SPM 1 - Coastal BW
SPM 1 - Coastal BW
SPM 1 - Coastal BW
^m^t^
U.S. Army
Coastal Engineering
Research Centei
DATA LIBRARY
Woods Hole Oceanographic Institution
SHORE PROTECTION
MANUAL
VOLUME I
I
m
i cr ( Chapters 1 Through 4
; J
s a
: -0
°
:
= o
i 1^
: m
S°a ;
1975
Second Edition
since 1930, conducted studies on shore processes and methods of shore protection. CERC continues an
extensive research and development program to improve both shore protection and offsliore engineering
techniques. The scientific and engineering aspects of coastal processes and offshore structures are in the
developmental stage and the requirement for improved techniques for use in design and engineering of
coastal structures is evident. This need was met in 1954, to the extent of available knowledge, by
publication of Teclmical Report Number 4, "Shore Protection, Planning and Design" (TR 4); revised
coastal zone starting with an introduction of coastal engineering, continuing with discussions of mechanics
of wave motion, wave and water level predicitons, and finally littoral processes.
Volume II translates the interaction of the physical environment and coastal structures into design
parameters for use in the solution of coastal engineering problems. It discusses planning, analysis, structural
features, and structural design as related to physical factors, and shows an example of a coastal engineering
problem which utilizes the technical content of material presented in all three volumes.
Volume III contains four appendixes including a glossary of coastal engineering terms, a list of symbols,
tables and plates, and a subject index.
R. A. Jachowski, Chief, Design Branch, Engineering Development Division, was the project engineer
responsible for preparation and assemblage of the text, under the general supervision of G. M. Watts, Chief,
Engineering Development Division. At the time of approval for publication by the Coastal Engineering
Research Board, LTC Don S. McCoy was Commander and Director and Thorndike Saville, Jr. was Technical
Director. Members of the Coastal Engineering Research Board were: MG John W. Morris (President),
MG Daniel A. Raymond, MG Ernest Graves, Jr., BG George B. Fink, Dean Morrough P. O'Brien, Dr. Arthur
T. Ippen, and Prof. Robert G. Dean. The board members were intimately involved in both the planning and
review of this manual.
Preparation of this manual includes the contribution, review and suggestions of numerous engineers,
scientists, technical and support personnel. Present members of the CERC staff who made significant
Ill
TABLE OF CONTENTS
VOLUME n
CHAPTER PAGE
5 PLANNING ANALYSIS
5.1 GENERAL 5-1
5.2 SEAWALLS. BULKHEADS, AND REVETMETS 5-3
5.3 PROTECTIVE BEACHES . ^ 5-7
5.4 SAND DUNES 5-21
5.5 SAND BYPASSING 5-24
5.6 GROINS 5-31
5.7 JETTIES 5-46
5.8 BREAKWATERS-SHORE-CONNECTED 5-49
5.9 BREAKWATERS-OFFSHORE 5-50
5.10 ENVIRONMENTAL CONSIDERATIONS 5-57
REFERENCES AND SELECTED BIBLIOGRAPHY 5-58
6 STRUCTURAL FEATURES
6.1 INTRODUCTION 6-1
6.2 SEAWALLS, BULKHEADS, AND REVETMENTS 6-1
6.3 PROTECTIVE BEACHES 6-16
6.4 SAND DUNES 6-36
6.5 SAND BYPASSING 6-54
6.6 GROINS 6-76
6.7 JETTIES 6-84
6.8 BREAKWATERS-SHORE-CONNECTED 6-88
6.9 BREAKWATERS-OFFSHORE 6-96
6.10 CONSTRUCTION MATERIALS 6-96
6.11 MISCELLANEOUS DESIGN PRACTICES 6-98
REFERENCES AND SELECTED BIBLIOGRAPHY 6-101
VOLUME III
APPENDIX
A GLOSSARY OF TERMS A-1
B LIST OF SYMBOLS B-1
C MISCELLANEOUS TABLES AND PLATES C-1
D SUBJECT INDEX D-1
IV
TABLE OF CONTENTS
VOLUME I
SECTION PAGE
VI
SECTION PAGE
vJi
SECTION PAGE
VIII
SECTION PAGE
ix
SECTION PAGE
FIGURE PAGE
XI
FIGURE PAGE
Xtl
FIGURE PAGE
XIII
FIGURE PAGE
XIV
FIGURE PAGE
4-1 Typical Profile Changes with Time, Westhampton Beach, N.Y.. 4-2
4-2 Three Types of Shoreline 4-3
4-3 Shoreline Erosion near Shipbottom, N.J 4-7
4-4 Shoreline Accretion and Erosion Near Beach Haven, N.J.. . . 4-8
4-5 Stable Shoreline Near Peahala, N.J 4-9
4-6 Fluctuations in Location of Mean Sea Level Shoreline on
Seven East Coast Beaches 4-12
4-7 Grain Size Scales 4-14
4-8 Example Size Distribution 4-17
4-9 Sand Size Distribution Along the U.S. Atlantic Coast. . . . 4-23
4-10 Mean Monthly Nearshore Wave Heights for Five Coastal
Segments 4-31
4-11 Mean Monthly Nearshore Wave Period for Five Coastal
Segments 4-32
4-12 Distribution of Significant Wave Heights from Coastal
Wave Gages for 1-year Records 4-34
4-13 Nearshore Current System Near LaJolla Canyon, California. . 4-44
4-14 Typical Rip Currents, Ludlam Island, N.J 4-46
4-15 Distribution of Longshore Current Velocities 4-47
4-16 Measured versus Predicted Longshore Current Speed 4-49
4-17 Coasts in the Vicinity of New York Bight 4-51
4-18 Three Scales of Profiles, Westhampton, Long Island 4-55
4-19 Unit Volume Change versus Time Between Surveys for Profiles
on South Shore of Long Island 4-59
4-20 Maximum Wave Induced Bottom Velocity as a Function of
Relative Depth 4-62
4-21 Maximum Bottom Velocity from Small Amplitude Theory .... 4-63
4-22 Initiation of Ripple Motion 4-66
4-23 Wave Conditions Producing Maximum Bottom Velocity of
0.5 ft/sec 4-67
4-24 Nearshore Bathymetry with Shore-Parallel Contours off
Panama City, Florida 4-68
4-25 Nearshore Bathymetry with Shore-Parallel Contours and
Linear Bars off Manasquan, N.J 4-69
4-26 Slow Accretion of Ridge-and-Runnel at Crane Beach, Mass.. . 4-76
4-27 Rapid Accretion of Ridge-and-Runnel - Lake Michigan .... 4-77
4-28 Typical Berm and Bar Profiles from Prototype Size
Laboratory Wave Tank 4-79
XV
FIGURE PAGE
XVI
LIST OF TABLES
TABLE PAGE
XVII
CHAPTER 1
INTRODUCTION
TO
COASTAL ENGINEERING
^AA^At<c~
HALEIWA BEACH, OAHU, HAWAII - 23 August 1970
CHAPTER 1
l-l
functional design of various structures. Chapter 7, "Structural
—
Design Physical Factors," treats tke effects of environmental forces
on the design of protective works. Chapter 8, "Engineering .Analysis
Case Study," presents a series of calculations for the preliminary
design of an offshore island facility in the mouth of the Delaware
Bay.
Where the land meets the ocean at a sandy beach, the shore has
natural defenses against attack by waves, currents and storms. First
of these defenses is the sloping nearshore bottom that causes waves to
break offshore, dissipating their energy over the surf zone. The pro-
cess of breaking often creates an offshore bar in front of the beach
that helps to trip following waves. The broken waves re-form to break
again, and may do this several times before finally rushing up the beach
foreshore. At the top of wave uprush a ridge of sand is formed. Beyond
this ridge, or crest of the berm, lies the flat beach berm that is
reached only by higher storm waves. A beach profile and its related
terminology are shown in Figure 1-1.
Winds blowing inland over the foreshore and berm move sand behind the
beach to form dunes. (See Figures 1-2 and 1-3.) Grass, and sometimes
bushes and trees, grow on the dunes, and the dunes become a natural levee
against sea attack. Dunes are the final natural protection line against
wave attack, and are also a reservoir for storage of sand against storm
waves
1-2
Figure 1-2. Sand Dunes along the South Shore of Lake Michigan
1.23 BARRIER BEACHES, LAGOONS AND INLETS
Diiring storms, strong winds generate high waves. Storm surge and
waves may raise the water level near the shore. If storm surge does
occur, large waves can then pass over the offshore bar formation without
breaking. If the storm occurs at high tide, storm surge super-elevates
the water, and some waves may break on the beach or even at the base of
the dunes. After a storm or storm season, natural defenses may again be
re-formed by normal wave and wind action.
Most of the sands of the beaches and nearshore slopes are normally
small, resistant rock particles that have traveled many miles from in-
land mountains. When the sand reaches the shore, it is moved alongshore
by waves and littoral currents. This alongshore transport is a constant
process, and great volumes may be transported. In most coastal segments
the direction of movement changes as direction of wave attack changes.
The moon, and to a lesser extent the sun, creates ocean tides by
gravitational forces. These forces of attraction, and the fact that the
sun, moon, and earth are always in motion with relation to each other,
cause waters of ocean basins to be set in motion. These tidal motions of
water masses are a form of very long period wave motion, resulting in a
rise and fall of the water surface at a point. There are normally two
tides per day, but some localities have only one per day.
1.32 WAVES
The familiar waves of the ocean 'are wind waves generated by winds
blowing over water. They may vary in size from ripples on a pond to
1-5
I
-6
large ocean waves as high as 100 feet. (See Figure 1-5.) Wind waves
cause most of the damage to the ocean coasts. Another type of wave,
the tsunami, is created by earthquakes or other tectonic disturbances
on the ocean bottom. Tsunamis have caused spectacular damage at times,
but fortunately, major tsunamis do not occur frequently.
Wind waves are of the type known as oscillatory waves, and are
usually defined by their height, length, and period. (See Figure 1-6.)
Wave height is the vertical distance from the top of the crest to the
bottom of the trough. Wavelength is the horizontal distance between
successive crests„ Wave period is the time between successive crests
passing a given point.
As waves propagate in water, only the form and part of the energy
of the waves move forward; the water particles remain.
The height, length, and period of wind waves are determined by the
fetch (the distance the wind blows over the sea in generating the waves),
the wind speed, the length of time the wind blows, and the deoay distance
(the distance the wave travels after leaving the generating area). Gener-
ally, the longer the fetch, the stronger the wind; and the longer the
time the wind blows, the larger the waves. The water depth, if shallow
enough, will also affect the size of wave generated. The wind simul-
taneously generates waves of many heights, lengths, and periods as it
blows over the sea.
Currents are created in oceans and adjacent bays and lagoons when
water in one area becomes higher than water in another area. Water in
the higher area flows toward the lower area, creating a current. Some
causes of differences in the elevation of the water surface in the oceans
are tides, wind, waves breaking on a beach, and streams. Changes in water
temperature or salinity cause changes in water density that may also pro-
duce currents.
1-7
li^Mirc l-fi. Large Wuve:; lii'caking ovci- u lireakwater
Wove Crest -^
Crest Lcngtti
Region
Stillwoler level
TrouQti Lcnqtii
Region d^ Oeptti
Oceon Bottom-
1-8
setup or storm surges are created by strong winds. The height of stomi
surge depends on wind velocity and dii'ection, fetch, water depth, ajid
ncarshore slope. In violent storms, storm surge may raise sea level at
the shore as much as 20 feet. In the United States, larger surges occur
on the Gulf coast because of the shallower and broader shelf off that
coast compared to the shelves off the Atlantic and Pacific coasts. Storm
surges may also be increased by a funneling effect in converging estuaries.
If water level rises :uid falls at an area, then water must flow into
and out of the area. Significant currents generated by tides occur at in-
lets to lagoons and bays or at entrances to harbors. At such constricted
places, tidal currents generally flow in when the tide is rising (flood
tide) and flow out as the tide falls (ebb tide). Exceptions can occur
at times of high river discharge or strong winds, or when density cur-
rents are an important part of the current system.
1-9
of the finer sizes of sand. Daytona Beach, Florida, is a good example
of a gently sloping beach composed of fine sand.
lo43 BREAKERS
Wind waves affect beaches in two major ways. Short steep waves,
which usually occur during a storm near the coast, tend to tear the
beach down. (See Figure 1-7.) Long swells, which originate from dis-
tant storms, tend to rebuild the beaches. On most beaches, there is a
constant change caused by the tearing away of the beach by local storms
followed by gradual rebuilding by swells. A series of violent local
storms in a short time can result in severe erosion of the shore if
there is not enough time between storms for swells to rebuild the
beaches. Alternate erosion and accretion of beaches may be seasonal
on some beaches; the winter storms tear the beach away, and the summer
swells rebuild it. Beaches may also follow long-term cyclic patterns.
They may erode for several years, and then accrete for several years.
l-IO
•Dune Crest
Crest :.:. I I
ACCRETION
Profile A
M.L.W.
Profile D - After storm wove ottock,
normol wove oction
ACCRETION
Profile A
Figure 1-7, Schematic Diagram of Storm Wave Attack on Beach and Dune
l-ll
transport direction can vary from season to season, day to day or hour
to hour. These reversals of transport direction are quite common for
most United States shores. Direction may vary at random, but in most
areas the net effect is seasonal.
Hurricanes or severe storms moving over the ocean near the shore may
greatly change beaches. Strong winds of a storm often create a storm surge.
This surge raises the water level and exposes to wave attack higher parts
of the beach not ordinarily vulnerable to waves. Such storms also generate
1-12
large, steep waves. These waves carry large quantities of sand from the
beach to the nearshore bottom. Land structures, inadequately protected
and located too close to the water, are then subjected to the forces of
waves and may be damaged or destroyed. Low- lying areas next to the ocean,
lagoons, and bays are often flooded by storm surge. Storm surges are
especially damaging if they occur concurrently with astronomical high tide.
Beach berms are built naturally by waves to about the highest eleva-
tion reached by normal storm waves. Berms tend to absorb the wave energy;
however, overtopping permits waves to reach the dunes or bluffs in back of
the beach and damage unprotected upland features.
When storm waves erode the berm and carry the sand offshore, the pro-
tective value of the berm is reduced and large waves can overtop the beach.
The width of the berm at the time of a storm is thus an important factor
in the amount of upland damage a storm can inflict.
During the early days of the United States, natural beach processes
continued to mold the shore as in ages past. As the country developed,
activity in the shore area was confined principally to harbor areas.
Between harbor areas, development along the shore progressed slowly as
small, isolated, fishing villages. As the national economy grew, im-
provements in transportation brought more people to the beaches. Gradu-
ally, extensive housing, commercial, recreational and resort developments
replaced fishing villages as the predominant coastal manmade features.
Examples of this development are Atlantic City and Miami Beach.
1-13
There are places where the beach has been gradually widened, as well
as narrowed, by natural processes over the years. This is evidenced by
lighthouses and other structures that once stood on the beach, but now
stand hundreds of feet inland.
While the sloping beach and beach berm are the outer line of defense
to absorb most of the wave energy, dunes are the last zone of defense in
absorbing the energy of storm waves that overtop the berm. Although dunes
erode during severe storms, they are often substantial enough to afford
complete protection to the land behind them. Even when breached by waves
of a severe storm, dunes may gradually rebuild naturally to provide pro-
tection during future storms. Continuing encroachment on the sea with
manmade development has often taken place without proper regard for the
protection provided by dunes. Large dune areas have been leveled to make
way for real estate developments, or have been lowered to permit easy
access to the beach. Where there is inadequate dune or similar protection
against storm waves, the storm waters may wash over low- lying land, mov-
ing or destroying everything in their path, as illustrated by Figure 1-8.
1-14
2
M
•H
0)
I—
(A
to
E
nj
Q
u
CO
I -15
1.54 BULKHEADS, SEAWALLS AND REVETMENTS
1.55 BREAKWATERS
1-16
shelter for a floating dredge to pump the impounded material across the
navigation opening back onto the downdrift beach. This method is used
at Channel Island Harbor near Port Hueneme, California.
1.56 GROINS
1.57 JETTIES
1-17
Jetties aid navigation by reducing movement of sand into the channel,
by stabilizing the location of the channel, and by shielding vessels from
waves. Sand is impounded at the updrift jetty, and the supply of sand to
the shore downdrift from the inlet is reduced, thus causing erosion of
that shore. Before the installation of a jetty, nature supplied sand by
transporting it across the inlet intermittently along the outer bar to the
downdrift shore.
1-18
Figure 1-9. Weir Jetty at Masonboro Inlet, North Carolina
1-19
I
-20
Sometimes structures must be provided to protect dunes, to maintain
a specific beach dimension, or to reduce nourishment requirements. In
each case, the cost of such structures must be weighed against the bene-
fits they would provide. Thus, measures to provide and keep a wider pro-
tective and recreational beach for a short section of an eroding shore
would require excessive nourishment without supplemental structures such
as groins to reduce the rate of loss of material from the widened beach.
A long, high terminal groin or jetty is frequently justified at the down-
drift end of a beach restoration project to reduce losses of fill into an
inlet and to stabilize the lip of the inlet.
Experience and study have demonstrated that sand from dunes, beaches,
and nearshore areas is the best material available naturally in suitable
form to protect shores. Where sand is available in abundant quantities,
protective measures are greatly simplified and reduced in cost. When
dunes and broad, gently sloping beaches can no longer be provided, it is
necessary to resort to alternative structures, and the recreational attrac-
tion of the seashore is lost or greatly diminished.
1-21
facilities are being considered that will trap the sand before it reaches
the canyon and transport it mechanically to a point where it can resume
normal longshore transport. Dune planting with appropriate grasses and
shmbs reduces landward windbome losses and aids in dune preservation.
1-22
CHAPTER 2
MECHANICS
OF
WAVE MOTION
LEO CARRILLO STATE BEACH, CALIFORNIA - July 1968
CHAPTER 2
2.1 INTRODUCTION
2.21 GENERAL
2-1
spray. When waves move out of the area where they are directly affected
by the wind, they assume a more ordered state with the appearance of
definite crests and troughs and with a more rhythmic rise and fall. These
waves may travel hundreds or thousands of miles after leaving the area in
which they were generated. Wave energy is dissipated internally within
the fluid by interaction with the air above, by turbulence on breaking,
and at the bottom in shallow depths.
Waves which reach coastal regions expend a large part of their energy
in the nearshore region. As the wave nears the shore, wave energy may be
dissipated as heat through turbulent fluid motion induced by breaking and
through bottom friction and percolation. While the heat is of little
concern to the coastal engineer, breaking is important since it affects
both beaches and manmade shore structures. Thus, shore protection measures
and coastal structure designs are dependent on the ability to predict wave
forms and fluid motion beneath waves, and on the reliability of such
predictions. Prediction methods generally have been based on simple waves
where elementary mathematical functions can be used to describe wave motion,
For some situations, simple mathematical formulas predict wave conditions
well, but for other situations predictions may be unsatisfactory for
engineering applications. Many theoretical concepts have evolved in the
past two centuries for describing complex sea waves; however, complete
agreement between theory and observation is not always found.
2-2
theories. The first finite amplitude theory, known as the trochoidal
theory, was developed by Gerstner (1802). It is so called because the
free surface or wave profile is a trochoid. This theory is mentioned
only because of its classical interest. It is not recommended for appli-
cation, since the water particle motion predicted is not that observed in
nature. The trochoidal theory does, however, predict wave profiles quite
accurately. Stokes (1880) developed a finite-amplitude theory which is
more satisfactory than the trochodial theory. Only the second-order Stokes'
equations will be presented, but the use of higher order approximations is
sometimes justified for the solution of practical problems.
the theories mentioned above as well as others. The mathematics used here
generally will be restricted to elementary arithmetic and algebraic opera-
tions. Emphasis is placed on selection of an appropriate theory in accord-
ance with its application and limitations.
2-3
which are difficult to describe in form or motion, and which may be com-
prised of several components are termed complex waves. Sinusoidal or
simple harmonic waves are examples of simple waves since their surface
profile can be described by a single sine or cosine function. A wave is
periodic if its motion and surface profile recur in equal intervals of
time. A wave form which moves relative to a fluid is called a progressive
wave; the direction in which it moves is termed the direction of wave
propagation. If a wave form merely moves up and down at a fixed position,
it is called a complete standing wave or a clapotis. A progressive wave
is said to be a wave of permanent form if it is propagated without experi-
encing any changes in free surface configuration.
(a) seas, when the waves are under the influence of wind in a
generating area, and
(b) swell, when the waves move out of the generating area and
are no longer subjected to significant wind action.
Seas are usually made up of steeper waves with shorter periods and
lengths, and the surface appears much more confused than for swell. Swell
behaves much like a free wave, i.e., free from the disturbing force that
caused it, while seas consist to some extent of forced waves, i.e., waves
on which the disturbing force is applied continuously.
2-4
Ocean waves are complex. Many aspects of fluid mechanics necessary
to a complete discussion have only a minor influence on solving most
coastal engineering problems. Thus, a simplified theory which omits most
of the complicating factors is useful. The assumptions made in developing
the simple theory should be understood, because not all of the assumptions
are justified in all problems. When an assumption is not valid in a
particular problem, a more complete theory should be employed.
(f) The particular wave being considered does not interact with
any other water motions.
The first three are acceptable for virtually all coastal engineering
problems. It will be necessary to relax assumptions (d) , (e), and (f)
for some specialized problems not considered in this Manual. Relaxing
the three final assumptions is essential in many problems, and is con-
sidered later in this chapter.
2-6
In applying assumption (g) to waves in water of varying depth encoun-
tered when waves approach a beach the local depth is usually used. This
can be rigorously justified, but not without difficulty, for most practical
cases in which the bottom slope is flatter than about 1 on 10. A progres-
sive wave moving into shallow water will change its shape significantly.
Effects due to viscosity and vertical velocity on a permeable bottom may
be measurable in some situations, but these effects can be neglected in
most engineering problems.
crest from the preceding trough) ; period T (the time for two successive
crests to pass a given point) ; and depth d (the distance from the bed
to the Stillwater level). (See Appendix B for a list of common symbols.)
2.231 Wave Celerity, Length and Period The speed at which a wave form
.
C = (2-1)
^
An expression relating the wave celerity to the wavelength and water depth
is given by
C = (2-2)
2-7
o
CO
o :
Q. :
>
o
<u
>
> •H
o „ o
> S in
X \ o
5 o M
o c= r- *- O
o U
o OJ X
o.
o ^^
c
(V
ro X I
II
.^ I"
II
o •H
o O
I
m
II II
I-
O l_
o C
•H
o
a>
+->
c
0)
£
0)
1—1
II
IM J)
B
u
E H
0)
a> o
a> 4h
o O
C
o
Hc
<4-l
a>
Q
u
3
W)
2-8
From Equation 2-1, it is seen that 2-2 can be written as
C = —
gT
2iT
,
tanh
/27rd
L
(2-3)
\ >
The values 2Tr/L and 2Tr/T are called the wave number k and wave angular
frequency w, respectively. From Equations 2-1 and 2-3 an expression
for wavelength as a function of depth and wave period may be obtained.
L = tanh (2-4)
2-n
Use of Equation 2-4 involves some difficulty since the unknown L, appears
on both sides of the equation. Tabulated values in Appendix C may be used
to simplify the solution of Equation 2-4.
Classification
subscript is omitted. (Ippen, 1966b, pp 21-240 If units of feet and
seconds are specified, the constant g/2Tr is equal to 5.12 ft/sec^ and
C
"
= ^
2-n
= 5.12 T (ft/sec) , (2-7)
and
If Equation 2-7 is used to compute wave celerity when the relative depth
is d/L = 0.25, the resulting error will be about 9 percent. It is evi-
dent that a relative depth of 0.5 is a satisfactory boundary separating
deepwater waves from waves in water of transitional depth. If a wave is
traveling in transitional depths. Equations 2-2 and 2-3 must be used with-
out simplification. Care should be exercised to use Equations 2-2 and 2-3
when necessary, that is, when the relative depth is between 1/2 and 1/25.
When the relative water depth becomes shallow, i.e., 2TTd/L < 1/4 or
d/L < 1/25, Equation 2-2 can be simplified to
C = yf^. (2-9)
2.232 The Sinusoidal Wave Profile The equation describing the free
.
17 = a cos I
— 1 = — cos I
— -— ) , (2-10)
(2-11)
2-10
If both sides of Equation 2-11 are multiplied by d/L, it becomes:
—
d d
= - tanh
, /27rd\
(2-12)
The term d/L^ has been tabulated as a function of d/L by Wiegel (1954),
and is presented in Appendix C on Table C-1. Table C-2 includes d/L as
a function of d/L^ iri addition to other useful functions such as 2ird/L
and tanh(2Trd/L) These functions simplify the solution of wave problems
.
SOLUTION:
d 600
1.1719.
Lo 512
d
->1.0
K L'
therefore
2-1
By Equation 2-1
L 512
T T
C = —
512
= 51.2 ft/sec.
For d = 10 feet
0.0195.
Z; ~ 512
- = 0.05692.
L
hence
10
= 176. r I . . ,
reet [transitional depth, since
, ,
—<
1 d
— l\
<—),
0.05692 \ Ad i-f A j
176
c = = 17.6 ft/sec.
10
H gT sinh[27r(z + d)/L]
w = — sin
2 L cosh (27rd/L)
occurs when 9 = 0, 2tt, etc., while the maximum horizontal velocity in the
negative direction occurs when 9 = tt. Sit, etc. On the other hand the
maximum positive vertical velocity occurs when 9 = tt/2, 5tt/2, etc., and
the maximum vertical velocity in the negative direction occurs when
e = 377/2, 777/2, etc. (See Figure 2-3.)
c
o
i-i
<U
•—<
<u
o
u
<
T5
C
If)
0)
u
o
o
o
to
I
CM
0)
2-14
hence
50
L = = 270 feet
0.1854
27:(50- 15)
u = 4.88 cosh [cos 60°] = 4.88 [cosh (0.8145)] (0.500).
270
From Table C-1 find
lird
= 0.8145 ,
and by interpolation
and
Therefore
Figure 2-3, a sketch of the local fluid motion, indicates that the
fluid under the crest moves in the direction of wave propagation and
returns during passage of the trough. Linear theory does not predict any
mass transport; hence the sketch shows only an oscillatory fluid motion.
*************************************
2.235 Water Particle Displacements Another important aspect of linear
.
2-15
mean particle position is considered to be at the center of the ellipse or
circle, then vertical particle displacement with respect to the mean
position cannot exceed one-half the wave height. Thus, since the wave
height is assigned to be small, the displacement of any fluid particle from
its mean position is small. Integration of Equations 2-13 and 2-14 gives
the horizontal and vertical particle displacement from the mean position,
respectively. (See Figure 2-4.)
Thus,
HgT^ cosh[27r(z + d)/L]
I = - sin (2-17)
47rL cosh (27rd/L)
—^
2ire
tanh
, 277d
.
Thus,
H
—
cosh [277(z + d)/L] .
/27rx 277t
— — (2-19)
; sin I
2 sinh(277d/L) \ L T
H
—
sinh [277(z + d)/L] /277X
—- \
277t
; cos I I. (2-20)
2 sinh (277d/L) \ L T /
— + ^r = 1, (2-21)
in which
H cosh[277(z+d)/L]
A = (2-22)
2 sinh (277d/L)
H sinh[277(z+d)/L]
^ " (2-23)
2 sinh(277d/L) '
2-16
(D
>
CO
<D
CI,
(D
Q
TD
C
cfl
c
o
o
a.
c
0)
S
o
u
U)
c
0)
E
(D
O
c«
i-H
cx
Q
(U
I—
o
ni
3
•H
2-17
Equation 2-21 is the equation of an ellipse with a major (horizontal)
semiaxis equal to A, and a minor (vertical) semiaxis equal to B. The
lengths of A and B are measures of the horizontal and vertical dis-
placements of the water particles. Thus, the water particles are predicted
to move in closed orbits by linear wave theory: i.e., each particle returns
to its initial position after each wave cycle. Morison and Crooke (1953),
compared laboratory measurements of particle orbits with wave theory and
found, as had others, that particle orbits were not completely closed. This
difference between linear theory and observations is due to the mass trans-
port phenomenon which is discussed in a subsequent section.
A=B = -e'^^/L
2
for->-.
L 2
(2-24)
B
SOLUTION :
2ir \ L ,
Equation 2-1,
T
Therefore, equating 2-1 and 2-3,
T 2n \ L r
and multiplying both sides by (2tt)2/lt
C 2 ~
cosh(27rd/L) ^""llT F;
since
T _ 1
L ~ C
Since
C = C
2n
.anh M
\ L J
2-19
and since,
L ~ cosh (27rd/L)
y y '
therefore,
T sinh (27rd/L) \ L T /
*************************************
************** EXAMPLE PROBLEM **************
GIVEN A wave in a depth of d = 40 feet, height of H = 10 feet, and a
:
FIND :
(c) For the deepwater conditions of (b) above, show that the particle
displacements are small relative to the wave height when z = - Lo/2.
SOLUTION :
d 40
0.0781 .
Lc 512
sinh (
— 1= 0.8394 ,
1
/27rd\
tanh I 1= 0.6430 .
H 1
A =
2 tanh(27rd/L) '
2-20
and Equation 2-23 reduces to
H
B = -
2
Thus,
10 1
A = = 7.78 feet ,
2 (0.6430)
H 10
B = -
2
= —
2
= 5.0 feet
When z = - d,
A = = = 5 96 feet
2sinh(27rd/L) 2(0.8394)
and, B = 0.
/27rz\
= —
27r(- 25)
^ = - 0.307
\L
,
I 512
thus
e-0.307 = 0.736 .
Therefore,
H
A=B= — 10.45
27:2/
e ^^ = (0.736) = 3.85 feet .
2 2
^o "512
r
Cc)
^
z = = = - 256 feet ,
2 2
L 512
H^
A = B ~ 2
e
2mL
f^
10.45
= -^^— (0.043) = 0.225
2
feet .
2-21
Thus the maximum displacement of the particle is 0.45 feet which is small
when compared with the deepwater height, H^ = 10.45 feet.
*************************************
2.236 Subsurface Pressure Subsurface pressure under a wave is the
.
cosh[27r(z+d)/L]
p = Pgr? Pgz ,
(2-28)
cosh (27Td/L)
since
H /27rx 27rtl
7? = — cos I —
2 \ L T ;
The ratio
cosh[27r(z+d)/L]
K^ = -^-^ . (2-29)
cosh(27rd/L)
2-22
The pressure response factor K, for the pressure at the bottom when
z = - d,
K, =
2
K = — 1
1
.^
cosh (ZTrd/L)
,,.-> ,
'
(2-31)
_ N(p+pgz)
r? - ,
(2-32)
where z is the depth below the SWL of the pressure gage, and N is a
correction factor equal to unity if the linear theory applies. Several
empirical studies have found N to be a function of period, depth, wave
amplitude and other factors. In general, N decreases with decreasing
period, being greater than 1.0 for long-period waves and less than 1.0 for
short-period waves.
FIND :The height of the wave H assuming that linear theory applies and
the average frequency corresponds to the average wave amplitude.
SOLUTION:
1 1
T = — = ^ _^^^ ~ 15 seconds
f (0.0666)
40
0.0347
K 1152
2-23
From Table C-1 o£ Appendix C, entering with d/L^,
d
0.07712
L
hence
40
L = ; = 519 feet ,
(0.07712)
and
cosh I 1= 1.1197
Since n = a = H/2 when the pressure is maximiom (under the wave crest) ,
Therefore,
H = 2 (2.47) ^ 5 feet
Note that the tabulated value of K in Appendix C, Table C-1, could not
be used since the pressure was not measured at the bottom.
2.237 Velocity of a Wave Group The speed with which a group of waves
.
or a wave train travels is generally not identical to the speed with which
individual waves within the group travel. The group speed is termed the
group velocity, C_; the individual wave speed is the phase velocity or
wave celerity given by Equations 2-2 or 2-3. For waves propagating in
deep or transitional water with gravity as the primary restoring force,
the group velocity will be less than the phase velocity. (For those waves
propagated primarily under the influence of surface tension, i.e., capil-
lary waves, the group velocity may exceed the velocity of an individual
wave.
2-24
with slightly different wavelengths and periods, The equation of the
water surface is given by:
H /27rx 27rt\
+ n, = — cos I
' 2 \ L.
v^'n^^Vz ^envelope
-0.2 -0.1 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 I.I 1.2 1.3
X
~2
/J-2J-1
\~r r
\
)~
t
~2 \'~j J
——
/Jj_~Xl \
J
( offer Kinsman, 1965)
2-26
hence the group and phase velocities are equal. Thus in shallow water,
because wave celerity is fully determined by the depth, all component waves
in a wave train will travel at the s^nie speed precluding the alternate
reinforcing and cancelling of components. In deep and transitional water,
wave celerity depends on the wavelength; hence slightly longer waves travel
slightly faster, and produce the small phase differences resulting in wave
groups. These waves are said to be dispersive or propagating in a
dispersive mediioriy i.e. in a medium where their celerity is dependent on
wavelength
2.238 Wave Energy and Power . The total energy of a wave system is the
sum of its kinetic energy and its potential energy. The kinetic energy is
that part of the total energy due to water particle velocities associated
with wave motion. Potential energy is that part of the energy resulting
from part of the fluid mass being above the trough - the wave crest.
According to the Airy theory, if the potential energy is determined
relative to mean water level, and all waves are propagated in the same
direction, potential and kinetic energy components are equal, and the
total wave energy in one wavelength per unit crest width is given by
E
E = - = ^
pgH*
. (2-39)
2-27
energy flux per unit wave crest width transmitted across a plane perpen-
dicular to wave advance is
P = InC = EC .
(2-40)
g
Energy flux F is frequently called wave power and
47rd/L
1 +
sink (47rd/L)
which the energy is being transmitted and the direction of wave advance.
An energy balance for a region through which waves are passing will
reveal, that for steady state, the amount of energy entering the region
will equal the amount leaving the region provided no energy is added or
removed from the system. Therefore, when the waves are moving so that
their crests are parallel to the bottom contours.
Eo
************** EXAMPLE PROBLEM **************
GIVEN A deepwater oscillatory wave with a wavelength of L^ = 512 feet,
:
FIND :
(a) Derive a relationship between the wave height in any depth of water
and the wave height in deep water, assuming that wave energy per
unit crest width is conserved as a wave moves from deep water into
shoaling water.
(b) Calculate the wave height for the given wave when the depth is 10
feet.
(c) Determine the rate at which energy per unit crest width is trans-
ported toward the shoreline and the total energy per unit width
delivered to the shore in 1 hour by the given waves
SOLUTION :
(a) Since the wave crests are parallel to the bottom contours, refraction
does not occur, therefore H^ = H^. (See Section 2.3. WAVE
REFRACTION.)
1-1 2 n C
and since from Equations 2-3 and 2-6
—
C
= tanh
,
/27Td^
47:d/L
1 +
sink (477 d/L)
H 1 1
K. (2-44)
tanh (27rd/L) (47rd/L)
1 +
sinh (4rrd/L)
(b) For the given wave, d/L^ = 10/512 = 0.01953. Either from Table
C-1 or from an evaluation of Equation 2-44 above.
H
7 = 1.233
H
Therefore
(c) The rate at which energy is being transported toward shore is the
wave energy flux.
P = - E^C
o o
= nEC .
2 " " 2 8 2 8
ft-lb
P = 5120 per ft . of wave crest ,
sec
5120
P = 9.31 horsepower per ft. of wave crest .
550
2-30
This represents the expenditure of
5120
ft-lb
sec
X 3600 —
sec
hr
= 18.55 X
, .
10* ft-lbs
,,
,
*************************************
The mean rate of energy transmission associated with waves propagating
into an area of calm water provides a better physical description of the
concept of group velocity. An excellent treatment of this subject is given
by Sverdrup and Munk (1947) and is repeated here.
2-31
After the first complete stroke one wave will be
present and its energy is E/2. One period later this
wave has advanced one wave length but has left one-
half of its energy or E/4 behind. It now occupies a
previously undisturbed area to which it has brought
energy E/4. In the meantime, a second wave has been
generated, occupying the position next to the plunger
where E/4 was left behind by the first wave. The
energy of this second wave equals E/4 + E/2 = 3E/4.
Repeated applications of this reasoning lead to the
results shown in Table 2-1.
Series
number
n
With a large number of waves (a large n), energy decreases with
increasing m, and the leading wave will eventually lose its identity.
At the group center, energy increases and decreases rapidly - to nearly
maximum and to nearly zero. Consequently, an energy front is located
at the center wave group for deepwater conditions. If waves had been
examined for shallow rather than deep water, the energy front would
have been found at the leading edge of the group. For any depth, the
ratio of group to phase velocity (C„/C) generally defines the energy
front. Also, wave energy is transported in the direction of phase propa-
gation, but moves with the group velocity rather than phase velocity.
The engineer must define regions where various wave theories are
valid. Since investigators differ on the limiting conditions for the
several theories, some overlap must be permitted in defining the regions.
Le Mehaute (1969) presented Figure 2-7 to illustrate approximate limits
of validity for several wave theories. Theories discussed here are indi-
cated as are Stokes' third- and fourth-order theories. Dean (1973),
after considering three analytic theories, presents a slightly different
analysis. Dean (1973) and Le Mehaute (1969) agree in recommending cnoidal
theory for shallow-water waves of low steepness, and Stokes' higher order
theories for steep waves in deep water, but differ in regions assigned
to Airy theory. Dean indicates that tabulated stream function theory is
most internally consistent over most of the domain considered. For the
limit of low steepness waves in transitional and deep water, the differ-
ence between stream function theory and Airy theory is small. Additional
wave theories not presented in Figure 2-7 may also be useful in studying
2-33
^'^
0.001
itiil:]ll[ii#i#ll''f^i'l-|!in']-]1^iti^|tiMr!^1-lin^i^^
II I
0.01 0.02 0.04 0.06 0.1 0.2 0.4 0.6 0.8 1.0 2.0 4.0 6.0 8.0 100
d
(ft/sec2) (after Le Mehaute,l969)
2-35
wave phenomena. For given values of H, d and T, Figure 2-7 may be used
as a guide in selecting an appropriate theory. The magnitude of the
Ursell parameter Ud shown in the figure may be used to establish the
boiondaries of regions where a particular wave theory should be used.
The Ursell parameter is defined by
T 2 14
"« = y^ .
(2-45)
2 \L T
I
I
2
where a = H/2, for first- and second-orders, but a < H/2 for orders
higher than the second, and Bg, B3 etc. are specified functions of the
wavelength L, and depth d.
Linear theory considers only the first term on the right side of
Equation 2-46. To consider additional terms represents a higher order of
approximation of the free surface profile. The order of the approxima-
tion is determined by the highest order term of the series considered.
Thus, the ordinate of the free surface to the third order is defined by
the first three terms in Equation 2-46.
When the use of a higher order theory is warranted, wave tables, such
as those prepared by Skjelbreia (1959), and Skjelbreia and Hendrickson
(1962), should be used to reduce the possibility of numerical errors made
in using the equations. Although Stokes (1880) first developed equations
for finite amplitude waves, the equations presented here are those of
Miche (1944).
2-36
2.251 Wave Celerity, Length, and Surface Profile It can be shovm that, .
^ = 2- gT , /27rd'
C tanh (2-3)
2tt \ L
and
L = ^^ tanh
'-^ (2-4)
2n \ L
The above equations, corrected to the third order, are given by:
+ +
C = —
gT
271
,
tanh
/27rd^
\ L
1 +
T
5 2 cosh (47rd/L)
8 sinh" (27rd/L)
2 cosh^ (47rd/L)
,(2-47)
and
H
—
/27rx 27rt
'
cos I
2 \ L
(2-49)
+
cosh (27rd/L)
2 + cosh(47rd/L) cos
u.
(8L sinh^ (27rd/L) t)
For deep water, (d/L > 1/2) Equation 2-49 becomes,
H„ --. /4jrx
T? =
o
COS
J2-nx
I-..- —
2-ni ^ '^"o
+ cos — —
4fft\
(2-50)
2
I
components of the water particle velocities to the second order are given
by
(2-51)
2-37
w =
nU
—L C
sinh [27r(z
r~r-r:
sinh(27rd/L)
—+7r~^ —
d)/L]
sin 1
\
2itx
L
(2-52)
+ 1 i'^\ r
si"h [4n(z + d)/L] . Mttx
4 \l/ sinh'»(27rd/L) ''" \l
Second-order equations for water-particle displacements from their
mean position for a finite amplitude wave are:
?
+
471 cosh (27rd/L) ^"^ \ L T 8L sinh^ (2;rd/L)
(2-53)
and
HgT^ sinh [27r(z + d)/L] 2nx l-nt
t = —2 cos —
^
47rL cosh (27rd/L) \ L T
(2-54)
3 ttH^ sinh [47r(z + d)/L] Mttx
+ ~ n. ; cos
16 L sinh''(27rd/L) \L
2-38
2.254 Subsurface Pressure . The pressure at any distance below the fluid
surface is given by
+
3
8
ttH^
L
tanh(27rd/L)
sinhM27rd/L)
cosh [47r(z
sinh^ (27rd/L)
d)/L]
cos r^ - ^ (2-56)
'H
— 1 = 0.142 « - (2-57)
"olmax '
27rd\ , /27rd\
= 0.142 tanh (2-58)
2-39
choice of a theory for a particular problem. It should be kept in mind
that linear (or first-order) theory applies to a wave which is symmetrical
about Stillwater level and has water particles that move in closed orbits.
On the other hand, Stokes' second-order theory predicts a wave form that
is unsymmetrical about the Stillwater level but still symmetrical about a
vertical line through the crest, and has water particle orbits which are
open.
FIND :
(a) Compare the wave profiles given by the first- and second-order
theories
(b) What is the difference between the first- and second- order
horizontal velocities at the surface under both the crest and
trough?
(c) How far in the direction of wave propagation will a water particle
move from its initial position during one wave period when z = 0?
(d) What is pressure at the bottom under the wave crest as predicted
by both the first- and second-order theories?
(e) What is the wave energy per unit width of crest predicted by the
first-order theory?
SOLUTION :
H
7? = — cos 6 ,
where
(2iTX 27rt
L T
H ttH^ cosh(27rd/L)
V = — cos 6 + — ; 2 + cosh I cos 26
2 8L sinh^(27rd/L)
for
20
-d = = 0.1 ,
L 200
2-40
and from Table C-2
/27rd\
cosh t
1 = 1.2040 ,
sinh I 1 = 0.6705 ,
/47rd\
cosh = 1.8991 ,
Therefore
7?
= 2 COS0 + 0.48 cos 2d ,
at the crest
where n^ 2 ^"'^ ^t 2 ^^^ ^^^ values of n
(i.e. cos 9=1, cos 26 = 1) and trough (i.e. cos 6 = - 1,
cos 26 = 1) according to second-order theory.
2-41
CVJ
in
LJ
According to first-order theory, a crest occurs at z = H/2,
cos 6=1 and a trough at z = - H/2, cos = - 1. Equation 2-13
therefore implies
with
Entering Table C-2 with d/L = 0.10, find tanh (2TTd/L) = 0.5569.
From Equation 2-3 which is true for both first- and second-order
theories.
C^ = —
27r
tanh
(t)
27rd\
=
(32.2) (200) (0.5569)
27r
= 571 (ft/sec)'
or
C = 23.9 ft/sec.
As a consequence.
T 1
= 0.0418 sec/ft
L C
H
^=2 '
27r(z + d)
cosh = cosh[27r(0.11)] = 1.249
and when
H
z = — — '
2
27r(z + d)
cosh = cosh [27r(0.09)] = 1.164
4 1.164
u, , = - (32.2) (0.0418) = — 2.6 ft/sec
*'^
2 1.2040
2-44
Entering Table C-2 again, it is found that when
27r(z + d)
cosh = cosh [27r(0.1124)] = 1.260.
47r(z + d)
cosh = cosh [47r(0.1 124)] = 2.175
When
z = T?^
2 = ~ 1-52 feet
27r(z + d)
cosh cosh [27r(0.0924)] = 1.174
47r(z + d)
cosh cosh [47r(0.0924)] = 1.753
u^2 =
'''''
4
:;
2
(32.2) (0.0418) ——1.260
1.2040
- +
3
- -—
/
4 \200)
47r\^
(23.9)
2.175
-^^
0.202
= 3.6 ft/sec ,
''2
u,y =
4
(32.2) (0.0418) —1.174
1. 2040
+ -—
3 /
4 \200/
47r\2
(23 9)
1.753
0.202
2.0 ft/ sec
(c) To find the horizontal distance that a particle moves during one
wave period at z = 0, Equation 2-55 can be written as:
2-45
For the example problem, when
z = ,
cosh (47rd/L) C
AX(0) >= ^ (t)" sinh2(27rd/L) '
1.
when
27r(z + d)
z = — d, cosh = cosh(O) =1.0
Therefore
p
= (2) (32.2) (4)
——
1
- (2) (32.2) (- 20) = 107 + 1288 = 1395 lbs/ ft^
M J. (^U^
3 '^(4)' (0.5569) 1
- ,o^ .oo ..
(2) (32.2) (1)
8 200 (0.6705)' (0.6705)'
2-46
Thus, second-order theory predicts a pressure,
(e) Using Equation 2-38, the energy in one wavelength per unit width
of crest given by the first-order theory is:
^^„^^ft-lbs
E = ^
pgtfL
=
(2) (32.2) (4)^200)
^^^-^^ i^jL_:^ .'
= 25,800 —r- •
8 8 ft
developed by Korteweg and DeVries (1895) The term anoidal is used since
.
2-4-
significance, it is used to express the relationships between the various
wave parameters. Tabular presentations of the elliptic integrals and
other important functions can be obtained from the above references. The
ordinate of the water surface Xg measured above the bottom is given by
Jt
+ Hcn^ 2K(k)^-| U (2-59a)
where y^ is the distance from the bottom to the wave trough, en is the
elliptic cosine function, K(k) is the complete elliptic integral of the
first kind, and k is the modulus of the elliptic integrals. The argument
of cn^ is frequently denoted simply by ( ), thus. Equation 2-59a above
can be written as
y^ + H cnM )
(2 -59b)
The distance from the bottom to the wave trough. yt' as used in
Equations 2-59a and b, is given by
yj
d d
H
d
—
16d'
3V
— ,
K(k) [K(k)
,
- E(k)] + 1-7Hd (2-60)
where is the distance from the bottom to the crest and E(k) is the
complete elliptic integral of the second kind. Wavelength is given by
'I6d^
L = kK(k) ,
(2-61)
3H
kK(k)
(2-62)
H E(k)\
1 +
y? K(k)
p = Pg (y^
- y)
(2-63)
2-48
that is, the pressure distribution may be assumed to vary linearly from
pgyg at the bed to zero at the surface.
Figures 2-9 and 2-10 show the dimensionless cnoidal wave surface
profiles for various values of the square of the modulus of the elliptic
integrals k^, while Figures 2-11 through 2-15 present dimensionless
plots of the parameters which characterize cnoidal waves. The ordinates
of Figures 2-11 and 2-12 should be read with care, since values of k^
are extremely close to 1.0 (k^ = l-lO"^ = 1-0. 1 = 0.99) It is the
.
FIND :
(a) Using cnoidal wave theory, find the wavelength L and compare this
length with the length determined using Airy theory.
(b) Determine the celerity C. Compare this celerity with the celerity
determined using Airy theory.
(c) Determine the distance above the bottom of the wave crest (y^)
and wave trough (y^)
SOLUTION:
(a) Calculate
H
T
d
= —
2.5
10
= 0.25
and
''
^ "' Jw - ^'''
2-49
c
o
o
3
in
o
o
\—
CL
a>
o
o
>*-
cn
>
o
•o
o
oc
a>
I
CM
a>
k.
3
o
c
o
o
cr
3
o
CO
o
a.
a>
o
o
3
CO
a>
>
o
oo
o
c
2-51
-10" 100
-10 10
-10"
-0.1
l-IO
40 60 100 200 400 600 1,000 2,000 4,000
J /g (oft«r WitstI, I9«0)
V d
2-52
to
^«.
X
^-^
c
o
T3
I
c
Q>
<u
CD
TJ
C
c
ro
T3
X
oc
o
CM
c
w
a>
aJ
m
(/>
Q.
c
o
_o
cm
ro
CM
a>
i_
3
<»
340 1
——
1
320
300
280
260
240
I
220
CM
200
180
160
140
120
» -o
100
80
601-
40
20
to
o
X
'^
Dc
o
c
Qi
a>
00
to
c
o
13
M6A
V (»l)3 I / I H ^'/ =
2-56
From Figure 2-11, entering with H/d and T/g/d', determine the square
of the modulus of the complete elliptical integrals, k^.
F = 1 - 10-^-9-y
or
(c) The percentage of the wave height above the SWL may be determined
from Figure 2-13. Entering the figure with L^H/d^ = 190, the
value of (y^ - d)/H is found to be 0.833 or 83.3 percent.
Therefore
Yc
= 0.833 H + d.
(y, - d)
+ 1 = 0.833
H
thus
y^
= (0.833- 1) (2.5) + 10 = 9.583 ft.
The results for the wave celerity determined under (b) above
can now be checked with the aid of Figure 2-15. Calculate
0.261 .
The difference between this number and the 18.38 ft/sec calculated
under (b) above is the result of small errors in reading the
curves
H
d + H sech^ (x - Ct)
or
3 H_
1? = H sech^ (x - Ct) (2-64)
4 d^
where the origin of x is at the wave crest. The volume of water within
the wave above the still water level per unit crest width is
16 %
V = d' H (2-65)
2-59
An equal amount of water per unit crest length is transported
forward past a vertical plane that is perpendicular to the direction of
wave advance. Several relations have been presented to determine the
celerity of a solitary wave; these equations differ depending on the
degree of approximation. Laboratory measurements by Daily and Stephan
(1953) indicate that the simple expression
C = Vg (H + d)" ,
(2-66)
where M and N are the functions of H/d shown on Figure 2-16, and y
is measured from the bottom. The expression for horizontal velocity u,
is often used to predict wave forces on marine structures sited in shallow
water. The maximum velocity u^nox' occurs when x and t are both equal
to zero; hence,
CN
~^^'
"'"'« " 1 + cos(My/d)
E = ^^pgH3/2d3/2 , (2-70)
and the pressure beneath a solitary wave depends upon the local fluid
velocity as does the pressure under a cnoidal wave; however, it may be
approximated by
P = Pg(y,-y). (2-71)
2-60
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
LI
2-61
when the water particle velocity at the wave crest becomes equal to the
wave celerity. This occurs when
(2-72)
2.31 INTRODUCTION
Equation 2-2 shows that wave celerity depends on the depth of water
in which the wave propagates. If the wave celerity decreases with depth,
wavelength must decrease proportionally. Variation in wave velocity occurs
2-62
along the crest of a wave moving at an angle to underwater contours because
that part of the wave in deeper water is moving faster than the part in
shallower water. This variation causes the wave crest to bend toward align-
ment with the contours, (See Figure 2-17.) This bending effect, called
refraction, depends on the relation of water depth to wavelength. It is
analogous to refraction for other types of waves such as, light and sound.
The decrease in wave celerity with decreasing water depth can be con-
sidered an analog to the decrease in the speed of light with an increase in
2-63
Figure 2-17. Wave Refraction at Westhampton Beach, Long Island, New York
2-64
the refractive index of the transmitting medium. Using this analogy,
O'Brien (1942) suggested the use of Snell's law of geometrical optics for
solving the problem of water-wave refraction by changes in depth. The
validity of this approach has been verified experimentally by Chien (1954),
Ralls (1956), and Wiegel and Arnold (1957). Chao (1970) showed analyti-
cally that Format 's principle and hence Snell's law followed from the
governing hydrodynamic equations, and was a valid approximation when applied
to the refraction problem. Generally, two basic techniques of refraction
analysis are available - graphical and numerical. Several graphical pro-
cedures are available, but fundamentally all methods of refraction analyses
are based on Snell's law.
" 2-n
In this equation, the velocity C^, does not depend on depth; therefore
in those regions deeper than one-half the wavelength (deep water) , refrac-
tion by bathymetry will not be significant. Where the water depth is
between 1/2 and 1/25 the wavelength (transitional water) , and in the region
where the water depth is less than 1/25 the wavelength (shallow water),
refraction effects may be significant. In transitional water, wave velocity
2-65
must be computed from Equation 2-2; in shallow water, tanh(2Trd/L) becomes
nearly equal to 2TTd/L and Equation 2-2 reduces to Equation 2-9.
^2 _ gd or C = (gd)*^ . (2-9)
Both Equations 2-2 and 2-9 show the dependence of wave velocity on depth.
To a first approximation, the total energy in a wave per unit crest width
may be written as
(2-38)
8
It has been noted that not all of the wave energy E is transmitted
forward with the wave; only one-half is transmitted forward in deep water.
The amount of energy transmitted forward for a given wave remains nearly
constant as the wave moves from deep water to the breaker line if energy
dissipation due to bottom friction (K^ = 1.0), percolation and reflected
wave energy is negligible.
^o = \KK^o' (2-73)
P = nbEC, (2-74)
\/b„\/C„\
(2-75)
(2-76)
2-66
and combining Equations 2-75 and 2-76,
H_
(2-77)
Ho
even increments of d/L^, and Table C-2 for even increments of d/L).
where
From this equation, a template may be constructed which will show the
angular change in a that occurs as an orthogonal passes over a
particular contour interval, and construct changed-direction orthogonal.
Such a template is shown in Figure 2-18. In application to wave refrac-
tion problems, it is simplest to construct this template on a transparent
material.
2-67
<J>a)^-^r)^f^vK>cJ — o
o o oo o
Refraction may be treated analytically at a straight shoreline with
parallel offshore contours, by using Snell's law directly:
is the angle between the deepwater wave crest and the shoreline.
For example, if cxq = 30° and the period and depth of the wave
are such that C/C^ = 0.5, then
cos a = 0.968
and
cosa^ = 0.866
cosaA'^ / 0.866 \^
= 0.945
cos a / \ 0.968
Figure 2-19 shows the relationships between a, a^, period, depth, and
K^ in graphical form.
Charts showing the bottom topography of the study area are obtained. Two
or more charts of differing scale may be required, but the procedures are
identical for charts of any scale. Underwater contours are drawn on the
chart, or on a tracing paper overlay, for various depth intervals. The
depth intervals chosen depend on the degree of accuracy desired. If
overlays are used, the shoreline should be traced for reference. In
tracing contours, small irregularities must be smoothed out, since bottom
features that are comparatively small in respect to the wavelength do not
affect the wave appreciably.
2-69
(O
and suitably spaced orthogonals are drawn perpendicular to this wave front
and parallel to the chosen direction of wave approach. Closely spaced
orthogonals give more detailed results than widely spaced orthogonals.
These lines are extended to the first depth contour shallower than L^/2
where L^ (in feet) = 5.12 T^.
T = 10 seconds
Templote Orthogonal Line
Contour
"-X
20'"^
'"'^Wc.-r-l- /~
bO \^
Incoming Orttiogonol
Turning Point-
20 ^^
II- 1071
Tangent to Mid-
Contour -^
^o \<^ 1.0461
Incoming Orthogonal
NOTE: The template has been turned obout R until the volue '/r ; 1.045
intersects the tangent to the mid-contour. The templote "orthogonal line
off ot some point " B"on the incoming orthogonol which is equidistant
from the two contours along the incoming and outgoing orthogonals.
2-72
(c) Rotate the template about the turning point until the C2/C2 value
corresponding to the contour interval being crossed intersects the tangent
to the midcontour. The orthogonal line on the chart now lies in the direc-
tion of the turned orthogonal on the template (Figure 2-20 bottom)
(d) Place a triangle along the base of the template and construct
a perpendicular to it so that the intersection of the perpendicular with
the incoming orthogonal is midway between the two contours when the dis-
tances are measured along the incoming orthogonal and the perpendicular
(See Point B in Figure 2-20 bottom) . Note that this point is not neces-
sarily on the midcontour line. This line represents the turned orthogonal;
2-73
J = Distance between contours at turning points,*
R = Distance along orttfogonal
T = 12 seconds
Lo= 737 ft.
Acc = 20-25
2-74
toward deep water. In such cases, a sheaf or fan of orthogonals may be
projected seaward in directions some 5 or 10 degrees apart. See Figure
2-22a. With the deepwater directions thus determined by the individual
orthogonals, companion orthogonals may be projected shoreward on either
side of the seaward projected ones to determine the refraction coefficient
for the various directions of wave approach. (See Figure 2 -22b.)
2.326 Computer Methods for Refraction Analysis . Harrison and Wilson (1964)
developed a method for the numerical calculation of wave refraction by use
of an electronic computer. Wilson (1966) extended the method so that, in
addition to the numerical calculation, the actual plotting of refraction
diagrams is accomplished automatically by use of a computer. Numerical
methods are a practical means of developing wave refraction diagrams when
an extensive refraction study of an area is required, and when they can
be relied upon to give accurate results. However, the interpretation of
computer output requires care, and the limitations of the particular scheme
used should be considered in the evaluation of the results. For a dis-
cussion of some of these limitations, see Coudert and Raichlen (1970).
For additional references, the reader is referred to the works of Keller
(1958), Mehr (1962), Griswold (1963), Wilson (1966), Lewis, et al., (1967),
Dobson (1967), Hardy (1968), Chao (1970), and Keulegan and Harrison (1970),
in which a number of available computer programs for calculation of refrac-
tion diagrams are presented. Most of these programs are based on an algo-
rithm derived by Munk and Arthur (1951) and, as such, are fundamentally
based on the geometrical optics approximation. (Fermat's Principle.)
2-75
u
C
o
H
P
o
ni
^(
a>
ai
a>
P-
X
I
C
uu
O
O
t/)
(N
I
(Nl
2-76
Beach Line
'Breakers
»*JL-'X«'L^xxx,xxxr;x)i»^*»* _x1
Contours Contours
Orthogonals Orthogonals
(a) (b)
Figure 2-24. Refraction by a Submarine Ridge (a) and Submarine
Canyon (b)
2-77
diverge, resulting in low heights at the shore. (b^i/b less than 1.0.)
Similarly, heights will be greater at a headland than in a bay. Since
the wave energy contained between two orthogonals is constant, a larger
part of the total energy expended at the shore is focused on projections
from the shoreline; consequently, refraction straightens an irregular
coast. Bottom topography can be inferred from refraction patterns on
aerial photography. The pattern in Figure 2-17 indicates the presence
of a submarine ridge.
2-78
Pierson (1951), Longuet-Higgins (1957), and Kinsman (1965), have sug-
gested a solution to the ocean-wave refraction problem. The sea surface
waves in deep water become a number of component monochromatic waves, each
with a distinct frequency and direction of propagation. The energy spec-
trum for each component may then be found and the conventional refraction
analysis techniques applied. Near the shore, the wave energy propagated
in a particular direction is approximated as the linear sum of the spectra
of wave components of all frequencies refracted in the given direction from
all of the deepwater directional components.
The work required for this analysis, even for a small number of indi-
vidual components, is laborious and time consuming. More recent research
by Borgman (1969) and Fan and Borgman (1970), has used the idea of direc-
tional spectra which may provide a technique for solving complex refraction
problems more rapidly.
2.41 INTRODUCTION
This process is also similar to that for other types of waves, such
as light or sound waves.
2-79
Figure 2-26a. Wave Incident on a Breakwater
No Diffraction
2-80
the shore is affected by diffraction caused by naturally occurring changes
in hydrography. An aerial photograph illustrating the diffraction of
waves by a breakwater is shown in Figure 2-27.
(1962), diffraction diagrams have been prepared which, for a uniform depth
adjacent to an impervious structure, show lines of equal wave height re-
duction. These diagrams are shown in Figures 2-28 through 2-39; the graph
coordinates are in units of wavelength. Wave height reduction is given in
terms of a diffraction coefficient K' which is defined as the ratio of a
wave height H, in the area affected by diffraction to the incident wave
height H^, in the area unaffected by diffraction. Thus, H and H^ are
determined by H = K'H^.
The diffraction diagrams shown in Figures 2-28 through 2-39 are con-
structed in polar coordinate form with arcs and rays centered at the struc-
ture's tip. The arcs are spaced one radius -wavelength unit apart and rays
15 apart In application, a given diagram must be scaled up or down so
.
2-81
Figure 2-21 . Wave Diffraction at Channel Islands Harbor Breakwater,
California
2-82
r——
C
<
>
o
o
00
o06 O
<4-i
•H
Q
u
3
•H
2-84
2-85
2-86
C
<
0)
>
C
o
o
U
4-1
0)
>
to
I
CM
0)
3
bO
2-87
o
iT)
O
00
<
ID
>
03
o
O
6
U
c
o
•H
o06 p
o
(D
o
o
o
o.
Q.
< I
> U
o
•H
o
c
o
«3 O
2-88
t—<
c
<
>
o
o
00
c
o
4-1
•H
Q
>
to
I
<N
D
U
3
•H
2-89
o
o
(N
C
o
• H
Q
CM
V
3
00
•H
2-90
C
<
>
o
LO
to
OS
o
o06 H
+J
O
nj
U
1-1
4-1
•H
Q
>
to
3
00
2-91
C
<
>
B
U
o
• H
+J
O
nJ
u
4-1
4^
>
00
^0
3
H00
2-93
o
diagram overlay template to correspond to the hydrographic chart being
used. In constructing this overlay, first determine how long each of its
radius-wavelength units must be. As noted previously, one radius -wave length
unit on the overlay must be identical to one wavelength on the hydrographic
chart. The next step is to construct and sketch all overlay rays and arcs
on clear plastic or translucent paper. This allows penciling in of the
scaled lines of equal K for each angle of wave approach that may be
considered pertinent to the problem. Thus, after studying the wave field
for one angle of wave approach, K lines may be erased for a subsequent
analysis of a different angle of wave approach.
The diffraction diagrams in Figures 2-28 through 2-39 show the break-
water extending to the right as seen looking toward the area of wave dif-
fraction; however, for some problems the structure may extend to the left.
All diffraction diagrams presented may be reversed by simply turning the
transparency over to the opposite side.
Template Overlay
Breakwater
O.I. I. '././. 1. 1. 1. '.I. '.I. J. '7^
I.
f'/'/'/'r /=z= -•-X
Wave Crests
2-95
FIND: The wave height at a point P having coordinates in units of wave-
length of X = 3 and y = 4. (Polar coordinates of x and y are r=5 at 53°.)
SOLUTION :
d.
180'
Wave Crests
Direction of Wave Approach
2-97
2.422 Waves Passing a Gap of Width Less than Five Wavelengths at Normal
Incidence The solution of this problem is more complex than that for a
.
2-98
u
o
*->
cd
s
u
—
/
CIS <N
U II
B OQ
^-^
nj
OO c/l
•r-l 4->
Q 00
c
t-> >
4-1 O
4-1
T3 O
N
•H
^
4->
(L>
(U nJ
L3 O
u
3
2-99
y/L
..0 2 4- 6 8 10 (2 14 16 18 20
GAP
B'UIL K =1.145 K:|.0 K'=0.8 K'=0.6 K'eO.4
" DIRECTION OF ( Johnson, 1952 )
INCIDENT WAVE
Figure 2-45. Contours of Equal Diffraction Coefficient
Gap Width =1.41 Wave Lengths (B/L = 1.41)
K'«0.4
( Johnson, 1952)
2-101
PI DIRECTION OF , ( Johnson, 1952)
', INCIDENT WOVF
(Johnson, 1952)
2-102
GAP
B=2.50L K =0.6
K'=l.2 K'= 1.0 K'=0.8
I
I DIRECTION OF ,
* Johnson, 1952)
ra INCIDENT WAVE (
( Johnson, 1952)
2-103
(Johnson, 1952)
DIRECTION OF
^
INCIDENT WAVE
GAP K'sl.O K'sl.O K'« 1.282 K't|.2
B:SL
( Johnson, 1952)
Figure 2-52. Contours of Equal Diffraction Coefficient
Gap Width = 5 Wave Lengths (B/L = 5)
2-104
Template Overlays
Wave Crests
Figure 2-53. Diffraction for a Breakwater Gap of Width > 5L (B/L > 5)
YZZZZZZZZZZZZZ.
Breakwater
Wave Crests
( Johnson, 1952)
2-105
= O»o
B = IL
V,
10
0=30
B=IL
10
12 4 6 8 10 12 14 16 le 20
y/L
-.
2
07 75"
I
B=IL '
10
12
( Johnson, 1952)
2-109
diagram shoreward to the breakwater; (b) at this point, constructing a
diffraction diagram carrying successive crests three or four wavelengths
shoreward, if possible; and (c) with the wave crest and wave direction
indicated by the last shoreward wave crest determined from the diffraction
diagram, constructing a new refraction diagram to the breaker line. The
work of Mobarek (1962) on the effect of bottom slope on wave diffraction
indicates that the method presented here is suitable for medium-period
waves. For long-period waves the effect of shoaling (Section 2.32) should
be considered. For the condition when the bottom contours are parallel to
the wave crests, the sloping bottom probably has little effect on diffrac-
tion. A typical refraction-diffraction diagram and the method for deter-
mining combined refraction-diffraction coefficients are shown in Figure
2-59. When a wave crest is not of uniform height, as when a wave is under-
going refraction, a lateral flow of energy - wave diffraction - will occur
along the wave crest. Therefore diffraction can occur without the wave
moving past a structure although the diffraction effects are visually more
dramatic at the structure.
2.51 GENERAL
2-110
b| Lines of Equal Diffraction Coefficient (k')
Breakwater
X\\\\\\\\\\V^^^\\\\S\^V\^^^^^^^
Wave Crests
2-III
December 1952
Figure 2-60. Wave Reflection at Hamlin Beach, New York
2-JI2
and the transmission coefficient will depend upon the geometry and compo-
sition of a structure and the incident wave characteristics such as wave
steepness and relative depth d/L, at the structure site.
(2-10)
which reduces to
27rx 2-nt
n = H. cos cos -— (2-79)
'
L T
Equation 2-79 represents the water surface for a standing wave or olapotis
which is periodic in time and in x having a maximum height of 2H^ when
both cos(2ttx/L) and cos(2TTt/T) equal 1. The water surface profile as a
function of 2ttx/L for several values of 27rt/T are shown in Figure 2-61.
There are some points (nodes) on the profile where the water surface
2-113
2-114
remains at the SWL for all values of t and other points (antinodes) where
the water particle excursion at the surface, is 2 H^ or twice the incident
wave height. The equations describing the water particle motion show that
the velocity is always horizontal under the nodes and always vertical under
the antinodes. At intermediate points, the water particles move along
diagonal lines as shown in Figure 2-61. Since water motion at the anti-
nodes is purely vertical, the presence of a vertical wall at any antinode
will not change the flow pattern described since there is no flow across
the vertical barrier and equivalently, there is no flow across a vertical
line passing through an antinode. (For the linear theory discussion here,
the water contained between any two antinodes will remain between those
two antinodes.) Consequently, the flow described here is valid for a
barrier at 2ttx/L = (x = 0) since there is an antinode at that location.
gT^
L = ^- tanh
2jr
,
—
/27rd\
\ L
/
,
47^4
T = (2-80)
gtanh (jrd/^)
The next possible resonant mode occurs when there is one complete wave in
the basin (barriers at x = and x = 2ir) and the next mode when there are
3/2 waves in the basin (barriers at x = and x = 37t/2, etc. In general,
Jlg = jL/2, where j = 1, 2, In reality, the length of a natural or
manmade basin 2.^, is fixed and the wavelength of the resonant wave con-
tained in the basin will be the variable; hence,
L = —
in
j
= 1,2,----, (2-81)
2-115
substituting Equation 2-81 into the expression for the wavelength; there-
fore,
47.^3 V2
= , j = 1,2,---- (2-82)
^J jg tanh (7rjd/4)
24
T.
]
= -r
j
— 1
r-w
(gd)'^
j = 1,2, (small values) . (2-83)
44' 1
'^i
=
O^ (^^ j
= 1, 2, • • • • (small values) .
(2-84)
Note that higher modes occur when there are 3, 5, ...., 2^ - 1, etc.,
quarters of a wave within the basin.
2-116
************** EXAMPLE PROBLEM **************
GIVEN: Lake Erie has a mean depth of d = 61 feet and its length % is
220" miles or 116,160 feet.
4
211,
1
T 1
2(116,160)
T. =
» 1 [32.2(61)]^^
*************************************
Note: Additional discussion of seiching is presented in Section 3.84.
The amount of wave energy reflected from a beach depends upon the
roughness, permeability and slope of the beach in addition to the steep-
ness and angle of approach of incident waves. Miche (1951) assumed that
the reflection coefficient for a beach x> could be described as the
product of two factors by the expression,
X = X,X2 , (2-85)
where x-, depends on the roughness and permeability of the beach and is
independent of the slope, while Xz depends on the beach slope and the
wave steepness.
2-117
to 0.6 has been recommended for rough slopes and step-faced structures
The second factor X2 is given by
01 o (2-86a)
01 o
(2-86b)
'\L'o \ o /max
I
H,
(2-87)
in which 3 is the angle the beach makes with the horizontal (tan B =
beach slope) and Hq/^q is the incident wave steepness in deep water.
,
FIND :
(h) What will be the steepest incident wave almost totally reflected
from the given revetment?
SOLUTION. Calculate:
tanj3 =
(ofttr Micht, 1951)
cotangent ^ -
tangent )S
max 0.5 -
X,-
2
/.«»/,n«onf
3
-
4
I
56789(ofUr Micht,
10
1951)
cotangent p
fl - -tt
tongent p
Figure 2-63. Xg Versus Beach Slope for Various Values of Hq/Lq
2-119
Entering Figure 2-63 with cot B = 5, and using the curve for Ho/ho -
0.01, a value o£ "
X2 ~ 0-41 is found. Assuming that since the beach
is impermeable, Xi = 0.8 and
X = X, X2 = 0.8(0.41) = 0.33 .
The steepest incident wave which will be nearly perfectly reflected from
the given revetment is, from Figure 2-62,
0.005 .
I— = 0.000014.
\ ol max
Hence
X = X, Xj = 0.8(0.0014) = 0.0011,
or the height of the reflected wave is about 0.1 percent of the incident
wave height
*************************************
2.6 BREAKING WAVES
Ho
— = 0.142 =*
1
- . (2-88)
which occurs when the crest angle as shown in Figure 2-64 is 120 This .
limiting steepness occurs when the water particle velocity at the wave
crest just equals the wave celerity; a further increase in steepness would
2-120
result in particle velocities at the wave crest greater than the wave
celerity and, consequently, instability.
When a wave moves into shoaling water, the limiting steepness which
it can attain decreases, being a function of both the relative depth d/L,
and the beach slope m, perpendicular to the direction of wave advance.
A wave of given deepwater characteristics will move toward a shore until
the water becomes shallow enough to initiate breaking, this depth is
usually denoted as d]-, and termed the breaking depth. Munk (1949) derived
several relationships from a modified solitary wave theory relating the
breaker height H^, the breaking depth d^,, the unrefracted deepwater
wave height H^, and the deepwater wavelength L^. His expressions are
given by
1
(2-89)
3.3(HyL^)H
and
= 1.28, (2-90)
H.
The ratio H^/H^ is frequently termed the breaker height index. Subse-
quent observations and investigations by Iversen (1952, 1953), Galvin
(1969), and Coda (1970) among others, have established that H2,/H^ and
dy/U-jj depend on beach slope and on incident wave steepness. Figure 2-65
shows Coda's empirically derived relationships between H^j/H^ and H^/L^
for several beach slopes. Curves shown on the figure are fitted to widely
scattered data; however they illustrate a dependence of H^j/H^^ on the
beach slope. Empirical relationships between d^/H^, and H^^/gT^ for
various beach slopes are presented in Figure 2-66. It is recommended
that Figures 2-65 and 2-66 be used, rather than Equations 2-89 and 2-90,
for making estimates of the depth at breaking or the maximum breaker
2-121
2-122
o
(VJ
height in a given depth since the figures take into consideration the
observed dependence of d^^/H^, and H^j/H^ on beach slope. The curves
in Figure 2-66 are given by
^
_± ^ '
(2-91)
Hfc b-(aHj,/gT^)
where a and b are functions of the beach slope m, and may be approxi-
mated by
a = 1.36g(l-e-i5'") C2-92)
1-56 (2-93)
b =
(l + e-19.5m)
FIND : The breaker height Hh and the depth db at which breaking occurs,
h' b \''''
— = Kp =
/
2-124
Figure l-dl . Spilling Breaking Wave
2-125
Figure 2-69. Surging Breaking Wave
2-126
hence,
% 5.25
= 0.010
Lo 5.12(10)^
From Figure 2-65, entering with H^/L^ = 0.010 and intersecting the curve
for a slope of 1:20 (m = 0.05) results in H^/H^ = 1.65. Therefore
^b 8.66
0.00269 ,
gT^ 32.2(10)2
^b
= 0.90
Thus djy = 0.90 (8.66) =7.80 feet, and therefore the wave will break
when it is approximately 7.80/(0.05) = 156 feet from the shoreline,
assuming a constant nearshore slope. The initial value selected for
the refraction coefficient should now be checked to determine if it is
correct for the actual breaker location as found in the. solution. If
necessary, a corrected value for the refraction coefficient should be
used and the breaker height recomputed. The example wave will result
in a plunging breaker. (See Figure 2-65.)
2-127
REFERENCES AND SELECTED BIBLIOGRAPHY
ARTHUR, R.S., MUNK, W.H., and ISSACS, J.D., "The direct Construction of
Wave Rays," Transaotions of the American Geophysical Union^ Vol. 33,
No, 6, 1952, pp. 855-865.
BLUE, F.L., JR. and JOHNSON, J.W., "Diffraction of Water Waves Passing
through a Breakwater Gap," Transaotions of the American Geophysical
Union, Vol. 30, No. 5, Oct. 1949, pp. 705-18.
BOUSSINESQ, J., "Essai sur la Theorie des Eaux Courantes," Mem. divers
Savants a L'Academie des Science, 32:56, 1877.
COUDERT, J.F., and RAICHLEN, F., "Wave Refraction Near San Pedro Bay,
California," Journal of the Waterways, Harbors and Coastal Engineering
Division, ASCE, WW3, 1970, pp. 737-747.
2-129
DAILY, J.W. and STEPHAN, S.C.,JR., "Characteristics of a Solitary Wave,"
Transactions of the Amerioan Society of Civil Engineers ^ Vol. 118, 1953,
pp. 575-587.
DANEL, P., "On the Limiting Clapotis," Gravity Waves^ Circ. No. 521,
National Bureau of Standards, Washington, D.C. ,1952, pp. 35-38.
EAGLESON, P.S. and DEAN, R.G., "Small Amplitude Wave Theory," Estuary and
Coastline Hydrodynamics , A.T. Ippen, ed., McGraw-Hill, New York, 1966,
pp. 1-92.
ESTEVA, D. and HARRIS, D.L., "Comparison of Pressure and Staff Wave Gages
Records," Proceedings of the l2th Conference on Coastal Engineering
Washington, D.C, 1971.
2-130
FAN, S.S., "Diffraction of Wind Waves," HEL-1-10, Hydraulics Laboratory,
University of California, Berkeley, Aug. 1968.
GALVIN, C.J., JR., "Breaker Travel and Choice of Design Wave Height,"
Journal of the Waterways and Harobrs Division^ ASCE, WW2, No. 6569,
1969, pp. 175-200.
GRACE, R.A., "How to Measure Waves," Ocean Industry, Vol. 5, No. 2, 1970,
pp. 65-69.
2-131
HOM-MA, M., HORIKAWA, K. , and KOMORI, S., "Response Characteristics of
Underwater Wave Gauge," Proceedings of the lOth Conference on Coastal
Engineering J Vol. 1, Tokyo, 1966, pp. 99-114.
HUNT, "Winds, Wind Setup and Seiches on Lake Erie," U.S. Army
I. A.,
Engineer District, Lake Survey, 1959.
IPPEN, A.T., "Waves and Tides in Coastal Processes," Journal of the Boston
Society of Civil Engineers^ Vol. 53, No. 2, 1966a., pp. 158-181.
IPPEN, A. and KULIN, G., "The Shoaling and Breaking of the Solitary Wave,"
Proceedings of the 6th Conference on Coastal Engineering ^ Grenoble,
France, 1954, pp. 27-49.
IVERSEN, H.W., "Laboratory Study of Breakers," Gravity WaveSj Circ. No. 521,
National Bureau of Standards, Washington, D.C., 1952.
KAPLAN, K., "Effective Height of Seawalls," Vol. 6, Bui. No. 2, U.S. Army,
Corps of Engineers, Beach Erosion Board, Washington, D.C., 1952.
2-132
KEULEGAN, G.H. and HARRISON, J., "Tsunami Refraction Diagrams by Digital
Computer," Journal of the Waterways, Harbors and Coastal Engineering
Division, ASCE, Vol. 96, WW2, No. 7261, May 1970, pp. 219-233.
KINSMAN, B., Wind Waves, their Generation and Propagation on the Ocean
Surface, Prentice Hall, New Jersey, 1965.
KORTEWEG, D.J. and DE VRIES, G., "On the Change of Form of Long Waves
Advancing in a Rectangular Canal, and on a New Type of Long Stationary
Waves," Philosophical Magazine, 5th Series, 1895, pp. 422-443.
LEE, C, "On the Design of Small Craft Harbors," Proceedings of the 9th
Conference on Coastal Engineering, Lisbon, Portugal, Jun 1964.
LEWIS, R.M., BLEISTEIN, N., and LUDWIG, D., "Uniform Assymptolic Theory of
Creeping Waves," Communications on Pure and Applied Mathematics, Vol. 20,
No. 2, 1967.
2-133
LONGUET-HIGGINS, M.S., "Mass Transport in the Boundary Layer at a Free
Oscillating Surface," Vol. 8, National Institute of Oceanography,
Wormley, Surrey, England, 1960.
McCOWAN, J., "On the Solitary Wave," Philosophical Magazine^ 5th Series,
Vol. 32, No. 194, 1891, pp. 45-58.
lation by the Beach Erosion Board, Vol. 7, Bui. No. 2, Apr. 1953.)
MORISON, J.R. and CROOKE, R.C., "The Mechanics of Deep Water, Shallow
Water, and Breaking Waves," TM-40, U.S. Army, Corps of Engineers,
Beach Erosion Board, Washington, D.C., 1953.
2-134
MUNK, W.H., "The Solitary Wave Theory and Its Application to Surf Problems,"
Annals of the New York Aoademy of Soienoes, Vol. 51, 1949, pp. 376-462.
MUNK, W.H. and ARTHUR, R.S., "Wave Intensity Along a Refracted Ray,"
Symposium on Gravity WaveSy Circ. 521, National Bureau of Standards,
Washington, D.C., 1951.
PALMER, R.Q., "Wave Refraction Plotter," Vol. 11, Bui. No. 1, U.S. Army,
Corps of Engineers, Beach Erosion Board, Washington, D.C., 1957.
PATRICK, D.A. and WIEGEL, R.L., "Amphibian Tractors in the Surf," Frooeed-
ings of the First Conference on Ships and Waves , Council on Wave
Research and Society of Naval Architects and Marine Engineers, 1955.
PIERSON, W.J., JR., "The Accuracy of Present Wave Forecasting Methods with
Reference to Problems in Beach Erosion on the New Jersey and Long
Island Coasts," TM-24, U.S. Army, Corps of Engineers, Beach Erosion
Board, Washington, D.C., 1951.
PIERSON, W.J., JR., "Observing and Forecasting Ocean Waves," H.O. No. 603,
U.S. Naval Oceanographic Office, Washington, D.C., 1965.
PLATZMAN, G.W., and RAO, D.B., "The Free Oscillations of Lake Erie," TR-8,
Department of Geophysical Sciences, University of Chicago, for the
U.S. Weather Bureau, Sep. 1963.
2-135
POCINKI, L.S., "The Application of Conformal Transformations to Ocean Wave
Refraction Problems," Department of Meteorology and Oceanography, New
York University, New York, 1950.
RALLS, G.C., "A Ripple Tank Study of Wave Refraction," Journal of the
Waterways and Harbors Division^ ASCE, Vol. 82, WWl, No. 911, Mar. 1956.
SAVILLE, T., JR., and KAPLAN, K., "A New Method for the Graphical
Construction of Wave Refraction Diagrams," Vol. 6, Bui. No. 3, U.S.
Army, Corps of Engineers, Beach Erosion Board, Washington, D.C., 1952,
pp. 23-24.
SILVESTER, R., "Design Wave for Littoral Drift Model," Journal of the
Waterways and Harbors Division, ASCE, Vol. 89, WW3, No. 360, Aug. 1963.
SKJELBREIA, L. and HENDRICKSON, J. A., Fifth Order Gravity Wave Theory and
Tables of Functions, National Engineering Science Company, 1962.
2-136
STOKES, G.C., "On the Theory of Oscillatory Waves," Mathematical and
Physical Papers^ Vol. 1, Cambridge University Press, Cambridge, 1880.
SVERDRUP. H.U. and MUNK, W.H., "Wind, Sea, and Swell: Theory of Relations
for Forecasting," TR-1, H.O. No. 601, U.S. Naval Hydrographic Office,
Washington, D.C., 1947, p. 7.
WIEGEL, R.L. and ARNOLD, A.L., "Model Studies of Wave Refraction," TM-103,
U.S. Army, Corps of Engineers, Beach Erosion Board, Washington, D.C.
Dec. 1957.
WILLIAMS, L.C., "CERC Wave Gages," Tll-30, U.S. Army, Corps of Engineers,
Coastal Engineering Research Center, Washington, D.C, 1969.
WILSON, W.S., "A Method for Calculating and Plotting Surface Wave Rays,"
TM-17, U.S. Army, Corps of Engineers, Coastal Engineering Research
Center, Washington, D.C, Feb. 1966.
WILTON, J.R., "On Deep Water Waves," Philosophical Magazine, 6th Series,
1914, pp. 385-394.
WORTHINGTON, H.W. and HERBICH, J.B., "A Computer Program to Estimate the
Combined Effects of Refraction and Diffraction of Water Waves,"
Sea Grant Publication, No. 219, Texas A§M University, Aug. 1970.
2-137
CHAPTER 3
WAVE AND
WATER LEVEL
PREDICTIONS
OCEAN CITY, NEW JERSEY - 9 March 1962
CHAPTER 3
3.1 INTRODUCTION
3-
The problem of obtaining wind information for wave hindcasting is
discussed, and specific instructions for estimating wind parameters are
given.
3-2
The significant wave period obtained by visual observations of waves
is likely to be the average period of 10 to 15 successive prominent waves.
When determined from gage records, the significant period is apt to be the
average period of the subjectively estimated most prominent waves, or the
average period of all waves whose troughs are below and whose crests are
above the mean water level, (zero up crossing method).
When the heights of individual waves on a wave record are ranked from
the highest to lowest, the frequency of occurrence of waves above any given
value is given to a close approximation by the cumulative form of the
Rayleigh distribution. This fact can be used to estimate the average
height of the one-third highest waves from measurements of a few of the
highest waves, or to estimate the height of a wave of any arbitrary
frequency from a knowledge of the significant wave height. According to
the Rayleigh distribution function, the probability that the wave height
H is more than some arbitrary value of H referred to as H is given
by
'""'
P(H > H) = e ^ (3-1)
where Hmis is a parameter^of the distribution, and P(H > H) is the number
n of waves larger than H divided by the total number N of waves in
the record. Thus P has the form n/N. The value Hrms is called the
root-mean- square height and is defined by
It was shown in Section 2.238, Wave Energy and Power, that the total energy
per unit surface area is given by
E = '-fil
The average energy per unit surface area for a number of waves is given by
Pg 1 N
3-5
By making the substitutions
^^'""'
f[(H-AH) < H < (fi+ AH)] = l^—\ He . (3-6)
The height of the wave with any given probability n/N of being exceeded
may be determined approximately from curve a in Figure 3-5 or from the
equation.
(3-7)
3-6
0.000
0.0005
0.001
0.005
~ 0.0
O
o
a>
>
0.05
From fifteen 15-minute records
3
E containing a total of 2,342 waves
S (3,007 waves calculated)
1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6 2.8 3.0 3.2
1.42
Scaled Heigtit (H/Hrms)
Figure 3-3. Theoretical and Observed Wave-Height Distributions.
Observed distributions for 15 individual IS-minute
observations from several Atlantic coast wave gages
are superimposed on the Rayleigh distribution curve
3-7
0.000
0.0005
0.001
Q. 0.005
- 0.0
o
.a
o
a>
>
0.05
3 From seventy-two 15-minute samples
E
containing a total of 11,678 waves
(15,364 calculated waves)
0.5
Hg Theory
J__L
1.0 1.2 1.4 I 2.0 2.2 2.4 2.6 2.8 3.0 3.2
A
1.42
Scaled Heigtit (H/Hrms)
3-8
The average height of all waves with heights greater than H (H) can be
obtained from the equation.
"^'"^
/ h' e
'
' d H
H
H(H) = (3-8)
/ H eVHrmJ dH
H
In the analysis system used by CERC from 1960 to 1970/ and whenever
digital recordings cannot be used, the average period of a few of the best
formed waves is selected as the significant wave period. An estimated
number of equivalent waves in the record is obtained by dividing the
duration of the record by this significant period. The highest waves
are then ranked in order with the highest wave ranked 1. The height of
the wave ranked nearest 0.135 times the total number of waves is taken as
the significant wave height. The derivation of this technique is based on
the assumption that the Rayleigh distribution law is exact. Harris (1970)
showed that this procedure agrees closely with values obtained by more
rigorous procedures which require the use of a computer. These procedures
are described in Section 3.23, Energy Spectra of Waves.
FIND :
3-10
SOLUTION : 1? = Hg = 10 feet
Hs = 1.416 H^,
rms
or
(a) From Figure 3-5, curve b, it is seen that for P = 0.1 (10 percent)
^^ = 1.80(7.06) =
« 1.80; Hjo = 1.80 H 12.7 feet, say 13 feet.
rms
"/
2.36; H^ = 2.36 H^^^ = 2.36 (7.06) = 16.7 ft., say 17 ft.
rtns
Note that;
^10 12.7
or H,„
^<^
= 1.27 H.
H$ 10
and
^ _
~
1^ or H^ = 1.67 H^
H, 10
*************************************
Goodknight and Russell (1963) analyzed wave-gage observations taken
on an oil platform in the Gulf of Mexico during four hurricanes. They
found agreement adequate for engineering_application between such important
parameters as Hg, Hio, ihiax, ^rms^ and H, although they did not find
consistently good agreement between measured wave-height distributions
and the entire Rayleigh distribution. Borgman (1972) substantiates this
conclusion from wave observations from other hurricanes. These findings
are consistent with Figures 3-3, and 3-4, based on wave records recently
obtained by CERC from shore-based gages. The CERC data include waves from
both extratropical storms and hurricanes.
N
77(t)= S a. cos(w.-0.) , (3-11)
J
j= 1 J '
where n(t) is the departure of the water surface from its average posi-
tion as a function of time, aj is the amplitude, toj is the frequency,
and (jjj is the phase of the j*^ wave at the time t = 0. The values
of oj are arbitrary, and cj may be assigned any value within suitable
limits. In analyzing records, however, it is convenient to set
- ZiTJ/D,
(iijj where j is an integer and D is the duration of the obser-
vation. The dij will be large only for those w^- that are prominent in
the record. When analyzed in this manner, the significant period may be
defined as D/j where j
, is the value of j corresponding to the
largest a^-.
It was shown by Kinsman (1965) , that the average energy of the wave
train is proportional to the average value of [n(t)]^. This is identical
to a^ where o is the standard deviation of the wave record. It can
also be shown that
1 ^
o' = - S a?. (3-12)
and
H^ « 4a
then
or
3-12
Thus
E(co)da; . (3-16)
where k = Ztt/L, and 9^- is the angle between the x axis and the
direction of wave propagation, and ^a is the phase of the j^^ wave
at The energy density E(e,u)
t = 0. represents the concentration of
energy at a particular wave direction 6 and frequency w
, therefore ,
27r oo
=
// E(0,a))dwd9 . (3-18)
3-13
Period (Seconds)
2.5 10 2.5
+ 0.4
iiJ
e
> + 0.2
o
—
o
o
0.0
OjO 0.2 0.4 0-0 2 0.4
Frequency (Hertz)
Savannah Coast Guard Light Tower
Period (Seconds)
10 2.5 10 2.5
0.4
T = 10.8 Sec
H, 6.9 Ft
liJ
> 0.2
o
o
I-
0.0 ^-«
0.0 0.2 4
Frequency (Hertz)
Nags Head, N.C.
Figure 3-6. Typical Wave Spectra from the Atlantic Coast.
The ordinate scale is the fraction of total
wave energy- in each frequency band of 0.011
Hertz, (one Hertz is one cycle/second). A
linear frequency scale is shown at the bottom
of each graph and a nonlinear period scale
at the top of each graph.
3-14
3.3 WA\"E FIELD
Riillips (1957) shewed that the turbulence assrciated with thf fl:-
of wind near the water v.c>uld create traveling pressure pulses. Thtir
pulses generate waves traveling at a speed appropriate to the diaensirn;
of the pressure pulse. Wave grc«t;h ry t-i5 7r::e55 i; - - f-
"
the waves are short and ^liien their speed is luenticu. _i-. : .;:
. t
of the wind velocity in the directirn cf wave travel. The err in; ii:a
1
pressure pulses is real, but is only ii:_i ;r. -i eniieth as _^rge ^s t^ie
original theor>' indicated.
3-15
Miles (1957) showed that the waves on the sea surface must be
matched by waves on the bottom surface of the atmosphere. The speed
of air and water must be equal at the water surface. Under most meteoro-
logical conditions, the air speed increases from near to 60 - 90 percent
of the free air value within 66 feet (20 meters) of the water surface.
Within a shear zone of this type, energy is extracted from the mean flow
of the wind and transferred to the waves. The magnitude of this transfer
at any frequency is proportional to the wave energy already present at
that frequency. Growth is normally most rapid at high frequencies. The
energy transfer is also a complex function of the wind profile, the
turbulence of the air stream, and the vector difference between wind and
wave velocities.
The theories of Miles and Phillips predict that waves grow most
rapidly when the component of the wind speed in the direction of wave
propagation is equal to the speed of wave propagation.
Shorter waves grow most rapidly. Those waves which propagate obliquely
to the wind are favored, for they are better matched to the component of
the wind velocity in the direction of wave propagation than those moving
parallel to the wind. Thus, the first wave pattern to appear for short
fetches and durations consists of two wave trains forming a rhombic
pattern with one diagonal along the direction of the mean wind.
3-16
3.32 VERIFICATION OF WAVE HINDCASTING
Bunting and Moskowitz (1970) and Bunting (1970) have compared fore-
cast wave heights with observations, using the same model with comparable
results.
Most coastal areas of the United States are so situated that most of
the waves reaching them are generated in water so deep that depth has no
effect on wave generation. In many of these areas, wave characteristics
3-17
o o
77^
. • \ •
may be determined by first analyzing meteorological data to find deep-
water conditions. Then, by analyzing refraction (Section 2.32, General -
Refraction by Bathymetry.), the changes in wave characteristics as the
wave moves through shallow water to the shore may be found. In other
areas, in particular along the North- Atlantic coast, wKere the bathymetry
is complex, refraction procedure results are frequently difficult to inter-
pret, and the conversion of deepwater wave data to shallow-water and near-
shore data becomes laborious and sometimes inaccurate.
Along the Gulf coast and in many inland lakes, generation of waves by
wind is appreciably affected by water depth. In addition, the nature and
extent of transitional and shallow-water regions complicate ordinary re-
fraction analysis by introducing a bottom-friction factor and associated
wave energy dissipation.
Wind reports from ships at sea are generally estimates based on the
appearance of the waves, the drifting of smoke, or the flapping of flags
although some are anemometer measurements. Actually, even if all ships
were equipped with several aneometers, the wind field over the sea would
still not be known in sufficient detail or precision to permit full
exploitation of modem theories for wave generation.
3-20
Horizontal pressure gradients arise in the atmosphere primarily
because of density differences, which in turn are generated primarily by
temperature differences. Wind results from nature's efforts to eliminate
the pressure gradients, but is modified by many other factors.
1 dp
U = -T
* pt /
an
. (3-19)
where U^ is the wind speed, p^ the density of the air, f the coriolis
parameter, f = 2w sine}), where to = 7.292 x 10"^ radians/second and (f)
3-21
3.41 ESTIMATING THE WIND CHARACTERISTICS
(a) Estimate the mean surface wind speed and direction, as dis-
cussed in Section 3.4, Wind Information Needed for Wave Prediction;
When forecasting for oceans or other large bodies of water, the most
common form of meteorological data used is the synoptic surface weather
chart. (^Synoptic means that the charts are drawn by analysis of many
individual items of meteorological data obtained simultaneously over a
wide area.) These charts depict lines of equal atmospheric pressure,
called isobars. Wind estimates at sea, based on an analysis of the sea-
level atmoshperic pressure are generally more reliable than wind observa-
tions because pressure, unlike wind, can be measured accurately on a
moving ship. Pressures are recorded in millibars, 1,000 dynes per square
centimeter. One thousand millibars (a bar) equals 29.53 inches of mercury
and is 98.7 percent of normal atmoshperic pressure.
Scattered about the chart are small arrow shafts with a varying num-
ber of feathers or barbs. The direction of a shaft shows the direction
of the wind; each one-half feather represents a unit of 5 knots (2.5
meters/second) in wind speed. Thus, in Figure 3-9 near the point 35°N.
latitude, 135°W. longitude, there are three such arrows, two with 5h
feathers which indicate a wind force of 31 to 35 knots (15 to 17.5 meters/
second) , and one with 3 feathers indicating a force of 26 to 30 knots
(13 to 15 meters/second)
On an actual chart, much more meteorological data than wind speed and
direction are shown for each station. This is accomplished by the use of
3-22
o
LO
U
(U
£i
O
t->
O
O
CNJ
M
O
tn
o
o
(-1
3
CO
to
(U
•rt
3-23
coded symbols, letters, and numbers placed at definite points in relation
to the station dot. A sample model report, showing the amount of informa-
tion possible to report on a chart, is shown in Figure 3-10. Not all of
the data shown on this plot are included in each report, and not all of
the data in the report are plotted on each map.
Geostrophic wind speeds are generally higher than surface wind speeds.
The following instructions, U.S. Fleet Weather Facility Manual (1966), are
recommended for obtaining estimates of the surface wind speeds over the
open sea from the geostrophic wind speeds:
Over oceans, the surface winds generally cross the isobars toward low
pressure at an angle of 10° to 20°.
3-24
Wind speed ( 23 to 27 knots ) Cloud type (Altostrotus)
ff
Cloud type (Dense cirrus
Itches)
Barometric pressure in
sky covered I
yu Post weather (Rain)
Height at base of cloud iddle cloud
(2 = 300 to 599 feet) ^ (6 = 7 or 8 tenths) Amount of precipitation
pp (45= 0.45 inches)
3-25
I
Ap
Ur
/'nf An
P 1013.3 mb
1. 247X10"' gm/cm'
f : Coriolis parameter : 2 wsin
0.3
30 35 40 45 50 55 60 70
Degrees Latitude
(offer Bretschneider, 1952a)
Figure 3-11. Geostrophic Wind Scale
3-26
Table 3-1. Correction for Sea-Air Temperature^
Sea Temperature minus
Air Temperature
with this precision. For practical wave predictions it is usually satis-
factory to regard the wind speed as reasonably constant if variations do
not exceed 5 knots (2.5 meters/second) from the mean. A coastline upwind
from the point of interest always limits a fetch. An upwind limit to the
fetch may also be provided by curvature or spreading of the isobars as
indicated in Figure 3-12 (Shields and Burdwell, 1970), or by a definite
shift in wind direction. Frequently the discontinuity at a weather front
will limit a fetch, although this is not always so.
1004
1004 '9°° 996
996
1000
1004
., 1012 1016
1020 1016
1006
1012
Estimates of the duration of the wind are also needed for wave predic-
tion. Computed results, especially for short durations and high wind speeds
may be sensitive to differences of only a few minutes in the duration. Com-
plete synoptic weather charts are prepared only at 6-hour intervals. Thus
interpolation between charts to determine the duration may be necessary.
Linear interpolation is adequate for most uses, and, when not obviously
incorrect, is usually the best procedure.
3-28
3.43 FORECASTS FOR LAKES, BAYS, AND ESTUARIES
FIND: Estimate the significant wave height Hg, and the significant
wave period Tg.
(b) The wind transfers a unit amount of energy to the water along
the central radial in the direction of the wind and along any other radial
an amount modified by the cosine of the angle between the radial and the
wind direction.
3-30
( U.S. Army, B.E.B. Tech. Memo No 132,1962)
3-32
(c) Waves are completely absorbed at shorelines.
Fetch distances determined in this manner usually are less than those
based on maximum straight- line distances over open water. This is true
because the width of the fetch places restrictions on the total amount of
energy transferred from wind to water until the fetch width exceeds twice
the fetch length.
-flCw^/w*)
E(co)da; = (ag^/co*) e " dco ,
(3-20)
3-33
fully arisen sea. The assumption of a universal form for the fully arisen
sea permits the computation of other wave characteristics such as total
wave energy, significant wave height, and period of maximum energy. The
equilibrium state between wind and waves rarely occurs in the ocean, and
may never occur for higher wind speeds.
A more general model may be constructed by assuming that the sea is
calm when the wind begins to blow. Integration of the equations governing
wave growth then permits the consideration of changes in the shape of the
spectrum with increasing fetch and duration. If enough wave and wind
records are available, empirical data may be analyzed to provide similar
information. Pierson, Neumann, and James (1955) introduced this type of
wave prediction scheme based almost entirely on empirical data. Inoue
(1966, 1967) repeated this exercise in a manner more consistent with the
Miles-Phillips theory using a differential equation for wave growth. Inoue
was a member of Pierson 's group when this work was carried out, and his
prediction scheme may be regarded as a replacement for the earlier
Pierson-Neuman-James (PNJ) wave prediction model. The topic has been
extended by Silvester and Vongvisessomjai (1971) and others.
These simplified wave prediction schemes are based on the implicit
assumption that the waves being considered are due entirely to a wind
blowing at constant speed and direction for an overwater distance called
the fetch and for a time period called the duration.
In principle it would be possible to consider some effects of variable
wind velocity hy tracing each wave train. Once waves leave a generating
area and become swell, the wave energy is then propagated according to the
group velocity. The total energy at a point, and the square of the signif-
icant wave height could be obtained by adding contributions from individual
wave trains. Without a computer, this procedure is too laborious, and
theoretically inaccurate.
A more practical procedure is to relax the restrictions implied by
derivation of these schemes. Thus wind direction may be considered con-
stant if it varies from the mean by less than some finite value, say 30°.
Wind speed may be considered constant if it varies from the mean by less
than ± 5 knots (2.5 meters/second) or h barb on the weather map. This
assumption is not much greater than the uncertainty inherent in wind
reports from ships. In this procedure, average values are used and are
assumed constant over the fetch area, and for a particular duration.
The theoretical spectra for the partially arisen sea can be used
to develop formulas for such wave parameters as total energy, significant
wave height and period of maximum energy density.
Similar formulas can also be developed empirically from wind and wave
observations. A quasi-empirical - quasi-theoretical procedure was used by
Sverdrup and Munk (1947) to construct the first widely used wave predic-
tion system. The Sverdrup-Munk prediction curves were revised by
Bretschneider (1952a, 1958) with additional empirical data. Thus, this
prediction system is often called the Sverdrup-Munk-Bretschneider (SMB)
method. It is the most convenient wave prediction system to use when a
limited amount of data and time are available.
3-34
3.51 SMB METHOD FOR PREDICTING WAVES IN DEEP WATER
0.42
^ = 0.283 tanh 0.0125 (3-21)
U^
,0.25
'gF
= 1.20 tanh 0.077 (3-22)
2nU
and.
^
= Kexp B In + Db (3-23)
U
where
exp (x) =
In loge
K 6.5882
A 0.0161
B 0.3692
C 2.2024
and
D = 0.8798
With these relationships, the significant wave height H^, and significant
wave period T^, at the end of a fetch may be estimated when wind speed,
fetch length, and duration of wind over the fetch are given. Estimation
of wind parameters is discussed in Section 3.4, Wind Information Needed
for Wave Prediction, following the evaluation of simplified prediction
techniques. In using Figures 3-15 or 3-16 the estimated wind velocity U,
the fetch length F, and the estimated wind duration t, when a fetch
first appears on a weather chart, are tabulated. Figure 3-15 or 3-16 is
then entered with the value of U, using the scale on the left if U is
in knots, or the scale on the right if U is in statute miles per hour.
This U line is then followed from the left side of the graph across to
its intersection with the fetch length or F line, or the duration t
line, whichever comes first.
3-35
( jnoH J8d s»|iW^ P88ds puiiw
~ in <\t o a> u> ~
^
( jnoH Jad S8|iw ) paads puim
f s
************** EXAMPLE PROBLEM *************
GIVEN : A wind speed U = 35 knots (40 mph) and a duration t = 10 hours
SOLUTION :
(a) Enter Figure 3-15 from the left side at U = 35 knots and move
horizontally across the figure from the left toward the right, until
intersecting the dashed line for a duration of 10 hours that comes
before the line indicating a fetch length of 200 nautical miles. At
the 10-hour duration line F = 92 nautical miles; this is the minimum
fetch F^ for this case.
When a series of surface synoptic weather charts (Fig. 3-9) are used
to determine wave patterns, the values of U, F and t can be tabulated
for the first chart. For the same fetch on a later chart drawn for a time
Z, after the first chart, U, F, and t are again tabulated. Using
the subscript 2 to refer to those of the second chart and subscript 1
to refer to those of the first chart, if U2 = U-^ the above procedures
,
3-38
the energy imparted to the waves by U^ , with a minimiim duration t^^ + Z/2
for a minimum fetch F^^ + F/2, does not change, U2 will be assumed to
impart energy to waves which already contain energy due to Uj^.
Plotted on Figures 3-15 and 3-16 are dotted lines of constant H^T'^ which
are considered lines of constant wave energy. To a first approximation,
deepwater wave energy is given by
U2
.
3-39
with either U2 or F2 whichever comes first from the left side of the
figure. If U comes first, Z/2 is added to the duration at this point,
and the U2 ordinate is followed to either this new duration or to the
F2 whichever is first from the left side of the chart. (Compare with the
preceding paragraph.) At this point, Hp2' ^F2' hn2> ^^^ ^m2 inclusive,
are read off. If the constant energy line had intersected F2 before U2,
it is only necessary to drop down along the F2 abscissa to its intersec-
tion ^'ith U2, and at this point read H^2> '^F2> ^2 ^^^ ^m2- (This pro-
cedure could be used for many cases in which U2 is greater than Uj.)
dicts that waves due to a constant wind blowing over an unlimited fetch
for an unlimited duration will eventually reach limiting height and period
distributions beyond which growth will not continue. In Figure 3-16 the
limit of this state is delineated by the line labelled maximum condition.
To the right of this line, it is assumed that any energy transport to the
waves by the wind is compensated by wave breaking, hence no wave growth
occurs
3.52 EFFECTS OF MOVING STORMS AND A VARIABLE WIND SPEED AND DIRECTION
3-40
STATION A^
oo;
.NEW YORK/
40"
STATION C| 69"
72"
40° 15'
73°45* "AREA 5
AREA 6
^
.V
-AREA 8
A\
AV
3-4
Service Command (1970) provides summaries of shipboard wave observations.
Summary of Synoptic Meteorological Observations (SSMO) for the hatched
areas indicated in Figure 3-17. Cumulative distribution functions for
wave heights as determined by both hindcasting techniques and the ship-
board observations are given in Figure 3-18. The average of the two
forecasting methods agrees reasonably well with the shipboard observations.
Figures 3-19 and 3-20 are used to estimate wave characteristics after
the waves have left the fetch area but are still travelling in deep water.
With Figure 3-19, and given H^, T^, F^ and D (the decay distance), it is
possible to compute the ratios
—z —
decayed wave height
fetch wave height
= — D
Hj,
, and
decayed wave
—-:
fetch
period
—
wave period
= —D
Tp
Water depth affects wave generation. For a given set of wind and
fetch conditions, wave heights will be smaller and wave periods shorter if
generation takes place in transitional or shallow water rather than in deep
water. Several forecasting approaches have been made; the method given by
Bretschneider as modified using the results of Ijima and Tang (1966) is
presented here. Bretschneider and Reid (1953) consider bottom friction
and percolation in the permeable sea bottom.
3-42
(sja)9y^) igbidH 3aom
OloOr^^DlOTrOCVJ— Oe
PT
aoiasd 3AVM sAiivtau
3-44
4600
4400
4200
4000
3800
3600
3400 .^
E
3200 ^
3000 i2
2800 »
2600 -z
2400 S-
2200 "
2000
1800
1600
1400
1200
1000
800
600
400
200
10 12 14 16 22
Wave period T^ , at end of decay (seconds)
3-45
in which wave energy is added due to wind stress and subtracted due to
bottom friction and percolation. This method uses deepwater forecasting
relationships originally developed by Sverdrup and Munk (1947) and revised
by Bretschneider (1951) to determine the energy added due to wind stress.
Wave energy lost due to bottom friction and percolation is determined from
the relationships developed by Bretschneider and Reid (1953). Resultant
wave heights and periods are obtained by combining the above relationships
by numerical methods. The basic assumptions applicable to development of
deepwater wave-generation relationships (Bretschneider, 1952b) as well as
development of relationships for bottom friction loss (Putnam and Johnson,
1949) and percolation loss (Putnam, 1949) apply.
and
gd \ 0.375
= 1.20 tanh 0.833 tanh ,(3-26)
2-nU ,U^
3-46
-H = 2.5ft. Tr- 4 ^o
^ 25 sec
—
^1
1 I
! I^-L M
15 2 25 3 4 5 6 7 8 9 10 1.5 2 25 3 4 5 6 7 8 9 10
X 1,000 Fetch (F) feet X 10,000
T-5 25sec^
H;4.5ft.~.^ T = 5.05ec^ /
3-47
H = 70ft
1=6 sec.
H = 5 75 ft.
3-48
/T=4 5sec. T = 5.5 sec.\ T = 6.0 sec.^ ,H = 8.0 ft.
= 4.0 secV 1=5 setv.,^^ T = 6.5 sec.'
H-8.0 ft.>
3-49
,H-90 ft. H = IOft
= 7 sec
15 2 2 5 3 4 5 6 7 8 9 10 15 2 2 5 3 4 5 6 7 8 9 10
X 1,000 Fetch (F) feet X 10,000
H-IOOft-, H-II.Oft.
T = 6-5 sec.^ fT= 7 1
3-50
T-7.0 secs
T=6 sec. H = ILO ft.N \
't:6 5 sec
3-51
SOLUTION : From Figure 3-27 for constant depth, d = 35 feet,
for
F = 80,000 feet,
and
U = 50 mph.
Then
H = 5.9 feet, say 6 feet,
and
T = 5.1 seconds, say 5 seconds.
***************************** ***k**f*
3.62 DECAY IN LAKES, BAYS, AND ESTUARIES
peaks
3-52
+ 0.4
An indication of the distribution of waves throughout a hurricane can
be obtained by plotting composite charts of shipboard wave observations.
The position of a report is determined by its distance from the storm
center and its direction from the storm track. Changes in storm intensity
and shape are often small enough to permit all observations obtained during
a period of 24 to 36 hours to be plotted on a single chart. Several plots
of this type from Pore (1957) are given in Figure 3-32. Additional data
of the same type have been presented by Arakawa and Suda (1953) Pore (1957)
,
U = Kr for r < R ,
KR^ (3-27)
U = for r > R ,
r
3-54
STORM DIR STORM DIR '\
.V
•
X \'
\10 \'»
,.V'
a CONNIE 1955 >. s
AUG n 0730E THRU
_!^'>- '5/^/1. JL'"^- C CONNIE 1955 AUG 12 0730E
..',
?'^V Y
+-H— I
(tH
\
—I— t^'-
.^'
ivii:
^'' '»i 1 I
>K° ;^
r
/• , i to n
-^^^
•<:v i3'
X'A.
rtf;
/, \" '•
K'
'A . ...
f DIANE 1955
\ ,./
AUG16. 1330E THRU
AUG 17 0730E
4»
I
*30
L
MO ->AiWO
1. /.
IBO laj 60 H \
y a, 10,
/
„ ".'Z /^v\* ^ -vx T.;v v^
"-•
Ml-"
'/
Figure 3-32. Compositive Wave Charts. The wave height in feet is plotted
beside the arrow indicating direction from which the waves
came. The length of the arrow is proportional to the wave
period. Dashed arrow indicates unknown period. Distances
are marked along the radii at intervals of 60 nautical miles.
3-55
A more widely used model was given by Myers (1954). A concise mathe-
matical description of this model is given by Harris (1958) as follows:
(3-28)
ZZ^^e-"/^
1,2
'0
10 20 30 40 50 60 70 10 20 30 40 SO (O
Distance From Pressure Center (Statute Miles) DIatonce From Wind Center (Statute Miles)
and
RAp
0.104 a Vp
T, = 8.6 e
200 +
1 (3-32)
s
vuT
3-57
where
H^ = deepwater significant wave height in feet
where H^ is in feet (derived from empirical data showing that the wave
steepness H/T^ will be about 0.22).
3-58
-M to
3-59
************** EXAMPLE PROBLEM **************
GIVEN Consider a hurricane at latitude 350N. with R = 36 nautical miles,
:
SOLUTION :
RAp _ 36(2.31)
= 0.832,
100 100
and
eO.832 = 2.30
then
0.208 X 1 X 26
H^ = 16.5 eO-832 1 +
V9I.6
3-60
Using Equation 3-32
0.104 X 1 X 26
0.416
Ts = 8.6 e 1 +
V91.6'
It should be noted that computing the values of wave height and period
to three significant figures does not imply the degree of accuracy of the
method; it is done to reduce the computational error.
for the 36.8-foot wave is: Tg = 2.13 /36.8' = 12.9 seconds, say 13 seconds.
36
t = = 1.38 hours = 4,970 seconds; (3-37)
26
3-61
The most probable maximum waves can be obtained by using
Hn = 0-707 H^ - (3-39)
^ log^ .
j 30?
Hj = 0.707(59.4) 100.4 feet, say 100 feet.
Assuming that the 100-foot wave occurred, then the most probable
second highest wave is obtained by setting n = 2, the third from n = 3, etc.
30F
—
Hj = 0.707(59.4) •Jlogg
I
— = 94.1 feet, say 94 feet;
303
= 0.707(59.4)
H3
-3 v-,..^ ,
i
./log
\l^^i>e -^
= 90.2 feet, say 90 feet.
H,
Fe = (3-40)
0.0555 U,
where
and
59.4
Fe = = (11.69) = 137 nautical miles.
0.0555(91.6)
3-62
10
9 ^r"^] 1
1-
8
7
X
<
1.0
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
For the remaining part of this problem, either the value of F^„ equal
to 220 nautical miles as determined from Figure 3-15 for U^ = 91.6,
Hq = 59.4 can be used with the deepwater forecasting curves, or else
Equation 3-40 can be used, as modified,
To = 2.13 V?^ .
Fg is defined below.
The latter being a numerical method is easier to use and more accurate
than the graphical method of using forecasting curves
The procedure for computing wind waves over the Continental Shelf
will be illustrated by using the bottom profile off the mouth of the
Chesapeake Bay and the standard project hurricane developed for the
Norfolk area. The storm surge computed for the standard project hurricane
and 2.5 feet of astronomical tide are added to the mean low water depths
to obtain the total water depth for wave generation. Refraction is
neglected in this example, i.e., K^ = 1.0. The results of these computa-
tions are given in Table 3-4 followed by examples and explanations.
Column 2, d^, is the depth in feet referred to mean low water at the
shoreward end of each section, denoted by X of Column 1.
3-65
o
<u
00
c
o
U
>
o
c/)
>
OS
E
o
U
^'
Column 7 is the deepwater significant wave height H^ and is
obtained from Equation 3-40;
(fH K AX
^ ,- ,
= A (Column 11).
H^ H^
F ' = where U^j = 91.6 knots (moving hurricane).
(0.0555 U^ 5.08
3-67
Column 17 is the shoaling coefficient Kg2 related to the values of
(To) /da (Column 16).
N = —T/ = '
Tj
,
where \ = R/Vp
P
=
26
= — 1.38 hours or 4,970 seconds.
Compute
To (14)^
Column 9, =^ = = 1.73.
dj, 113
304 HK
Column 11, A = — _ = —rr7zzz~
(304 (42.9) (0.924)
(113)^
^^^,
= 0.943.
(d.p)^
Column 12, lO = 0.89 (from Figure 3-35) for values of =^ == 1.73 and
A = 0.943.
Column 16, —
d-
= 1.58
2
Column 18, H = 0.919 (38.2) = 35.1 feet, which is the shallow-water height
for depth d^ = 110 feet, corresponding to MLW of 104 feet.
4,970
Column 19, N = ^— = 377 ,
or the total
, ,
number r 1-
or waves applicable to
, ,
o
steady-state significant wave of H = 35.1 feet, say 35 feet
Column 20, H_^ = 35.1(0.707) Vlog 377' = 60.4 feet, say 60 feet
The moving fetch model of Wilson (1955) has been adapted for computer
usage by Wilson (1961) The basic equations were modified by Wilson (1966)
.
The Bretschneider (1959) model for hurricane wave prediction was modified
by Bretschneider (1972) Borgman (1972) used the results of Wilson (1957)
.
The focus now changes from wave prediction to water level fluctuations
in oceans and other bodies of water which have periods substantially longer
than those associated with surface waves. Several known physical processes
combine to cause these longer-term variations of the water level.
3-69
The first five have periods that range from a few minutes to a few
days; the last two have periods that range from semi-annually to many
years. Although important in long-term changes in water elevations,
climatological and secular variations are not discussed here.
3-70
3.81 ASTRONOMICAL TIDES
Data concerning tidal ranges along the seacoasts of the United States
are given to the nearest foot in Table 3-5. Spring ranges are shown for
areas having approximately equal daily tides; diurnal ranges are shown
for areas having either a diurnal tide or a pronounced diurnal inequality.
Detailed data concerning tidal ranges are published annually in Tide Tables,
U.S. Department of Commerce, National Ocean Survey.
3.82 TSUNAMIS
3-71
DAr 10 II 12 13 14 15 16 17 18 19
LuMf dita: mu. S. declirntion, 9th; ipogee, lOlb: lul quirter, 13th: on equitor, 16th; new moon. 20Ui; peri|M,
3-72
DAY 10 11 12 13 14 15 16 17 18 19 20
Ft
SAN FRANCISCO
^^^^^^^^^^^
6
4
2
-2
SEATTLE
12
10
8
6
4
2
-2
18
16
14
12
10
8
6
4
2
-2
32
30
28
26
24
22
20
18
16
14
12
10
-2
lunat (Jala: max S. declination, 9th, apogee, lOlh, last quartet, 13tli; on equalor, 16lh, new moon. 20th, perigee,
3-73
Table 3-5. Tidal Ranges
Station
Tide Gage Record Showing Tsunami
HONOLULU. HAWAII
May 23-24, 1960
3-75
Theoretical and applied research dealing with tsunami problems has
been greatly intensified since 1960. Preisendorfer (1971) lists more than
60 significant theoretical papers published since 1960. The list does not
include observational papers concerned with the warning system.
Table 3-6. Fluctuations in Water Levels - Great Lakes System (1860 through 1973).
Lake
1 , 1 Nnr
the Great Lakes, is subject to greater wind-induced surface fluctuations,
that is, wind setup, than any other Lake. Wind setup is discussed in
Section 3.86, Storm Surge and Wind Setup.
T =—If (3-43)
4V
3-80
observations are insufficient to provide quantitative trends. (Savage,
1957; Fairchild, 1958; Dorrestein, 1962; Galvin and Eagleson, 1965.) A
laboratory study by Saville (1961) indicated that for waves breaking on
a slope there would be a decrease in the mean water level relative to the
Stillwater level just prior to breaking, with a maximum depression or set-
down at about the breaking point. This study also indicated that from the
breaking point the mean water surface slopes upward to the point of inter-
section with the shore and has been termed wave setup. Wcwe setup is
defined as that super -elevation of the mean water level caused by wave
action alone. This phenomenon is related to a conversion of kinetic
energy of wave motion to a quasi-steady potential energy.
in which Sj^ is the setdown at the breaking zone, T is the wave period,
Hq is the deepwater significant wave height, dj^ is the depth of water
at the breaker point and g is gravity. The laboratory data of Saville
(1961) gives somewhat larger values than those obtained by use of Equation
3-46.
3-81
Equation 3-49 provides a conservative estimate for wave setup at the
shore. The difference between laboratory data and theory, however, is not
likely to exceed the uncertainties of field data.
************** EXAMPLE PROBLEM **************
GIVEN : Hh = 20 feet, T = 12 seconds.
S^ = 0.19
gTV J
Sj^ = 0.19
landmass such as southern Florida and moves offshore. High offshore winds
in this case can cause the water level to drop several feet.
3-83
nautical miles from the center. From the zone of maximum wind to the
periphery of the hurricane, the pressure continues to increase; however,
the wind speed decreases. The atmospheric pressure within the eye is the
best single index for estimating the surge potential of a hurricane. This
pressure is referred to as the centvdl pressure index (CPI) .Generally
for hurricanes of fixed size, the lower the CPI, the higher the wind speeds,
Hurricanes may also be characterized by other important elements, such as
the radius of maximum winds (R) which is an index of the size of the storm,
and the speed of forward motion of the storm system (Vj.) . A discussion of
the formation, development and general characteristics of hurricanes is
given by Dunn and Miller (1964)
Extratropical storms that occur along the northern part of the east
coast of the United States accompanied by strong winds blowing from the
northeast quadrant are called northeasters. Nearly all destructive north-
easters have occurred in the period from November to April; the hurricane
season is from about June to November. A typical northeaster consists of
a single center of low pressure and the winds revolve about this center,
but wind patterns are less symmetrical than those associated with hurri-
canes.
levels will depart from normal during a storm depends on several factors.
The factors are related to the:
Several distinct factors that may be responsible for changing water levels
during the passage of a storm may be identified as:
(a) astronomical tides
(b) direct winds
(c) atmospheric pressure differences
(d) earth's rotation
(e) rainfall
(f) surface waves and associated wave setup
(g) storm motion effects
The elevation of setup or setdown in a basin depends on storm inten-
sity, path or track, overwater duration, atmospheric pressure variation,
speed of translation, storm size, and associated rainfall. Basin charac-
teristics that influence water-level changes are basin size and shape,
and bottom configuration and roughness. The size of the storm relative
to the size of the basin is also important. The magnitude of storm
surges is shown in Figures 3-41 and 3-42. Figure 3-41 shows the differ-
ence between observed water levels and predicted astronomical tide levels
during Hurricane Carla (1961) at several Texas and Louisiana coast tide
stations. Figure 3-42 shows high water marks obtained from a storm survey
made after Hurricane Carla. Harris (1963b) gives similar data from other
hurricanes
3-84
3.864 Initial Water Level . Water surfaces on the open coast or in en-
closed or semienclosed basins are not always at their normal level prior
to the arrival of a storm. This departure of the water surface from its
normal position in the absence of astronomical tides, referred to as an
initial water level, is a contributing factor to the water level reached
during the passage of a storm system. This level may be 2 feet above
normal for some locations along the U.S. Gulf coast. Some writers refer
to this difference in water level as a forerunner in advance of the storm
due to initial circulation and water transport by waves particularly when
the water level is above normal. Harris (1963b) on the other hand, indi-
cates that this general rise may be due to short-period anomalies in the
mean sea level not related to hurricanes. Whatever the cause, the initial
water level should be considered when evaluating the components of open-
coast storm surge. The existence of an initial water level preceeding the
approach of Hurricane Carla is shown in Figure 3-41 and in a study of the
synoptic weather charts for this storm. (Harris, 1963b.) At 0700 hours
(Eastern Standard Time) 9 September 1961, the winds at Galveston, Texas
were about 10 mph, but the open coast tide station (Pleasure Pier) shows
the difference between the observed water level and astronomical tide to
be above 2 feet. Rises of this nature on the open coast can also affect
levels in bays and estuaries.
There are other causes for departures of the water levels from normal
in semienclosed and enclosed basins, such as the effects of evaporation
and rainfall. Generally, rainfall plays a more dominant role since these
basins are affected by direct rainfall and can be greatly affected by
rainfall runoff from rivers. The initial rise caused by rainfall is due
to rains preceding the storm; rains during the passage of a storm have a
time-dependent effect on the change in water level.
3-85
90*
Fort Point
Pier 21
Pleoture Pier
«S»J
-40° 40"
Figure 3-41. Storm Surge and Observed Tide Chart. Hurricane C:irla,
7-12 September 1961. Insert Maps for Freeport and
Galveston, Texas, Areas
3-86
Time Eye of Hurricane
Neorest to Stotion
September, 1961
j
Table 3-8. Highest and Lowest Water Levels — Continued
Table
1969, in excess of 20 feet MSL over many miles of the open Gulf Coast,
with a peak value of 24 feet MSL near Pass Christian, Mississippi. High
water levels in excess of 12 feet MSL on the open coast and 20 feet within
bays were recorded along the Texas coast as the result of Hurricane Carla,
September, 1961. Water levels above 13 feet MSL were recorded in the
Florida Keys during Hurricane Donna, 1960.
The notation and the coordinate scheme employed are shown schematic-
ally in Figure 3-43. D is the total water depth at time t, and is
given by D = d + S, where d is the undisturbed water depth and S is
the height of the free surface above or below the undisturbed depth re-
sulting from the surge. The Cartesian coordinate axes, x and y, are in
the horizontal plane at the undisturbed water level and the z axis is
directed positively upward. The x axis is taken normal to the shoreline
(positive in the shoreward direction), and the y axis is taken alongshore
(positive to the left when facing the shoreline from the sea).
3-92
W77777777777777777777777777Z777777777;77777777Z777Z
PROFILE
LAND
^ JZ.
JiL -^ SL J^
SHORELINE-
SEA
PLAN VIEW
3-93
The differential equations appropriate for tropical or extratropical
storm surge problems on the open coast and in enclosed and semienclosed
basins are as follows;
»M^, »M^
H; ^ ^
^
as
^
^
a, ,„ il ,
^
II' _ :^ , W,P (3-50)
9t ax 3y 3x ax 3x p p ^
'
o s
c a
C «
" c
Hi = angular velocity of earth
(7.29 X 10"^ radians/second);
<f)
= geographical latitude;
"^bx' "^bii
~ ^ ^"^ y components of bottom stress;
Equations 3-50 and 3-51 are approximate expressions for the equations
of motion and Equation 3-52 is the continuity relation for a fluid of
constant density. These basic equations provide, for all practical pur-
poses, a complete description of the water motions associated with nearly
horizontal flows such as the storm surge problem. Since these equations
satisfactorily describe the phenomenon involved, a nearly exact solution
can only be obtained by using these relations in complete form.
3-95
warranted by the size or importance of the problem. These methods are
recommended only if a computer is available. A brief description of
these methods and references to them follows.
(1) Storm Surge on the Open Coast . Ocean basins are large and
deep beyond the shallow waters of the Continental Shelf. The expanse of
ocean basins permits large tropical or extratropical storms to be situated
entirely over water areas allowing tremendous energy to be transferred
from the atmosphere to the water. Wind-induced surface currents, when
moving from the deep ocean to the coast, are impeded by the shoaling
bottom, causing an increase in water level over the Continental Shelf.
3-96
Onshore winds, cause the water level to begin to rise at the edge of the
Continental Shelf. The amount of rise increases shoreward to a maximum
level at the shoreline. Storm surge at the shoreline can occur over long
distances along the coast. The breadth and width of the surge will depend
on the storm's size, intensity, track and speed of forward motion as well
'
as the shape of the coastline, and the offshore bathymetry. The highest
water level reached at a location along the coast during the passage of a
storm is called the maximum surge for that location; the highest maximum
surge is called the peak surge. Maximum water levels along a coast do not
necessarily occur at the same time. The time of the maximum surge at one
location may differ by several hours from the maximum surge at another
location. The variation of maximum surge values and their time for many
locations along the east coast during Hurricane Carol, 1954, are shown in
Figure 3-44. This hurricane moved a long distance along the coast before
making landfall, and altered the water levels along the entire east coast.
The location of the peak surge relative to the location of the landfall
where the eye crosses the shoreline depends on the seabed bathymetry, wind-
field, configuration of the coastline, and the path the storm takes over
the shelf. For hurricanes moving more or less perpendicular to a coast
with relatively straight bottom contours, the peak surge will occur close
to the point where the region of maximum winds intersects the shoreline,
approximately at a distance R, to the right of the storm center. Peak
surge is generally used by coastal engineers to establish design water
levels at a site.
Attempts to evaluate storm surge on the open coast, and also in bays
and estuaries, when obtained entirely from theoretical considerations,
require verification particularly when simplified models are used. The
surge is verified by comparing the theoretical system response and computed
water levels with those observed during an actual storm. The comparison is
not always simple, because of the lack of field data. Most water-level
data obtained from past hurricanes were taken from high water marks in low-
lying areas some distance inland from the open coast. The few water-level
recording stations along the open coast are too widely separated for satis-
factory verification. Estimates of the water level on the open coast from
levels observed at inland locations are unreliable, since corrective adjust-
ments must be made to the data, and the transformation is difficult.
3-97
90" 80<» 70°
'^•B'ar Hartior
'"Portland
'Portsmouth
•"^Boston
I
Sandy Hook
Philadelphia*
Baltimore*
Annapolis*'
31
Atlantic City
+ 40° —
Morehead Citv
Wilmington,
Southport
Charleston.
30
/
Savannah*.
29
./
S"
;302
+ 27 + 30° —
Key:
Development Stage ••••••
Tropical Stoge ^^^i
Hurricane Stoge •^-^
Frontal Stage «»*^»
OPosition at 0700 EST
•Position at 1900 EST
90" 70°
I
Figure 3-44. Storm Surge Chart. Hurricane Carol, 30, 31 August 1951,
Insert Map for New York Harbor
3-98
Aug 30 Aug 31 Aug 28 Aug. 29 Aug 30 Aug. 31 SepI
Figure 3-44. Storm Surge Chart. Hurricane Carol, 30, 31 August 1951.
Insert Map for New York Harbor -- Continued
3-99
maximum wind R; the minimiom central pressure of the hurricanes p ;
the forward speed of the hurricane V„, while approaching or crossing
the coast; and the maximum sustained wind speed W, 30 feet above the
mean water level.
Based on this analysis, the U.S. Weather Bureau (now the National
Weather Service) and U.S. Army Corps of Engineers jointly established
specific storm characteristics for use in the design of coastal struc-
tures. Because the parameters characterizing these storms are specified
from statistical considerations and not from observations, the storms
are termed hypothetical huvvioanes or hypo-hurriaanes . The parameters
for such storms are assumed constant during the entire surge generation
period. Graham and Nunn (1959) have developed criteria for a design storm
where the central pressure has an occurrence probability of once in 100
years. This storm is referred to as the standard project hurr-ioane (SPH)
The mathematical model used for predicting the wind and pressure fields in
the SPH is discussed in Section 3.72, Model Wind and Pressure Fields for
Hurricanes. The SPH is defined as a "hypo-hurricane that is intended to
represent the most severe combination of hurricane parameters that is
reasonably oharaateristia of a region excluding extremely rare combina-
tions." Most coastal structures built by the U.S. Army Corps of Engineers
that are designed to withstand or protect against hurricanes are based on
design water associated with the SPH.
Two simple methods are presented here for estimating storm surge on
the open coast: one a quasi-static numerical scheme and the other a nomo-
graph method. These methods should never be used for estimating the surge
in bays, estuaries, rivers or in low-lying regions landward of the normal
open-coast shoreline. Neither method is entirely satisfactory for all
cases, but for many problems both appear to give reasonable solutions.
The use of each method is illustrated by estimating the peak open- coast
storm surge for an actual hurricane. The peak surges thus calculated are
compared to the surge computed by a complete two-dimensional numerical
model for the same storm.
3-100
(a) Quasi-Static Method for Prediction of Hurricane Surge .
TTiismethod for determining open-coast storm surge is based on theoretical
approximations of the governing hydrodynamic equations originally proposed
by Freeman, Baer, and Jung (1957). The term quasi-static method is used
here to emphasize that this method should be restricted to slow-moving,
large-scale storm systems. This method is called the Bathys trophic Storm
Tide Theory and, unlike earlier one-dimensional theories, some of the
effects of longshore flow and the apparent Coriolis force are considered.
Such an approximation of the theory can be described as a quasi-static
method in which a numerical solution is obtained by successively integrat-
ing wind stresses over the Continental Shelf from its seaward edge to the
shore for a predetermined interval of time.
gD —
3x
= fV + -^ (3-53)
ay _ ^sy - '^by
(3-54)
3t p
3-101
» o
A —
> o
o ° o
•H
D _ 4->
m » O
2 »- "- §
O o o
X
o
> TJ Q. (^
o u
z •H
o
p
M
>s
'P
cd
CQ
^1
.2
(A
in
n
o
p<
in
iJ
g
tn
Q>
u
^
o
u,
(4-1
0)
•H
[L,
2 §
II n
I- ». Q- o« ^
3-102
the bottom contours for valid computations. The above assumption also
implies that there is no flooding landward of the shore; thus, there is
a deficiency in the method when substantial flooding occurs.
The bottom contours of the actual seabed are rarely straight and para-
llel; however, the traverse line can often be oriented so that it is nearly
perpendicular to the contours in an average way. For complex offshore
bathymetry, such an approximation would be invalid. For storms moving
more or less perpendicular to a coastline, the traverse line can be taken
through, or anywhere to the right of, the region of maximum winds, but
never to the left of this region. Many other factors such as the angle
of approach of a storm, the coastline configuration, and inertial effects,
limit the use of such a simple approach.
The bottom and surface shear stresses are assumed to vary according
to:
T
^by KVIVI
(bottom shear stress) (3-55)
sx
^-
~p = kW^ cos e
3-103
The wind stress coefficient is based on one given by Van Dom (1953) and
others which is assiomed to be a function of wind speed; thus,
k = X ,
for W < W^ (3-58)
as
ax
are used to denote discrete points in space and time, respectively. The
quantities Ax and At are allowed to be nonuniform spacings along the
traverse line and time increments, respectively. The finite difference
forms of Equations 3-61, 3-62, and 3-63 are then given as follows:
Ax n+ l
(3-65)
^^^- If +f v"""^
(As,)r.v! \
(3-66)
1
+
B.- + B.>i)" + (B.- + B.^,) n l
At + V + 'A
(3-67)
l + K|V«,^JAt(D-)-^^
where A and B are the kinematic forms of the wind stress given by:
A = k W cos (3-68)
B == k W^ sin (3-69)
The total water depth at the mid-interval between two time steps is
given by:
„n „n + l
*
^A ^A
+ Js
+ (s. + s),;,^
i 2 ^
(3-70)
where
Sg = initial setup
S^ = astronomical tide
3-105
The initial setup refers to the water level at the time the storm surge
computations are started. An approximate relationship giving the atmos-
pheric pressure setup in feet when pressure is expressed in inches of
mercury is
The total depth at the new time level is evaluated by the relation:
d,.+ d,.^,
Dn++'A1 + s, + s^"^+ (s. + sj;,,'A
(3-72)
n+l
(Sap), + (s^p),.,]
The total water level rise at the coast is the sum of all component water
level changes resulting from the meteorological storm plus those components
which are not related to the storm. Hence the total setup is given by
where Sj/ is a component due to the wave setup in the region shoreward of
the breaker line and is related to the breaker height by Equation 3-49.
(See Section 3.85, Wave Setup.) The setup component S^ accounts for the
setup resulting from local conditions such as bottom configuration, coast-
line shape, or other flows influencing the system, such as flows from bay
inlets or rivers. This component can only be estimated from a full under-
standing of the influence of topographic and hydrographic features not
considered in the numerical calculations. Contributions made by the
various setup components to the total surge are shown in Figure 3-46.
Equations 3-65, 3-66, and 3-67 can be rewritten in a more useful form
by introducing dimensional constants for terms frequently appearing in the
equations. Hence,
C,1 Ax
n+l
(As.)r;4 A. + ^.l) (3-74)
i+'A
C2 Ax
(-^)::i (sm0),. + (sin0).^^] V^^,)^ (3-75)
D'
n+l
i+'A
20 \ /
18
E
^o 16
O
w 14
z
5 12
o
o
o
>
100
200
300
a.
a>
O
400
E
o
mo 500
600
The values of the dimensional constants Cj, C2, and C3 in Equations
3-74, 3-75 and 3-76 depend on the units used in performing the calculations.
Table 3-9 gives the dimensions of the variables used in the numerical
scheme in three systems of units and the corresponding value of the
constants for each system. The first column of units is given in the
metric system while the other two are given in mixed units in the English
system.
Parameters
It is customary to assume that the system is initially in a state of
equilibrium and V = 0, and S is uniform along the traverse line at the
start of the computational scheme. Thus, computations should be initiated
for conditions prior to the arrival of strong winds over the Continental
Shelf. Although the real system would seldom, if ever, be in a complete
initial state of equilibrium, errors in assuming it to be are of little
consequence in the computational scheme after several time steps in the
calculations, because the effects of the forcing functions will eventually
predominate
Tide records from the region affected by the storm show the mean water
level to be about 1.2 feet above normal before being affected by the storm.
This value is taken to represent the initial water level. The range of
astronomical tides in this location is about 1.6 feet. A constant value of
0.8-foot above MSL is used in the computations; the final surge hydrograph
at the coast can be subsequently corrected to account for the predicted
variations of the tide. Variations in the initial water level and astro-
nomical tides may be added algebraically to the storm surge calculations
without seriously affecting the final results. The atmospheric pressure
difference, P„ - P„. is needed for evaluating the pressure setup com-
ponent, S^p. Data given in HUR 7-113 (1969) suggest that p^ = 26.73
inches of mercury and p^ = 29.92 inches of mercury are representative for
the hurricane. All of these values are assumed to remain constant for the
calculations
3-109
3-110
a m
o
o
u
3 in
O,
0)
3
ti
%
to
n o (D
JJ ON
<
in
D
E
o
V.^ ^ i
1
o
o
This quasi-linear computational scheme can he used for manual com-
putations; however, since the calculations are repetitive, they can be
performed more efficiently by us,ing a digital computer. When carried out
manually, the technique is laborious, tedious and subject to error.
Bodine (1971) gives a computer program based on the numerical scheme
presented here. Other programs bas«d on similar niomerical schemes using
the quasi-static methods have been developed.
The first nomogram (Fig. 3-51) permits an estimate of the peak surge
Sj, generated by an idealized hurricane with specified CPI and radius of
3-115
3ur|
O *-4 _
T f f f f f
'"J- 00 r^ cxD
'•^ <*i -^ in
13 I I I I
IS
<
+ t4-
°S IS
s
o"
S.S-
II
^ m -'
« ^ „ c^
<
^ -^ »o
o •— c>i
il < '
< E
« PJ ^ -H
^1 r- ^
vo
\f)
ph
r-
^h
~T~ t r^
>3
via 6
-» ^ so
— . ^ <s
n Q
o r-
o^ — ^ ^ *
tn .-' «
— r- ts
+
^5
Q O tn c^
w -: w
c^ <s o
a;
. -q * -^
.- « o
D O .X ri-
M -- "
sun
3-117
VD
in
o
•H
a-
1/1
uO
c
o
I
to
a;
•H
18
30 40 50 60 70 80 90 100 110 120 130 140
Ap (Pressure Drop)mbs. (from jeiesnionski, 1972)
3-119
maximum wind that is moving perpendicular to a shoreline with a speed of
15 MPH. This nomogram indicates there is a critical storm size as reflec-
ted by the radius of maximum winds, R. For a given pressure drop greater
than zero, the highest peak surge is produced for a critical value of R
equal to 30 miles and any value of R greater or less than this value
results in a lesser value of the peak surge.
A second factor F^ given in Figures 3-52 and 3-53 adjusts for the
effects of variations in bathymetric characteristics along the gulf and
Atlantic coasts. A third factor Fm given in Figure 3-54, adjusts for
the effects of storm speed and the angle with which the storm track
intercepts the coast
Sp = h Fs ^M (3-78)
Vp = 13 knots
3-120
laaj ui ijidso
o
o
at
,082
9092
oOfrE
4 6 8 10 12
Computed Peak Surge, Sp (Feet above MSL)
3-124
SOLUTION: Because of the units used in the various nomograms, the units
of the parameters are converted to
3.19(1000)
Ap^ = 3.19 in. He.
^
= = 108 milhbars .
29.53
Sj «s 22 feet (MSL)
F^ ^ 1.23 .
^M * 0.97 .
3-125
Jelesnianski (1972) has calculated the open coast surge for
Hurricane Camilla with a full two-dimensional mathematical model. The
maximum surge envelope along the coast, based on computations from this
model, is shown in Figure 3-56 where the zero distance corresponds to
the point of landfall of the hurricane eye. This figure shows that
the peak surge estimated by this method is about 25+ feet MSL.
Thus, by three independent estimates, it has been found that the peak
surge is about 25 feet MSL which corresponds approximately to that
observed (U.S. Geological Survey) at Pass Christian, Mississippi of
24.2 feet MSL. It would be expected that a slightly higher peak water
elevation occurred because Pass Christian is located a few miles left of
the position where the maximum winds made landfall.
Once the wind field is determined, estimation of the storm surge may
be determined by methods based on the complete basic formulas or the quasi-
static method given. The nomogram method cannot be used, since this scheme
is based on the hurricane parameters.
3-126
o
in a turning of the height contours of the lake surface. However, the
turning of the contours lagged behind the wind so that for a time the
wind blew parallel to the water level contours instead of perpendicular
to them. Contour lines of the lake surface from 1800 hours on 26 August
to 0600 hours on 27 August 1949 are shown in Figure 3-57. The map con-
tours for 2300 hours on 26 August show the wind blowing parallel to the
highest contours at two locations. (Haurwitz, 1951), (Saville, 1952),
(Sibul, 1955), (Tickner, 1957), and (U.S. Army, Corps of Engineers, 1955).)
Recorded examples of wind setup on the Great Lakes are available from
the U.S. Lake Survey, National Ocean Survey, and NOAA. These observations
have been used for the development of theoretical methods for forecasting
water levels during approaching storms and for the planning and design of
engineering works. As a result of the need to predict unusually high
stages on the Great Lakes, numerous theoretical investigations have been
made of wind setup for that area. (Harris, 1953), (Harris and Angelo,
1962), (Platzman and Rao, 1963), (Jelesnianski, 1958), (Irish and
Platzman, 1962), and (Platzman, 1958, 1963, 1965, and 1967),)
Water level variations in an enclosed basin cannot be estimated satis-
factorily if a basin is irregularly shaped, or if natural barriers such as
islands affect the horizontal water motions. However, if the basin is
simple in shape and long compared to width, then water level elevations
may be reasonably calculated using the hydrodynamic equations in one
dimension. Thus if the motion is considered only along the x-axis (major
axis), and advection of momentum, pressure deficit, astronomical effects
and precipitation effects are neglected, then Equations 3-50 and 3-52
reduce to
3U 98 1
9s aiu
(3-80)
at
~ ax
If it is further assumed that steady state exists, then Equation 3-79 becomes
dS 1
i^s+'^b)
(3-81)
dx pgD
The bottom stress is taken in the same direction as the wind stress, since
for equilibrium conditions the flow near the bottom is opposite to flow
induced by winds in the upper layers. Theoretical development of this
wind setup equation was given by Hellstrom (1941), Keulegan (1951), and
others. The mechanics of the various determinations have differed some-
what, but the resultant equation has been about the same. This wind
setup equation is expressed as:
k'np^W'F
AS = COS0 (3-82)
PgD
3-128
53 - *'
0»
3? > So - 3 . !iifi
ssi'
where
= air density
p^
W = Wind velocity
F = Fetch length
CW^F
AS = cos 6 (3-83)
D
3-130
When a basin enclosed with vertical sides has a well-defined major
axis and a gradually varying cross section, it is possible to use a
dynamic one-dimensional, computational model for evaluating fluctuations
in water level resulting from a forcing mechanism such as wind stress,
and also to account for some of the effects of a varying cross section.
The validity of such a model depends on the behavior of the storm system
and the geometric configuration of the basin.
Equations 3-79 and 3-80, when the varying width b of the basin is
introduced, can be written in terms of the volume flow rate Q(x,t) as
9Q 9S b ,
as _ 1 aq (3-85)
9t b dx
where x is taken along the major axis of the basin, and for any time t,
A is the cross-section area, and D is the average depth A/b or
D = d + S.
Various schemes have been proposed for evaluating the water level
changes in an enclosed body of water by using the differential equations.
(Equations 3-84 and 3-85.) The formulation of the problem and the numer-
ical scheme given here is from Bodine, Herchenroder and Harris (1972)
The surface stress and bottom stress terms are taken identical to the
terms given for the quasi-static method for open-coast surge. (See
Equations 3-55 and 3-56, Section 3-865b(l) (a) .) The stress terms, in
terms of the volume flow rate, become
^ _ (3-86)
A finite difference representation of Equations 3-88 and 3-85 may be
expressed in the form
(3-89)
At n+ 1
(3-90)
where
G = 1 + (3-91)
{^I'A-'^ls/S (V/.+V3A)
The value of G is greater than unity for most flow conditions except
for the case when the flow vanishes (Q = 0). The subscripts and super-
scripts i and n are used to denote discrete points in space and time,
respectively. A schematic of the grid system used is shown in Figure 3-58.
It is seen that S at the new time level (t + At) is first evaluated
based on the known values of Q, D, S, A and b(x) lying on triangle (1)
and subsequently followed by an evaluation of S at the new time level
based on known values lying on triangle (2). The solutions at successive
time levels are obtained by a marching process with Q evaluated at the
new time level for all integer steps along x; S is evaluated for all
mid-integer steps along x. Width as a function of x, b(x), is taken
constant for all t and the total depth D is permitted to vary with
time. Thus the cross-section area of the basin A is a function of
distance along the major axis of the basin and a function of time.
The scheme requires that for numerical stability, the time increment
specified for successive calculations At be taken less than Ax/i/g^ax>
where Dmax is the maximum depth (d + S) anticipated in the basin during
passage of a storm system. (Abbott, 1966.)
3-132
CM
0>
O
I
X
e
c
D
~
— e
o
6
0)
+J
M
X
CO
X
•H
00
I
4)
HE
<
X
<a
3-133
For such a recursive type calculation, computer computations reduce
the effort required. However, since this is a one-dimensional formulation
of the problem, it is possible to carry out the necessary computations
manually.
The March 1955 storm on Lake Erie is used to demonstrate the scheme.
A plan view of Lake Erie is shown in Figure 3-59. The width and cross-
section area of the Lake are shown in Figure 3-60, mean bottom profile in
Figure 3-61, and wind and wind direction data for the March 1955 storm in
Figure 3-62.
GIVEN: X
o
_l
<
to
3-135
20 40 60 80 100 120 140 160 180 200 220 240 260
Distonce from Toledo ( Statute Miles)
(after Plotzmon ond Roo, 1963)
3-136
20 40 60 80 100 120 140 160 180 200 220 240 260
Distance from Toledo ( Statute Miles)
3-137
FIND: Wind setup at the new time level.
4KAt|Q;'^^ I
G = 1 +
(D?.:^ + D?+3/J"{b,.^j,+b,>3/J
= i
oi
*^ ~ "^ "^
[(0.0162)2 + (0.0132)2] (26.3 + 21.0)
and
gAt
'^
2Ax {^i+'A ^i+3/2) {^i+'A ^1 + 3/2)"
^^ =
QLV
1.0129
~ [0.070 + 0.0274-0.0463] = 0.0513 mi? /hr.
3-139
and finally, the water level at this discrete point in space and time
is given by Equation 3-90 as
At / _ _ \n+l
^"4 = C«-^(Q,-Q«)' i+'A
The wind setup hydrograph for the ends of the lake as determined by
computer is shown in Figure 3-63. The storm winds blew in the general
direction toward Buffalo at the northeastern end of the lake as indi-
cated by the setdown at Toledo and the setup at Buffalo. The spatial
steps Ax taken are quite large; smaller increments in Ax would give
a more accurate estimate. Calculations with the mathematical model
were initiated with the system in a calm state, i.e., Q = S = 0.
The wind setup profile along the major axis of the lake determined
by the numerical scheme is shown in Figure 3-64 for three time periods
during the storm. The nodal point where the computed water surface
crosses the Stillwater level occurs near the center of the lake, but
this nodal point can vary with time.
3-140
in
o
t—
o
e
I—
O
-^-
en
I
o
T3
(1>
T3
C
o
o
o
M—
CD
O
t—
o
o
X
CO
T3
c
ro
CO
ro
3-141
1/3
o
•p
CO
mu
0)
•H
o
a>
CO
to
O
U
3
bo
•H
[I.
oo (O <\j O cvj ^ CO o
I I I I
(499j)dnias puiM
3-142
extensive low- lying areas surrounding them which may be subject to exten-
sive flooding during severe storm conditions. Also, these basins are
usually shallow, and on the windward side of the basin substantial bottom
areas can become exposed. Thus, during the passage of a storm system,
many semienclosed basins have time-dependent moving boundaries. The
water present at any time in a bay or estuary is also dependent upon the
sea state because of the interrelation between these two bodies of water.
Because of the above characteristics of semienclosed basins, a simplified
approach does not usually provide a satisfactory estimate of water motions,
and methods based on full two-dimensional dynamic approaches should be
employed.
3-143
REFERENCES AND SELECTED BIBLIOGRAPHY
ARAKAWA, H., and SUDA, K. , "Analysis of Winds, Wind Waves, and Swell Over
the Sea to the East of Japan During the Typhoon of September 26, 1935,"
Monthly Weather Review ^ Vol. 81, No, 2, Feb. 1953, pp. 31-37.
BELLAIRE, F.R., "The Modification of Warm Air Moving Over Cold Water,"
Proceedings 3 Eight Conference on Great Lakes Research, 1965.
BODINE, B.R., "Storm Surge on the Open Coast: Fundamentals and Simpli-
fied Prediction," TM-35, U.S. Army, Corps of Engineers, Coastal
Engineering Research Center, Washington, D.C., May 1971.
BODINE, B.R., HERCHENRODER, B.E., and HARRIS, D.L., "A Numerical Model
for Calculating Storm Surge in a Long Narrow Basin with a Variable
Cross Section," Report in preparation, U.S. Army, Corps of Engineers,
Coastal Engineering Research Center, Washington, D.C. , 1972.
BOWEN, A.J.,INMAN, D.L., and SIMMONS, V.P., "Wave 'Setdown' and Set-up,"
Journal of Geophysical Research, Vol. 73, No. 8, 1968.
3-145
BRETSCHNEIDER, C.L., "Hurricane Design-Wave Practice," Journal of the
Waterways and Harbors Division^ ASCE, Vol. 83, WW2 , No. 1238, 1957.
BRETSCHNEIDER, C.L., and REID, R.O., "Change in Wave Height Due to Bottom
Friction, Percolation and Refraction," S4th Annual Meeting of American
Geophysical Union, 1953.
3-146
CHRYSTAL, G., "On the Hydrodynamical Theory of Seiches, with a Biblio-
graphical Sketch," Transactions of the Royal Society of Edinburgh,
Robert Grant and Son, London, Vol. XLI, Part 3, No. 25, 1905.
COLLINS, J.I., and VIEW4AN, M.J., "A Simplified Empirical Model for Hurri-
cane Wind Fields," Offshore Technology Conference, Vol. 1, Preprints,
Houston, Tex., Apr. 1971.
CONNER, W.C., KRAFT, R.H., and HARRIS, D.L., "Empirical Methods of Fore-
casting the Maximum Storm Tide Due to Hurricanes and Other Tropical
Storms," Monthly Weather Review, Vol. 85, No. 4, Apr. 1957, pp. 113-116.
COX, D.C., ed., Proceedings of the Tsunami Meetings Associated with the
10th Paoifia Science Congress, Honolulu, Hawaii, 1961.
DARWIN, G.H., The Tides and Kindred Phenomena in the Solar System,
W.H. Freeman, San Francisco, 1962, 378 pp.
DEFANT, A., Ebbe und Flut des Meeres, der Atmosphare, und der Erdfeste,
Springer, Berlin-Gottingen-Heidelberg, 1953.
DOODSON, A.T., and WARBURG, H.D., Admiralty Manual of Tides, Her Majesty's
Stationery Office, London, 1941, 270 pp.
DRAPER, L., "The Analysis and Presentation of Wave Data - A Plea for Uni-
formity," Proceedings of the 10th Conference on Coastal Engineering,
ASCE, New York, 1967.
DUNN, G.E., and MILLER, B.I., Atlantic Hurricanes y Rev. ed. , Louisiana
State University Press, New Orleans, 1964.
EWING, J. A., "The Generation and Propagation of Sea Waves," Dynamic Waves
in Civil Engineering, Wiley, New York, 1971, pp. 43-56.
3-147
EWING, M., PRESS, F., and DONN, W.L., "An Explanation of the Lake Michigan
Wave of 26 June, 1954," Science, Vol. 120, 1954, pp. 684-686.
FREEMAN, J.C, JR., BAER, L., and JUNG, C.H., "The Bathystrophic Storm
Tide," Journal of Marine Research, Vol. 16, No. 1, 1957.
GUTHRIE, R.C, "A Study of Ocean Wave Spectra for Relation to Ships Bending
Stresses," Informal Report No. IR 71-12, U.S. Naval Oceanographic Office,
Washington, D.C, 1971.
HARRIS, D.L. "Wind Tide and Seiches in the Great Lakes," Proceedings of
the Fourth Conference on Coastal Engineering, ASCE, 1953, pp. 25-51.
3-148
HARRIS, D.L., "An Interim Hurricane Storm Surge Forecasting Guide,"
National Hurricane Research Project, Report No. 32, U.S. Weather
Bureau, Washington, D.C. , 1959.
HARRIS, D.L., "Coastal Flooding by the Storm of March 5-7, 1962," Pre-
sented at the Amerioan Meteorological Society Annual Meeting, Unpub-
lished Manuscript, U.S. Weather Bureau, 1963a.
HARRIS, D.L., "A Critical Survey of the Storm Surge Protection Problem,"
The 11th Symposium of Tsunami and Storm Surges, 1967, pp. 47-65.
HARRIS, D.L., and ANGELO, A.T., "A Regression Model for the Prediction
of Storm Surges on Lake Erie," Proceedings, Fifth Conference on Great
Lakes Research, Toronto, 1962.
HASSE L., and WAGNER, V., "On the Relationship Between Geostrophic and
Surface Wind at Sea," Monthly Weather Review, Vol. 99, No. 4, Apr.
1971, pp. 255-260.
HAURWITZ, B., "The Slope of Lake Surfaces under Variable Wind Stresses,"
TM-25, U.S. Army, Corps of Engineers, Beach Erosion Board, Washington,
D.C, Nov. 1951.
HELLSTROM, B., "Wind Effects on Lakes and Rivers," Nr. 158, Handlingar,
Ingeniors Vetenskaps Akadamien, Stockholm, 1941.
HIDY, CM., and PLATE, E.J., "Wind Action on Water Standing in a Labora-
tory Channel," Journal of Fluid Mechanics, Vol. 25, 1966, pp. 651-687.
3-149
HWANG, L., and DIVOKY, D. , "Breaking Wave Setup and Decay on Gentle
Slopes," Prooeedings of the 12th Coastal Engineering Conferenae^
ASCE, Vol. 1, 1970.
INOUE, T. , "On the Growth of the Spectrum of a Wind Generated Sea Accord-
ing to a Modified Miles-Phillips Mechanism," Geophysical Sciences Lab-
oratory Report No. TR-66-6, Department of Meteorology and Oceanography,
New York University, N.Y. , 1966.
INOUE, T., "On the Growth of the Spectrum of a Wind Generated Sea Accord-
ing to a Modified Miles-Phillips Mechanism and its Application to Wave
Forecasting," Geophysical Sciences Laboratory Report No. TR-67-5,
Department of Meteorology and Oceanography, New York University,
N.Y., 1967.
ISAACS, J.D., and SAVILLE, T. JR., "A Comparison Between Recorded and
Forecast Waves on the Pacific Coast," Annals of the New York Academy
of Sciences , Vol. 51, Art. 3, "Ocean Surface Waves," 1949, pp.
502-510.
JACOBS, S.L., "Wave Hindcasts vs. Recorded Waves," Final Report 06768,
Department of Meterology and Oceanography, University of Michigan,
Ann Arbor, Mich., 1965.
3-150
JELESNIANSKI, C.P., "SPLASH-Special Program to List Amplitudes of Surges
from Hurricanes," National Weather Service, Unpublished Manuscript,
National Oceanic and Atmospheric Administration, Rockville, Md. , 1972.
LONGUET-HIGGINS, M.S., and STEWART, R.W., "A Note on Wave Setup," Journal
of Marine Researoh, Vol. 21(1), 1963, pp. 4-10.
3-151
MASCH, F.D., et al , "A Numerical Model for the Simulation of Tidal Hydro-
dynamics in Shallow Irregular Estuaries," Technical Report HYD 12,6901
Hydraulic Engineering Laboratory, University of Texas, Austin, Tex.,
1969.
MILES, J.W., "On the Generation of Surface Waves by Shear Flows," Journal
of Fluid Meohaniosj Vol. 3, 1957, pp. 185-204.
MIYAZAKI, M., "A Numerical Computation of the Storm Surge of Hurricane
Carla 1961 in the Gulf of Mexico," Technical Report No. 10, Depart-
ment of Geophysical Sciences, University of Chicago, Chicago, 111.,
1963.
MORTIMER, C.H., "Spectra of Long Surface Waves and Tides in Lake Michigan
and at Green Bay, Wisconsin," Proceedings, Eight Conference on Great
Lakes Eesearohj 1965.
MOSKOWITZ, L., "Estimates of the Power Spectrums for Fully Developed Seas
for Wind Speeds of 20 to 40 Knots," Journal of Geophysical Research,
Vol. 69, No. 24, 1964, pp. 5161-5179.
MUNK, W.H., "Proposed Uniform Procedure for Observing Waves and Interpret-
ing Instrument Records," Wave Project, Scripps Institute of Oceanogra-
phy, LaJolla, Calif., 1944.
PETERSON, K.R., and GOODYEAR, H.V., "Criteria for a Standard Project North-
easter for New England North of Cape Cod," National Hurricane Research
Project, Report No. 68, U.S. Weather Bureau, Washington, D.C., 1964.
PIERSON, W.J., JR., and MOSKOWITZ, L,, "A Proposed Spectral Form for
Fully Developed Wind Seas Based on the Similarity Theory of S.A.
Kitaigorodskii," Journal of Geophysical Research, Vol. 69, No. 24,
1964, pp. 5181-5190.
PIERSON, W.J., JR., NEUMANN, G., and JAMES, R.W., "Practical Methods for
Observing and Forecasting Ocean Waves by Means of Wave Spectra and
Statistics," Publication No. 603, U.S. Navy Hydrographic Office,
Washington, D.C., 1955.
3-152
PLATZMAN, G.W., "A Numerical Computation of the Surge of 26 June 1954 on
Lake Michigan," Geophysics j Vol. 6, 1958.
PLATZMAN, G.W. , and RAO, D.B., "Spectra of Lake Erie Water Levels,"
Journal of Geophysical Research^ Vol. 69, Oct. 1963, pp. 2525-2535.
PLATZMAN, G.W., and RAO, D.B., "The Free Oscillations of Lake Erie,"
Studies on Oceanography ^ University of Washington Press, Seattle,
Wash., 1965, pp. 359-382.
PORE, N.A., and CUMMINGS, R.A., "A Fortran Program for the Calculation of
Hourly Values of Astronomical Tide and Time and Height of High and Low
Water," Technical Memorandum WBTM TDL-6, U.S. Department of Commerce,
Environmental Science Services Administration, Washington, D.C., 1967.
PORE, N.A., and RICHARDSON, W.S., "Interim Report on Sea and Swell Fore-
casting," Technical Memorandum TDL-13, U.S. Weather Bureau, 1967.
PORE, N.A., and RICHARDSON, W.S. , "Second Interim Report on Sea and Swell
Forecasting," Technical Memorandum TDL-17, U.S. Weather Bureau, 1969.
3-153
PUTNAM, J. A., and JOHNSON, J.W., "The Dissipation of Wave Energy by
Bottom Friction," Transactions of the American Geophysical Union^
Vol. 30, No, 1, 1949, pp. 67-74.
REID, RoO., "Short Course on Storm Surge," Summer Lectures, Texas A^M
University, College Station, Tex., 1964.
REID, R.O., and BODINE, B.R., "Numerical Model for Storm Surges Galveston
Bay," Journal of the Waterways and Harbors Division^ ASCE, Vol. 94,
No. WWl, Proc. Paper 5805, 1968, pp. 33-57.
ROLL, H.U., and FISHER, F., "Eine Kritische Bemerkung zum Neumann-
Spektrum des Seeganges," Deut. Hydrograph. Z, , Vol. 9, No. 9, 1956.
SAVAGE, R.P., "Model Tests of Wave Runup for the Hurricane Protection
Project," The Bulletin of the Beach Erosion Boards Vol. 11, July 1957,
pp. 1-12.
SAVILLE, T., JR., "Wind Setup and Waves in Shallow Water," TM-27, U.S.
Army, Corps of Engineers, Beach Erosion Board, Washington, D.C.,
June 1952.,
SAVILLE, T., JR., "The Effect of Fetch Width on Wave Generation," TM-70,
U.S. Army, Corps of Engineers, Beach Erosion Board, Washington, D.C.,
Dec. 1954.
SHEMDIN, O.H., and HSU, E.T., "The Dynamics of Wind in the Vicinity of
Progressive Water Waves," Technical Report No. 66, Department of
Civil Engineering, Stanford University, Stanford, Calif., 1966.
SHIELDS, G.C., and BURDWELL, G.B., "Western Region Sea State and Surf
Forecaster's Manual," Technical Memorandum WR-51, U.S. Weather
Bureau, 1970.
3-154
SIBUL, 0., "Laboratory Study of Wind Tides in Shallow Water," TM-61, U.S.
Army, Corps of Engineers, Beach Erosion Board, Washington, D.C., Aug.
1955.
SIMPSON, R.B., and ANDERSON, D.V., "The Periods of the Longitudinal Sur-
face Seiche of Lake Ontario," Proaeedings , Seventh Confevenoe on Great
Lakes Research , 1964.
SVERDRUP, H.U., JOHNSON, M.W., and FLEMING, R.H., The Oceans; Their
Physios t Chemistry, and General Biology; Prentice-Hall, New York, 1942.
SVERDRUP, H.U., and MUNK, W.H., "Wind, Sea, and Swell: Theory of Rela-
tions for Forecasting," Publication No. 601, U.S. Navy Hydrographic
Office, Washington, D.C. , Mar. 1947.
SYMONS, T.M., and ZELTER, B.D., "The Tsunami of May 22, 1960 as Recorded
at Tide Stations, Preliminary Report," U.S. Dept. of Commerce, Coast
and Geodetic Survey, Washington, D.C, 1960.
U.S. ARMY, CORPS OF ENGINEERS, "Waves and Wind Tides in Shallow Lakes
and Reservoirs, Summary Report, Project CW-167," South Atlantic-
Gulf Region, Engineer District, Jacksonville, Fla. , 1955.
3-155
U.S. DEPARTMENT OF COMMERCE, NATIONAL OCEANIC AND ATMOSPHERIC ADMINI-
STRATION, Tide Tables 3 East Coast, North and South America, National
Ocean Survey, Rockville, Md. , 1973.
U.S. FLEET WEATHER FACILITY, "Forecasting Ocean Waves and Surf," U.S.
Naval Air Station, Oceanographic Services Section, Operations Depart-
ment, San Diego, Calif., 1966.
U.S. WEATHER BUREAU, "Preliminary Analysis of Surface Wind Field and Sea-
Level Pressures of Hurricane Camille (August 1969)," Hurricane Research
Report, HUR 7-113, U.S. Weather Bureau, Washington, D.C, 1969.
U.S. WEATHER BUREAU, "Revised Surface Wind Field (30 ft) of Hurricane
Camille in Gulf of Mexico (August 1969)," Hurricane Research Report,
HUR 7-113A, U.S. Weather Bureau, Washington, D.C, 1970.
WILSON, B.W., "Hurricane Wave Statistics for the Gulf of Mexico," TM-98,
U.S. Array, Corps of Engineers, Beach Erosion Board, Washington, D.C,
June, 1957.
3-156
WILSON, B.W., "Deep Water Wave Generations by Moving Wind Systems,"
Journal of the Waterways and Harbors Division^ ASCE, Vol. 87,
No. WW2, Proc. Paper 2821, May, 1961, pp. 113-141.
WILSON, B.W., "Design Sea and Wind Conditions for Offshore Structures,"
Proaeedings of Offshore Exploration Conferenoe^ Long Beach, Calif.,
1966, pp. 665-708.
WILSON, B.W., WEB, L.M., and HENDRICKSON, J. A., "The Nature of Tsunamis,
Their Generation and Dispersion in Water of Finite Depth," National
Engineering Science Co., Washington, D.C, 1962.
3-157
CHAPTER 4
LITTORAL
PROCESSES
'^
\
'./•»-
,^^
.'
NORTH CAROLINA COAST - 13 February 1973
CHAPTER 4
LITTORAL PROCESSES
4.1 INTRODUCTION
4.11 DEFINITIONS
4.112 Areal View . Figure 4-2 shows three generalized charts of different
U.S. coastal areas, all to the same scale. Figure 4-2a shows a rocky coast,
well-indented, where sand is restricted to local pocket beaches. Figure
4-2b shows a long straight coast with an uninterrupted sand beach. Figure
4-2c shows short barrier islands interrupted by inlets. These are some of
the different coastal configurations which reflect differences in littoral
processes and the local geology.
4.121 Waves . Water waves are the principal cause of most shoreline
changes. Without wave action on a coast, most of the coastal engineering
4-1
Datum = MTL = Mean Tide Level
Initial Survey 11 October 1962
- MTL-Beach
Intersect
(Surveys did not always reach MTL)
20
a>
a>
o 10
4-2
Rhode Island
(See USCaCS Chart I2IQ).
OCEAN
Northeastern Florida
(See use 8 GS Chart 1246)
BAY
OCEAN
b. Straight Barrier Island Shoreline
BAY
4-3
problems involving littoral processes would not occur. A knowledge of
incident waves, or of surf, is essential for coastal engineering planning,
design, and construction.
4.122 Currents. Water waves induce an orbital motion in the fluid beneath
them. (See Section 2.23.) These are not closed orbits, and the fluid
experiences a slight wave-induced drift, or mass transport. Magnitude and
direction of mass transport are functions of elevation above bottom and
wave parameters (Equation 2-55), and are also influenced by wind and tem-
perature gradients. The action of mass transport, extended over a long
period, can be important in carrying sediment onshore or offshore, par-
ticularly seaward of the breaker position.
4-4
zone is the principal cause of longshore currents - currents that flow
parallel to the shoreline within the surf zone. These longshore currents
are largely responsible for the longshore sediment transport.
There is some mean exchange between the water flowing in the surf
zone and the water seaward of the breaker zone. The most easily seen of
these exchange mechanisms are the rip currents (Shepard and Inman, 1950),
which are concentrated jets of water flowing seaward through the breaker
zone.
Currents induced by storm surges (Murray, 1970) are less well known
because of the difficulty in measuring them, but their effects are un-
doubtedly significant.
4.124 Winds . Winds act directly on beaches by blowing sand off the
beaches (deflation) and by depositing sand in dunes. (Savage and Wood-
house, 1968.) Deflation usually removes the finer material, leaving
behind coarser sediment and shell fragments. Sand blown seaward from
the beach usually falls in the surf zone; thus it is not lost, but is
introduced into the littoral transport system. Sand blown landward from
the beach may form dunes, add to existing dunes, or be deposited in
lagoons behind barrier islands.
4-5
1967; Keulegan, 1951; and van Dom, 1953.) These surface currents drift
in the direction of the wind at a speed equal to 2 to 3 percent of the
wind speed. In hurricanes, winds generate surface currents of 2 to 8
feet per second. Such wind-induced surface currents toward the shore
cause significant bottom return flows which may transport sediment sea-
ward; similarly, strong offshore winds can result in an offshore surface
current, and an onshore bottom current which can aid in transporting
sediment landward.
4-6
m
u
1-5
s
ID
2
e
o
t->
+->
o
•H
(U
2
C
o
•H
in
o
<L>
to
I
•H
4-7
m
U
(U
s
z
c
0)
>
o
ca
o
H
tn
o
U
C
o
•H
P
<U
^^
O
O
4)
c
U
O
4-8
u
o
IX
•H
i-l
0>
o
CO
-s
w
3
•H
4-9
Table 4-1 . Seasonal
than are typical of Atlantic coast beaches. (Urban and Galvin, 1969;
Zeigler and Tuttle, 1961.) Available data indicate that the greatest
changes on the profile are in the position of the beach face and of the
longshore bar - two relatively mobile elements of the profile. Beaches
change in plan view as well. Figure 4-6 shows the change in shoreline
position at seven east coast localities as a function of time between
autumn 1962 and spring 1967.
The long-term changes shown in Figures 4-3, 4-4, and 4-5 illustrate
shorelines of erosion, accretion, and stability. Long-term erosion or
accretion rates are rarely more than a few feet per year in horizontal
motion of the shoreline, except in localities particularly exposed to
erosion, such as near inlets or capes. Figure 4-5 indicates that shore-
lines can be stable for a long time. It should be noted that the erod-
ing, accreting, and stable beaches shown in Figures 4-3, 4-4, and 4-5
are on the same barrier island within a few miles of each other.
Net longshore transport rates along ocean beaches range from near
zero to 1 million cubic yards per year, but are typically 100,000 to
500,000 cubic yards per year. Such quantities, if removed from a 10- to
20-mile stretch of beach year after year, would result in severe erosion
problems. The fact that many beaches have high rates of longshore trans-
port without unusually severe erosion suggests that an equilibrium condi-
tion exists on these beaches, in which the material eroded is balanced by
the material supplied; or in which seasonal reversals of littoral trans-
port replace material previously removed.
4.21 CLASSIFICATION
o
o
(O
o
J=
CO
>
_l
o
a>
W
c
o
a>
c
o
4-12
Unified Soil Classification is the principal classification used by engi-
neers. The Wentworth classification is the basis of a classification
widely used by geologists, but is becoming more widely used by engineers
designing beach fills.
millimeters.
4-13
Wentworth Scale
(Size Description)
(c) The normal distribution is described by its mean and standard
deviation. Since the distribution of sand size is approximately lognormal,
then individual sand size distributions can be more easily described by
units based on the logarithm of the diameter rather than the absolute diam-
eter. Comparison with the theoretical lognormal distribution is also a
convenient way of characterizing and comparing the size distribution of
different samples.
Of these three advantages, only (a) is unique to the phi units. The
other two, (b) and (c) , would be valid for any unit based on the logarithm
of size.
(c) The median diameter can be easily obtained without phi units.
(d) Phi units are dimensionless, and are not usable in physically
related quantities where grain size must have units of length such as
grain-size, Reynolds number, or relative roughness.
The median (Mj) and the mean (M) define typical sizes of a sample of
littoral materials. The median size, N^ in millimeters, is the most com-
mon measure of sand size in engineering reports. It may be defined as
M^ = d,„ (4-2)
where dso is the size in millimeters that divides the sample so that half
the sample, by weight, has particles coarser than the dso size. An equiv-
alent definition holds for the median of the phi size distribution, using
the symbol M^,), instead of M^f.
4-15
Inraan, 1952; Folk and Ward, 1957, McCammon, 1962.) These formulas are
averages of 2, 3, 5, or more symmetrically selected percentiles of the
phi frequency distribution, such as the formula of Folk and Ward.
M0 = (4-3)
^
where i|) is the particle size in phi units from the distribution curve
at the percentiles equivalent to the subscripts 16, 50 and 84 (Fig. 4-8);
4>^ is the size in phi units that is exceeded by x percent (by dry weight)
of the total sample. These definitions of percentile (after Griffiths, 1967,
p. 105) are known as graphic measures. A more complex method - the method
of moments - can yield more precise results when properly used.
The median and the mean describe the approximate center of the sedi-
ment size distribution. In the past, most coastal engineering projects
have used only this size information. However, for more detailed design
calculations of fill quantities required for beach restoration projects
(Sections 5.3 and 6.3), it is necessary to know more about the size dis-
tribution.
Since the actual size distributions are such that the log of the size
is approximately normally distributed, the approximate distribution can be
described (in phi units) by the two parameters that describe a normal dis-
tribution - the mean and the standard deviation. In addition to these
two parameters (mean and standard deviation), skewness and kurtosis
describe how far the actual size distribution of the sample departs from
this theoretical lognormal distribution.
'84 ^16
oa.
'4>
= : ,
(4-4)
where i^iqi^ is the sediment size, in phi units, that is finer than 84 per-
cent by weight, of the sample. If the sediment size in the sample actually
has a lognormal distribution, then a± is the standard deviation of the
4-16
99.9
99.8
99.5
o
oo
c
0)
o
k_
«
Q.
>
E
o
I I I
J I I I I
1.5 1.0 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.15 0.1 0.08 0.06
Diameter (mm)
4-17
sediment in phi units. For perfectly sorted sediment, 0^ = 0. For
typical well-sorted sediments, o, =» 0.5.
The degree by which the phi size distribution departs from symmetry
is measured by the skewness (Inman, 1952), as
^<t>
" ^d4> (4-5)
* a<l>
Currently, median grain size is the most commonly reported sand size
characteristic, probably because there are only limited data to show the
usefulness of other size distribution parameters in coastal engineering
design. However, the standard deviation (Equation 4-4) must also be
given as a parameter for use in beach fill design. (See Section 5,3;
Krumbein and James, 1965; Vallianos, 1970; Berg and Duane, 1968.)
4-18
Table 4-2. Density of Littoral Materials
4.241 Rock. Exposed rock along a shore indicates that the rate at which
sand is supplied to the coast is less than the potential rate of sand
transport by waves and currents. Reaction of a rocky shore to wave attack
is determined by the structure, degree of lithification, and ground water
characteristics of the exposed rock, and by the severity of the wave
climate. Protection of eroding cliffs is a complex problem involving
geology, rock mechanics, and coastal engineering. Two examples of these
problems are the protection of the cliffs at Newport, Rhode Island (U.S.
Army, Corps of Engineers, 1965) and at Gay Head, Martha's Vineyard,
Massachusetts. (U.S. Army, Corps of Engineers, 1970.)
4-21
coastlines, active coral reefs are restricted to southern Florida, Hawaii,
Virgin Islands and Puerto Rico. On Some of the Florida coast, reeflike
structures are produced by sabellariid worms. (Kirtley, 1971.) Organic
reefs stabilize the shoreline and sometimes affect navigation.
From the eastern tip of Long Island, New York, to the southern tip
of Florida, the littoral materials are characteristically sand with median
diameters in the range of 0.2 to 0.6 millimeter (2.3 to 0.7 phi). This
material is mainly quartz sand. In Florida, the percentage of calcium
carbonate in the sand tends to increase going south until, south of the
Palm Beach area, the sand becomes predominantly calcium carbonate. Size
distributions for the Atlantic coast, compiled from a number of sources,
are shown in Figure 4-9. (Bash, 1972.) Fine sediments and organic sedi-
ments are common minor constituents of the littoral materials on these
coasts, especially in South Carolina and Georgia. Beach rock and coquina
are common at shallow depths along the Atlantic coast of Florida.
4.252 Gulf Coast . The Gulf of Mexico coasts of Florida, Alabama, and
Mississippi are characterized by fine white sand beaches and by stretches
of swamp. The swampy stretches are mainly in Florida, extending from
Cape Sable to Cape Romano, and from Tarpon Springs to the Ochlockonee
River. (Shepard and Wanless, 1971, p. 163.)
4-22
1.6 -0.68
1.4 -0.49
1.2 -0.26
0.0
0.8 0.32
0.6 0.64
0.4 .32
|l|i
« 0.6 0.64
E 0.4
o 1.32
0.2 !!\i,
2.32 =
•a
<i>
1
1.0
0.8
0.6
0.4 —
0.2
4.253 Pacific Coast . Sands on the southern California coast range in
size from 0.1 to 0.6 millimeter (3.3 to 0.7 phi). (Emery, 1960, p. 190.)
The northern California coast becomes increasingly rocky, and coarser
material becomes more abundant. The Oregon and Washington coasts include
considerable sand (Bascom, 1964), with many rock outcrops. Sand-sized
sediment is contributed by the Columbia River and other smaller rivers.
4.254 Alaska . Alaska has a long coastline (47,300 miles), and is corres-
pondingly variable in littoral materials. However, beaches are generally
narrow, steep, and coarse-grained; they commonly lie at the base of sea-
cliffs. (Sellman, et al., 1971, p. D-10.) Quartz sand is less common and
gravel more common here than on many other U.S. coasts.
4.256 Great Lakes . The U.S. coasts of the Great Lakes vary from high
bluffs of clay, shale, and rock, through lower rocky shores and sandy
beaches, to low marshy clay flats. (U.S. Army Corps of Engineers, National
Shoreline Study, August 1971, North Central Division, p. 13.) The litto-
ral materials are quite variable. Specific features are discussed, for
example, by Bowman (1951); Hulsey (1962); Davis (1964-65); Bajorunas and
Duane (1967); Berg and Duane (1968); Saylor and Upchurch (1970); Hands
(1970); and Corps of Engineers (1953, 1971).
(d) variation of (a), (b) , and (c) , with horizontal and vertical
position on the site, and
(e) possible variation in (a), (b) , (c) , and (d) with time.
4-24
The extent of sampling depends on the importance of littoral materials
as related to the total engineering problem. The sampling program should
specify:
(d) vertical location and type of sampled volume (e.g. surface layer
or vertical core)
Beaches typically show more variation across the profile than along
the shore, so sampling to determine variation in the littoral zone should
usually be along a line perpendicular to the shoreline.
For reconnaissance sampling, a sample from both the wetted beach face
and from the dunes is recommended. More extensive samples could be ob-
tained at constant spacings across the beach or at different locations on
the beach profile. Spacings between sampling lines are determined by the
variation visible along the beach or by statistical techniques.
4-25
ballpoint ink on plastic strip (plastic orange flagging commonly used by-
surveyors). Some information may be preprinted by rubber stamp on the
plastic strip using indelible laundry ink. The advantage is that the
label can be stored in the bag with the wet sample without the label
deteriorating or the information washing or wearing off.
to the U.S. Standard series. These standard sieve openings vary by a fac-
tor of 1.19 from one opening to the next larger (by the fourth root of 2,
or quarter phi intervals), e.g., 0.25, 0.30, 0.35, 0.42, and 0.50 milli-
meters (2.00, 1.75, 1.50, 1.25, 1.00 phi). The range of sieve sizes used,
and the size interval between sieves selected can be varied as required.
Typical beach sand can be analyzed adequately using sieves with openings
ranging from 0.062 to 2.0 millimeters (4.0 to -1.0 phi), in size incre-
ments increasing by a factor of 1.41 (half-phi intervals).
4-26
There are numerous types of settling tubes; the most common is the
visual accumulation tube (Colby and Christensen, 1956), of which there
are also several types. The type now used at CERC (the rapid sediment
analyzer or RSA) works in the following way:
Waves arriving at the shore are the primary cause of sediment trans-
port in the littoral zone. Higher waves break further offshore, widening
the surf zone and setting more sand in motion. Changes in wave period or
height result in moving sand onshore or offshore. The angle between the
crest of the breaking wave and the shoreline determines the direction of
the longshore component of water motion in the surf zone, and usually the
lon^chore transport direction. For these reasons, knowledge about the
wave climate - the combined distribution of height, period, and direction
through the seasons - is required for an adequate understanding of the
littoral processes of any specific area.
4-27
4.32 FACTORS DETERMINING LITTORAL WAVE CLIMATE
4.322 Effect of Bottom Topography . As waves travel from deep water, they
change height and direction because of refraction, shoaling, bottom fric-
tion, and percolation. Laboratory experiments indicate that height and
apparent period are also changed by nonlinear deformation of the waves in
shallow water.
4-28
Such differences are often visible on aerial photos. Photos may-
show two or more distinct wave trains in the nearshore area, with the
wave train most apparent offshore decreasing in importance as the surf
zone is approached, (e.g., Harris, 1972.) The difference appears to be
caused by the effects of refraction and shoaling on waves of different
periods. Longer period waves, which may be only slightly visible off-
shore, may become the most prominent waves at breaking, because shoaling
increases their height relative to the shorter period waves. Thus, the
wave period measured from the dominant wave offshore may be less than the
wave period measured from the dominant wave entering the surf zone when
two wave trains of unequal period reach the shore at the same time.
A storm near the coastline will influence wave climate with storm
surge and high seas; a storm offshore will influence wave climate only
by swell. The relation between the meteorological severity of a storm
and the resulting beach change is complicated. (See Section 4.35.)
Storms are not uniformly distributed in time or space: storms vary
seasonally and from year to year; storms originate more frequently in
some areas than in others; and storms follow characteristic tracks deter-
mined by prevailing global circulation and weather patterns.
4.331 Mean Value Data on U.S. Littoral Wave Climates . Wave height and
period data for some localities of the U.S. are becoming increasingly
available (e.g., Thompson and Harris, 1972), but most localities still
lack such data. However, wave direction is difficult to measure, and
consequently direction data are rarely available.
4-29
The quality and quantity of available data often do not justify
elaborate statistical analysis. Even where adequate data are available,
a simple characterization of wave climate meets many engineering needs.
While mean values of height and, to a lesser degree, period are useful,
data on wave direction are generally of insufficient quality for even
mean value use. Table 4-4 compiles mean annual wave heights collected
from a number of wave gages and by visual observers along the coasts of
the United States. Visual observations were made from the beach of waves
near breaking. The gages were fixed in depths of 10 to 28 feet.
In Figure 4-10 the minimum monthly mean littoral zone wave height
averaged for the California, Oregon, and Washington coasts exceeds the
maxirmm mean littoral zone wave height averaged for the other coasts.
This difference greatly affects the potential for sediment transport in
the respective littoral zones, and should be considered by engineers
when applying experience gained in a locality with one nearshore wave
climate to a problem at a locality with another wave climate.
4-30
fO OJ —
(4^) iM6!9H 9ADM dJogsJoaN Amiuo|/^ ud9^
0)
>
u
o
U
GO
c
•H
TJ
3
i-H
o
c
O
•H
>
o
to
cd V)
<u +J
Z C
<D
i-H S)
x; <u
l-> CO
c
o -•
S rt
§So
0)
u
3
M
CVJ — O0>C0NU>lO ro
4-32
Tabic 44. Mean Wave Height at Coaslal Localities of Conterminous United States
o
o
O)
CC
i_
o
CD
>-
I
01
O)
en
o
O
>
to
o
o
o
e
o
X
>
o
o
c
CO
c
o
3
J3
to
3
C7>
00 to 't CM
4-34
Section 3.23 and by Kinsman (1965), National Academy of Sciences (1963),
and Neumann and Pierson (1966).
fH>h.) = ^ (4-6)
H = H . + a . (4-7)
1.61 H —0.61 H
s s
4-35
When data are not available for a specific beach, the wave climate
can be estimated by extrapolating from another location, after correcting
for differences in coastal exposure, winds, and storms.
On the east, gulf, and Great Lakes coasts, local winds are often
highly correlated with the direction of longshore currents. Such wind
data are available in "Local Climatological Data" sheets published monthly
by the National Weather Service, National Oceanographic and Atmospheric
Agency (NOAA) for about 300 U.S. weather stations. Other NOAA wind-data
sources include annual summaries of the Local Climatological Data by sta-
tion (Local Climatological Data with Comparative Data), and weekly sum-
maries of the observed weather (Daily Weather Maps), all of which can be
ordered from the Superintendent of Documents, U.S. Government Printing
Office, Washington, D.C. 20402.
4-36
4.35 EFFECT OF EXTREME EVENTS
Most storms move large amounts of sand from the beach offshore, but
after the storm, the lower waves that follow tend to restore this sand to
the face of the beach. Depending on the extent of restoration, the storm
may result in little permanent change. Depending on the path of the storm
and the angle of the waves, a significant amount of material can also be
moved alongshore. If the direction of longshore transport caused by the
storm is opposite to the net direction of transport, the sand will prob-
ably be returned in the months after the storm and permanent beach changes
effected by the storm will be small. If the direction of transport before,
during, and after the storm is the same, then large amounts of material
could be moved by the storm with little possibility of restoration. Suc-
cessive storms on the same beach may cause significant transport in oppo-
site directions, (e.g. Everts, 1973.)
There are some unique events that are only accomplished by catastro-
phic storms. The combination of storm surge and high waves allows water
to reach some areas not ordinarily attacked by waves. These extreme con-
ditions may result in the overtopping of dunes and in the formation of
washover fans and inlets. (Morgan, et al., 1958; Nichols and Marston,
1939.) Some inlets are periodically reopened by storms and then sealed
by littoral drift tremsported by normal wave action.
4-37
The wave climate at a particular beach also determines the effect
a storm will have. In a high-energy climate, storm waves are not much
larger than ordinary waves, and their effects may not be significant. An
example of this might be northeasters occurring at Cape Cod. In a low-
energy wave climate, where transport volumes are usually low, storm waves
can move significant amounts of sand, as do hurricanes on the gulf coast.
In October 1963, the worst storm in the memory of the Eskimo people
occurred over an ice-free part of the Arctic Ocean, and attacked the coast
near Barrow, Alaska. (Hume and Schalk, 1967.) Detailed measurements of
some of the key coastal areas had been made just before the storm. Freeze-
up just after the storm preserved the changes to the beach until surveys
could be made the following July. Most of the beaches accreted 1 to 2
feet, although Point Barrow was turned into an island. According to Hume
and Schalk, "The storm of 1963 would appear to have added to the Point the
sediment of at least 20 years of normal longshore transport." Because of
the low-energy wave climate and the short season in which littoral pro-
cesses can occur at Barrow, this storm significantly modified the beach.
A study of two hurricanes, Carla (1961) and Cindy (1963), was made by
Hayes (1967). He concluded that "the importance of catastrophic storms
as sediment movers cannot be over-emphasized," and observed that, in low-
energy wave climates, most of the total energy is expended in the near-
shore zone as a series of catastrophes. In this region, however, the rare
"extreme" hurricane is probably not as significant in making net changes
as the more frequent moderate hurricanes.
4-38
A hurricane (unnamed) coincided with spring tide on the New England
coast on 21 September 1938. Property damage and loss of life were both
high. A storm of this magnitude was estimated to occur about once every
150 years. A study of the beach changes along a 12-mile section of the
Rhode Island coast (Nichol and Marsten, 1939) showed that most of the
changes in the beach profile were temporary. The net result was some
cliff erosion and a slight retrogression of the beaches.
There is only slight exchange of fluid between the offshore and the
surf zone. Onshore -offshore flows take place in a number of ways, which
at present are not fully understood.
4-39
of the breaker crest. Since it is this wave-induced water particle motion
that causes the sediment to move, it is useful to know the length of the
elliptical path traveled by the water particles and the maximum velocity
and acceleration attained during this orbit.
The basic equations for water-wave motion before breaking are dis-
cussed in Chapter 2. Quantitative estimates of water motion are possible
from small -amplitude wave theory (Section 2.23), even near breaking where
assiomptions of the theory are not valid. (Dean, 1970; Eagleson, 1956.)
Equations 2-13 and 2-14, in Section 2.234 give the fluid-particle velocity
components u, w in a wave where small -amplitude theory is applicable.
(See Figure 2-3 for relation to wave phase and water particle accelera-
tion.)
HT^gd
2A = -^^^ , C4-10)
The term under the radical is the wave speed in shallow water.
FIND ;
(d) Compare the maximum horizontal velocity ^cuc» with the wave
speed, C.
4-40
solution:
HT Vgd^
2A =
2ffd
27r(2)
2d
1^32.2(2)^
umax = TTT;
T '^^^
= per second
2.0 feet ^
2(2)
—
2A =
L
3.2
40.1
= 0.08 .
^max 2.0
-^^^ = = 0.25.
C 8.0
*************************************
Although small -amplitude theory gives a fair understanding of many
wave-related phenomena, there are important phenomena that it does not
predict. Observation and a more complete analysis of wave motion show
that particle orbits are not closed. Instead, the water particles
advance a little in the direction of the wave motion each time the wave
4-41
passes. The rate of this advance is the mass transport vetoaity; (Equa-
tion 2-55, Section 2.253). This velocity becomes important for sediment
transport, especially for sediment suspended above ripples seaward of the
breaker.
("wax/ J))
"(-d)='-^' . (4-12)
where C is the wave speed given by Equation 2-3. Equation 2-55, and
thus Equation 4-12, does not include allowance for return flow which
must be present to balance the mass transported in the direction of wave
travel. In addition, the actual distribution of the time-averaged net
velocity depends sensitively on such external factors as bottom character-
istics, temperature distribution, and wind velocity. (Mei, Liu, and Carter,
1972.) Most observations show the time-averaged net velocity near the
bottom is directed toward the breaker region from both sides. (See Inman
and Quinn, (1952), for field measurements in surf zone; Galvin and Eagle-
son, (1965) for laboratory observations; and Mei, Liu and Carter (1972,
p. 220), for comprehensive discussion.) However, both field and labora-
tory observations have shown that wind-induced bottom currents may be
great enough to reverse the direction of the shoreward time-averaged
wave-induced velocity at the bottom when there are strong onshore winds.
(Cook and Gorsline, 1972; and Kraai, 1969.)
During most of the wave cycle in shallow water, the particle velo-
city is approximately horizontal and constant over the depth, although
right at breaking there are significant vertical velocities as the water
is drawn up into the crest of the breaker. The maximum particle velocity
under a breaking wave is approximated by solitary wave theory (Equation
2-66) to be
where (H+d) is the distance measured from crest of the breaker to the
bottom.
4-42
steepness (height-to-length ratio) decreases. (Galvin, 1967.) Of the
four breaker types, spilling breakers most closely resemble the solitary
waves whose speed is described by Equation 4-13. (Galvin, 1972.) Spill-
ing breakers differ little in fluid motion from unbroken waves (Divoky,
LeMehaute, and Lin, 1970), and thus tend to be less effective in trans-
porting sediment than plunging or collapsing breakers.
The most intense local fluid motions are produced by plunging break-
ers. As the wave moves into shallower depths, the front face begins to
steepen. When the wave reaches a mean depth about equal to its height,
it breaks by curling over at the crest. The crest of the wave acts as a
free-falling jet that scours a trough into the bottom. At the same time,
just seaward of the trough, the longshore bar is formed, in part by sedi-
ment scoured from the trough and in part by sediment transported in rip-
ples moving from the offshore.
4-43
Scripps Canyon
2 December 1948
Wave Period 15 Seconds
Wave from WNW
-.25
.25-.50KN
.50- 1.0 KN
>l Knot
— Hh = 4.5
k
n
s^^"
(from Shepard and Inman, 1950)
4-44
4.432 Diffuse Return Flow . Wind- and wave-induced water drift, pres-
sure gradients at the bottom due to setup, density differences due to
suspended sediment and temperature, and other mechanisms produce patterns
of motion in the surf zone that vary from highly organized rip currents
to broad diffuse flows that require continued observation to detect. Dif-
fuse return flows may be visible in aerial photos as fronts of turbid
water moving seaward from the surf zone. Such flows may be seen in the
photos reproduced in Sonu (1972, p. 3239).
4.441 Velocity and Flow Rate . Longshore currents flow parallel to the
shoreline, and are restricted mainly between the zone of breaking waves
and the shoreline. Most longshore currents are generated by the long-
shore component of motion in waves that obliquely approach the shoreline.
4-45
Figure 4-14. Typical Rip Currents, Ludlam
Island, New Jersey
4-46
mostly on breaker height. The outer edge of the surf zone is determined
by the breaker position. Since waves break in water depths approximately
proportional to wave height, the width of the surf zone on a beach in-
creases with wave height. This increase in width increases the cross
section of the surf zone.
2400 T T T
2000
10
c
o
Total of 5591 Observations
^o 1600
March -December 1968
>
w
O 1200
Xi 800
E
ij -3 -2 -I 1 2 3 4 5
Longshore Current Velocity, (feet per sec)
Longshore current velocity varies both across the surf zone (Longuet-
Higgins, 1970b) and in the longshore direction (Calvin and Eagleson, 1965).
Where an obstacle to the flow, such as a groin, extends through the surf
zone, the longshore current speed downdrift of the obstacle is low, but
it increases with distance downdrift. Laboratory data suggest that the
current takes a longshore distance of about 10 surf widths to become fully
developed. These same experiments (Calvin and Eagleson, 1965) suggest that
the velocity profile varies more across the surf zone at the start of the
flow than it does downdrift where the flow has fully developed. The ratio
of longshore current speed at the breaker position to longshore current
speed averaged across the surf zone varied from about 0.4 where the flow
started to about 0.8 or 1.0 where the flow was fully developed.
4-47
4.442 Velocity Prediction . The variation in longshore current velocity
across the surf zone and along the shore, and the uncertainties in vari-
ables such as the surf zone hydrography, make prediction of longshore
current velocity uncertain. There are three equations of possible use
in predicting longshore currents: Longuet-Higgins (1970); an adaptation
from Bruun (1963); and Galvin (1963). All three equations require co-
efficients identified by comparing measured and computed velocities, and
all three show about the same degree of agreement with data. Two sets of
data (Putnam, et al., 1949, field data; Galvin and Eagleson, 1965, labora-
tory data) appear to be the most appropriate for checking predictions.
a. Longuet-Higgins .
where
'2
0.694 r(2i3)
"^
M, = / (4-15)
y
According to Longuet-Higgins (1970a, p. 6788) , v^ is the longshore cur-
rent speed at the breaker position, r is a mixing coefficient which
ranges between 0.17 (little mixing) and 0.5 (complete mixing), but is
commonly about 0.2; g is the depth-to-height ratio of breaking waves
in shallow water taken to be 1.2 and fj? is the friction coefficient,
taken to be 0.01. Using these values, M,= 9.0.
4-48
'
o
a>
T3
a>
a>
o.
CO
o
C7>
o
0)
CO
O
a>
b. Bruun (1963 as Modified) .
where v^ is the mean velocity in the surf zone where the flow is fully
developed, and M2 involves a friction factor of the Chezy kind (see
Galvin, 1967, p. 297.)
c. Galvin (1963) .
where v^ is the mean velocity in the surf zone where the flow is fully
developed, and K is a coefficient depending on breaker height-to-depth
ratio and the ratio of trough depression on breaker height. To a good
approximation, K may be taken as 1.0. (Galvin and Eagleson, 1965.)
4.45 SUMMARY
4.51 INTRODUCTION
4-50
erosion, accompanied by longshore transport which has carried sand along
the coast to supply the barriers and spits extending across the bays.
The primary agent producing this erosion and transport is the action of
waves impinging on the shore.
Cb) What is the trend of shoreline migration over short and long
time intervals? (Needed for design of coastal structures, including
navigation channels.)
(c) How far seaward is sand actively moving? (Needed in the design
of sewage outfalls and water intakes.)
(e) What is the average shape, and the expected range of shapes,
for a given beach profile? (Needed for design of groins, beach fills,
navigation structures and flood protection.)
4-52
4.512 Zones of Transport . Littoral transport occurs in two modes: hed-
toad transport J the motion of grains rolled over the bottom by the shear
of water moving above the sediment bed; and suspended-load transport j the
transport of grains by currents after the grains have been lifted from
the bed by turbulence.
Both modes of transport are usually present at the same time, but it
is hard to distinguish where bedload transport ends and suspended-load
transport begins. It is more useful to identify two zones of transport
based on the type of fluid motion initiating sediment motion: the off-
shore zone where transport is initiated by wave-induced motion over rip-
ples, and the surf zone where transport is initiated by the passing break-
er. In either zone, net sediment transport is the product of two pro-
cesses: the periodic wave-induced fluid motion that initiates sediment
motion, and the superimposed currents (usually weak) which transport the
sediment set in motion.
As depth decreases to a value several times the wave height, the veloc-
ity distribution with time changes from approximately sinusoidal to a dis-
tribution that has a high shoreward component associated with the brief
passage of the wave crest, and lower seaward velocities associated with
the longer time interval occupied by the passage of the trough. As the
shoreward water velocity associated with the passing crest decreases and
begins to reverse direction over a ripple, a cloud of sand erupts upward
from the lee (landward) side of the ripple crest. This cloud of sand
drifts seaward with the seaward flow under the trough. At these shallow
depths, the distance traveled by the cloud of suspended sediment is two
or more ripple wavelengths, so that the sand concentration at a point
above the ripples usually exhibits at least two maximums during the pass-
age of the wave trough. These maximums are the suspension clouds shed by
the two nearest upstream ripples. The approach of the next wave crest
reverses the direction of the sand remaining suspended in the cloud. The
landward flow also drags material shoreward as bedload.
4-53
the subject of Section 4.4, include longshore currents and mass-transport
currents in the on shore -offshore direction. It is possible to have ripple
forms moving shoreward while residual currents above the ripples carry
suspended sediment clouds in a net offshore direction. Information on the
transport of sediment above ripples is given in Bijker (1970), Kennedy and
Locher (1972), and Mogridge and Kamphuis (1972).
b. Surf Zone. The stress of the water on the bottom due to turbu-
lence and wave- induced velocity gradients moves sediment in the surf zone
with each passing breaker crest. This sediment motion is both bedload
and suspended- load transport. Sediment in motion oscillates back and
forth with each passing wave, and moves alongshore with the longshore
current. On the beach face, the landward termination of the surf zone,
the broken wave advances up the slope as a bore of gradually decreasing
height, and then drains seaward in a gradually thinning sheet of water.
Frequently, the draining return flows in gullies and carries sediment to
the base of the beach face.
4-54
30
V
20
10
a. Shelf Profiles. The shelf profile is typically a smooth, concave-
up curve showing depth to increase seaward at a rate that decreases with
distance from shore. (bottom profile in Figure 4-18.) The smoothness of
the profile may be interrupted by other superposed geomorphic features,
such as linear shoals. (Duane, et al., 1972.) Data for shelf profiles
are usually obtained from charts of the National Ocean Survey (formerly,
U.S. Coast and Geodetic Survey).
The foreshore is that part of the beach extending from the highest
elevation reached by waves at normal high tide seaward to the ordinary
low water line. The foreshore is usually the steepest part of the beach
profile. The boundary between the backshore and the foreshore may be
the crest of the most seaward berm, if a berm is well developed. The
seaward edge of the foreshore is often marked by an abrupt step at low
tide level.
4-56
terrace (seaward from the foreshore, if the low-tide terrace is absent)
are the longshore troughs and longshore bars.
4.514 Profile Accuracy . Beach and nearshore profiles are the major
source of data for engineering studies of beach changes; sometimes lit-
toral transport can be estimated from these profiles. Usually, beach
and nearshore profiles are measured at about the same time, but differ-
ent tecliniques are needed for their measurement. The nearshore profile
is usually measured from a boat or amphibious craft, using an echo sounder
or leadline, or from a sea sled. (Kolessar and Reynolds, 1965-66; and
Reimnitz and Ross, 1971.) Beach profiles are usually surveyed by standard
leveling and taping techniques.
Closure error arises from the assumption that the outer ends of
nearshore profiles have experienced no change in elevation between two
successive surveys. Such an assumption is often made in practice, and
may result in significant error. An uncompensated closure error of 0.1
foot, spread over 1,000 feet at the seaward end of a profile, implies a
change of 3.7 cubic yards per time interval per foot of beach front where
4-57
the time interval is the time between successive surveys. Such a volume
change may be an important quantity in the sediment budget of the litto-
ral zone.
A fourth source of error comes from assuming that the measured beach
profiles (which are only an instantaneous picture), represent a long-term
condition. Actually, beach and nearshore profiles change rapidly in re-
sponse to changing wave conditions, so that differences between succes-
sive surveys of a profile may merely reflect temporary differences in
bottom elevation caused by storms and seasonal changes in wave climate.
Such fluctuations obliterate long-term trends during the relatively short
time available to most engineering studies. This fact is illustrated for
nearshore profiles by the work of Taney (1961a, Appendix B) who identified
and tabulated 128 profile lines on the south shore of Long Island that had
been surveyed more than once from 192 7 to 1956. Of these, 47 are on
straight shorelines away from apparent influence by inlets, and extend
from Mean Low Water (MLW) to about -30 feet MLW. Most of these 47 pro-
files were surveyed three or more times, so that 86 separate volume
changes are available. These data lead to the following conclusions:
(b) Gross volume changes (the absolute sums of the 86 volume changes)
are far greater than net voliome changes (the algebraic sums of the 86 vol-
lome changes). The gross volume change for all 86 measured changes is 8,113
cubic yards per foot; the net change is -559 cubic yards per foot (loss in
volume)
(c) The mean net change between surveys, averaged over all pairs of
surveys, is -559/86 or -6.5 cubic yards per foot of beach. The median
time between surveys is 7 years, giving a nominal rate of volume change
of about -1 cubic yard per year per foot.
4-58
500-
>>
to
>»
V
(jt>- -^
>
k.
3
c oo o
«
I I I I ll I I I I I I I
100 200
J.
.c 8
u
«
E
_3
O
>
c
13
500-
Time Between Surveys, months
( based on doto from Toney, 1961 a )
Figure 4-19. Unit Volume Change Versus Time Between Surveys for Profiles
on South Shore of Long Island. Data from Profiles Extending
from MSL to about the -30 depth Contour.
4-59
For coastal engineering, the accuracy of shelf profiles is usually
less critical than the accuracy of beach and nearshore profiles. Gener-
ally, observed depth changes between successive surveys of the shelf do
not exceed the error inherent in the measurement. However, soundings
separated by decades suggest that the linear shoals superposed on the
profile do show small but real shifts in position. (Moody, 1964, p. 143.)
Charts giving depths on the continental shelves may include soundings
that differ by decades in date.
Sand level changes in the beach and nearshore zone may be measured
quite accurately from pipes imbedded in the sand. (Inman and Rusnak, 1956;
Urban and Galvin, 1969; and Gonzales, 1970.)
For typical beach sediment, size is the only property that varies
greatly. However, quantitative evaluation of the size effect is usually
lacking. A gross indication of a size effect is the accumulation of
coarse sediment in zones of maximum wave energy dissipation, and deposi-
tion of fine sediment in areas sheltered from wave action. (e.g. King,
1972, pp. 302, 307, 426.) Sorting by size is common over ripples (Inman,
1957) and large longshore bars (Saylor and Hands, 1970). Field work on
size effects in littoral transport does not permit definite conclusions.
(King, 1972, p. 483; Inman, Komar, and Bowen, 1969; Castanho, 1970; Ingle,
1966, Figure 112; Yasso, 1962; and Zenkovich, 1967a.)
4-60
Durability (resistance to abrasion, crushing, and solution) is usu-
ally not a factor within the lifetime of an engineering project. (Kuenen,
1956; Rusnak, Stockman, and Hofmann, 1966; and Thiel, 1940.) Possible
exceptions may include basaltic sands on Hawaiian beaches (Moberly, 1968),
some fragile carbonate sands which may be crushed to finer sizes when sub-
ject to traffic, (Duane and Meisburger, 1969, p. 44), and carbonate sands
which may be soluble imder some conditions. (Bricker, 1971.) In general,
recent information lends further support to the conclusion of Mason (1942)
that, "On sandy beaches the loss of material ascribable to abrasion . . .
max ( j-> TT , ^
y '^>
,
(4-19)
H sinh 2^<y^
which is plotted in dimensionless form in Figure 4-20 and for common values
of wave period in Figure 4-21. Figures 4-20 and 4-21 give maximum bottom
particle velocity, ^imax{-d) > ^^ ^ function of depth, wave period, and
local wave height. This prediction by linear theory for the offshore zone,
and the related results from solitary-wave theory for the zone near break-
ing, agree fairly well with measurement in the field. (Inman and Nasu,
1956; and Cook and Gorsline, 1972, Figures 5 and 6.).
4-61
I.Oi)|- II 1 !
1
; t 1
'Q
0.8
0.6
0.5
0.4
0.3
0.2
0.1
0.08
0.061-
J>
0.05
- 0.04g
o 0.03
>
I 0.02
0.01
0.0081-
0.006
0.005
0.004
0.003 =
0.002
0.001
5000
4000
3000
18 20 22 24
(dimensionless)
4-63
************** EXAMPLE PROBLEM ***************
GIVEN: A wave in depth d = 200 feet, with period T = 9 seconds, and a
maximum bottom velocity umaxr-d) 2: 1.0 foot per second.
FIND : The minimxim wave height that will create the given bottom velocity.
SOLUTION : Calculate
Lo =
lit
= 5.12(9)2
= 415 feet .
and
200
Lo 415
= 0.482
u T
= 0.30
H ,
"ma.C(_jjT
H
0.30
H =(l)(2) = 3ofeee.
0.30
Thus a 30-foot minimum wave height with a 9-second wave period is needed
to create a bottom velocity equal to or greater than 1 foot per second
in 200 feet of water. Alternatively, a curve for a 9-second period can
be interpolated in Figure 4-21 and uma«(-tf)T/H can be read from the
curve's intersection with the 200-foot depth.
*************************************
As a wave moves into shallower water, both bottom velocity and size
of water-particle orbit increase. For some low velocities where the hori-
zontal orbit is small, sand moves as individual grains rolling across the
surface, but for most conditions, once quartz sand begins to move, ripples
form (Kennedy and Falcon, 1965; Carstens, et al., 1969; Cook, 1970). Thus,
the initiation of sediment motion is usually indicated by the formation
of sediment ripples.
4-64
Figure 4-22, from Inman (1957) and including data from two earlier
studies, shows the velocity needed to start motion in a sediment bed of
a given grain size. These results are in general agreement with other
studies relating critical velocity to grain size. Also shown in this
figure are maximum velocities above which ripples tend to be smoothed
off, in qualitative agreement with conditions for bed forms in unidirec-
tional flows. (Southard, 1972.)
a. Bathymetry. Dietz (1963) and others point out that waves rework
nearshore sands, smoothing out irregularities by longshore and onshore-
offshore transport. This smoothing produces a quasi -equilibrium surface
in the nearshore zone which forms relatively straight contours, nearly
parallel to the shoreline.
Following the idea of Dietz (1963), the depth below which shore-
parallel contours give way to irregular contours is assumed to mark the
local transition between the nearshore zone where sands are moved by the
waves in significant quantities and the offshore zone where sand is moved
in lesser quantities. Possible exceptions to this shore-parallel contour
rule are the contours around river deltas and inlets, or where reefs and
ledges occur in the nearshore zone.
4-65
o
•H
4-1
^
(1)
rH
CX
ex.
•H
m
o
C
o
•H
•H
4J
•H
C
CM
CM
01
U
3
(08S/U) Tl /^ilooiSA
4-66
1000
' 1
i
—
r-
800 E
600
500 : :.. .
400
300
200
100
80
- 60
S 50
"'-
.C
40
i- 30
20
10
6
5
4
2
e5'>4e
30'I0
2000 4000
Bathymetry, such as that in Figures 4-24 and 4-25, suggests that the
depth to the deepest shore-parallel contour is usually constant along the
shore for distances of several miles, but that this depth may vary with
longshore distances of about 10 miles. (See Figure 4-25.) The depth to
the deepest shore-parallel contour may depend on the contour spacing, but
this is not important if contour intervals are small relative to the total
depths involved. In general, the deepest shore-parallel contour is between
15 and 60 feet. In most localities, this depth is somewhat deeper than
that at which nearshore profiles are presumed to close-out. These near-
shore contours probably reflect longer term adjustment to extreme storms
that occur rarely during the typical time interval between repetitive
nearshore profile surveys.
1965) suggest that littoral sands grade seaward into finer materials before
the relatively coarse sands of the shelf are reached. In some places the
boundary between the coarser shelf sediment and this finer material is quite
sharp. (Pilkey and Frankenberg, 1964.) The finer material is currently
4-68
4-69
interpreted as bounding the seaward edge of sediment moved by waves in sig-
nificant quantities. This band of finer material suggests that there is
little exchange between littoral and shelf sands in most places.
Sand transport from the nearshore zone is more likely. Surveys show
that sand in the nearshore zone does move, although it is difficult to meas-
ure direction of motion. The presence of shore-parallel contours along most
open shores is evidence that the waves actively mold the sand bottom in the
nearshore zone. However, the time scale of transport in the nearshore zone
may be relatively slow.
4-70
notable rapid rearrangement of a profile is by storm waves, especially
during storm surge (Section 3.8) which enables the waves to attack at
higher elevations on the beach, (see Figure 1-7.)
The part of the beach washed by runup and runback is the beach face.
Under normal conditions, the beach face is contained within the fore-
shore, but during storms, the beach face is moved shoreward by the cut-
ting action of the waves on the profile. The waves during storms are
steeper, and the runback of each wave on the beach face carries away more
sand than is brought to the beach by the runup of the next wave. Thus
the beach face migrates landward, cutting a scarp into the berm. (See
Figure 1-7.)
In moderate storms, the storm surge and accompanying steep waves will
subside before the berm has been significantly eroded. In severe storms, or
after a series of moderate storms, the backshore may be completely eroded,
after which the waves will begin to erode the coastal dunes, cliffs, or main-
land behind the beach.
Duration and storm surge (Columns 9 and 10) are consistent with storm
damage data (Columns 11 through 16, although the effect is influenced by
the choice of profiles included in each study. The December 1970 storm
4-71
includes all profiles surveyed. The two February 1972 storms at Jones
Beach include all profiles away from the influence of inlets, reducing
the number of profiles from 15 in the December 1970 storm to 10 in the
February 1972 storms. The November 1953 storm gives relatively high
storm damage which may partly result from the long time interval between
pre-storm survey and the storm. (See Columns 1 and 3.) These results
are also affected by the fact that they omit some profiles "believed to
be influenced by the presence of a seawall or a bulkhead," (Caldwell,
1959, p. 4.)
Although the data in Table 4-5 are not completely comparable, the
results do suggest that the average volume of sand eroded above mean sea
level from beaches about 5 or more miles long has a certain range of val-
ues. A moderate storm may remove 4 to 10 cubic yards per foot of beach
front above MSL; an extreme storm (or a moderate storm that persists for
a long time) may remove 10 to 20 cubic yards per foot; rare storms that
are most destructive in beach erosion due to a combination of intensity,
duration, and orientation may remove 20 to 50 cubic yards per foot. These
values are average for beaches 5 to 10 miles or more long, and they are
compatible with other, less complete, data for notable storms. (Caldwell,
1959; Shuyskiy, 1970; and Harrison and Wagner, 1964.) For comparative
purposes, a berm 100 feet wide at an elevation of 10 feet MSL contains 37
cubic yards per foot of beach front, a quantity that would be adequate
except for extreme storms.
Extreme values of erosion may be more useful than mean values for
design. Column 17 of Table 4-5 suggests that the ratio of the most ero-
ded above-MSL-profile to the average profile for east coast beaches ranges
from about 1.5 to 6. If the average erosion per profile is based only on
those profiles showing net erosion, then this ratio is probably between
1.5 and 3.
The storm surveys also show that the shoreline on many beaches may
prograde seaward even though the profile as a whole loses volume, or
vice versa. This possibility suggests caution in interpreting aerial
photos of storm damage. (Everts, 1973.)
4-74
b. Beach Recovery . The typical beach profile left by a severe storm
is a simple, con cave -upward curve extending seaward to low tide level or
below, (See top of Figure 4-26.) The sand that has been eroded from the
beach is deposited mostly as a ramp or bar in the surf zone that exists at
the time of the storm. Immediately after the storm, beach repair begins
by a process that has been documented in detail, (e.g., Hayes, 1971; Davis,
et al., 1972; Davis and Fox, 1972; and Sonu and van Beek, 1971.) Sand that
has been deposited seaward of the shoreline during the storm begins moving
landward as a sandbar with a gently sloping seaward face and a steeper land-
ward face. (See Figure 4-26.) These bars have associated lows (riinnels)
on the landward side and occasional drainage gullies across them. (King,
1972, p. 339.) These systems are characteristic of post-storm beach accre-
tion under a wide range of wave, tide, and sediment conditions. (Davis,
et al., 1972.) They are sometimes called ridge -and- runnel systems.
The rate at which the berm builds up or the bar migrates landward to
weld onto the beach varies greatly, apparently in response to: wave con-
ditions, beach slope, grain size, and the length of time the waves work
on the bars. (Hayes, 1971.) Compare the slow rate of accretion at Crane
Beach in Figure 4-26 (mean tidal range 9 feet, spring range 13 feet) with
the rapid accretion on the Lake Michigan shore in Figure 4-2 7 (tidal range
less than 0.25 foot).
4-75
Summer Accretion 29 May -7 September 1967
Station CBA, Crane Beach
Ipswich, Massachusetts
^»
-29 Moy
\ -I 3 June
\
^ __M_LW_
**
— I July
\ .
>•
,\ I July
-20 July
\
-24 July
-"3 Aug
\ ^.
_8 Aug
14 Aug
\
\
\ 22 Aug
10 n 30 Aug
;
5 —
'•
7Sept„LW
4-76
o
The ideal result of post-storm beach recovery is a wide backshore
that will protect the shore from the next storm. Beach recovery may be
prevented when the period between successive storms is too short. Main-
tenance of coastal protection requires knowledge of the necessary width
and elevation of the backshore appropriate to local conditions, and
adequate surveillance to determine when this natural sand reservoir is
diminished to a point where it may not protect the backshore during the
next storm.
4.525 Bar-Berm Prediction . High, steep waves scour the beach, eroding the
foreshore into a simple con cave -upwards profile. The material eroded from
the beach is deposited offshore as a longshore bar. Waves of low steepness
tend to push sand onto the beach, usually as migrating longshore-bar sys-
tems which eventually become part of the beach. In contrast to the concave-
up eroded profile discussed previously, the accreted profile is concave-
downward. Idealized eroded and accreted profiles (measured in a prototype-
scale wave tank) across the beach and nearshore zone are shown in Figure
4-28.
creted profile and that a bar characterizes an eroded profile. (See Figure
4-28.) Tliis picture is somewhat idealized. A sharp berm crest between
backshore and foreshore is often lacking, and on some beaches the berm is
absent, so that the top of the foreshore reaches the dune or cliff line.
Berms are illustrated in Figures 4-1 and 4-27.
4-78
, .
developed in tideless seas such as Lake Michigan (Fig. 4-27). (Saylor
and Hands, 1970.) Keulegan (1948) found that the ratio of depth of long-
shore trough to depth of bar was approximately 1.69 in laboratory experi-
ments, but most field measurements showed less depth difference, averag-
ing about 1.23 (based on MLLW) for 276 measurements from the Scripps pier.
(Shepard, 1950.) According to Shepard, bars are not significant on slopes
steeper than 4° (1 on 14).
4-80
c. Dimensionless Fall Time. Prediction of accreted (berm) or erod-
ed (bar) profiles is possible using a dimensionless fall-time parameter
Ho
F^
O
= ,
(4-20)
VrT
Since values of V^ range from about 0.066 foot per second (2 centi-
meters per second) for fine sand to 0.49 foot per second (15 centimeters
per second) for coarse sand (Fig. 4-31), ¥q ranges from about 0.25 for
low swell on coarse sand beaches to 10 or more for storm waves on fine
sand beaches. Such values suggest that typical field and laboratory con-
ditions define a range of conditions within which the importance of sus-
pended sediment in the surf zone may vary from significant to negligible.
4-81
0Q t^^^H?^:r':T[-i-t-^4-::||;::ra i -<i'r: l i
ll;j» !
0.001 0.002 0.004 0.01 0.02 0.04 0.07 0.1 0.2 0.4 0.7 I
4-82
\KJKJ
a|039 dAdi9 pjopuois Jd|Aj^
^m (Dr^ooOkOCM^too^wcMinciifloOin
— — — — cvj<Mc\jm»o^^io*o
4.526 Slope of the Foreshore . The foreshore is the steepest part of the
beach profile. The equilibrium slope of the foreshore is a useful design
parameter, since this slope, along with the berm elevation, determines
minimum beach width.
The slope of the foreshore tends to increase as the grain size in-
creases. (U.S. Army, Beach Erosion Board, 1933; Bascom, 1951; and King,
1972, p. 324.) This relationship between size and slope is modified by-
exposure to different wave conditions (Bascom, 1951; and Johnson, 1956);
by specific gravity of beach materials (Nayak, 1970; and Dubois, 1972);
by porosity and permeability of beach material (Savage, 1958); and prob-
ably by the tidal range at the beach. Analysis by King (1972, p. 330)
suggests that slope depends dominantly on sand size, and also signifi-
cantly on an unspecified measure of wave energy.
The inverse relation between slope and wave height is partly caused
by the relative frequency of steep or high eroding waves which produce
gentle foreshore slopes and low accretionary post-storm waves which pro-
duce steeper beaches. (See Figures 4-1, 4-26, and 4-2 7.)
The relation between foreshore slope and grain size shows greater
scatter in the laboratory than in the field. However, the tendency for
slope of the foreshore to increase with decreasing mean wave height is
supported by laboratory data of Rector (1954, Table 1). In this labora-
tory data, there is an even stronger inverse relation between deepwater
steepness, Wq/Lo, and slope of the foreshore than between Wq a^d the
slope.
4-85
adois
lO
o
CVJ
CO
"~
adois
o
OJ
lO
00
d
o
(c) Slope of the foreshore tends to decrease with increasing wave
height, again with scatter.
sand) moved in the littoral zone imder action of waves and currents. The
rate, Q, at which littoral drift is moved parallel to the shoreline is
the longshore transport rate. Since this movement is parallel to the
shoreline, there are two possible directions of motion, right and left,
relative to an observer standing on the shore looking out to sea. Move-
ment from the observer's right to his left is motion toward the left, in-
dicated by the subscript It. Movement toward the observer's right is
indicated by the subscript rt.
%=%t- Q£. •
C4-22)
7 = —
^1rt
, (4-23)
(1 +7)
—
Q„ = Q„
^^
(4-24)
4-88
Longshore transport rates are usually given in units of volume per
time (cubic yards per year in the U.S.). Typical rates for oceanfront
beaches range from 10^ to 10^ cubic yards per year. (See Table 4-6.)
These volume rates typically include about 40 percent voids and 60 per-
cent solids.
At present, there are four basic methods to use for the prediction
of longshore transport rate:
2. If rates from nearby sites are unknown, the next best way to
predict transport rates at a site is to compute them from data showing
historical changes in the topography of the littoral zone (charts, sur-
veys, and dredging records are primary sources).
4.532 Energy Flux Method . Method 3 is based on the assumption that long-
shore transport rate, Q, depends on the longshore component of energy flux
in the surf zone. The longshore energy flux in the surf zone is approxi-
mated by assuming conservation of energy flux in shoaling waves, using
small -amplitude theory, and then evaluating the energy flux relation at
the breaker position. The energy flux per unit length of wave crest, or,
equivalently, the rate at which wave energy is transmitted across a plane
of unit width perpendicular to the direction of wave advance is (from
Section 2.238, combining Equations 2-39 and 2-40):
P = EC = ^ H2 C . (4-25)
4-89
Table 4-6. Longshore
If the wave crests make an angle, a with the shoreline, the energy flux
in the direction of wave advance per unit length of beaah is
P cos a = —— C
PgH^
cos a , (4-26)
8 *
—
?£ = P cos a sin a = Pg
-^ H 1 C cos a sin a ,
8 *
Pp = 77 H^ C sin 2a . (4-27)
* lo *
= (4-28)
Pfi.
ll^S '^2ai, .
where H^j and a^ are the wave height and direction at breaking and C
is the wave speed from Equation 2-3, evaluated in a depth equal to 1.28
Hi.
4-91
4" 6" 8' 10' 20° 60° 80° 100°
Oeepwater Wave Angle, Qo, degrees
4-92
************** EXAMPLE PROBLEM ***************
GIVEN A breaking wave with height, Hj, = 4 feet; period T = 7 seconds.
:
FIND ;
(b) The angle the wave made with the shoreline when it was in deep
water, Uq.
SOLUTION : Calculate
Hfe 4.0
-— = r = 0.00254 .
Enter Figure 4-34 to the point where the line labeled aj, = 6" crosses
H2,/gT2 = 0.00254 and read Pji/pg^ "^ = 6.7 x lO"** from the left axis
"f
and a(-) = 18° from the bottom axis.
Pg = 0.00067 pg2 Hi T ,
and
«o = 18° •
*************************************
For offshore conditions, the group velocity is equal to one-half the
deepwater wave speed C^, where C^ is given by Equation 2-7, and the
refraction coefficient, K^, can be determined by the methods of Section
2.32. Hence,
Figure 4-35 also presents the longshore component of wave energy flux
in a dimensionless form, P£,/(pg^ H^ T), as a function of deepwater wave steep-
ness, H^/gT^, and the angle the wave crest makes with the shoreline in
either deep water, a^, or at the breaker line, a^,. Refraction by
straight parallel bottom contours is again assumed. As illustrated by
4-93
E
oo
ID
(i,°H,H/''d
4-94
the example problem below. Figure 4-35 can be used to determine the
longshore component of wave energy flux when deepwater wave height, H^,
period, T, and deepwater wave angle, a^, are known.
FIND:
(b) The angle the wave makes with the shoreline when it breaks [assum-
ing refraction is by straight, parallel bottom contours).
SOLUTION: Calculate,
Ho 5.0
= 0.0032
gT^ 32.2 (7.0)
Enter Figure 4-35 with a^, = 25° and H^/gT^ = 0.0032, read P^/pg^H^ T =
1.5 X 10"^. This corresponds with a breaker angle a^, = 9.5° which is
obtained by interpolation between the dashed curves of constant aj,.
Tlierefore,
Pj = (0.0015)pg^H2T ,
and
afc = 9.5° .
**************************************
Equations 4-25 and 4-28 are valid only if there is a single wave
train with one period and one height. However, most ocean wave condi-
tions are characterized by a variety of heights with a distribution usu-
ally described by a Rayleigh distribution. (See Section 3.22.) For a
Rayleigh distribution, the correct height to use in Equation 4-28 or in
the formulas shown in Table 4-8 is the root-mean-square amplitude. How-
ever, most wave data are available as significant heights, and coastal
engineers are used to dealing with significant heights.
4-95
value of the exact energy flux for sinusoidal wave heights with a Rayleigh
distribution. Since this means that P^g is proportional to energy flux
and not equal to it, P^s is referred to as the longshore energy flux
factor in Table 4-8 and the following sections.
For the purpose of this section, it is assumed that the shoaling co-
efficient. Kg, for nearshore breaking waves is equivalent to the breaker
height index, H^^/Ho', found from observation. (See Figure 2-65.)
4-96
Table 4-7. Longshore Energy Flux.Pg , for a Single Periodic Wave in
Any Specified Depth. (Four Equivalent Expressions from
Small-Amplitude Theory)
Equation
Table 4-9. Assumptions for Pj Formulas in Table 4-8.
E = (pg4)/8 « SHg
b. Group velocity equals wave speed at breaking, and breaking speed is given by
solitary wave theory according to the approximation. (Galvin, 1967, Equation 11.)
c. a can be replaced by ay
a. Same as lb above
if (cosaj,)'/^ = 1.0
a. Same as la above
b. Same as lb above
c. Same as 3a above
d. Cos ttjj = 1.0
NOTE: Constants evaluated for foot-pound-second units. Small-amplitude theory is assumed valid in
deep water. Nearshore contours are assumed to be straight and parallel to the shoreline.
4-98
10'
I
t• •
10'
o
O
ce ,0
rio^
o
a.
10
-1 0'
10" 10 10 10' 10"
Figure 4-36. Longshore Transport Rate Versus Energy Flux Factor for Field
and Lab Conditions
4-99
10^
a.
ro
T5
10*
o
O
o
(A
.c
I/)
c
o
10'
changes may reduce the value of P^g when deepwater wave height statis-
tics are used as a starting point for computing P^g. (Walton, 1972;
Bretschneider and Reid, 1954; and Bretschneider, 1954.)
The Equations 4-35 through 4-38 in Table 4-8 are related to the Equa-
tion for Eq previously recommended for use with this method (Caldwell,
1956, Equations 5 and 6; or the equations in Figure 2-22, page 175, CERC
Technical Report No. 4, 1966 edition) by a constant
The relation between Q and P^g in Figures 4-36 and 4-37 can be
approximated by
Q = (7.5X103)Pj^ (4-40)
4-101
transport rate, only the (H^j, a^) data and Figure 4-38, or the (H^, a^)
data and Figure 4-39 are needed. If the shoaling coefficient is signifi-
cantly different from 1.3, multiply the Q obtained from Figure 4-39 by
the factor 0.88 vlCg. (See Table 4-9, Assumption 2d.)
4-102
Lij 1
Qo iQO 20° 30° 40° 50° 60° 70° 80° 90°
Deepwater Angle, Of
q > ( degrees )
4-104
Table 4-10. Deepwater Wave Heights, in Percent by Direction, off East-Facing
Coast of Inland Sea
Compass Direction
sector of wave directions. If ^(.a^) is evaluated at a^ = 45° (NE or
SE in the example problem), it will have a value 12 percent higher than
the average value for FCa^) over 45° sector bisected by the NE or SE
directions. Thus, if the data warrant a high degree of accuracy. Equa-
tion 4-43 should be averaged by integrating over the sector of directions
involved.
4-107
Galvin (1972) showed that when field values of longshore transport
rate are plotted against mean annual breaker height from the same locality,
a curve
Q = 2 X 10= H^ ,
(4-44)
forms an envelope above almost all known pairs of (Q, Hj,) , as shown in
Figure 4-40. Here, as before, Q is given in units of cubic yards per
year; Hjy in feet.
Figure 4-40 includes all known (Q, Hj,) pairs for which both Q and
H2j are based on at least 1 year of data, and for which Q is considered
to be the gross longshore transport rate, Q^, defined by Equation 4-21.
Since all other known (Q, H^,) pairs plot below the line given by Equation
4-41, the line provides an upper limit on the estimate of longshore trans-
port rate. From the defining equations for Q^ and Q^, any line that
forms an upper limit to longshore transport rate must be the gross trans-
port rate, since the quantities Qrt> Q£t> ^^^ Qn» ^^ defined in Section
4.531, are always less than or equal to Q^^.
4-108
0^
a> 10
o.
cr.
o
CO
CO 10
c
o
10
Table 4-12. Example Estimate of Gross Longshore Transport Rate for Shore of Inland Sea
(1)
Table C-1 Appendix C) evaluated at the breaker position and i^s^g i^
the shoaling coefficient evaluated at the wave gage:
H. - H (4-45)
In this example, wave gage statistics have been used for illustrative
purposes. However, visual observations of breakers, such as those listed
in Table 4-4, would be even more appropriate since Equation 4-44 has been
"calibrated" for such observations. On the other hand, hindcast statis-
tics would be less satisfactory than gage statistics due to the uncertain
effect of nearshore topography on the transformation of deepwater statis-
tics to breaker conditions.
4.61 BACKGROUND
4-III
X
o
OS
O
c
o
•H
t^
6
•H
X
o
u
0)
CO
o
ex
•H
OQ
o
•H
•H
4-112
Island crest elevation is determined by the nature of the sand form-
ing the beach, and by the waves and water levels of the ocean. The beach
and waves interact to determine the elevation of the limit of wave runup -
the primary factor in determining island crest elevation. Normally the
island crest elevation is almost constant over long sections of beach.
However, duneless barrier island crest elevations vary with geographical
area. For example, the crest elevation typical of Core Banks, North
Carolina, is about +6 feet MSL; +4 feet MSL is typical for Padre Island,
Texas; +11 feet MSL is typical for Nauset Beach, Massachusetts.
Landward of the upper limit of wave uprush or berm crest are the back-
shore and the deflation plain. This area is shaped by the wind, and in-
frequently by the flow of water down the plain when the island crest is
overtopped by waves. (e.g., Godfrey, 1972.) Obstructions which trap
wind-transported sand cause the formation of dunes in this area. (See
discussion in Section 6.4 Sand Dunes.) Beachgrasses which trap wind-
transported sand from the beach and the deflation plain are the major
agent in creating and maintaining foredunes.
Foredunes, the line of dunes just behind a beach, have two primary
functions in shore processes. First, they prevent overtopping of the
island during some abnormal sea conditions. Second, they serve as a
reservoir for beach sand.
able protection against wave overtopping. (See Section 5.3 and 6.3.)
4-113
O—
this source if the lagoon is less than 5 miles wide. Where the lagoon
is wider (especially 10 miles or greater) flooding from the lagoon side by
wind setup should be investigated before large dune construction projects
are undertaken.
4.622 Reservoir of Beach Sand . During storms, erosion of the beach occurs
and the shoreline recedes. If the storm is severe, waves attack and erode
the foredunes and supply sand to the beach; in later erosion stages, sand
is supplied to the back of the island by overwash. (Godfrey, 1972.)
Volumes of sand eroded from beaches during storms have been estimated
in recent beach investigations. Everts (1973) reported on two storms dur-
ing February 1972 which affected Jones Beach, New York. The first storm
eroded an average of 27,000 cubic yards per mile above mean sea level for
the 9-mile study area; the second storm (2 weeks later) eroded an average
of 35,000 cubic yards of sand per mile above mean sea level at the same
site. Losses at individual profiles ranged up to 120,000 cubic yards per
mile. Davis (1972) reported a beach erosion rate on Mustang Island, Texas,
following Hurricane Fern (September 1971), of 12.3 cubic yards per linear
foot of beach for a 1,500-foot stretch of beach (about 65,000 cubic yards
per mile of beach). On Lake Michigan in July 1969, a storm eroded an aver-
age of 3.6 cubic yards per linear foot of beach (about 29,000 cubic yards
per mile) from an 800-foot beach near Stevensville, Michigan. (Fox, 1970.)
Because much of the eroded sand is usually returned to the beach by wave
action soon after the storm, these volumes are probably representative of
temporary storm losses.
4-115
front major foredune systems in different geographical locations. For
example, fine sand beaches front a massive foredune system on Mustang
Island, Texas, and coarse sand beaches front dunes on the Cape Cod spits,
4.71 INTRODUCTION
4-116
Sediment budget studies have been presented by Johnson (1959), Bowen
and Inman (1966), Vallianos (1970), Pierce (1969), and Caldwell (1966).
Point sources or point sinks are sources or sinks that add or sub-
tract sand across a limited part of a control volume boundary. A tidal
inlet often functions as a point sink. Point sources or sinks are gener-
ally measured in units of volume per year.
Line sources or tine sinks are sources or sinks that add or subtract
sand across an extended segment of a control volume boundary. Wind trans-
port landward from the beaches of a low barrier island is a line sink for
the ocean beach. Line sources or sinks are generally measured in units of
volume per year per lonit length of shoreline. To compute the total effect
of a line source or sink, it is necessary to multiply this quantity by the
total length of shoreline over which the line source or sink operates.
Q^ is a point source
Q^ is a point sink
q^ is a line source
q^ is a line sink
4-117
Table 4-13. Classification of Elements in the Littoral Zone Sediment Budget
Location of
Source or Sink
4.713 Sediment Budget Boundaries .A sediment budget is used to identify
and quantify the sources and sinks that are active in a specified area.
By so doing, erosion or deposition rates are determined as the balance
of known sinks and sources. Boundaries for the sediment budget are deter-
mined by the area imder study, the time scale of interest, and study pur-
poses. In a given study area, adjacent sand budget compartments (control
volumes) may be needed with shore-perpendicular boundaries at significant
changes in the littoral system. For example, compartment boundaries may
be needed at inlets, between eroding and stable beach segments, and between
stable and accreting beach segments. Shore-parallel boundaries are needed
on both the seaward and landward sides of the control volumes. They may be
established wherever needed, but the seaward boundary is usually established
at or beyond the limit of active sediment movement, and the landward bound-
ary beyond the erosion limit anticipated for the life of the study. The
bottom surface of a control volume should pass below the sediment layer
that is actively moving, and the top boundary should include the highest
surface elevation in the control volume. Thus, the budget of a particular
beach and nearshore zone would have shore parallel boundaries landward of
the line of expected erosion and at or beyond the seaward limit of signifi-
cant transport. A budget for barrier island sand dunes might have a bound-
ary at the bay side of the island and the landward edge of the backshore.
4.721 Rivers . It is estimated that rivers of the world bring about 3.4
cubic miles or 18.5 billion cubic yards of sediment to the coast each year
(voliime of solids without voids). (Stoddard, 1969; from Strakhov, 1967.)
Only a small percentage of this sediment is in the sand size range that is
common on beaches. The large rivers which account for most of the volume
of sediment carry relatively little sand. For example, it is estimated
(Scruton, 1960) that the sediment load brought to the Gulf of Mexico each
year by the Mississippi River consists of 50 percent clay, 48 percent silt,
and only 2 percent sand. Even lower percentages of sand seem probable for
other large river discharges. (See Gibbs, 1967, p. 1218, for information
on the Amazon River.) But smaller rivers flowing through sandy drainage
areas may carry 50 percent or more of sand. (Chow, 1964, p. 17-20.) In
southern California, sand brought to the coast by the floods of small
rivers is a significant source of littoral material. (Handin, 1951; and
Norris, 1964.)
4-119
Offshore Sink
WATER
Qa
I
*- Q4
IN
r r
Sediment Budget Boundary Top of Cliff
LAND
Budget Boundory
ri Recession
\ \
I \
> \,*i-First Survey Profile
Most of the sediment carried to the coast by rivers is deposited
in comparatively small areas, often in estuaries where the sediment is
trapped before it reaches the coast. (Strakhov, 19670 The small frac-
tion of sand in the total material brought to the coast and the local
estuarine and deltaic depositional sites of this sediment suggest that
rivers are not the immediate source of sediment on beaches for much of
the world's coastline. Many sources of evidence indicate that sand-
sized sediment is not supplied to the coasts by rivers on most segments
of the U.S. Atlantic and Gulf coasts. Therefore, other sediment sources
must be important.
Figure 4-44 shows three equivalent volumes, all indicating a net ero-
sion of hAx„ To the right in Figure 4-44 is a typical beach profile.
The dashed line profile below it is the same as the solid line profile.
The horizontal distance between solid and dashed profiles is Ax, the
horizontal retreat of the profile due to (assumed) uniform erosion. The
unit volume loss, hAx, between dune base and depth to seaward limit is
equivalent to the unit volume indicated by the slanted parallelogram on
the middle of Figure 4-44. The unit volume of this parallelogram, hAx,
is equivalent to the shaded rectangle on the left of Figure 4-44. If the
vertical distance, h is 40 feet, and Ax = 1 foot of horizontal erosion,
then the unit volume lost is 40/27, or 1.5 cubic yards per foot of beach
front
4-121
«
erosion of the nearshore bottom and beaches, but in the control volume
defined, this transport takes place within the control volume. Transport
from the offshore has been treated as a line source.
4-123
beach fill is subject to uncertainties: the source of the dredged sand
often contains significant but variable quantities of finer materials that
are soon lost to the littoral zone; the surveys of both the borrow area
and the replenished area are subject to uncertainty because sediment trans-
port occurs during the dredging activities; and in practice only limited
efforts are made to obtain estimates of the size distribution of fill
placed on the beach. Thus, the resulting estimate of the quantity of
suitable fill placed on the beach is uncertain. More frequent sampling
and surveys could help identify this significant element in many sedi-
ment budgets.
4.732 Overwash . On low barrier islands, sand may be removed from the
beach and dune area by overwashing during storms. Such rates may average
locally up to 1 cubic yard per year per foot. Data presented by Pierce
(1969) suggest that for over half of the shoreline between Cape Hatteras
and Cape Lookout, North Carolina, the short term loss due to overwash was
0.6 cubic yard per year per foot of beach front. Figure 4-46 is an aerial
view of overwash in the region studied by Pierce (1969). Overwash does
not occur on all barrier islands, but if it does, it may function as a
source for the beach on the lagoon side.
from transport in the littoral zone as backshore deposits and dune areas
along the shore. Depending on the frequency of severe storms, such sand
may remain in storage for intervals ranging from months to years. Back-
shore deposition can occur in hours or days by the action of waves after
4-124
Figure 4-45. Sediment Trapped Inside Old Drim Inlet,
North Carolina
4-125
( I November 1971
4-126
storms. Dune deposits require longer to form - months or years - because
wind transport usually moves material at a lesser rate than wave transport.
I£ the immediate beach area is the control volume of interest, and budget
calculations are made based on data taken just after a severe storm, allow-
ance should be made in budget calculations for sand that will be stored in
berms through natural wave action. (See Table 4-5.)
4-127
( 14 September 1969)
4-128
4.736 Deflation . The loose sand that forms beaches is available to be
transported by wind. After a storm, shells and other objects are often
found perched on pedestals of sand left standing after the wind eroded
less protected sand in the neighborhood. Such erosion over the total
beach surface can amount to significant quantities. Unstabilized dunes
may form and migrate landward, resulting in an important net loss to the
littoral zone. Examples include some dunes along the Oregon coast (Coo-
per, 1958), between Pismo Beach and Point Arguello, California (Bowen
and Inman, 1966); central Padre Island (Watson, 1971); and near Cape Hen-
lopen, Delaware (Kraft, 1971). Typical rates of transport due to wind
range from 1 to 10 cubic yards per year per foot of beach front where
wind transport is noticeable. (Cooper, 1958; Bowen and Inman, 1966;
Savage and Woodhouse, 1968; and Gage, 1970.) However average rates prob-
ably range from 1 to 3 cubic yards per year per foot.
4.738 Mining and Dredging . From ancient times, sand and gravel have
been mined along coasts. In some countries, for example Denmark and
England, mining has occasionally had undesirable effects on coastal settle-
ments in the vicinity. Sand mining in most places has been discouraged by
legislation and the rising cost of coastal land, but it still is locally
important. (Magoon, et al., 1972.) It is expected that mining will become
more important in the offshore area in the future. (Duane, 1968, and Fisher,
1969.)
Such mining must be conducted far enough offshore so the mined pit
will not act as a sink for littoral materials, or refract waves adversely,
or substantially reduce the wave damping by bottom friction and percolation.
Material is also lost to the littoral zone when dredged from naviga-
ble waters (channels and entrances) within the littoral zone, and the
dredged material is dumped in some area outside of the littoral zone.
These dump areas can be for land fill, or in deep water offshore. This
action has been a common practice, because the first costs for some
dredging operations are cheaper when done this way.
4-129
(24 Moy 1972)
4-130
4.74 CONVECTION OF LITTORAL MATERIALS
While relative changes in sea level do not directly enter the sedi-
ment budget process, the net effect of these elevation changes is to move
the shoreline either landward (relative rise in sea level) or seaward
(relative fall in sea level). It thus can result in the appearance of a
gain or loss of sediment volume.
4-131
of elements in the sand budget varies with locality and with the bound-
aries of the particular littoral control volume. (These elements are
classified as point or line sources or sinks in Table 4-13, and the budget
is summarized in Equation 4-48.)
LAND OCEAN
Beach Replenishment
Submarine Canyon
Rivers
Dredging
Dune and Backshore
Storage
Winds
(supply and deflation)
Calcium Carbonate
(production and loss)
Mining
Longshore Currents
Tidal Currents
Longshore Winds
The erosion of beaches and cliffs and river contributions are the
principal known natural sources of beach sand in most localities. Inlets,
4-132
Table 4-14. Sand Budget of the Littoral Zone
lagoons and deep water in the longshore direction comprise the principal
known natural sinks for beach sand. Of potential, but usually unknown,
importance as either a source or a sink is the offshore zone seaward of
the beach.
(a) An eroding beach 4.4 miles long at root of spit that is 10 miles
long. Beaches on the remainder of the spit are stable. (See
Figure 4-50a.)
(b) A uniform recession rate of 3 feet per year along the eroding
4o4 miles.
(c) Depth of lowest shore parallel contour is -30 feet MSL, and
average dune crest elevation is 15 feet MSL.
(i) The 1.3-mile long beach to the right of the jettied inlet will
stabilize (no erosion) and realign with y changing to 3.5.
(j) The accumulation at the end of the spit will continue to grow
at an average annual rate of 400,000 cubic yards per year after
the proposed inlet is constructed.
4-134
(a) Site Sketch
Reach 4
10 -
FIND:
(b) After-inlet erosion rate of the beach to the left of the inlet.
SOLUTION Divide the beach under study into four sand budget compart-
:
4-136
for the sand budget calculation. The average annual volume of material
contributed to the littoral system per foot of eroding beach reaches
2 and 3 is:
n+ -
%(2) -
n+
^5(5)
-^-=
- 1,A
(15 + 30) (3)
Yl
= 5.0 yd?/yryft. .
Then, from Equation 4-46 the total annual contribution of the eroding
beaches to the system can be determined as:
^{(2)
"*"
^1(3)
" (1-3 f"i-
+ 3-1 "^i-) (5280 ft./mi.) (5 yd?/yr./ft.)
= 1.16 X 105yd?/yr.
284,000 yd.3/yr.
\{1,2)^
= 318,000 yd.Vyr.
= 4000,000 yd?/yr.
4-137
From Equation 4-24
then
Q..
=
'^
1+7 '
and
%> Q-rt' ^lt> ^^^ %. ^°^ each reach are shown in Figure 4-50b.
The gross transport rate across the inlet with the new y = 3.5 using
Equation 4-24, is:
Qn(2.5)(l+T)
= (284,000 yd?/yr.)
%{2,3) f^J ,
%{2,3)
= 512,000 yd?/yr. .
Q2 = 77,000 yd?/yr.
^(3)
" 400,000 yd?/yr. + 77,000 yd?/yr. - 284,000 yd?/yr. ,
{^H3))
4-138
Nourishment needed to maintain historic erosion rate on Reach 3 beach
is:
^5(5)
"
\3) ('i'3i3))
after inlet - b^^^ before inlet .
(qj^^^)
= 11 1,000 yd?/yr. .
Qj^^^
Q^, Qj^^, and Q^^ for the after-inlet condition are computed using
Equation 4-24 and related equations. The after-inlet sand budget is
shown in Figure 4-50c.
*************************************
4.8 ENGINEERING STUDY OF LITTORAL PROCESSES
4-139
Historical records of shoreline changes are usually in the form of
charts, surveyed profiles, dredging reports, beach replenishment reports
and aerial photos. As an example of such historical data. Figure 4-52
shows the positions of the shoreline at Sandy Hook, New Jersey, during
six surveys from 1835 to 1932. Such shoreline changes are useful for
computing longshore transport rates. The Corps of Engineers maintains,
in its District and Division offices, survey, dredging, and other reports
relating to Corps projects. Charts may be obtained from various federal
agencies including the Defense Mapping Agency Hydrographic Center, Geolog-
ical Survey, National Ocean Survey, and Defense Mapping Agency Topographic
Center. A map called "Status of Aerial Photography," which may be obtained
from the Map Information Office, Geological Survey, Washington, D.C. 20242,
shows the locations and types of aerial photos available for the U.S., and
lists the sources from which the photos may be requested. A description of
a coastal imagery data bank can be found in the interim report by Szuwalski
(1972).
Other kinds of data usually available are wave, tide, and meteoro-
logical data. Chapter 3 discusses wave and water level predictions; Sec-
tion 4.3 discusses the effects of waves on the littoral zone; and Sub-
section 4.34 presents methods of estimating wave climate and gives pos-
sible sources of data. These referenced sections indicate the wave, tide,
and storm data necessary to evaluate coastal engineering problems.
Charts show the coastal exposure of a study site, and since exposure
determines the possible directions from which waves reach the coast, expo-
sure also determines the most likely direction of longshore transport.
4-140
Figure 4-52. Growth of Sandy Hook, New Jersey, 1835-1932
4-141
about the net transport might be made, if only wave patterns of this photo
are analyzed. The possibility of seasonal or storm-induced reversals in
sediment transport direction should be investigated by periodic inspections
or aerial photos of the sand accumulation at groins and jetties.
(4 December 1967)
Figure 4-53. Transport Directions at New Buffalo Harbor Jetty on
Lake Michigan
4-142
(21 Moy 1972;
4-143
(29 August 1972)
4-144
of a convergent nodal zone. (Watson, 1971.) Nodal zones o£ divergence
are more common than nodal zones of convergence, because longshore trans-
port commonly diverges at exposed shores and converges toward major gaps
in the ocean shore, such as the openings of New York Harbor, Delaware Bay,
and Chesapeake Bay.
( 14 September 19691
4-146
Detailed study of inlets through barrier islands on the U.S. Atlan-
tic and gulf coasts shows that the shape of the shoreline at an inlet can
be classified in one of four characteristic planforms. (See Figure 4-58,
adapted from Galvin, 1971,) Inlets with overlapping offset (Fig. 4-59)
occur only where waves from the updrift side dominate longshore transport.
Wliere waves from the updrift side are less dominant, the updrift offset
(Fig. 4-59) is common. Where waves approach equally from both sides, in-
lets typically have negligible offset. (See Figure 4-60.) IVhere the sup-
ply of littoral drift on the updrift side is limited, and the coast is
fairly well exposed, a noticeable downdrift offset is common as, for exam-
ple, in southern New Jersey and southern Delmarva. (See Hayes, et al.,
(1970.) These planform relations to littoral processes have been found
for inlets through sandy barrier islands, but they do not necessarily hold
at inlets with rocky boundaries. The relations hold regionally, but tem-
porary local departures due to inlet migration may occur.
4-147
Qrt Q/t
Overlapping Offset
Adequate
updrift source j^
Updrift Offset
Downdrift Offset
Inadequate
updrift source
Qrt Qit
Negligible Offset
(Galvin,l97l )
4-148
(14 September 1969)
Figure 4-59. Fire Island Inlet, New York - Overlapping Offset
BAY
fluctuations of sand level at a site (Everts, 1973), and thus provides
data useful in beach fill and groin design.
Methods for obtaining beach and nearshore profiles, and the accu-
racy of the resulting profiles are discussed in Section 4.514.
Tracers are particles which react to fluid forces in the same manner
as particles in the sediment whose motion is being traced, yet which are
physically identifiable when mixed with this sediment. Ideally, tracers
must have the same size distribution, density, shape, surface chemistry,
and strength as the surrounding sediment, and in addition have a physical
property that easily distinguishes them from their neighbors.
4-150
One advantage of natural tracers is their tendency to "average" out
short-term trends and provide qualitatively accurate historical background
information on transport. Their use requires a minimum amount of field
work and a minimum number of technical personnel. Disadvantages include
the irregularity of their occurrence, the difficulty in distinguishing the
tracer from the sediment itself, and a lack of quantitative control on
rates of injection. In addition, natural tracers are unable to reveal
short-term changes in the direction of transport and changes in material
sources.
Colored glass, brick fragments and oolitic grains are a few examples
of nonradioactive particles that have been used as tracers. The most com-
monly used stable tracer is made by coating indigenous grains with bright
colored paint or flourescent dye. (Yasso, 1962; Ingle, 1966; Stuvier and
Purpura, 1968; Kidson and Carr, 1962; and Teleki, 1966.) The dyes make
the grains readily distinguishable among large sample quantities, but do
not significantly alter the physical properties of the grains. The dyes
must be durable enough to withstand short-term abrasion. The use of paints
and dyes as tracer materials offers advantages over radioactive methods in
that they require less sophisticated equipment to tag and detect the grains,
and do not require licensing or the same degree of safety precautions. How-
ever, less information is obtained for the same costs, and generally in a
less timely matter.
4-151
by a number of methods. Radioactive material has been placed in holes
drilled in a large pebble. It has been incorporated in molten glass
which, when hardened, is crushed and resized (Sato, et al., 1962; and
Taney, 1963). Radioactive material has been plated on the surface of
natural sediments. (Stephens, et al., 1968.) Radioactive gas (krypton
85 and xenon 133) has been absorbed into quartz sand. (Chleck, et al.,
1963; and Acree, et al., 1969.)
(Q,,,Qe,,Q^,Q„,7).
defined by Equations 4-21, 4-22, and 4-23. If any two are known, the
remaining three can be obtained from the three equations.
4-152
Section 4.531 describes four methods for estimating longshore trans-
port rate, and Sections 4.532 through 4.535 describe in detail how to use
two of these four methods. (See Methods 3 and 4.)
4-153
(2) Sediment transport in the nearshore zone. Seaward of the
breakers, sand is set in motion by waves moving over ripples, either
rolling the sand as bed load, or carrying it up in vortices as suspended
load. The sand, once in motion, is transported by mean tidal and wind-
induced currents and by the mass transport velocity due to waves. The
magnitude and direction of the resulting sediment transport are uncer-
tain under normal circumstances, although mass transport due to waves
is more than adequate to return sand lost from the beach during storms.
It appears that bottom mass transport acts to keep the sand close to the
shore, but that some material, probably finer sand, escapes offshore as
the result of the combined wind- and wave-induced bottom currents.
(4) The slope of the foreshore. There is a tendency for the fore-
shore to become steeper as grain size increases, and to become flatter
as mean wave height increases. Data for this relation exhibit much scat-
ter and quantitative relationships are difficult to predict.
4-154
REFERENCES AND SELECTED BIBLIOGRAPHY
BAJORUNAS, L„, and DUANE, D.B., "Shifting Offshore Bars and Harbor Shoal-
ing," Journal of Geophysioal Research, Vol. 72, 1967, pp. 6195-6205.
BASCOM, W.N., "The Relationship Between Sand Size and Beach-Face Slope,"
Transactions of the American Geophysical Union, Vol. 32, No. 6, 1951.
BASCOM, W.N., "Waves and Beaches," Beaches, Ch. IX, Doubleday, New York,
1964, pp. 184-212.
BERG, D.W., and DUANE, D.B., "Effect of Particle Size and Distribution on
Stability of Artificially Filled Beach, Presque Isle Penninsula,
Pennsylvania," Proceedings of the 11th Conference Great Lakes Research,
April 1968.
BIGELOW, H.B., and EDMONDSON, W.T., "Wind Waves at Sea Breakers and Surf,"
H.Oo 602, U.S. Navy Hydrographic Office, Washington, D.C. , 1947.
4-155
BIJKER, E.W., "Bed Roughness Influence on Computation of Littoral Drift,"
Abstracts of the 12th Coastal Engineering Conference^ Washington, D.C.,
1970.
BODINE, B.R., "Hurricane Surge Frequency Estimated for the Gulf Coast of
Texas," Bl-26, U.S. Army, Corps of Engineers, Coastal Engineering
Research Center, Washington, D.C., Feb. 1969.
BOWEN, A.J., and INMAN, D.L., "Budget of Littoral Sands in the Vicinity
of Point Arguello, California," TM-19, U.S. Army, Corps of Engineers,
Coastal Engineering Research Center, Washington, D.C. , Dec. 1966.
BOWMAN, R.S., "Sedimentary Processes Along Lake Erie Shore: Sandusky Bay,
Vicinity of Willow Point," in "Investigations of Lake Erie Shore Ero-
sion," Survey 18, Ohio Geological Survey, 1951.
BRICKER, O.P., ed.. Carbonate Sediments, No. 19, The Johns Hopkins
University Studies in Geology, 1971, 376 pp.
BUMPUS, D.F., "Residual Drift Along the Bottom on the Continental Shelf
in the Middle Atlantic Bight," Limnology and Oceanography, 75th
Anniversary Vol., Supplement to Vol. X, 1965.
4-156
BYRNE, R„J., "Field Occurrences of Induced Multiple Gravity Waves,"
Journal of Geophysical Besearoh^ Vol. 74, No. 10, May 1969, pp.
2590-2596.
CALDWELL, J.M., "Wave Action and Sand Movement Near Anaheim Bay,
California," TM-68, U.S. Army, Corps of Engineers, Beach Erosion
Board, Washington, D.C., Feb. 1956.
COLBY, B.C., and CHRISTENSEN, R.P., "Visual Accumulation Tube for Size
Analysis of Sands," Journal of the Hydraulics Division, ASCE, Vol. 82,
NOo 3, June 1956.
COLONY, R„J., "Source of the Sands on South Shore of Long Island and the
Coast of New Jersey," Journal of Sedimentary Petrology, Vol. 2, 1932,
pp. 150-159.
4-157
COOPER, W.S., "Coastal Sand Dunes of Oregon and Washington," Memoir No.
72, Geological Society of America, June 1958.
CUMMINS, R.S., Jr., "Radioactive Sediment Tracer Tests, Cape Fear River,
North Carolina," MP No. 2-649, Waterways Experiment Station, U.S.
Army, Corps of Engineers, Vicksburg, Miss., May 1964.
DARLING, J.M., "Surf Observations Along the United States Coasts," Journal
of the Waterways and Harbors Division, ASCE, WWl, Feb. 1968, pp. 11-21.
DARLING, J.M., and DUMM, D.G., "The Wave Record Program at CERC," MP
1-67, U.S. Army, Corps of Engineers, Coastal Engineering Research
Center, Washington, D.C., Jan. 1967.
DAVIS, R.A., Jr., "Beach Changes on the Central Texas Coast Associated
with Hurricane Fern, September 1971," Vol. 16, Contributions in Marine
Science, University of Texas, Marine Science Institute, Port Aransas,
Tex., 1972.
DAVIS, R.A., Jr., and FOX, W.T., "Beach and Nearshore Dynamics in Eastern
Lake Michigan," TR No. 4, ONR Task No. 388-092/10-18-68, (414), Office
of Naval Research, Washington, D.C., June 1971.
DAVIS, R.A., Jr., and FOX, W.T., "Coastal Processes and Nearshore Sand
Bars," Journal of Sedimentary Petrology, Vol. 42, No. 2, June 1972,
pp. 401-412.
4-158
DAVIS, R.A., Jr., and FOX, W.T., "Four-Dimensional Model for Beach and
Inner Nearshore Sedimentation," The Jowmal of Geology , Vol. 80, No.
4, July 1972.
DEAN, R.G., "Heuristic Models of Sand Transport in the Surf Zone," Con-
ference on Engineering Dynamics in the Coastal Zone, 1973.
DEWALL, A.E., "The 17 December 1970 East Coast Storm Beach Changes,"
Unpublished Manuscript, U.S. Army, Corps of Engineers, Coastal
Engineering Research Center, Washington, D.C., 1972.
DEWALL, A.E., PRITCHETT, P.C., and GALVIN, C.J., Jr., "Beach Changes
Caused by a Northeaster Along the Atlantic Coast," Abstracts from the
Annual Meeting of the Geological Society of America, Washington, D.C.,
1971.
DEWALL, A.E., and RICHTER, J.J. , "Beach Changes at Boca Raton, Florida,"
Annual Meeting of the American Shore and Beach Preservation Associa-
tion, (presentation only) Fort Lauderdale, Fla. , Aug. 1972.
DIVOKY, D., LeMEHAUTE, B., and LIN, A., "Breaking Waves on Gentle Slopes,"
Journal of Geophysical Research, Vol. 75, No. 9, 1970.
DOLAN, R., "Barrier Dune System Along the Outer Banks of North Carolina:
A Reappraisal," Science, 1972, pp. 286-288.
4-159
DOLAN, R., PERM, J.C., and McARTHUR, L.S., "Measurements of Beach Process
Variables, Outer Banks, North Carolina," TR No. 64, Coastal Studies
Institute, Louisiana State University, Baton Rouge, La., Jan. 1969.
DRAPER, L., "Wave Activity at the Sea Bed Around Northwestern Europe,"
Marine Geology ^ Vol. 5, 1967, pp. 133-140.
DUANE, D.B., and JUDGE, C.W., "Radioisotopic Sand Tracer Study, Point
Conception, California," MP 2-69, U.S. Army, Corps of Engineers,
Coastal Engineering Research Center, Washington, D.C, May 1969.
DUBOIS, R.N., "Inverse Relation Between Foreshore Slope and Mean Grain
Size as a Function of the Heavy Mineral Content," Geological Society
of America Bulletin, Vol. 83, Mar. 1972, pp. 871-876.
EAGLESON, P.S., and DEAN, R.G., "A Discussion of 'The Supply and Loss of
Sand to the Coast' by J. Wo Jolinson," Journal of the Waterways and
Harbors Division, ASCE, Vol. 86, No. WW2, June 1960.
4-160
EATON, R„0., "Littoral Processes on Sandy Coasts," Proceedings of the
First Conference on Coastal Engineering ^ Council on Wave Research,
Engineering Foundation, Oct. 1950.
FISHER, CM., "Mining the Ocean for Beach Sand," Proceedings of Civil
Engineering in the Oceans II, ASCE, Dec. 1969.
4-161
FOLK, R.L., and WARD, W.C., "Brazos River Bar. A Study in the Signifi-
cances of Grain Size Parameters," Journal of Sedimentcay Petyyology,
Vol. 27, 1957, pp. 3-26.
GAGE, B.O., "Experimental Dunes of the Texas Coast," MP 1-70, U.S. Army,
Corps of Engineers, Coastal Engineering Research Center, Washington,
D.C., Jan. 1970.
GALVIN, C.J., Jr., "A Theoretical Distribution of Waiting Times for Tracer
Particles on Sand Bed," Vol. 1, Bulletin and Summary of Research Pro-
gress, Fiscal Year 1964, U.S. Army, Corps of Engineers, Coastal
Engineering Research Center, Washington, D.C., 1964.
GALVIN, CoJ. Jr., "A Selected Bibliography and Review of the Theory and
Use of Tracers in Sediment Transport Studies," Vol. I, Bulletin and
Summary Report of Research Progress, Fiscal Year 1964, U.S. Army,
Corps of Engineers, Coastal Engineering Research Center, Washington,
D.C., 1964.
GALVIN, C.J., Jr., "A Gross Longshore Transport Rate Formula," Proceedings
of the 13th Coastal Engineering Conference^ Vancouver, B.C., Canada,
July 1972.
GALVIN, C.J., Jr., "Wave Breaking in Shallow Water," ^aves on Beaches and
Resulting Sediment Transport^ Academic Press, March 1972.
GALVIN, C.J., "Wave Climate and Coastal Processes," Water Environments and
Human Needs^ A.T. Ippen, Ed., M.I.T. Parsons Laboratory for Water
Resources and Hydrodynamics, Cambridge, Mass., 1971., pp. 48-78.
GALVIN, C.Jo, and NELSON, R.A., "A Compilation of Longshore Current Data,"
MP 2-67, U.S. Army, Corps of Engineers, Coastal Engineering Research
Center, Washington, D.C, Mar. 1967.
4-162
GALVIN, C.J., and SEELIG, W.N., "Surf on U.S. Coastline," Unpublished
Research Paper, U.S. Army, Corps of Engineers, Coastal Engineering
Research Center, Washington, D.C., Aug. 1969.
GALVIN, C.J., and SEELIG, W.N., "Nearshore Wave Direction from Visual
Observation," Transactions of the American Geophysical Union^ Vol. 52,
No. 4, Apr. 1971.
GIBBS, R.J., "The Geochemistry of the Amazon River System: Part I. The
Factors that Control the Salinity and the Composition and Concentration
of the Suspended Solids," Geological Society of America Bulletin^ Vol.
78, Oct. 1967, pp. 1203-1232.
GILES, RoT., and PILKEY, O.H., "Atlantic Beach and Dune Sediments of the
Southern United States," Journal of Sedimentary Petrology ^ Vol. 35,
No. 4, Dec. 1965, pp. 900-910.
GOLDSMITH, V., COLONELL, J.M., and TURBIDE, P.W., "Forms of Erosion and
Accretion on Cape Cod Beaches," Proceedings of the 13th International
Conference on Coastal Engineering ^ Vancouver, B.C., July 1972.
GONZALES, W„R., "A Method for Driving Pipe in Beach Rock," Vol. Ill,
Bulletin and Summary of Research Progress, Fiscal Years 1967-69, U.S.
Army, Corps of Engineers, Coastal Engineering Research Center, Wash-
ington, D.C., July 1970.
HALL, J. v., and HERRON, W.J., "Test of Nourishment of the Shore by Off-
shore Deposition of Sand," TM-17, U.S. Army, Corps of Engineers, Beach
Erosion Board, Washington, D.C., June 1950.
4-163
HALLERMEIER, R.J., and GALVIN, C.J., "Wave Height Variation Around
Vertical Cylinders," Transactions of the Amerioan Geophysical Union
53rd Annual Meeting, 1972, p. 397.
HARRIS, D.L., "Wave Estimates for Coastal Regions," Shelf Sediment Trans-
port, edited by Swift, Duane, and Pilkey. Dowden, Hutchinson, and
Ross, Stroudsburg, Pa., 1972.
HARRISON, W., and WAGNER, K.A., "Beach Changes at Virginia Beach, Virginia,"
MP 6-64, U.S. Army, Corps of Engineers, Coastal Engineering Research
Center, Washington, D.C., Nov. 1964.
HAYES, M.O., GOLDSMITH, V., and HOBBS, C.H. Ill, "Offset Coastal Inlets,"
Proceedings of the 12th Coastal Engineering Conference, Sept. 1970,
pp. 1187-1200.
4-164
HERRON, W.J., and HARRIS, R.L., "Littoral Bypassing and Beach Restoration
in the Vicinity of Port Huenerae, California," Proceedings of the 10th
Conferenae on Coastal Engineering^ Tokyo, 1966, ASCE, United Engineer-
ing Center, New York, 1967.
HICKS, S.D., "On Classifications and Trends of Long Period Sea Level
Series," Shore and Beach^ Apr. 1972.
HUBBELL, D.W., and SAYRE, W.W., "Sand Transport Studies with Radioactive
Tracers," Journal of the Hydraulics Division^ ASCE, Vol. 90, No. HY3,
1965, pp. 39-68.
HUME, J.D., and SCHALK, M., "Shoreline Processes Near Barrow, Alaska: A
Comparison of the Normal and the Catastrophic," Arctic, Vol. 20, No. 2,
June 1967, pp. 86-103.
INGLE, J.Co, "The Movement of Beach Sand," Bevel. Sediment, Vol. 5, Elsevier,
Amsterdam, 1966.
4-165
INMAN, D.L., "Wave-Generated Ripples in Nearshore Sands," TM-100, U.S.
Army, Corps of Engineers, Beach Erosion Board, Washington, D.C., Oct.
1957.
INMAN, D.L., and Quinn, W.H., "Currents in the Surf Zone," Proceedings of
the Second Conference on Coastal Engineering, ASCE, Council on Wave
Research, Berkeley, Calif., 1952, pp. 24-36.
INMAN, D.L., and NASU, N., "Orbital Velocity Associated with Wave Action
Near the Breaker Zone," TM-79, U.S. Army, Corps of Engineers, Beach
Erosion Board, Washington, D.C., Mar. 1956.
INMAN, D.L., and RUSNAK, G.S., "Changes in Sand Level on the Beach and
Shelf at La Jolla, California," TM-82, U.S. Army, Corps of Engineers,
Beach Erosion Board, Washington, D.C., July 1956.
INMAN, D.L., and FRAUTSCHY, J.D., "Littoral Processes and the Development
of Shorelines," Proceedings of the Coastal Engineering Specialty Con-
ference (Santa Barbara) ^ ASCE, 1966, pp. 511-536.
INMAN, D.L., TAIT, R.J., and NORDSTROM, C.E., "Mixing in the Surf Zone,"
Journal of Geophysical Research^ Vol. 76, No. 15, May 1971, p. 3493.
IWAGAKI, Y, and NODA, H., "Laboratory Study of Scale Effect in Two Dimen-
sional Beach Processes," Proceedings of the Eighth Conference on Coastal
Engineering, Ch. 14, ASCE, 1962.
JAMES, W.R., "A Class of Probability Models for Littoral Drift," Proceedings
of the 12th Coastal Engineering Conference, Washington, D.C., Sept. 1970.
4-166
JOHNSON, J.W., "Sand Transport by Littoral Currents," Prooeedings of the
Fifth Hydraulios Conference^ Bulletin 34, State University o£ Iowa,
Studies in Engineering, 1953, pp. 89-109.
JOHNSON, J.W., "The Supply and Loss of Sand to the Coast," ASCE Journal,
Vol. 85, NOo WW3, Sept. 1959, pp. 227-251.
KAMEL, A.M., "Littoral Studies near San Francisco Using Tracer Techniques,"
1^-131, U.S. Army, Corps of Engineers, Beach Erosion Board, Washington,
D.C., Nov. 1962.
KENNEDY, J.F., and FALCON, M., "Wave -Gene rated Sediment Ripples," Hydro-
dynamics Laboratory Report 86, Department of Civil Engineering,
Massachusetts Institute of Technology, Cambridge, Mass., 1965.
67
KIDSON, C, and CARR, A. P., "Marking Beach Materials for Tracing Experi-
ments," Jonxmdl of the Hydrcculias Division^ ASCE, Vol. 88, July 1962.
KINSMAN, BLAIR, Wind Waves, Their Generation and Propagation on The Ocean
Surface, Prentice-Hall, Englewood Cliffs, N.J,, 1965.
KIRTLEY, D.W. , "Reef-Building Worms," Sea Frontiers, Vol. 17, No. 2, 1971.
KOLESSAR, M.A., and REYNOLDS, J.L., "The Sears Sea Sled for Surveying
in the Surf Zone," Vol. II, Bulletin and Sununary Report of Research
Progress, Fiscal years 1965-66, U.S. Army, Corps of Engineers, Coastal
Engineering Research Center, Washington, D.C., 1966.
KOMAR, PAUL D., and INMAN, DOUGLAS L., "Longshore Sand Transport on
Beaches, Journal of Geophysical Research, V. 75, No. 30, Oct 20, 1970,
pp. 5914-5927.
KRUMBEIN, W.C, "A Method for Specification of Sand for Beach Fills,"
114-102, U.S.Army, Corps of Engineers, Beach Erosion Board, Washington,
D.C., Oct. 1957.
4-168
KRUMBEIN, W.C, and JAMES, W.R., "A Lognormal Size Distribution Model for
Estimating Stability of Beach Fill Material," TM-16, U.S. Army, Corps
of Engineers, Coastal Engineering Research Center, Washington, D.C.,
Nov. 1965.
LEE, J., YANCY, T., and WILDE, P., "Recent Sedimentation of the Central
California Continental Shelf," Part A: Introduction and Grain Size
Data, HEL 2-28, College of Engineers, University of California,
Berkeley, Calif., Oct. 1970.
McCORMICK, C.L., "A Probable Cause for Some Changes in the Beach Erosion
Rates on Eastern Long Island," Unpublished Paper, Southampton College,
Long Island University, New York, 1971.
4-169
MAGOON, O.T,, HAUGEN, J.C., and SLOAN, R.L., "Coastal Sand Mining in
Northern California, U.S.A.," Pvoaeedings of the ISth Coastal
Engineering Research Conference, Vancouver, B.C., Canada, July
1972.
MASON, M.A., "Abrasion of Beach Sands," TM-2, U.S. Army, Corps of Engineers,
Beach Erosion Board, Washington, D.C., Feb. 1942.
MEI, C.C, LIU, P., and CARTER, T.G., "Mass Transport in Water Waves,
Part I: Theory - Part II: Experiments," Report No. 146, Ralph M.
Parsons Laboratory for Water Resources and Hydrodynamics, Massachu-
setts Institute of Technology, Apr. 1972.
4-170
MOORE, G.W., and COLE, A.Y., "Coastal Processes, Vicinity of Cape
Thompson, Alaska," in "Geologic Investigations of Cape Thompson, NW
Alaska - Preliminary Report," Trace Element Investigation Report 753,
U.S. Geological Survey, Washington, D.C., 1960.
MORSE, B., and GROSS, M.G., and BAJWES, C.A., "Movement of Seabed
Drifters Near the Columbia River," Journal of the Waterways and
Hcwbors Division, ASCE, Vol. 94, No. WWl, Paper No. 5817, Feb. 1968,
pp. 93-103.
MURRAY, S.P., "Bottom Currents Near the Coast During Hurricane Camille,"
Journal of Geophysical Research, Vol. 75, No. 24, Aug. 1970.
NATIONAL MARINE CONSULTANTS, INC., "Wave Statistics for Seven Deep Water
Stations Along the California Coast," prepared for Department of the
Army, U.S. Army, Corps of Engineers Districts, Los Angeles and San
Francisco, Calif., Dec. 1960a.
NATIONAL MARINE CONSULTANTS, INC., "Wave Statistics for Ten Most Severe
Storms Affecting Three Selected Stations off the Coast of Northern
California, Ehirmg the Period 1951-1960," prepared for U.S. Army
Engineer District, San Francisco, Calif., Dec. 1960b.
OTTO, G.H., "A Modified Logarithmic Probability Graph for the Interpreta-
tion of Mechanical Analyses of Sediments," Journal of Sedimentary
Petrology, Vol. 9, 1939, pp. 62-76.
4-171
PETTIJOHN, F.J., Sedimentary Rooks, Harper and Brothers, New York, 1957,
p. 117.
PUTNAM, J. A., MUNK W.H. , and TRAYLOR, M.A. , "The Prediction of Longshore
Currents," Transactions of the American Geophysical Union, Vol. 30,
1949, pp. 337-345.
RAMSEY, M.D., "The Relationship Between Mean Sand Size vs Local Foreshore
Slope for 166 New Jersey Sand Samples," Unpublished MFR, U.S. Army,
Corps of Engineers, Coastal Engineering Research Center, Washington,
D.C., Sept. 1971.
RAMSEY, M.D., and CALVIN, C.J., "Size Analysis of Sand Samples from
Three S. New Jersey Beaches," Unpublished Paper, U.S. Army, Corps
of Engineers, Coastal Engineering Research Center, Washington, D.C.,
Sept. 1971.
REIMNITZ, E., and ROSS, D.A., "The Sea Sled - A Device for Measuring
Bottom Profiles in the Surf Zone," Marine Geology, Vol. 11, 1971.
ROSS, D.A., "Atlantic Continental Shelf and Slope of the United States -
Heavy Minerals of the Continental Margin from Southern Nova Scotia to
Northern New Jersey," Professional Paper 529-G, U.S. Geological Survey,
U.S. Government Printing Office, Washington, D.C., 1970.
4-172
RUSNAK, G.A., STOCKMAN, R.W. , and HOFMANN, H.A., "The Role of Shell
Material in the Natural Sand Replenishment Cycle of the Beach and
Nearshore Area Between Lake Worth Inlet and the Miami Ship Channel,"
CERC Contract Report (DA-49-055-CIV-ENG-63-12) , Institute of Marine
Sciences, University of Miami, Coral Gables, Fla. , 1966.
SATO, S., IJIMA, T. , and TANAKA, N. , "A Study of Critical Depth and Mode
of Sand Movement Using Radioactive Glass Sand," Proceedings of the
Eighth Conferenoe on Engineeving , Mexico City, 1962, pp. 304-323.
SAVILLE, T., Jr., "Model Study of Sand Transport Along an Infinitely Long
Straight Beach," Transactions of the American Geophysical Union, Vol.
31, 1950.
SAVILLE, T., Jr., and CALDWELL, J.M., "Accuracy of Hydro graphic Surveying
In and Near the Surf Zone," TM-32, U.S. Army, Corps of Engineers, Beach
Erosion Board, Washington, D.C., Mar. 1953.
SAYLOR, J.H., and HANDS, E.B., "Properties of Longshore Bars in the Great
Lakes," Proceedings of the 12th Conference on Coastal Engineering, Vol.
2, 1970, pp. 839-853.
4-173
SAYLOR, J.H., and UPCHURCH, S. , "Bottom Stability and Sedimentary Processes
at Little Lake Harbors, Lake Superior," U.S. Army, Corps of Engineers,
Lake Survey District Detroit, Mich., 1970.
SCRUTON, P.C., "Delta Building and the Deltaic Sequence," Recent Sedi-
mentSy Northwest Gulf of Mexico^ American Association of Petroleum
Geologists, 1960, pp. 82-102.
SHAY, E.A., and JOHNSON, J.W., "Model Studies on the Movement on Sand
Transported by Wave Action Along a Straight Beach," Issue 7, Ser. 14,
Institute of Engineering Research, University of California, Berkeley,
Calif., 1951.
SHEPARD, P.P., "Rise of Sea Level Along Northwest Gulf of Mexico," Recent
Sediments , Northwest Gulf of Mexico, American Association of Petroleum
Geologists, 1960b, pp. 338-344.
SHEPARD, P.P., and BUFFINGTON, E.G., "La Jolla Submarine Fan Valley,"
Marine Geology, Vol. 6, 1968, pp. 107-143.
SHEPARD, P.P., and DILL, R.F., ^'Submarine Canyons and Other Sea Valleys,"
Rand McNally, Chicago, 1966.
SHEPARD, P.P., and GRANT, U.S., IV, "Wave Erosion Along the Southern
California Coast," Bulletin of the Geologic Society of America, Vol.
58, 1947, pp. 919-926.
4-174
SHEPARD, F,P., MacDONALD, G.A., and COX, D.C., "The Tsunami of April 1,
1946," Bulletin of the Soripps Institute of Oceanography, Vol. 5,
No. 6, 1950, pp. 391-528.
SHUYSKIY, Y.D., "The Effect of Strong Storms on Sand Beaches of the Baltic
Eastern Shore," Ooeanology, Vol. 9, No. 3, 1970, p. 388.
SONU, C.J., and VAN BEEK, J.L., "Systemic Beach Changes on the Outer
Banks, North Carolina," Journal of Geology, Vol. 79, 1971, pp. 416-425.
STAFFORD, D.B., "An Aerial Photographic Technique for Beach Erosion Sur-
veys in North Carolina," TM-36, U.S. Army, Corps of Engineers, Coastal
Engineering Research Center, Washington, D.C., Oct. 1971.
STODDARD, D.R„, "World Erosion and Sedimentation," Water, Earth, and Man,
Barnes and Noble, 1969.
STONE, R.O., and SUMMERS, H.J., "Final Report, Study of Subaqueous and
Subaerial Sand Ripples," ONR Project No. N00014-67-A-0269-0002, Task
No. NR- 388-085, Report No. USG Geology 72-1, 1972.
4-175
SZULWALSKI, A., "Coastal Imagery Data Bank: Interim Report," MP-3- 72,
U.S. Army, Corps of Engineers, Coastal Engineering Research Center,
Washington, D.C., Nov. 1972.
TANEY, NoE., "Geomorphology of the South Shore of Long Island, New York,"
TM-128, U.S. Army, Corps of Engineers, Beach Erosion Board, Washington,
D.C., Sept. 1961a.
TANEY, N.E., "Littoral Materials of the South Shore of Long Island, New
York," TM-129, U.S. Army, Corps of Engineers Beach Erosion Board,
Washington, D.C., Nov. 1961b.
THOMPSON, E.F., and HARRIS, D.L., "A Wave Climatology for U.S. Coastal
Waters," Proceedings of Offshore Technology Conference, Dallas, Texas,
May, 1972.
THOREAU, H.Do, Cape Cod, Thicknor and Fields, Boston, Mass., 1865.
4-176
TUCKER, M.J., "Recent Measurement and Analysis Techniques Developed at
The National Institute of Oceanography," National Institute of
Oceanography 3 Collected Reprints, V. 11, No. 465, 1963. Reprinted
from Ocean Wave Spectra; Proceedings of Conference on Ocean Wave
Spectra, 1961, pp. 219-226.
URBAN, H.D., and CALVIN, C.J., "Pipe Profile Data and Wave Observations
from the CERC Beach Evaluation Program, January - March, 1968,"
MP 3-69, U.S. Army, Corps of Engineers, Coastal Engineering Research
Center, Washington, D.C., Sept. 1969.
U.S. ARMY, CORPS OF ENGINEERS, "Relation Between Sand Size and Slope of
the Foreshore," Interim Report, Beach Erosion Board, Washington, D.C.
1933.
U.So ARMY, CORPS OF ENGINEERS, "Survey of Ocean City Harbor and Inlet and
Sinepuxent Bay, Maryland," Engineer District, Baltimore, Md. , 1948b,
(Unpublished)
U.S. ARMY, CORPS OF ENGINEERS, "Atlantic Coast of New Jersey, Sandy Hook
to Bamegat Inlet, Beach Erosion Control Report on Cooperative Study,"
Engineer District, New York, N.Y., 1954b, (Unpublished).
4-177
U.S. ARMY, CORPS OF ENGINEERS, "Cliff Walk, Newport, Rhode Island, Beach
Erosion Control Study," Letter from Secretary of the Army, U.S. Govern-
ment Printing Office, Washington, 1965.
U.S. ARMY, CORPS OF ENGINEERS, "A Study of Methods to Preserve Gay Head
Cliffs, Martha's Vineyard, Massachusetts," Report prepared by Woodard-
Moorhouse and Associates, Inc., for the Engineer, New England Division,
Oct. 1970.
U.S. CONGRESS, "Beach Erosion Study, Lake Michigan Shore Line of Milwaukee
County, Wisconsin," House Document 526, 79th Congress, 2nd Session,
p. 16, 1946.
U.S. CONGRESS, "North Carolina Shore Line, Beach Erosion Study," House
Document 763, 80th Congress, 2nd Session, 1948.
U.S. CONGRESS, "Ocean City, New Jersey, Beach Erosion Control Study,"
House Document 184, 83d Congress, 1st Session, 1953a.
U.S. CONGRESS, "Cold Spring Inlet (Cape May Harbor), New Jersey," House
Document 206, 83d Congress, 1st Session, 1953b.
U.S. CONGRESS, "Racine County, Wisconsin," House Document 88, 83d Congress,
1st Session, 1953e.
U.S. CONGRESS, "Illinois Shore of Lake Michigan," House Document 28, 83d
Congress, 1st Session 1953f.
U.S. CONGRESS, "Waikiki Beach, Island of Oahu, T.H. , Beach Erosion Study,"
House Document No. 227, 83d Congress, 1st Session, 1953g.
U.S. CONGRESS, "Pinellas County, Florida," House Document 380, 83d, Congress,
2d Session, 1954a.
4-178
U.S. CONGRESS, "Port Hueneme, California," House Docioment 362, 83d,
Congress, 2d Session, 1954b.
WATTS, G.M., "A Study of Sand Movement at South Lake Worth Inlet, Florida,"
TM-42, U.S. Army, Corps of Engineers, Beach Erosion Board, Washington,
D.C., Oct. 1953b.
WATTS, G.M., "Behavior of Beach Fill at Ocean City, New Jersey," TM-77
U.S. Army, Corps of Engineers, Beach Erosion Board, Washington, D.C.,
Feb. 1956.
4-179
WIEGEL, R,L., Oaeanogrccphio Engineering^ Prentice-Hall, Englewood Cliffs,
N.J., 1964.
WOODARD, D.W., et al., "The Use of Grasses for Dune Stabilization Along
the Gulf Coast with Initial Emphasis on the Texas Coast," Report No.
114, Gulf Universities Research Corporation, Galveston, Tex., 1971.
WRIGHT, F., and COLEMAN, Jr., "River Delta Morphology: Wave Climate and
the Role of the Subaqueous Profile," Scienae, Vol. 176, 1972, pp. 282-
284.
ZEIGLER, J.M., and GILL, B., "Tables and Graphs for the Settling Velocity
of Quartz in Water Above the Range of Stokes Law," Ref. No. 59-36,
Woods Hole Oceanographic Institution, Woods Hole, Mass., July 1959,
ZEIGLER, J. Mo, and TUTTLE, S.D., "Beach Changes Based on Daily Measure-
ments of Four Cape Cod Beaches," Journal of Geology, Vol. 69, No. 5,
1961, pp. 583-599.
ZEIGLER, J.Mo, et al., "Residence Time of Sand Composing the Beaches and
Bars of Outer Cape Cod," Proceedings of the Ninth Conference on Coastal
Engineering J ASCE, Vol. 26, 1964, pp. 403-416.
4-180
O U. S. GOVERNMENT PRINTING OFFICE : 1976 O - 508-951 (Vol. I)