AdamsEtAl_ClimChng_20110317 - CLAS Users

advertisement
1
2
Effects of Climate Change and Wave Direction on Longshore Sediment
Transport Patterns in Southern California
3
4
Peter N. Adams
5
6
Department of Geological Sciences, University of Florida, Gainesville, FL 32611, Tel.: 352846-0825, Fax: 352-392-9294, email: adamsp@ufl.edu
7
8
Douglas L. Inman
9
Scripps Institution of Oceanography, University of California San Diego, La Jolla, CA 92093
10
11
Jessica L. Lovering
12
Department of Geological Sciences, University of Florida, Gainesville, FL 32611
13
14
-------------------------------------------------------------------------------------------------------------------
15
Submitted to Climatic Change: 02-27-2010
16
Reviews received: 01-10-2011
17
Revised and returned: 03-04-2011
18
Adams et al., revision for Climatic Change – Mar. 2011
19
20
Abstract
Changes in deep-water wave climate drive coastal morphologic change according to
21
unique shoaling transformation patterns of waves over local shelf bathymetry. The
22
Southern California Bight has a particularly complex shelf configuration, of tectonic origin,
23
which poses a challenge to predictions of wave driven, morphologic coastal change.
24
Northward shifts in cyclonic activity in the central Pacific Ocean, which may arise due to
25
global climate change, will significantly alter the heights, periods, and directions of waves
26
approaching the California coasts. In this paper, we present the results of a series of
27
numerical experiments that explore the sensitivity of longshore sediment transport
28
patterns to changes in deep water wave direction, for several wave height and period
29
scenarios. We outline a numerical modeling procedure, which links a spectral wave
30
transformation model (SWAN) with a calculation of gradients in potential longshore
31
sediment transport rate (CGEM), to project magnitudes of potential coastal erosion and
32
accretion, under proscribed deep water wave conditions. The sediment transport model
33
employs two significant assumptions: (1) quantity of sediment movement is calculated for
34
the transport-limited case, as opposed to supply-limited case, and (2) nearshore wave
35
conditions used to evaluate transport are calculated at the 5-meter isobath, as opposed to
36
the wave break point. To illustrate the sensitivity of the sedimentary system to changes in
37
deep-water wave direction, we apply this modeling procedure to two sites that represent
38
two different coastal exposures and bathymetric configurations. The Santa Barbara site,
39
oriented with a roughly west-to-east trending coastline, provides an example where the
40
behavior of the coastal erosional/accretional character is exacerbated by deep-water wave
-1-
Adams et al., revision for Climatic Change – Mar. 2011
41
climate intensification. Where sheltered, an increase in wave height enhances accretion,
42
and where exposed, increases in wave height and period enhance erosion. In contrast, all
43
simulations run for the Torrey Pines site, oriented with a north-to-south trending coastline,
44
resulted in erosion, the magnitude of which was strongly influenced by wave height and
45
less so by wave period. At both sites, the absolute value of coastal accretion or erosion
46
strongly increases with a shift from northwesterly to westerly waves. These results
47
provide some examples of the potential outcomes, which may result from increases in
48
cyclonic activity, El Niño frequency, or other changes in ocean storminess that may
49
accompany global climate change.
50
Keywords: Climate Change, Coastal Erosion, Longshore Sediment Transport, Southern
51
California
52
1 Introduction
53
In California, 80% of the state's residents live within 30 miles of the coast (Griggs et
54
al., 2005). To mitigate the effects of climate change on coastal communities, it is necessary
55
to assess the oceanographic and geomorphic changes expected within the coastal zone.
56
Effective planning for the future of the California coast will need to draw on climate models
57
that predict the forcing scenarios and coastal change models that predict the coast's
58
response.
59
Coastal landforms exhibit dynamic equilibrium by adjusting their morphology in
60
response to changes in sea level, sediment supply, and ocean wave climate. Global climate
61
change exerts varying degrees of influence on each of these factors. Proxy records indicate
-2-
Adams et al., revision for Climatic Change – Mar. 2011
62
that wave climate has influenced coastal sedimentary accretion throughout the Holocene
63
(Masters, 2006) and, although the causative links between climate change and severe
64
storms are not reconciled in the scientific literature (Emanuel, 2005; Emanuel et al., 2008),
65
it is well-accepted that changes in ocean wave climate (i.e. locations, frequency, and
66
severity of open ocean storms) will bring about changes in the locations and magnitudes of
67
coastal erosion and accretion in the future (Slott et al., 2006). Numerous studies indicate
68
that changes in ocean wave climate are detectable (Gulev and Hasse, 1999; Aumann et al.,
69
2008; Komar and Allan, 2008; Wang et al., 2009), but translating these open ocean changes
70
to nearshore erosional driving forces is complex, and requires an understanding of the
71
interactions between wave fields and the bathymetry of the continental shelf. Along the
72
southern California coast, from Pt. Conception to the U.S./Mexico border, the situation is
73
further complicated by the intricate shelf bathymetry and the presence of the Channel
74
Islands, which prominently interfere with incoming wave field (Shepard and Emery, 1941).
75
In this paper we investigate potential effects of changes in ocean wave climate
76
(wave height, period, and direction) on the magnitudes of coastal erosion and accretion at
77
two, physiographically different sites within the Southern California Bight (SCB). We
78
consider the physical setting of this location, describe our mathematical modeling
79
approach, and present the results of a series of numerical experiments that explore a range
80
of wave climates. Lastly, we discuss the implications of the modeled coastal behavior in
81
light of some possible scenarios of global climate change.
-3-
Adams et al., revision for Climatic Change – Mar. 2011
82
2 Geomorphic and Oceanographic Setting
83
For the purposes of this study, we consider the SCB to extend from a northwestern
84
boundary at Point Arguello (34.58˚ N, 120.65˚ W, Fig. 1) to the U.S.-Mexico border (32.54˚
85
N, 117.12˚ W, Fig. 1) south of San Diego. Tectonic processes along this active margin,
86
between the Pacific and North American plates, are responsible for shaping the shallow
87
ocean basins, continental shelf, and large-scale terrestrial landmasses (Christiansen and
88
Yeats, 1992). The edges of the plates on either side of the boundary have been folded and
89
fractured by transpressional plate motions, creating the high relief terrestrial landscape,
90
pocket beaches backed by resistant bedrock sea cliffs, narrow continental shelf, deeply
91
incised submarine canyons, and irregularly shaped submarine basins that are
92
characteristic of a collisional coasts, as classified by (Inman and Nordstrom, 1971). In
93
particular, the coastal mountain ranges and local shelf basins have been constructed by
94
crustal displacement along a network of subparallel strike-slip faults, which characterize
95
the plate interface (Hogarth et al., 2007). In general, these motions have resulted in the
96
highly irregular, complex bathymetry that makes up the California Borderlands (Legg,
97
1991), that feature the Channel Islands, as well as numerous submerged seamounts and
98
troughs (Fig. 1).
99
The wave climate of Southern California has been extensively studied since the
100
pioneering investigations that applied the theoretical relationships of wave transformation
101
to predict breakers and surf along the beaches of La Jolla, California (Munk and Traylor,
102
1947; Sverdrup and Munk, 1947). Buoys maintained by the National Oceanic and
103
Atmospheric Administration (NOAA) have greatly assisted understanding of deep-water
-4-
Adams et al., revision for Climatic Change – Mar. 2011
104
wave conditions within the SCB (O’Reilly et al., 1996). Monitoring efforts continued to be
105
improved through the development of the Coastal Data Information Page (CDIP) program,
106
Scripps Institution of Oceanography, which provides modeled forecasts at a number of
107
locations. Within the SCB, the presence of the Channel Islands (Fig. 1) significantly alters
108
the deep-water (open ocean) wave climate to a more complicated nearshore wave field
109
along the Southern California coast. The islands intercept waves approaching from almost
110
any direction and the shallow water bathymetry adjacent to the islands refracts and
111
reorients wave rays to produce a complicated wave energy distribution along the coast of
112
the Southern California mainland. Several studies have targeted the sheltering effect of the
113
Channel Islands within the SCB and the complexity of modeling wave transformation
114
through such a complicated bathymetry (Pawka et al., 1984; O’Reilly and Guza, 1993;
115
Rogers et al., 2007). It has also been documented that wave reflection off sheer cliff faces in
116
the Channel Islands can be a very important process in the alteration of wave energy along
117
the mainland coast (O’Reilly et al., 1999). The resulting distribution of wave energy at the
118
coast consists of dramatic longshore variability in wave energy flux and radiation stress.
119
These factors are considered to be fundamental in generating the nearshore currents
120
responsible for longshore sediment transport and the maintenance of sandy beaches.
121
Some studies highlight evidence for changing storminess and wave climate in the northeast
122
Pacific Ocean (Bromirski et al., 2003; 2005). Recently, (Adams et al., 2008) examined a 50-
123
year numerical hindcast of deep-water, winter wave climate in the bight, to understand the
124
correlation of decadal-to-interannual climate variability with offshore wave fields. Their
125
study found that El Niño winters during Pacific Decadal Oscillation (PDO) warm phase have
126
significantly more energetic wave fields than those during PDO-cool phase, suggesting an
-5-
Adams et al., revision for Climatic Change – Mar. 2011
127
interesting connection between global climate change and coastal evolution, based on
128
patterns of storminess.
129
3 Model Description
130
The numerical model employed to evaluate potential coastal change consists of two
131
components: (1) a spectral wave transformation model, known as SWAN (Booij et al.,
132
1999), that calculates shoaling and refraction of a proscribed deep water wave field over a
133
defined bathymetric grid, and reports coastal wave conditions, and (2) an empirically-
134
derived longshore sediment transport formulation, referred to herein as CGEM (Coastal
135
Geomorphic Erosion Model), that utilizes the coastal wave conditions derived by the wave
136
transformation model to compute divergence of volumetric transport rates of nearshore
137
sediment, also known as divergence of drift (Inman, 1987; Inman and Jenkins, 2003). This
138
divergence of drift is the difference between downdrift and updrift volumetric transport
139
rates (sediment outflow minus sediment inflow), and represents the volume of
140
sedimentary erosion or accretion at a coastal compartment over the model time step. The
141
interaction between the two components of the model is shown schematically in Figure 2.
142
Two significant assumptions, and one model limitation, are invoked to simplify
143
calculations. First, wave transformation is calculated to the fixed 5-meter isobath, which is
144
usually seaward of the wave break point. Although the sediment transport model calls for
145
wave conditions at the break point, wave breaking proceeds over a breaker zone, that can
146
be several tens of meters wide, depending on the slope of the beach. Through several tests
147
of the SWAN wave model, we have determined that, under vigorous deep water wave
148
conditions (i.e. Hs=4-5 meters and T=14 s), wave breaking initiates over a depth of 9-10
-6-
Adams et al., revision for Climatic Change – Mar. 2011
149
meters with the percentage of wave breaking increasing shoreward. We consider the mean
150
water depth within this breaker zone (4.5-5 m) to be a reasonable representative value to
151
use for breakpoint conditions in our modeling procedure. Second, the longshore sediment
152
transport formulation assumes a transport-limited case, as opposed to a supply-limited
153
case. The transport-limited assumption results in the calculation of potential divergence of
154
drift, which would reflect the case of an inexhaustible supply of nearshore sediment. If
155
terrestrial sediment supply to the coast is limited, from decreased riverine inputs for
156
example, this assumption may be challenged. Lastly, the two components of the model are
157
not yet backward coupled, meaning that nearshore bathymetry is not updated after a
158
calculation of divergence of drift is conducted. Hence, the model should not be run over a
159
time series of changing wave conditions to simulate the evolutionary behavior of a coast.
160
In this paper, we present the results of instantaneous scenarios of divergence of drift for
161
individual sets of deep-water wave conditions.
162
3.1 Wave Transformation Model
163
SWAN, short for Simulating WAves Nearshore, is a 3rd generation, finite-difference,
164
wave model that operates on the principle of a wave action balance (energy density divided
165
by relative frequency) (Holthuijsen et al., 1993; Booij et al., 1999). In the absence of wind
166
forcing, this model requires only two general inputs to perform a computation of the wave
167
field throughout a region - bathymetry and deep-water wave conditions.
168
Ocean bathymetry exhibits a strong control on the direction and rate of wave energy
169
translation. When the water depth is shallower than half the wavelength, interaction of
170
wave orbitals with the sea floor causes shoaling transformation and refraction of waves, so
-7-
Adams et al., revision for Climatic Change – Mar. 2011
171
the spatial pattern of nearshore wave energy depends strongly on the distribution of
172
seafloor elevation. A review of some studies that investigated how bathymetric changes
173
influence shoreline change by altering shoaling and refraction patterns is provided by
174
(Bender and Dean, 2003). In this study, we used two bathymetric grids for the wave
175
transformation modeling, both obtained from the National Geophysical Data Center
176
(NOAA) U.S. coastal relief model grid database
177
(http://www.ngdc.noaa.gov/mgg/coastal/coastal.html). A spatially-coarse grid (30 arc-
178
sec) was used to evaluate wave conditions over the entire bight (32˚-35˚N, 121˚-117˚W),
179
providing 480 x 360 cell matrices of wave heights and directions covering approximately
180
124,000 km2, in which each value represents the conditions for a 0.72 km2 area of sea
181
surface. Two smaller, high resolution grids (3 arc-sec), referred to herein as “nests”, were
182
used to evaluate local wave conditions on a finer spatial scale. The nests named SntBrb and
183
TorPns, for portions of the Santa Barbara and Torrey Pines coasts respectively, each occupy
184
approximately 700 km2 within the domain of the main (coarse) grid (Fig. 3), resulting in
185
each cell of output representing wave conditions over a 0.0072 km2 area of sea surface.
186
At the western and southern margins of the coarse grid, deep-water wave
187
conditions are proscribed as boundary conditions, consisting of significant wave height,
188
peak period with a JONSWAP frequency distribution (Hasselmann et al., 1973), and peak
189
direction with a cosine square spread of 15 degrees. The wave field is computed from the
190
grid boundaries over the input bathymetry to the coast. Example results from a SWAN run
191
for the coarse grid are provided in Figure 3. From the SWAN output of the coarse grid, the
192
boundary conditions for the individual nests were obtained (examples provided in Figures
193
4 and 5). Because the goal of the investigation was to examine "snapshots" of longshore
-8-
Adams et al., revision for Climatic Change – Mar. 2011
194
distribution of erosion/accretion patterns resulting from various deep-water wave
195
scenarios, all SWAN runs conducted in this study were performed in stationary mode.
196
Temporally evolving wave fields, such as those produced during a storm were not
197
examined, thereby avoiding the problem of swell arrival timing often associated with
198
stationary runs (Rogers et al., 2007).
199
3.2 Longshore Sediment Transport Model
200
The output passed from the wave transformation model to the longshore sediment
201
transport model includes the complete oceanic grids of significant wave height and peak
202
directions at each “wet node” within the particular nest being examined. CGEM queries and
203
interpolates the SWAN output at a known set of locations within the nest that constitute
204
the 5-meter isobath, using an alongshore spacing of 100 meters. These longshore sets of
205
wave height and direction form the basis of the sediment transport computation, which
206
originates from a semi-empirical formulation originally proposed by Komar and Inman
207
(1970). This relationship was later modified and named the CERC formula (Rosati et al.,
208
2002), which has been used in several coastal evolution modeling studies recently (Ashton
209
et al., 2001; Ashton and Murray, 2006). The CERC formula has been tested and found to
210
provide results in close agreement with processed based models (Haas and Hanes, 2004).
211
In its most general form, the volumetric longshore sediment transport rate, Ql, can be
212
expressed as
Ql 
213
Il
( s   w )gN o

-9-
(1)
Adams et al., revision for Climatic Change – Mar. 2011
214
where Il is the immersed-weight transport rate (Inman and Bagnold, 1963), s and w are
215
densities of quartz sediment (2650 kg/m3) and seawater (1024 kg/m3), respectively, g is
216
the gravitational acceleration constant (9.81 m/s2), and No is the volume concentration of
217
solid grains (1 - porosity), set to 0.6, for all numerical experiments in this study. The
218
subscript l is used to represent the longshore component of any variable with which it is
219
associated.
220
3.2.1 Angle of Incidence
221
The angle of incidence is of primary importance in the calculation of longshore
222
sediment transport. If wave rays approach the beach at an angle perfectly orthogonal to the
223
trend of the coast, the longshore component of wave energy flux is zero, and there is no net
224
longshore current to drive longshore sediment transport. If wave rays approach the beach
225
at an oblique angle (somewhere between orthogonal and parallel), there is a component of
226
wave energy flux parallel to the shoreline, which drives longshore sediment transport.
227
Along the 5-meter isobath, the coastal orientation is computed by a downcoast-
228
moving sliding window computation, which fits a trendline to 5 adjacent points and reports
229
an azimuth value for the midpoint of the segment. The coastal orientation is subtracted
230
from the queried wave direction along the isobath to provide an angle of incidence.
231
3.2.2 Wave Energy Flux
232
The longshore component of wave energy flux is considered to provide the fluid
233
thrust required to move sediment under the influence of the breaking wave bore. The
234
governing equation is
-10-
Adams et al., revision for Climatic Change – Mar. 2011
235
1
Pl  ECn   w gH 2Cnsin  cos 
8
236
237
where E is wave energy density, C is nearshore wave celerity, which is depth controlled, n is

ratio of group to individual wave speed (~1 in shallow water and 1/2 in deep water), H is
238
nearshore (breaking) wave height,  is the angle of incidence, the trigonometric
239
components of which result from the tensor transformation of the onshore flux of
240
longshore directed momentum (LonguetHiggins and Stewart,1964). As noted above, we
241
estimate breaking wave height by evaluating this quantity at the 5-meter isobath. The
242
immersed weight transport rate is simply a scaled version of the longshore component of
243
wave energy flux
(2)
244
Il  KPl
245
246
where the scaling parameter, K, is set to 0.8 in this study. It is noted that Komar and Inman

(1970) tested this relationship at different locations, where beach sediment sizes differed,
247
but found that the relationship held, irrespective of sedimentary texture.
248
3.2.3 Divergence of Drift
249
After presenting the relationships for the longshore component of wave energy flux,
250
immersed weight transport rate, and volumetric longshore sediment transport rate, we can
251
show that a simple relationship for the divergence of drift is obtained by applying the
252
vector differential operator to the volumetric longshore sediment transport rate using a
253
dot product,
 Ql 
254

-11-
Ql
x
(3)
(4)
Adams et al., revision for Climatic Change – Mar. 2011
255
256
where x is the position along the coast or, in CGEM, position along the 5-meter isobath.
This quantity is, effectively, the calculated change in sediment volume over the
257
longshore reach dx, during the time interval that these wave conditions are applied.
258
Herein, we adopt the sign convention that divergence is defined as the net difference
259
between sediment inflow and sediment outflow, making positive divergence of drift (where
260
inflow exceeds outflow) result in accretion at a site, whereas negative divergence of drift
261
(where outflow exceeds inflow) results in erosion. The longshore pattern of divergence of
262
drift is therefore out of phase with longshore sediment transport, as expected. At
263
longshore positions where transport is increasing at the greatest rate (positively sloping
264
inflection points), divergence of drift is at a local minimum; where transport is decreasing
265
at the greatest rate (negatively sloping inflection points), divergence of drift is at a local
266
maximum; where transport is at a local minimum or maximum, divergence of drift should
267
be zero.
268
4 Numerical Experiments and Results
269
To provide insight on how climate change-driven alteration of deep water wave
270
conditions might affect the magnitude of erosion and accretion along the Southern
271
California coast, we conducted a series of controlled numerical experiments at two
272
physiographically-distinct, reaches of the Southern California coast, which we refer to as
273
the Santa Barbara (SntBrb nest) and Torrey Pines (TorPns nest) sites. The goal of these
274
experiments is to test the hypothesis that deep water wave direction exhibits critical
275
control on the longshore sediment transport patterns at coastal sites within the
276
bathymetrically complex SCB. Each of the locations chosen witness unique swell patterns
-12-
Adams et al., revision for Climatic Change – Mar. 2011
277
resulting largely from their relative orientations with respect to the deep water wave field
278
of the North Pacific Ocean, each site's local shelf bathymetry, and the blocking patterns that
279
the Channel Islands provide to each site (Fig. 3). The two sites represent end member
280
coastal orientations within the Bight; the SntBrb nest exhibits a west-to-east general
281
shoreline orientation, whereas the TorPns nest exhibits a north-to-south general shoreline
282
orientation. For each experiment, we conducted 104 SWAN-CGEM simulations that vary
283
deep water wave direction from 260˚ to 320˚ in 5˚ intervals for four pairs of wave height
284
period scenarios: (1) H=2 m, T=12 s, (2) H=2 m, T=16 s, (3) H=4 m, T=12 s, (4) H=4 m, T=16
285
s. These ranges span the distributions of deep water wave conditions documented for the
286
SCB by Adams et al. (2008).
287
4.1 Site 1 - Santa Barbara
288
The western end of Goleta Beach, adjacent the UCSB campus in southeastern Santa
289
Barbara county, California, has been the site of regular nourishment due to chronic sand
290
loss and the community desire to maintain a recreational beach. The site has witnessed
291
profound changes in beach width, morphology, and sediment volume over the past 30
292
years with anecdotal photo histories documenting wide, well-vegetated, sandy beaches in
293
the 1970's, which were fully inundated during the El Niño winters of 1982-83 and 1997-98
294
(Sylvester, 2010). Waves entering the Santa Barbara channel, as swell, have been modeled
295
by (Guza et al., 2001), and are regularly forecasted by the Coastal Data Information
296
Program, CDIP.
-13-
Adams et al., revision for Climatic Change – Mar. 2011
297
298
4.1.1 SWAN Transformed Wave Field
SWAN output from a moderate, westerly swell (H=2 m, T=12 s,  = 270˚) is shown in
299
Figure 4. The wave height distribution pattern for the entire SCB (Fig. 4A) illustrates the
300
blocking effect of the Channel Islands. Fig. 4B shows the decrease (by more than half) in
301
wave height as waves enter shallow water. Fig. 4C shows the significant amount of
302
refraction that occurs as waves approach the nearshore and the significant sheltering
303
experienced by the Goleta Beach site, herein referred to as SB-2 (blue star), as compared to
304
the exposed site SB-1 (red star) located immediately west of Goleta Point, 1 km from SB-2.
305
4.1.2 CGEM Results
306
It is instructional to observe the results of two CGEM simulations plotted along
307
shore in the vicinity of SB-1 and SB-2. Figure 6 shows the strong influence exerted by wave
308
direction at the Santa Barbara site. Output from two SWAN simulations are passed to
309
CGEM to examine longshore patterns of potential sediment transport rate and divergence
310
of drift. For each SWAN simulation, deep water wave height and period are set to 4.0 m
311
and 16 s, but the deep water wave direction is 320˚ (northwesterly) in case A,
312
representative typical La Niña storm wave conditions, as opposed to 270˚ (westerly) in
313
case B, representative of El Niño storm wave conditions, during which time the jet stream
314
occupies a more southern position than usual due to the anomalous atmospheric pressure
315
distribution (Storlazzi and Griggs, 2000). Comparison of the two deep water input cases is
316
as follows. Along the 5-meter isobath, the significant wave height for Case A
317
(northwesterly) is very small (<0.5 m), in comparison to deep water inputs conditions
318
(Hsig=4 m), everywhere along the 5 km reach. For Case B, the wave heights are
-14-
Adams et al., revision for Climatic Change – Mar. 2011
319
substantially higher, approximately 2 meters everywhere along the reach. The angle of
320
incidence varies for both Case A and Case B, but in the same pattern, developing a sizable
321
longshore component of wave energy flux. The longshore sediment transport rates vary in
322
much the same manner as wave heights for the two cases, with appreciable transport in
323
Case B, and negligible transport in Case A. The potential divergence of drift pattern for
324
Case B shows loss of sediment (erosion) at SB-1 (red star) and gain of sediment (accretion)
325
at SB-2 (blue star). Potential divergence of drift is negligible along the length of the 5 km
326
reach for Case A.
327
4.1.3 SntBrb Experiment
328
The 104 SWAN-CGEM simulations which were run for the SntBrb experiment
329
produced divergence of drift patterns which are reported for SB-1 and SB-2 in Figure 7.
330
This compendium of experiment results illustrates that the exposed SB-1 site experiences
331
increasing erosion as wave conditions become more westerly. This is in contrast to the
332
sheltered SB-2 site, which becomes more accretionary as wave direction becomes more
333
westerly. Increasing deep-water wave height causes enhancement of erosional or
334
accretional behavior at both SB-1 and SB-2, depending on the site tendency under milder
335
conditions. Increasing wave period causes enhanced erosion at SB-1, but causes decreased
336
accretion at SB-2.
337
4.2 Site 2 - Torrey Pines
338
Torrey Pines beach, located approximately 7 km north of Scripps Pier in La Jolla,
339
California, has been the site of many scientific inquiries in the field of coastal processes
340
(Thornton and Guza, 1983; Seymour et al., 2005; Yates et al., 2009). The relatively straight,
-15-
Adams et al., revision for Climatic Change – Mar. 2011
341
north-south trending reach resides within the Oceanside littoral cell and owes any
342
longshore variation in wave energy flux to the blocking effects of the Channel Islands,
343
rather than to complexities of nearshore bathymetry, save for the areas around the Scripps
344
and La Jolla submarine canyons in the southern portion of the TorPns nest.
345
4.2.1 SWAN Transformed Wave Field
346
As for the Santa Barbara site discussed above, we show the behavior of the Torrey
347
Pines site to a SWAN simulation for a moderate, westerly swell (H=2 m, T=12 s,  = 270˚) in
348
Figure 5. The demonstrable change in wave height visible around 33.05˚ north latitude in
349
Figure 5B is a result of waves penetrating through a window between Santa Catalina and
350
San Clemente Islands during periods of westerly swell. The general shore-normal
351
orientation of the wave field (for input conditions shown) promotes nearshore wave height
352
increase as a result of shoaling in the absence of refraction.
353
4.2.2 CGEM Results
354
Comparative examples of CGEM simulations from the TorPns nest for Cases A and B
355
(described above in Section 4.1.2) are given in Figure 8. Just as for the SntBrb nest, the
356
significant wave height for Case A along the 5-meter isobath in the vicinity of Torrey Pines
357
sites TP-1 and TP-2 is very small (<1.0 m). However, in the TorPns nest, this may be due to
358
blockage of waves by the Channel Islands rather than to severe refraction as in the SntBrb
359
nest. Angles of incidence for both case A and B at TorPns are small, approximately less than
360
+/- 5˚. For case B, between kilometer markers 23 and 25.7, angle of incidence is negative
361
which results in northward-directed longshore sediment transport pattern in this region,
362
as opposed to the southward directed transport elsewhere in the nest. For case A,
-16-
Adams et al., revision for Climatic Change – Mar. 2011
363
divergence of drift pattern is negligible everywhere within the 5 km span surrounding TP-1
364
and TP-2, whereas for case B, a strongly negative potential divergence of drift (erosion)
365
emerges at Torrey Pines Beach, near TP-1 and TP2.
366
4.2.3 TorPns Experiment
367
As for the SntBrb nest discussed in Section 4.1.3, the 104 SWAN-CGEM simulations,
368
which were run for the TorPns experiment produced divergence of drift patterns which are
369
reported for TP-1 and TP-2 in Figure 9. All simulations run for the Torrey Pines site
370
resulted in erosion. At TP-1, the peak in magnitude of potential divergence of drift occurs
371
when waves are just north of westerly (275˚ - 290˚). Increases in wave period slightly
372
increased erosion for 2 meter waves, and shifted the peak direction for maximum erosion
373
for 4 meter waves, at TP-1. At TP-2, wave height has a profound influence; doubling of the
374
deep water wave height causes potential divergence of drift to approximately triple across
375
the directional range. Increasing wave period, however, had negligible effect at site TP-2.
376
5 Summary and Implications
377
A numerical modeling procedure for assessing the patterns of littoral sediment
378
transport in Southern California has been presented. The procedure combines a spectral
379
wave transformation model with a calculation of gradients (divergence) in longshore
380
sediment transport rates, assuming transport-limited conditions. To illustrate some
381
specific coastal impacts resulting from climate change, we have applied this procedure to
382
two physically-distinct sites within the SCB. We conducted a sensitivity analysis at the two
-17-
Adams et al., revision for Climatic Change – Mar. 2011
383
study sites, whereby effects of variability in deep water wave direction were explored for
384
four wave height / wave period combinations.
385
This study demonstrates that the longshore sediment transport patterns in the
386
littoral zone, and therefore the locations of erosional hotspots, along the Southern
387
California coast are extremely sensitive to deep water wave direction. We speculate that
388
this sensitivity is due to two principle reasons: (1) the severe refraction required by
389
northwesterly waves in order to be incident upon south facing coasts (e.g. SntBrb nest),
390
which could also be considered a sheltering effect of Pt. Arguello, and (2) the sheltering
391
effects of the Channel Islands (blocking of incoming swells and additional refraction of
392
grazing waves) on west facing coasts (TorPns nest). Observations of wave interruption by
393
islands were made decades ago by Arthur (1951).
394
Although the cross-shore transport of sediment in the littoral zone is not specifically
395
addressed in these numerical experiments, we acknowledge its potential importance
396
during large wave events. Long-period swells, often associated with large wave events, will
397
increase refraction, which increases the cross-shore component of wave energy flux.
398
Associated high wave set-up can promotes off-shore transport and the temporary storage
399
of littoral sediment in offshore bars. Despite the fact that cross-shore transport can be
400
strong during events, the associated “erosion” is often temporary, as recovery proceeds
401
relatively rapidly after a large wave event via onshore transport of bar sediment.
402
It is noteworthy that the Santa Barbara wave direction experiment illustrated how
403
the divergence of drift at a site exposed to the open ocean (SB-1) experiences enhancement
404
of erosion as the wave field intensifies (increased wave heights and periods), whereas the
-18-
Adams et al., revision for Climatic Change – Mar. 2011
405
divergence of drift at a site locally sheltered by a headland (SB-2) experiences
406
enhancement of accretion as the deep water wave heights are increased. The fact that the
407
sheltered site experienced slight decreases in accretion for increased wave periods may be
408
due to the higher degree of refraction that the longer period swells undergo before
409
reaching the coast.
410
The Torrey Pines experiment reveals interesting behavior regarding the direction of
411
longshore transport and erosion. As shown in Figure 8, for a westerly wave field of strong
412
intensity (Hs = 4 m, T = 16 s, a = 270˚), the angle of incidence between position markers 23-
413
25.75 km is negative, but changes to positive between position markers 25.75-28 km. The
414
result of this change in angle of incidence is a change in direction of longshore transport
415
from northward to southward. However, divergence of drift is negative for this set of deep
416
water wave conditions at both TP-1, where transport direction is northward, and TP-2,
417
where transport direction is southward. This persistence of erosional character at this site,
418
irrespective of transport direction, suggests that continental shelf bathymetry may exert a
419
unique control on nearshore wave fields at this site.
420
It is observed that at both SB-1 (SntBrb nest) and TP-2 (TorPns nest), negative
421
divergence of drift (erosion) appears to operate under all deep water conditions simulated.
422
This brings up two questions: (1) Why does the coast at SB-1 protrude seaward relative to
423
the coast at SB-2, if the SB-1 shoreline is retreating under all simulated conditions? (2) Why
424
does the Torrey Pines coastline maintain a relatively straight appearance if the shoreline at
425
TP-2 is consistently retreating more rapidly than the shoreline at TP-1? Several
426
explanations are offered to address these discrepancies and we suspect the answer is a
-19-
Adams et al., revision for Climatic Change – Mar. 2011
427
combination of these. First, as mentioned above, the erosion/accretion portion of the
428
CGEM model assumes transport-limited conditions, meaning that deficiencies in sediment
429
supply are not considered to play a role in coastal landform evolution. If sediment supply
430
is limited at these sites, then the model may overestimate the magnitude of divergence of
431
drift. Second, the model only considers the movement of sediment alongshore at these
432
coastal sites and does not address the rocky cliffs that back the beaches along much of the
433
Southern California coast. During high wave conditions when sediment supply is limited, it
434
is quite likely that beach sand is temporarily stored offshore in bars and waves directly
435
impact the bedrock cliffs, whose retreat is not governed by Equation (4) above. To
436
simulate shoreline retreat of an exposed cliffed coast, bare bedrock cutting processes must
437
be adequately modeled. Neither of these caveats, however, changes the fact that the
438
gradients in potential longshore sediment transport patterns are highly sensitive to deep
439
water wave conditions, and particularly, to wave direction.
440
These results have implications regarding climate change, longshore sediment
441
transport patterns, and the distribution of hotspots of coastal erosion within the SCB. Our
442
numerical simulations illustrate a dramatic increase in absolute value of divergence of drift
443
as wave climate approaches westerly deep water wave directions. It has been documented
444
that during El Niño winters, waves entering the Southern California Bight tend to be more
445
westerly than during non El Niño winters (Adams et al., 2008). Therefore, model results
446
presented herein are consistent with the observations of severe coastal change in
447
California during the El Niño winters of 1982-83, and 1997-98 (Storlazzi et al., 2000).
448
Recent research investigating tropical cyclonic behavior during the early Pliocene (5-3 ma)
449
reveals a feedback that may serve to increase hurricane frequency and intensity in the
-20-
Adams et al., revision for Climatic Change – Mar. 2011
450
central Pacific during warmer intervals (Fedorov et al., 2010). This is relevant to Earth’s
451
current climate trend because the early Pliocene is considered a possible analogue to
452
modern greenhouse conditions. Federov et al. (2010) provide results of numerical
453
simulations of tropical cyclone tracks in early Pliocene climate, which illustrate a dramatic
454
poleward shift in sustained hurricane strength within the eastern Pacific Ocean. This
455
pattern results in increased westerly storminess, implying that warmer climates will cause
456
the SCB to witness greater absolute values of divergence of drift. In other words, a more
457
volatile coastline, exhibiting higher magnitude erosion and accretion, might be expected as
458
a result of increased frequency of strong westerly waves.
459
-21-
Adams et al., revision for Climatic Change – Mar. 2011
460
References
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
Adams P, Inman D, Graham N (2008) Southern california deep-water wave climate:
Characterization and application to coastal processes. Journal of Coastal Research
24(4): 1022–1035
Arthur R (1951) The effect of islands on surface waves. Bulletin of the Scripps Institution of
Oceanography p 6no1
Ashton A, Murray A (2006) High-angle wave instability and emergent shoreline shapes: 1.
modeling of sand waves, flying spits, and capes. Journal of Geophysical ResearchEarth surface 111(F4):F04,011
Ashton A, Murray A, Arnoult O (2001) Formation of coastline features by large-scale
instabilities induced by high-angle waves. Nature 414(6861):296–300
Aumann H, Ruzmaikin A, Teixeira J (2008) Frequency of severe storms and global warming.
Geophysical Research Letters 35(19) 1516
Bender C, Dean R (2003) Wave field modification by bathymetric anomalies and resulting
shoreline changes: a review with recent results. Coastal Engineering 49(1-2):125–153
Booij N, Ris R, Holthuijsen L (1999) A third-generation wave model for coastal regions. 1.model description and validation. Journal of Geophysical Research 104(C4):7649–
7666
Bromirski P, Flick R, Cayan D (2003) Storminess variability along the california coast:
1858–2000. Journal of Climate 16(6):982–993
Bromirski P, Cayan D, Flick R (2005) Wave spectral energy variability in the northeast
pacific. J Geophys Res 110:C03,005
Christiansen, R.L. and Yeats, R.S., 1992, Post-Laramide geology of the U.S. Cordilleran
region, in Burchfiel, B.C., Lipman, P.W., and Zoback, M.L., eds., The Cordilleran orogen:
Conterminous U.S.: The geology of North America [DNAG] Vol. G-3: Geological Society
of America, p. 261–406.
Emanuel K (2005) Increasing destructiveness of tropical cyclones over the past 30 years.
Nature 436(7051):686–688
Emanuel K, Sundararajan R, Williams J (2008) Hurricanes and global warming. Bull Amer.
Meteorol Soc 89(3):347–367
Federov, A. V., Brierley, C. M., and Emanuel, K. (2010) Tropical cyclones and permanent El
-22-
Adams et al., revision for Climatic Change – Mar. 2011
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
Niño in the early Pliocene epoch. Nature v. 463, pp. 1066-1070.
Griggs GB, Patsch K, Savoy LE (2005) Living with the changing california coast p 540
Gulev S, Hasse L (1999) Changes of wind waves in the north atlantic over the last 30 years.
International Journal of Climatology 19(10):1091–1117
Guza R, O’Reilly W, Region USMMSPO, of California U, Institute SBMS (2001) Wave
prediction in the santa barbara channel: final technical summary, final technical
report. US Dept of the Interior, Minerals Management Service, Pacific OCS Region p 16
Haas, K. and Hanes, D. M., Process based modeling of total longshore sediment transport.
(2004) Journal of coastal research vol. 20 (3) pp. 853-861
Hasselmann K, Barnett T, Bouws E, Carlson H, Cartwright D, Enke K, Ewing J, Gien- app H,
Hasselmann DE, Meerburg A, Mller P, Olbers D, Richter K, Sell W, Walden H (1973)
Measurements of wind-wave growth and swell decay during the joint north sea wave
project (jonswap). Ergnzungsheft zur Deutschen Hydrographischen Zeitschrift Reihe
8(12):95
Hogarth L, Babcock J, Driscoll N, Dantec N, Haas J, Inman D, Masters P (2007) Long-term
tectonic control on holocene shelf sedimentation offshore la jolla, california. Geology
35(3):275
Holthuijsen, L.H., Booij, N. and Ris, R.C., 1993. A spectral wave model for the coastal zone.
Proceedings of the 2nd International Symposium on Ocean Wave Measurement and
Analysis, New Orleans, LA, pp. 630-641.
Inman D, Jenkins SA (2003) Accretion and erosion waves on beaches. Encyclopedia of
Coastal Science pp 1–4
Inman D, Nordstrom C (1971) On the tectonic and morphologic classification of coasts. The
Journal of geology 79(1):1–21
Inman DL (1987) Accretion and erosion waves on beaches. Shore & Beach 55(3&4):61–66
Inman DL, Bagnold RA (1963) Littoral processes. The Sea, Volume 6 pp 529–553
Komar P, Allan J (2008) Increasing hurricane-generated wave heights along the us east
coast and their climate controls. Journal of coastal research 24(2):479–488
Komar P, Inman D (1970) Longshore sand transport on beaches. Journal of Geophysical
Research-Oceans 70(30):5914–5927
Legg MR (1991) Developments in understanding the tectonic evolution of the California
continental borderland. From Shoreline to Abyss: Contributions in Marine Geology in
Honor of Francis Parker Shepard (Society for Sedimentary Geology):291–312
-23-
Adams et al., revision for Climatic Change – Mar. 2011
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
Longuet-Higgins M, Stewart R (1964) Radiation stresses in water waves; a physical
discussion, with applications. Deep Sea Res 11(4):529–562
Masters P (2006) Holocene sand beaches of southern california: Enso forcing and coastal
processes on millennial scales. Palaeogeography, Palaeoclimatology, Palaeoecology
232(1):73–95
Munk W, Traylor M (1947) Refraction of ocean waves: a process linking underwater
topography to beach erosion. The Journal of geology
O’Reilly W, Guza R (1993) Comparison of two spectral wave models in the southern
california bight. Coastal Engineering 19(3):263–282
O’Reilly W, Herbers T, Seymour R, Guza R (1996) A comparison of directional buoy and
fixed platform measurements of pacific swell. Journal of Atmospheric and Oceanic
Tech- nology 13(1):231–238
O’Reilly W. C., R. T. Guza, and R. J. Seymour, Wave Prediction in the Santa Barbara Channel,
Proc. 5th California Islands Symposium, Mineral Management Service, Santa Barbara
CA , March 29–31, 1999.
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
Pawka S, Inman D, Guza R (1984) Island sheltering of surface gravity waves: model and
experiment. Continental Shelf Research 3:35–53
Rogers W, Kaihatu J, Hsu L, Jensen R, Dykes J, Holland K (2007) Forecasting and hindcasting waves with the swan model in the southern california bight. Coastal
Engineering 54(1):1–15
Rosati J, Walton TL, Bodge K (2002) Longshore sediment transport. Coastal Engineer- ing
Manual, Part III, Coastal Sediment Processes Chapter 2-3, Volume Engineer Manual
1110-2-1100 U.S. Army Corps of Engineers
Seymour R, Guza R, O’Reilly W, Elgar S (2005) Rapid erosion of a small southern california
beach fill. Coastal Engineering 52:151–158
Shepard FP, Emery KO (1941) Submarine topography off the california coast: Canyons and
tectonic interpretation. US Geological Survey Special Paper p 171
Slott J, Murray A, Ashton A, Crowley T (2006) Coastline responses to changing storm
patterns. Geophysical Research Letters 33(18):L18,404
Storlazzi C, and Griggs G. (2000) Influence of El Niño–Southern Oscillation (ENSO) events
-24-
Adams et al., revision for Climatic Change – Mar. 2011
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
on the evolution of central California's shoreline. Geological Society of America
Bulletin vol. 112 (2) pp. 236
Storlazzi C, Willis C, Griggs G (2000) Comparative impacts of the 1982-83 and 1997-98 el
nin ̃o winters on the central california coast. Journal of Coastal Research 16(4):1022–
1036
Sverdrup HU, Munk WH (1947) Wind, sea, and swell: Theory of relations for forecasting. US
Navy Dept, Hydrographic Office, HO Pub (601):1–44
Sylvester AG (2010) Ucsb beach: 41 years of waxing and waning.
http://wwwgeolucsbedu/faculty/sylvester/UCSBbeacheshtml
Thornton E, Guza R (1983) Transformation of wave height distribution. J Geophys Res
88(10):5925–5938
Wang X, Swail V, Zwiers F, Zhang X, Feng Y (2009) Detection of external influence on trends
of atmospheric storminess and northern oceans wave heights. Climate Dynamics
32(2):189–203
Yates M, Guza R, O’Reilly W, Seymour R (2009) Seasonal persistence of a small southern
california beach fill. Coastal Engineering 56:559–564
-25-
Adams et al., revision for Climatic Change – Mar. 2011
616
Figures
617
618
619
620
621
622
623
624
Figure 1. Hillshade view of a 30-arc second digital elevation model of Southern California
Bight with modern shoreline shown as a white line. Note complex appearance of high relief
terrestrial landscape and irregular submarine basin bathymetry. Digital elevation data
obtained from the NOAA National Geophysical Data Center Coastal Relief Model
(http://www.ngdc.noaa.gov/mgg/coastal/startcrm.htm). The mean sea level shoreline
and 5-m isobath were interpolated from this data set as well.
625
-26-
Adams et al., revision for Climatic Change – Mar. 2011
626
627
628
629
630
Figure 2. Schematic diagram of numerical modeling procedure used in this study, showing
relationship between SWAN and CGEM components.
-27-
Adams et al., revision for Climatic Change – Mar. 2011
631
632
633
634
635
636
Figure 3. Example wave height and direction output from SWAN wave transformation
model over the entire coarse grid of the Southern California Bight. Locations of the two
study sites explored in this study are shown in boxes.
-28-
Adams et al., revision for Climatic Change – Mar. 2011
637
638
639
640
641
642
643
644
645
646
647
648
Figure 4. Example wave height and direction output from SWAN wave transformation
model for input conditions of Hsig =2 m, T =12 s, and  =270˚. A. Wave height map showing
results over entire Southern California Bight and location of SntBrb nest (box). B. Wave
height map showing results over the nested Santa Barbara grid (SntBrb), approximately 50
km of west-east trending coast, and location of region of interest (box). C. Detailed wave
height map showing wave direction vectors, bathymetric contours (10 meter contour
interval shown in thin white lines), location of sites SB-1 (red star) and SB-2 (blue star).
Location of 5-meter isobath shown in thick white line. Wave heights on A., B., and C. are
plotted with respect to the same colorbar whose units are meters.
-29-
Adams et al., revision for Climatic Change – Mar. 2011
649
650
651
652
653
654
655
656
657
658
659
660
Figure 5. Example wave height and direction output from SWAN wave transformation
model for input conditions of Hsig =2 m, T =12 s, and  =270˚. A. Wave height map showing
results over entire Southern California Bight and location of TorPns nest (box). B. Wave
height map showing results over the nested Torrey Pines grid (TorPns), approximately 40
km of north-south trending coast, and location of region of interest (box). C. Detailed wave
height map showing wave direction vectors, bathymetric contours (10 meter contour
interval shown in thin white lines), location of sites TP-1 (red star) and TP-2 (blue star).
Location of 5-meter isobath shown in thick white line. Wave heights on A., B., and C. are
plotted with respect to the same colorbar whose units are meters.
-30-
Adams et al., revision for Climatic Change – Mar. 2011
661
662
663
664
665
666
667
668
669
Figure 6. Example CGEM output along 5-meter isobath within the nested Santa Barbara
Grid (SntBrb) for two sets of deep water wave conditions, which differ only in wave
direction. Blue lines show results of deep water conditions Hsig=4 m, T=16 s, and =320˚.
Red lines show results of deep water conditions Hsig =4 m, T =16 s, and  =270˚. Red and
blue stars show locations of SB-1 and SB-2, used for numerical experiments. LST is an
abbreviation of longshore sediment transport.
-31-
Adams et al., revision for Climatic Change – Mar. 2011
670
671
672
673
674
675
Figure 7. Compendium of potential divergence of drift results from 104 SWAN-CGEM
model simulations for the SntBrb nest at sites SB-1 and SB-2. Positive values of divergence
of drift represent accretion and negative values represent erosion.
-32-
Adams et al., revision for Climatic Change – Mar. 2011
676
677
678
679
680
681
682
683
684
685
Figure 8. Example CGEM output along 5-meter isobath within the nested Torrey Pines
Grid (TorPns) for two sets of deep water wave conditions, which differ only in wave
direction. Blue lines show results of deep water conditions Hsig=4 m, T=16 s, and =320˚.
Red lines show results of deep water conditions Hsig =4 m, T =16 s, and  =270˚. Red and
blue stars show locations of TP-1 and TP-2, used for numerical experiments. LST is an
abbreviation of longshore sediment transport.
-33-
Adams et al., revision for Climatic Change – Mar. 2011
686
687
688
689
690
Figure 9. Compendium of potential divergence of drift results from 104 SWAN-CGEM
model simulations for the TorPns nest at sites TP-1 and TP-2.
-34-
Download