You are on page 1of 22

Marine Geology, 45 (1982) 41--62

Elsevier Scientific Publishing Company, Amsterdam -- Printed in The Netherlands 41

DYNAMICS OF A HIGH-ENERGY DISSIPATIVE SURF ZONE

L.D. WRIGHT', R.T. G U Z A 2 and A.D. SHORT I


1Coastal Studies Unit, Dept. of Geography, University of Sydney, Sydney, N.S.W. 2006
(Australia)
2Shore Processes Laboratory, Scripps Institution of Oceanography, La JoUa, CA 92093
(U.S.A.)
(Received September 24, 1980; revised and accepted December 29, 1980)

ABSTRACT

~Vright, L.D., Guza, R.T. and Short, A.D., 1982. Dynamics of a high-energy dissipative
surf zone. Mar. Geol., 45: 41--62.

Pressure and horizontal current (u, v) time series were measured at different positions
across the inner 150 m of a wide ( ~ 5 0 0 m) surf zone of a microtidal high wave-energy
beach. Incident waves had average heights of 3--4 m with maxima of 5 m and periods of
12 to 15 sec. Bores of broken waves diminished in height at a nearly constant rate as they
progressed across the surf zone. The ratio, % of bore height H to local water depth h was
everywhere less than 1 for even the highest bores and was on the order of 0.40 for the
significant bores at incident wave frequencies. Rip circulation was weak or absent but a
moderate longshore current was present. Shore-normal flows were vertically segregated
with strong net onshore flows prevailing just below the surface accompanied by weaker
net seaward flows near the bed. Spectra of water surface oscillations, ~ as determined
from pressure, u, and v reveal that most of the energy in the inner surf zone was at
infragravity frequencies (periods greater than 30 sec). Shoreward decay of wave energy
at incident wave frequencies was accompanied by shoreward growth of infragravity
energy. Near the beach the infragravity oscillations had heights on the order of 1 m.
Cross-spectra show that the infragravity oscillations were standing in the shore-normal
direction. From the relative magnitudes o f infragravity versus incident wave currents, it
is inferred that the surf beat may be an order of magnitude more important than incident
waves to the transport of sediment in the inner surf zone.

INTRODUC~ON

It is clear from recent literaturethat beach morphologies and surf zone


dynamics differdramatically between the two extremes of steep, reflective
beach systems and flatdissipativesystems (Huntley and Bowen, 1975a;
Guza and Inman, 1975; Wright et al.,1979a, b; Short, 1979a, b, 1980;
Wright, 1980). Between the two extremes are four intermediate states
characterized by different scales of bar-trough and rhythmic longshore
topographies and rip cells(Wright et al.,1979a). The high reflection co-
efficientsfor incident waves and subharmonic resonances which are typical
of reflectivebeaches (Guza and Davis, 1974; Guza and Inman, 1975;
Huntley and Bowen 1975b, 1979; Wright et al.,1979a), and are often

0025-3227/82/0000--0000/$02.75 © 1982 Elsevier Scientific Publishing Company


42

present in certain regions of intermediate states (Wright et al., 1979a), are


generally absent on dissipative beaches (Huntley and Bowen, 1975a; Guza
and Bowen, 1977). Spilling breakers, pronounced infragravity oscillations
(or "surf beat"), and the presence of multiple parallel longshore bars have
been reported as c o m m o n features of dissipative surf zones (Huntley and
Bowen, 1975a; Short, 1975; Huntley, 1976; Sasaki et al., 1977; Wright et
al., 1979a, b; Holman et al., 1979).
To date, most sets of field data on surf zone dynamics have been obtained
from reflective, intermediate or low-energy dissipative beaches. For obvious
reasons, field measurements in wide, high-energy surf zones have been
limited. However, the processes which act under such conditions are of
considerable importance since high-energy dissipative beaches are dynamically
and morphologically analogous to the "storm profiles" of seasonally varying
beach systems. The purpose of this paper is to report observations of surf
characteristics, inshore circulation, and low frequency oscillations from a
set of experiments conducted on a perennially high-energy dissipative beach.

EXPERIMENT SITE AND METHODS

The experiments were conducted near Goolwa, adjacent to the Coorong


Region of South Australia over the period 28 January--3 February, 1980.
The experiment site (Fig.l) was situated near the northwestern end of an
uninterrupted depositional bight 210 km in length. The site was 10 km from
the nearest headland and approximately 6 km west of the Murray River
m o u t h which constitutes the only interruption to the otherwise continuous
long, straight beach system. The beach is composed of fine sand (mean
diameter 2.9 ¢ or 0.13 mm) consisting of approximately 50% CaCO3. The
beach is exposed to the west-coast swell environment of the Southern
Ocean (e.g. Davies, 1973) where the prevailing westerly winds blowing over
a very long fetch generate high~nergy southwesterly swell. The dominant
incident swell arrive essentially normal to the shore and experience minimal
refraction. The only long-term wave statistics are from lighthouse observa-
tions which show the wave climate to be dominated by tong-period (12--16
sec) southwesterly swell. Significant swell height exceeds 4 m for 6% of the
time and exceeds 2 m for 68% of the time (Short and Hesp, 1980). Average
spring tide range is 0.8 m; however, the experiments were carried o u t during
neap tides when the range was only 0.4 m.
The high-energy swell climate and abundant fine sand result in an extremely
dissipative beach system. Beach profiles exhibit very little spatial or temporal
variation. During the period of observations, beach profiles were uniform
alongshore within 200 m of the site. Rip cells, beach cusps, and rhythmic
surf-zone topography were absent during the experiment period. The average
height of the breakers at the seaward-most break was a b o u t 3 m but with
maxima of 5 m as estimated by theodolite from shore and measured from
a boat just seaward of the breaker zone. The surf zone was 400 to 500 m
wide with at least two subtle b u t continuous longshore bars separated by
43

.......... .iiiiiiiiiiiiiii~,. . :.

% ~ ~iliiiiii!iiiiiiii

\'%
Fig.l. Location and offshore bathymetry of the study site.

shallow troughs (Figs.2 and 3). The subaerial beach and inner 150 m of the
surf zone (to the limit of surveys) had an average gradient, tan/3, of 0.011
(0.38 °) and a slightly concave upwards profile (Figs. 3 and 4).
The field instrumentation system used in the experiments has been
described in detail by Bradshaw et al. (1978). Time series of pressure were
obtained with strain-gage pressure transducers (Shaevitz Model P-741).
Horizontal flow velocities (u, v) were measured b y means of orthogonally
m o u n t e d miniature bidirectional duct-impeller flow meters (Sonu et al.,
1974; Teleki, 1976). Sensors were sampled simultaneously at intervals of
1 sec. Sensor positions for different runs are shown in Fig.4. Owing to the
extreme width of the surf zone direct measurements o f wave time series
were confined to the inner 150 m of the surf zone.
44

Fig.2. Aerial view of the Goolwa surf zone near the experiment site. Distance from the
subaerial beach to the outer break point is 500 m. Bars are beneath "white" water.

BORE CHARACTERISTICS

Several closely related indices of the d ~ e of surf zone reflectivity versus


dissipativeness and of likely breaker type exist in the literature. Guza and
Inman (1975) and Guza and Bowen (1977) used a surf-scaling parameter of
the form:
e = aw~/g tan2~ (1)
where a is the runup amplitude, ¢oi is the incident wave radian frequency
(2n/T), g is the acceleration of gravity, and ~ is the average beach slope.
They showed t h a t for r e f l e c t i v e conditions conducive to surging breakers
and subharmonic edge wave generation to prevail e should have values less
than ~ 2.0--2.5. High values of e imply dissipative conditions with the degree
of dissipativeness increasing with increasing e. Wright et al. (1979) used e
with ab, the breaker amplitude replacing runup amplitude to differentiate
the reflective vs dissipative nature of different beach states; this definition
is followed here. Battjes (1975) used a different but equivalent form:
tan ~ (~),/~
}= - (2)
(Hb/Lo) ,12

where Hb is breaker height and Lo is deep-water wave length = g T2/2~. Galvin


(1972) studied, experimentally, the relationship between breaker type and
45

Fig.3. Ground view of the Goolwa surf zone at the experiment site showing multiple,
long-crested bores.

+3
2/2/00;GOOLWA,RUNSGSLGS2
+2 ~ ~ I/2/00; G00LWA RUNS G4 GS,
31/I/00;GOOLWA,RUNSGI,G2,G3

~ ro
0 ~ FM--FM--FM ~FMFM- FM~
TD ~u l • '
I F~M..IFM ! "TaFM .FM
-I C , 4 , f ~ i • ' FM.

MOUNT IVK:WJNT;=IVIOIILE ~ "~' ~ ,~U~,u''~'


~2
P~T MOBILE GI 2 "~.'4:56 ST~TW:)NARY
MO~N T I ---- MOCJNT
-io -~o -io -3o-2o-,b 6 16 2o 3o 4b ~ 6o ~o 8o go ,oo .'o ,~o 13o ,4o
METRES

Fig.4. Cross-sectional profile o f the beach and inner surf zone of the experiment site and
locations o f instrument stations. FM indicates f l o w meters, TD indicates pressure trans-
ducers.
46

the parameter:
B b = ( H b / g T 2) tan{J = e tan~/2~ 2 (3)
and found that breakers were of the spilling type when Bb > 0.068.
Battjes (1975) showed that Galvin's data were also applicable to eq.2 and
the spilling criterion is ~ < 0.4 which corresponds to e > 20. In the case of
the Goolwa surf zone where ~ -~ 0.01, Hb ~ 3, and T = 12--15 sec, the
values of these three criteria were: e = 250--400, ~ = 0.085--0.11 and Be =
0.135--0.212. Accordingly breakers observed at Goolwa were exclusively
of the spilling type.
Waves were almost normally incident: the long-crested swell broke by
spilling evenly and virtually simultaneously along crest lengths of up to
500 m. After breaking, the bores retained their long crested forms as they
advanced shoreward (Fig.3). Usually 8--12 bores were present in the surf
zone at any one time. Typical of dissipative surf zones in general, bores at
incident wave frequency progressively decreased in height as they
propagated shoreward accompanied by growth of long-period "surf-beat"
oscillations as illustrated by the time series from different stations shown
in Fig.5.
Variations in bore height at different positions across the surf zone were
estimated from the total variance of 7, a~ in the frequency band
0.033--0.333 Hz (3 < T ,~ 30 sec) using the expression (following Guza and
Thornton, 1980):
Hs = 4o~
where H S is significant height, roughly equal to the height of the highest
1 / 3 of the bores. Two different values of o~ were used: one value was
computed directly from pressure time series and the other was predicted
from the variance of shore normal velocity u, using linear wave theory as
discussed by Guza and T h o r n t o n (1980). The results from different depths
within the surf zone are shown in Table I and Fig.6. Each value of Hs
plotted on the graph was c o m p u t e d from 1800 consecutive observations
made at 1-see intervals (i.e. from runs of 30 rain.). The graph indicates
reasonably close correspondence between Hs values determined from u
with those based on pressure. This suggests that, despite the strong
turbulence and high dissipativeness of the Goolwa surf zone, local relation-
ships between u and pressure conform to linear theory so far as total
variance is concerned. This is consistent with the field observations which
Guza and T h o r n t o n (1980) made on Torrey Pines Beach, California, in a
much narrower surf zone with smaller e (6--60). However, the close
correspondence between pressure and u at s p e c i f i c f r e q u e n c i e s as observed
at Torrey Pines was n o t seen at Goolwa. We are uncertain as to the extent
to which this is due to flow meter response; our ducted flow meters give
only partial response to oscillation periods less than about 6 sec.
An important parameter in studies of breaking waves and surf is the ratio
= H b / h where HD is the local breaker or bore height and h is local depth.
In saturated surf zones, Hb is limited by depth such that 7 remains constant
47

/I

05

to--o

q~ 05

05

0 - to =24 secs X =3 5 m h --O 7 5

, ~ , , I , I , , J I J ~ o ~ I , A , , J
0 50 I~ 150 ~0 2~
seconds
Fig.5. Sample segments of ~ (pressure) time series from three separate stations across the
surf zone (Run G4, 1 Feb. 1980). To facilitate comparisons between the three stations,
the origins of the mid and inner series have been shifted by constant time intervals
roughly equal to the bore travel time (A t) between stations. Individual bores are manifest
as spikes in the record.

and Hb decreases progressively shoreward (within the surf zone) as depth


decreases. This ratio is important not only for predicting surf height relative
to depth but also in equations relating setup at the beach to breaker height
(e.g. Bowen et al., 1968; Bowen, 1969) and for predicting longshore currents
(e.g. Komar 1975, 1976). Bowen et al., (1968) made laboratory observations
of 7 with monochromatic waves of varying steepness in laboratory surf zones
75 cm to 170 cm wide. They found 7 to range from 0.88 to 1.28 with the
48

TABLE I

Bore h e i g h t s a n d ~ values

Run xs h Hs* Hs** Hmax* ms* Ts** 7max*


G1 120 1.90 0.74 0.61 0.99 0.39 0.32 0.52
G2 120 1.90 0.73 0.75 0.96 0.38 0.39 0.50
G3 85 1.20 0.46 0.44 -- 0.38 0.36 --
G4 130 1.88 0.89 -- 1.2 0.47 -- 0.64
G4 65 1.23 0.56 0.34 1.0 0.45 0.28 0.85
G4 35 0.78 0.36 0.35 0.4 0.46 0.44 0.49
G5 130 1.80 0.90 -- 1.3 0.50 -- 0.72
G5 75 1.25 0.57 0.42 0.8 0.46 0.34 0.64
G5 35 0.75 0.35 0.46 0.7 0.47 0.61 0.93
G6 130 1.72 0.84 -- 1.0 0.49 -- 0.58
G6 50 0.92 0.39 0.32 0.5 0.42 0.35 0.54
G6 35 0.67 0.25 0.27 0.3 0.37 0.40 0.45

m e a n 0.44 0.39 0.62

* C o m p u t e d f r o m pressure.
* * C o m p u t e d f r o m u using linear t h e o r y .
( e a c h value is based o n a r e c o r d l e n g t h o f 30 rain.)
x, h a n d H values in m.

lowest values accompanying spilling waves. The majority of subsequent


studies of surf zone dynamics have assumed 0,8 < 7 < 1.2 based largely on
l a b o r a t o r y d a t a ; i t is a c o m m o n "rule of thumb" t o a s s u m e 7 ~ 1. U n t i l
recently there have been limited field data to substantiate or refute this

/.o ~ 05

0.8 3<T<30 sec


I"
xx

i t I l I I I L I i,I
0.2 06 1.0 /'4 1.8 ?'0
DEPTH (h) METRES

Fig.6. Plot of significant bore h e ~ t s , Hs (As computed from pressure and current time
series) against local water depth h. Data points are equivalent to those presented in
Table I.
49

assumption. Suhayda and Pettigrew (1977) observed 7 values as low as 0.6 in


the mid region of a low-energy dissipative natural surf zone. Weishar and
Byrne (1979) report an average value of 7 = 0.78 from a natural beach on
which the majority of the breakers were plunging. It should be noted that
both of those studies involved analyses of photographic records and are
likely to be biased towards the highest breakers.
From Fig.6 and Table I it is clear that at Goolwa, 7, at least within the
region of observations and as pertains to significant bore heights (Hs), was
substantially below the "conventional" monochromatic values with a mean
7s of 0.42. Some a m o u n t of the reduction in 7s is due to the statistical
nature of a wave spectrum (Battjes, 1974). However, even when a 7m~x value
was c o m p u t e d using the highest individual wave for each of four 30 minute
runs at each of 3 locations, the resultant average value was only 0.65.
Although Ts and 7m~ were low, the constancy of 7 between stations
indicates that the surf was saturated.
The low observed 7 values suggest that setup against the beach may be
considerably lower than would be predicted by conventional methods.
Of relevance here are the shore-normal gradients in energy density, E:
E = pgH~/16 (4)
(where p is water density, note that the constant 1/16 is used with H s
instead of the constant of 1/8 which is conventionally used with mono-
chromatic waves since we assume that H~ = V ~ Hrn~) and radiation stress
(excess m o m e n t u m flux), S ~ which in shallow water is:
Sxx = (3/2) E = (3/32) p g H :s (5)
(e.g. Longuet Higgins and Stewart, 1964). Since H decreases shoreward
in the surf zone, E and S ~ also decrease. The shoreward decrease in Sx~
is balanced by a seaward<iirected pressure gradient due to setup in mean
water level ff and:
dSxx dff
dx - --pgh - ~ (6)

on a plane beach of constant/3 the water surface slope is also constant and
is (Bowen et al., 1968):
d~ 1
dx = tan/3 (1 + 1672/3) (7)

For Goolwa, dff/dx would be about 0.00038 assuming a constant bed slope
of tan/3 --0.011. For a 500 m wide surf zone this would suggest a m a x i m u m
setup, ff max at the beach face due to incident wave dissipation of ~ 1 9 cm.
However,/3 was not exactly constant at Goolwa but varied between 0.01 up
to a local m a x i m u m of 0.02. It should also be noted here that it is possible
that 7 was n o t constant seaward of the outermost observation station and
m a y have increased somewhat with proximity to the outer break. In such a
50

case ~ max and d~/dx would have been larger than the estimates given above.
The main point here however, is that the above value of ~ max is much lower
than the value of 1.13 m that would be predicted following the normal
engineering practice (e.g. Svendsen and Jonnson, 1976):
3
~ma~ :'~THb; 7 4" 1 , / t 5 : 3 m (8)

A final point concerning bore behaviour relates to the phase speeds of the
bores within the surf zone. Cross-spectra between outer and mid, and mid
and inner, pressure sensor pairs were calculated for each of 3 runs of 1800
sec length. Typical pressure sensor spectra are shown in Fig.7a. Phase
differences changed linearly with frequency, in the wind wave band,
indicating a non-dispersive system. For each run and sensor pair, mean wave
travel times were calculated using the average travel time of bores in 9
frequency bands in the range 0.04--0.11 ttz as estimated from phase angles
between successive stations. Most points were significant at the 95% level;
two were discarded.
Table II indicates the observed phase speeds C, the phase speed~as
predicted from the mean, (gh) 1/2 between two stations, the ratio C/(gh) ~n
and the ratio C/[gh (1 + 7/2)] ~/2 where 7/2 represents a correction factor
for bore-crest depth. The actual phase speeds of bores propagating
across the surf zone were moderately different from the phase speeds
predicted by linear long-wave theory. It is clear that the ratio ~/(gh) 1/2
was substantially larger over the outer sectors than closer to shore.
There would appear to be a progressive retardation of bore propagation
by factors other than diminishing depth alone as the bores decrease in
height and approach the shore. These results are similar to those obtained
by means of a bore-by-bore comparison of records such as that shown in
Fig.5.
With respect to phase speeds, the Goolwa data contrast with data from
lower-energy and less dissipative surf zones. From observations on Torrey
Pines Beach, T h o r n t o n and Guza (in press) found 0.9 < C / ( g h ) 1/2 < 1.1 using
C values calculated in a fashion similar to those presented in Table II.

TABLE II

Observed and predicted bore phase speeds

Ax (gh) 112 at C= Ax/At C](gh) Ip ~/[gh(l + ~/2)]!P


(m) (m/sec) (sec) (m/sec)

G4 outer-mid 65 3.69 14.2 4.59 1.24 1.13


G5 outer-mid 55 3.84 11.3 4.92 1.28 1.16
G6 outer-mid 80 3.60 20.05 3.98 1.11 1.01
G4 mid-inner 30 3.11 12.11 2.47 0.79 0.72
G5 mid-inner 40 3.10 17.0 2.34 0.76 0.69
G6 mid-inner 25 2.78 8.7 1.73 0.62 0.56
51

/01

A.

--- RUN G4, /NNER' STATION, 7~

i....

.I i0-1

...... - 'd v ~. / ~" \ IX 1%, VFREOUENCK Hz


/0-2
20080 2o /6 /o e 5 PERIOD, sec
ioo 50

/0 2

B.

RUN G4, MID-STATION, U


............. RUN G4, MID-STATION, V
--~ RUN G4, /NNER STATION, LL
lO /

k..

i IO 0

.....
k

i
iO-/
..'"'""/"-/"'"'"::"i
...... J
•..;'...._...._....... .

,.""\........,._
~"~ i\
oo2 o~ "" - FREOUENCY, H z
O'Ol 003 005 00~ O'l Oll~ 0"15 ~: ~o 2
10-2
200 5O40 20 /6 ]0 e 5 PERIOD, sec
IOO

Fig.7. Power spectra of ~, u and v for three stations as computed for 200 lags from 1800
data points recorded at 1 sec intervals (Run G4, 1 Feb. 1980). Outer, mid, and inner
stations were located respectively at 135 m, 65 m, and 35 m seaward of the swash zone.

S u h a y d a and P e t t i g r e w ( 1 9 7 7 ) p e r f o r m e d a w a v e - b y - w a v e analysis o f a
p h o t o g r a p h i c record o b t a i n e d f r o m a l o w - e n e r g y surf z o n e w i t h plunging
breakers. T h e y f o u n d t h a t a q u a n t i t y m ~' C / ( g h ( 1 + 7 / 2 ) ) 1/2 decreased f r o m
values greater t h a n I in t h e o u t e r s u r f z o n e t o a m i n i m u m value o f a b o u t
0.8 in t h e m i d - s u r f z o n e and s u b s e q u e n t l y b e c a m e m u c h greater t h a n 1 again
52

within the inner surf zone and lower swash zone. Since the Goolwa data
were obtained within the inner 1/3 of the surf zone, they would appear to
display a trend opposite to that described by Suhayda and Pettigrew. How-
ever, Suhayda and Pettigrew (1977) and Thornton and Guza (in press) were
observing bores which constituted the major spectral energy. At Goolwa the
bores at inner stations were comparatively minor high-frequency features
riding on energetically dominant surf beat (Fig.7). Furthermore, as will be
discussed shortly, surf-beat amplitudes in the vicinity of the inner Goolwa
stations were on the order of the mean water depth. It is thus quite possible
that the surf beat was responsible for the pronounced phase-speed retardation
in the inner surf zone.

I N F R A G R A V I T Y OSCILLATIONS

The most dramatic feature of the Goolwa surf was the prominence of
large infragravity or "surf b e a t " oscillations (Fig.5). Power spectra, for
(pressure) at the outer (x = 130 m) ~, u and v at the mid (x = 65 m) and
and u at the inner (x = 35 m) stations are presented in Fig.7. The spectra
show abundant energy throughout the infragravity frequency range
(T > 30 sec) b u t with the most prominent peaks occurring in the
frequency band 0.0067--0.017 Hz (60--150 sec). Phase angles of ~/2
between V and u accompanied by high coherence for the frequencies
lower than 0.02 Hz (T > 50) indicate that those oscillations were standing
in the shore normal direction. This is further supported by phase angles
of u for V at outer and inner stations, which suggests a standing wave with
a nodal position lying between the two stations. The higher infragravity
frequencies (30 sec < T < 50 sec) have confused phase relationships and
may have been partially progressive.
The tendency for shoreward dissipation of incident waves to be
accompanied by shoreward growth of lower-frequency infragravity motions
is roughly apparent from Fig.7. This is better illustrated b y Fig.8 which
shows the significant surf beat heights at different stations c o m p u t e d
from the total variance of V and u for all frequencies less than 0.033 Hz
(T > 30 sec). The plot of absolute Hs (T > 30 sec) would seem to imply
relative constancy of surf beat height between outer and inner stations.
However, this effect is the result of the somewhat arbitrary definition
of infragravity motion as consisting of all periods greater than 30 sec,
including those between 30 and 50 sec which were generally non-standing
and experienced shoreward decay. If those periods were eliminated, the
graph would display a progressive shoreward increase. A shoreward increase
in relative height of the surf beat is strongly evident from the curve of the
ratio Hs (T > 30 sec)/Hs (T < 3 0 sec) which rises from the vicinity of 1 at
the outer station to over 2 at the inner station. Standing surf beat was
thus clearly the dominant mode of oscillation in the inner surf zone where
its significant height averaged a b o u t 65 cm with height maxima of a b o u t
1 m. However, since these infragravity oscillations were standing and since
53

/0
A o

®
0.8
® x ® X

06
X X l
X

0.4
H s, T > 30 sec
0.2 x Computed from Pressure Spectra
® Computed from /1 Spectra
i I
0 O2 0!4 06 08 l!O /2 lie /s z!o
2.4 D E P T H (m)

22 X
\
\
B. \®
BO
\ ®
\
~ 1.8 \
x \
\
I.~ /.6 \
x \ ®
x \ ®
\
\
1"4I \
\
\
~ 1.2 \
\
H s (T>3OseCyHs (T<3Osec)
kA
®
x \ \
x x \ \ ®
\
\ ®
\
~8
X xX
0.6 X

I I I I I I I I I
0 "2 04 0"6 0"8 I'0 12 /'4 1"6 1"8 20
DEP TH (m)

Fig.8. Significant height of surf-beat (all periods greater than 30 see) at different depths
across the surf zone. A. Absolute significant surf-beat height. B. Significant surf beat
height (T > 30 sec) expressed as a ratio relative to the significant wind wave and swell
height (T < 30 sec).

the first antinode must be at the beach face, the vertical oscillations there
would have been greater than those at the inner station which was 35 m
seaward of the intersection of mean water level with the beach.
Consistent with observations of Huntley et al. (1977) that runup primarily
expresses the standing motions, these oscillations strongly dominated visual
54

runup on the beach face, although small bores of incident waves penetrated
up the beach superimposed on surf beat crests.
The observation that the Goolwa surf beat was primarily standing is
consistent with other observations of dissipative beaches (e.g. Suhayda,
1974; Huntley, 1976; Holman et al., 1979; Wright et al., 1979) and with
arguments that standing waves of infragravity frequency may control
positions of parallel longshore bars (Short, 1975). Two alternative models
for the standing surf beat exist: they must be either leaky mode standing
waves (e.g. Suhayda, 1974} or trapped edge waves (e.g. Huntley, 1976). In
the leaky mode case, reflected energy is reradiated back to sea without
being trapped inshore and, for normally incident oscillations there is no
shore-parallel motion or amplitude variation. In the edge wave case,
reflected energy is trapped inshore by refraction resulting in shore parallel
as well as shore normal oscillations (e.g. Bowen and Inman, 1969). The
wave number ke (= 2~/Le where Le is longshore edge wave length) of the
longshore oscillation is related to edge wave radian frequency ~e (= 2n/Te
where Te is edge wave period) by Eckart's (1951) dispersion relation for
plane beach profiles:
¢o~ = g k e (2n + 1) tan~ (9)
where n is the mode number (n = 0, 1, 2, 3 . . . ) designating the number
of offshore nodes (zero crossings). Normal to the beach, the amplitudes
of leaky and trapped mode standing waves vary from maxima at the beach
face, manifest as runup (Huntley et al., 1977) to zero at a succession of
nodes which are separated by antinodes of progressively diminishing
amplitude. For leaky mode standing waves, standing edge waves, and
longshore progressive edge waves the velocity potentials ¢ are given res-
pectively by:
ag
¢ = - - J0 [2(¢o 2 x / g tan~) 1/2 ] sin wt (leaky mode) (10)
¢O

¢ = -ag
- e k e x Ln (2kex) sin wt cos key (standing edge wave) (11)
CO

q~ = ~ e--keXLn (2kex) sin (keY - - COet) (progressive edge wave) (12)


G9

where a is amplitude at the beach face, J0 is the zero order Bessel function
and L~ is the La Guerre polynomial of order n. The associated velocities u
and v and water surface elevations 77 are u = O¢/~x, v = O¢/ay, ~ = - - ( l / g ) B ¢ / b t .
Figure 9A, B shows a three-dimensional plot of water surface elevations
of a leaky m o d e standing wave and a mode 3 progressive edge wave each of
100,sec period and having shoreline amplitudes of 50 cm as they should
appear over a surf zone b e d slope simit~ to that which prevailed at Goolwa
(tan~ = 0.011). In actual fact many different frequencies were present so
that in reality the pattern w o u l d have been much more c o m p l e x than those
shown in Fig.9.
55

zero crossmg
j no#e~ 1 ~e

Z, 05

_/ 5000m

o" ..z/ J x 3ooom


0.5

t,~ 0 25

~ ~ . ..3

250 500 750 IO00


METRES

Fig.9. Predicted three dimensional and shore-normal variations in water surface elevations,
~, associated with a 100-sec standing surf-beat, similar to that observed at Goolwa, on a
slope, tan#, o f 0.011. A. Leaky m o d e standing wave. B. Mode 3 edge wave. C. Shore-
normal variations and nodal--antinodal alternations for leaky mode and n = 0, ! , 2, and 3
edge waves.
56

From Fig.9C it is evident that, in the shore normal direction, water


surface variations for leaky modes are similar to those predicted for higher
mode (n > 2) edge waves within the region of observations at Goolwa
although the two solutions became increasingly different at greater distances
from the beach. Unfortunately, the spatial coverage of observation stations
was insufficient to demonstrate which of the two models was applicable to
the Goolwa situation. However, there is mounting evidence to suggest that
at least the longshore components of surf-beat motions are edge waves
(Huntley, 1976; Huntley et al., in press). Both leaky and trapped models
predict shore-normal alternations between nodes and antinodes and are
capable of accounting for the occurrence of multiple parallel longshore
bars which increase in spacing with increasing distance seaward. For edge
waves to produce straight parallel bars (as opposed to crescentic bars) the
edge waves should be progressive rather than standing alongshore (Huntley,
1976). For the natural situation where there are multiple standing wave
frequencies, the offshore succession of nodes and antinodes should appear
at different locations for different frequencies. Such an effect provides the
most obvious explanation for the "valleys" separating infragravity peaks
(Fig.7). However, this does not rule out the possibility that preferential
amplification of certain discrete frequencies may occur under some circum-
stances owing either to frequency selection by the topography, or to
incident wave conditions or a combination of both. Preferential amplifica-
tion was not present in the Goolwa data.

SURF-ZONE CURRENTS AND IMPLICATIONS F O R SEDIMENT TRA N SPO RT

From the point of view of geological processes such as sediment transport


and morphologic change, it is the strength and direction of horizontal
currents that are most important. The surf zone currents which were
observed at Goolwa fall into three general categories: (1) oscillatory flows
due to the incident waves; (2) oscillatory flows due to infragravity oscilla-
tions (surf-beat}; and (3} net circulation. Together, these three types of flows
resulted in strong instantaneous near b o t t o m current velocities with maxima
on the order of 2 m sec -1 at all stations.
B o t t o m current velocities show a shoreward decrease in swell and wind
wave energy accompanied by a shoreward strengthening of low-frequency
surf-beat flows similar to that shown in Fig.5 for pressure. The dominance
of surf-beat oscillations in near-bottom currents is obvious in the sample
current time series shown in Fig.10.
These shoreward increases in infragravity current speeds and decreases
in incident wave currents are quantified by F i g . l l which shows significant
velocity, us (calculated from on-offshore current variance) as a function of
water depth. Three separate values are plotted: a value for the total absolute
Us c o m p u t e d from the total variance encompassing all periods greater than
3 sec; a value for currents related only to oscillations of incident wave and
neighbouring frequencies (30 > T > 3 sec); and a value for current oscillations
57

u (Shore-normGI flow) at I10 cm above bed


: *53 cm sec-I Shoreward

~-+1

A u (Shore normal flow) at 45 abovebed B.

t',
."
3 --2 0 = - 9 cm sec't Seaward /
C.

I
tt (Pressure transducer)
o . . . . . . . . . . . . . . '
I00 200 0 400 500 600 700 900
SECONDS
GOOLWA RUN G2 17.20 31st.JAN.'80
Fig.10. Shore-normal current (u) time series from near-surface and near-bottom sensors
at x = 120 m seaward of the swash zone, h = 1 . 9 m . ( R u n G2, 31 Jan. 1980).

of infragravity frequency (T > 30 sec). The lines on the graph are simply
least-squares fits to the data points constructed to facilitate recognition
o f trends; they do not represent any theory. From the graph, the total us
(all frequencies) at the inner station is seen to be about 1.2 times as large
as at the outer station. A pronounced shoreward decrease in incident wave
contribution was accompanied by an even more dramatic increase in the
strength of infragravity currents with the result that, at the inner station
current velocities associated with infragravity oscillations were 2.2--2.5
times stronger than those induced by incident waves.
The majority of m o d e m approaches to marine sediment transport
generally follow Bagnold (e.g. Bagnold, 1963, 1966) in considering the
rate of sediment transport to depend on some function of the cube of the
instantaneous velocity. Hence the relative importance of the surf beat to
sedimentary processes is considerably greater than indicated by velocity
ratios alone. In terms of relative sand transporting (or agitating) ability,
total b o t t o m currents at the inner station were a b o u t 1.8 times as
effective as those at the outer station if a simple cubic dependence on u,
is assumed. Similarly, at the inner station, the sediment moving capacity
o f surf-beat related currents was 10--15 times greater than that of the
58

/-8

[6

X
/-4
×

®
/2

q~
/Oi
~ k ~ a 3o

: ~ 0.8

06
. I

0.4
x -- Us, all frequencies
• = Us, T > 3 0 sec o ®- Volues estimated from
0"2 Pressure Spectro
=Us, 3 < T < 3 0 sec

I I I I A I I I I I

02 0-4 06 0"8 I'0 12 1"4 16 /8 2"0


DEPTH (h} METRES

Fig.11. Significant shore-normal near-bottom (z = 45 em above bed) current velocities


(Us) at different depths across the surf zone. Values shown are computed for all
frequencies, for wind wave frequencies (T< 30 sec), and surf-beat frequencies
(T> 30 sec).

incident waves. The effect of the infragravity frequencies on suspended


sediment concentration below the lower limit of backwash was strongly
evident in a short time series obtained on 2 February, 1980, using a
Monitek Model 350/136 turbidimeter situated 10 cm above the bed (Fig.12)
The record shows pronounced turbidity maxima separated by intervals of
60--100 sec occurring in association with low-frequency backwash maxima.
Unfortunately, data logger malfunction prevented us from obtaining a
sufficiently long record for meaningful spectral analysis. Nevertheless, the
time series shown in Fig.12 is sufficient to demonstrate that surf beat
dominates suspended sediment concentrations in the lower s w a s h zone/
inner surf zone. Moderately long periods of suspended sediment concentra-
tion maxima in the inner surf zone have also been reported by Brennink-
meyer (1976).
The most distinctive feature of the net, time-averaged c~culation was the
pronounced vertical segregation o f shoreward versus seaward transports.
59

n TURBIDITY

ok
/Z (pressure)

÷lt onshoreSWASH VELOCITY

, , , , I , , . • i . . . . J 1 . . . . I . . . . I . . . . I . . . . I . . . . 1

50 I00 /50 250 3~ 350 400 450


TIME hreco,'~s]

Fig.12. T i m e series of turbidity f r o m c o n t i n u o u s recording using a Monitek Model 3 5 0 /


136 turbidimeter and associated pressure and swash velocities. Pressure and current
sensors were situated at a slightly higher elevation than the turbidimeter. Gaps s h o w n in
u and n records represent bad data due to a c o m b i n a t i o n o f periodic emergence and
fouling o f the f l o w meter b y the slurry-like high sediment concentrations in backwash.
(Run G5-1, 2 Feb. 1980).

From Figs. 10 and 13 it can be seen that net shoreward transport was strong
just below the surface while seaward transport prevailed near the bed. This
vertical segregation of flow is apparently a persistent feature of the Goolwa
surf zone and is well known by local fishermen who utilize it when deploying
their nets to the outer surf zone. It was quickly discovered by our field team
that in order to install instruments at the outermost stations it was necessary
to crawl over the bed using SCUBA assisted by seaward flow. Although
vertically segregated flows similar to that shown in Fig.13 have often been
referred to in text books and in the older literature, many field observations

? .... ? .... ,0 ~.c-f


GOOLWA , S , A 31st, JAN 1980 AVERSE $HOREWAI~D~ SEAWARD VELOCITY
,o,~=-===ffi,,v
'*3 F ~ N E T (lin~eovero~tcl]FLOW

DISTANCE SEAWARD METRES

Fig.13. Cross sectiona] current structure of the Goo|wa surf zone (Runs G-l, G-2, G-3,
31 Jan. 1980).
60

{e.g. Komar, 1976) have suggested that horizontal segregation in the form
o f rip cells is much more common. It may reasonably be inferred that
vertical flow segregation is a characteristic of highly dissipative beaches
but is unlikely to prevail on beaches intermediate between the dissipative
and reflective extremes.
There is relatively little to say about the longshore current except to note
that it set toward the east with net time averaged speeds of 15--20 cm sec -~
It is not possible to relate this to oblique wave incidence since any obliquity
which may have been present was imperceptible. Longshore current velocity
maxima of over 1 m sec -' were recorded but these were associated with
infragravity oscillations (see v spectrum, Fig.7).

CONCLUSIONS

The foregoing is intended to provide a documentation of processes


observed in one high-energy and highly dissipative surf zone. Any generaliza-
tions must be purely tentative at this stage. It must also be emphasized that
these results are only applicable to extremely dissipative surf zones. There
is a comparative abundance of data from reflective and intermediate beach
states which shows that the dynamics of such beaches are different in some
respects from those reported in this paper.
The significance of the Gootwa observations lies in the fact that t h e y
were made under conditions which are morphodynamicaUy similar to those
which probably prevail during storm wave attack on beaches in more
moderate energy environments. Furthermore the Goolwa surf zone must
afford the nearest approximation to the equilibrium condition associated
with high energy waves acting on fine sand beaches since the wave
conditions during the observations were similar to those which persist
more or less year round and since successive beach surveys during the
experiment and at other times (e.g. Short, 1980) showed n ~ i g i b l e change.
It can be inferred t h a t conditions described are those toward which other
fine sand beaches strive to adjust during high energy events.
The comparative conformity of bore phase speeds and local ~ and u
relationships to linear theory, even within such a highly turbulent surf zone,
offers encouragement for predictive modelling of sand transport. However,
the observation that -f was substantially lower than is conventionally
assumed has an important bearing on prediction of bore height and setup
and suggests that 7 may depend on degree of dissipativeness (i.e. on e).
Probably the most important point is that, in the inner surf zone and on
the intertidal beach, it is the surf beat not the individual incident waves
t h a t dominate beach face runup, b o t t o m currents, and restflting sediment
transport.

ACKNOWLEDGEMENTS

This study has been supported by the U.S. Office of Naval Research,
Coastal Sciences Program, Task NR 388-157, Grant N-00014-80-G-0001
61

(LDW and ADS) and Contract N-000014-75-C-0300 (RTG) and by the


Australian Research Grants Committee. Skilled and competent field assis-
tance was provided by M.P. Bradshaw, F.C. Coffey, M.O. Green, G. Holmes,
B. Kjerfve, J. Mackaness, and C. Shaw. M.O. Green also provided extensive
assistance with data reduction and analysis. Figures were prepared by J.
Roberts and M. Rigney.

REFERENCES

Bagnold, R.A., 1963. Mechanics of marine sedimentation. In: M.N. Hill (Editor), The
Sea, V.3: The Earth Beneath the Sea. Wiley, New York, N.Y., pp. 507--582.
Bagnold, R.A., 1966. An Approach to the sediment transport problem from general
physics. U.S. Geol. Surv. Prof. Pap. 442--1, 37 pp.
Battjes, J.A., 1974. Computation of set-up, longshore currents, run-up and overtopping
due to wind generated waves. Communications on Hydraulics, Delft University of
Technology, Rep. 74--2.
Battjes, J.A., 1975. Surf Similarity. Proc. Coastal Engineering Conf., 14th, Copenhagen,
pp.466--479.
Bowen, A.J., 1969. Rip currents I: theoretical investigations. J. Geophys. Res., 74:
5467--5478.
Bowen, A.J. and Guza, R.T., 1978. Edge waves and surf beat. J. Geophys. Res., 83:
1913--1920.
Bowen, A.J., Inman, D.L. and Simmons, V.P., 1968. Wave set-down and set-up. J.
Geophys. Res., 73: 2569--2577.
Bradshaw, M.P., Chappell, J., Hales, R.S. and Wright, L.D., 1978. Field Monitoring and
Analysis of Beach and Inshore Hydrodynamics. Proc. Aust. Coastal and Ocean
Engineering Conf., 4th, (Adelaide), 1978, pp.171--175.
Brenninkmeyer, B., 1976. Sand fountains in the surf zone. In: R.A. Davis and R.L.
Ethington (Editors), Beach and Nearshore sedimentation. Soc. Econ. Paleontol.
Mineral. Spec. Publ., 24: 69--91.
Davies, J.L., 1973. Geographical Variations in Coastal Development, Hafner, New York,
N.Y., 204 pp.
Eckart, C., 1951. Surface waves in water of variable depth. Scrips Inst. Oceanogr., SIO
Wave Report 100, 99 pp.
Galvin, C.J., 1972. Wave breaking in shallow water. In: R.E. Meyer (Editor), Waves on
Beaches. Academic Press, New York, N.Y., pp.413--456.
Guza, R.T. and Bowen, A.J., 1977. Resonant interactions from waves breaking on a
beach. Proc. Int. Conf. Coastal Eng., 15th, pp.560--579.
Guza, R.T. and Davis, R.E., 1974. Excitation of edge waves by waves incident on beach.
J. Geophys. Res., 79(9) : 1285--1291.
Guza, R.T. and Inman, D.L., 1975. Edge waves and beach cusps. J. Geophys. Res.,
80(21): 2997--3012.
Guza, R.T. and Thornton, E.B., 1980. Local and shoaled comparisons of sea surface
elevations, pressures, and velocities. J. Geophys. Res., 85: 1524--1530.
Holman, R.A., Huntley, D.A. and Bowen, A.J., 1979. Infragravity waves in storm
conditions. Proc. Coastal Eng. Conf., 16th, pp.268--284.
Huntley, D.A., 1976. Long period waves on a natural beach. J. Geophys. Res., 81:
6441--6449.
Huntley, D.A. and Bowen, A.J., 1975a. Comparison of the hydrodynamics of steep
and shallow beaches. In: J.R. Hails and A. Cart (Editors), Nearshore Sediment
Dynamics and Sedimentation. Wiley, New York, N.Y., 316 pp.
Huntley, D.A. and Bowen, A.J., 1975b. Field observations of edge waves and their
effect on beach material. J. Geol. Soc. London, 131: 69--81.
62

Huntley, D.A., Guza, R.T. and Bowen, A.J. 1977. A universal form for shoreline run-up
spectra? J. Geophys. Res., 82: 2577--2581.
Huntley, D.A., Guza, R.T. and Thornton, E.B., in press. Observations of progressive edge
waves at surf beat periods. J. Geophys. Res.
Komar, P.D., 1975. Nearshore currents: generation by obliquely incident waves and long-
shore variation in breaker height. In: J.R. Hails and A. Carr (Editor). Sediment
Dynamics and Sedimentation. Wiley, New York, N.Y., pp.17--45.
Komar, P.D., 1976. Beach Processes and Sedimentation. Prentice-Hall, Englewood, N.J.,
429 pp.
Longuet-Higgins and Stewart, 1964. Radiation stress in water waves: a physical discussion
with applications. Deep-Sea Res., 11: 529--562.
Sasaki, T. Horikawa, K. and Hotta, S., 1977. Nearshore Current on a Gently Sloping
Beach. Proc. Int. Conf. Coastal Eng., 15th, 626--644.
Short, A.D., 1975. Multiple offshore bars and standing waves. J. Geophys. Res., 80(27):
3838--3840.
Short, A.D., 1979a. Wave power and beach stages: a global model. Proc. Int. Conf.
Coastal Eng., 16th, pp.1145--1162.
Short, A.D., 1979b. Three dimensional beach stage model. J. Geol., 87: 553--571.
Short, A.D., 1980. Beach response to variations in breaker height. Proc. Int. Conf.
Coastal Eng., 17th, Sydney.
Short, A.D. and Hesp, P., 1980. Coastal engineering and morphodynamics assessment
of the coast within the South East Coast Protection District, South Australia. Fin.
Rep., Coast Protection Board, Adelaide, 234 pp.
Sonu, C.J., Pettigrew, N. and Fredericks, R.G., 1974. Measurements of swash profile
and orbital m o t i o n on the beach. In: Ocean Wave Measurement and Analysis. Am.
Soc. Civ. Eng. 1: 621--638.
Suhayda, J.N., 1974. Standing waves on beaches. J. Geophys. Res., 79: 3065--3071.
Suhayda, J.N. and Pettigrew, N., 1977. Observations of wave height and wave celerity
in the surf zone. J. Geophys. Res., 82: 1419--1424.
Svendsen, I.A. and Jonsson, I.G., 1976. Hydrodynamics of Coastal Regions. Technical
University of Denmark, 282 pp.
Teleki, P.G., Musialouski, F.R. and Prins, D.A., 1976. Measurement techniques of
coastal waves and currents. U.S. Army CERC Misc. Rep. N o . 7 6 - - p . l l .
Thornton, E.B. and Guza, R.T., in press. Wave transformation measured on a shallow
sloping beach. J. Geophys. Res.
Weishar, L.L. and Byrne, R.J., 1979. Field study of breaking wave characteristics. Proc.
Int. Conf. Coastal Eng., 16th, Hamburg, pp. 487--506.
Wright, L.D., Chappell, J., Thorn, B.G., Bradshaw, M.P. and Cowell, P., 1979a. Morpho-
dynamics of reflective and dissipative beach and inshore systems: Southeastern
Australia. Mar. Geol., 32: 105--140.
Wright, L.D., Thorn, B.G. and ChappeU, J., 1979b. Morphodynamic Variability of High
Energy Beaches. Proc. Int. Conf. Coastal Eng., 16th, p p . l l 8 0 - - 1 1 9 4 .
Wright, L.D., 1980. Modes of beach cut in relation to surf-zone morphodynamics. Proc.
Int. Conf. Coastal Eng., 17th, Sydney.

You might also like