pushed or jacked into the seaoor if an adequate reaction force can be applied. For a satisfactory degree
of safety against failure in bearing of foundation piles, jacking loads must be two to three times the design load. The actual jacking of the piles
can be accomplished by a number of systems. A rack-and-pinion system may be used, with the rack being an integral part of the pile and
running its entire length. A chain acted on by a chain jack or a cable acted on by a hydraulic cable puller may be used, with the chain or cable
applying load to the top of the pile. A short-stroke hydraulic jack equipped with a means of gripping the wall of a pile may also be used.
G-6.6.4 Jetted Piles. Jetting is used to place piles primarily in cohesionless soils. The piles are pushed or lowered into the soil area, which
has been greatly weakened by jetting. The jetting action is generally conned to the inside of a pile or to portions of the outside of the pile
several diameters above its tip. Jetting can also be used in a form of reverse circulation in which both air and water are forced down a pipe
inside or outside the pile. The air-water mixture helps to lift the displaced soil materials to the surface of the soil.
G-36 G-36

Você também pode gostar