You are on page 1of 12

Global and Planetary Change 39 (2003) 53 – 64

www.elsevier.com/locate/gloplacha

Effects of climatic and anthropogenic change on small


mountainous rivers: the Salinas River example
Katherine L. Farnsworth*, John D. Milliman
Virginia Institute of Marine Science, School of Marine Science, College of William and Mary, PO Box 1346, Gloucester Point, VA 23062, USA
Received 15 September 2001; accepted 18 December 2002

Abstract

Over the past 50 years rivers have discharged about 18  109 t of sediment annually to the global ocean, mountainous rivers
with small drainage basins playing a particularly key role. Mean annual values, however, often mask short- and longer term
variations resulting from climatic and anthropogenic factors. Discharge from large rivers generally displays relatively little inter-
annual variation, whereas long-term discharge and sediment delivery from smaller rivers often are determined by infrequent
events. The lack of adequate monitoring unfortunately means that these event-driven deliveries to the coastal ocean are usually
poorly documented and the magnitude of their impact often insufficiently understood.
The Salinas River provides one example of the natural and anthropogenic variability occurring in small river systems, as
well as the problems associated with understanding sediment transport and discharge based on limited sediment-load data. With
half of the historical (70 years) sediment load being delivered in less than 5 weeks, the Salinas is an event-dominated river on
which the construction of dams and withdrawal of water have severely changed (and continue to change) its natural flow
regime. The use of outdated sediment measurements leads to potentially large errors in load estimates, thereby underscoring the
need for newer and more extensive data, particularly during infrequent events.
D 2003 Elsevier Science B.V. All rights reserved.

Keywords: Sediment; Anthropogenic change; River

1. Introduction Meade, 1983; Milliman and Syvitski, 1992; Ludwig


and Probst, 1998), a disproportional amount being
Rivers form the major link between the terrestrial discharged by rivers draining active mountains in
and marine systems, discharging more than 35  103 southern Asia, Oceania, and North and South Amer-
km3 of water and 20  109 t of suspended and ica. The importance of small mountainous rivers only
dissolved solids annually (Milliman and Farnsworth, became obvious when the database was sufficient
in preparation). Suspended sediment delivery to the (e.g., Li, 1976; Griffiths and Glasby, 1985) to indicate
global ocean has been calculated to be between 15 and the extreme sediment yields from high-standing
18  10 9 t/year (Holeman, 1967; Milliman and islands. Holeman (1967) essentially ignored the sedi-
ment flux from Oceania, whereas we presently esti-
mate that these high-standing islands discharge more
* Corresponding author. Tel.: +1-804-684-7267. than 40% of the sediment reaching the global ocean
E-mail address: farnswor@vims.edu (K.L. Farnsworth). (e.g., Milliman et al., 1999).

0921-8181/03/$ - see front matter D 2003 Elsevier Science B.V. All rights reserved.
doi:10.1016/S0921-8181(03)00017-1
54 K.L. Farnsworth, J.D. Milliman / Global and Planetary Change 39 (2003) 53–64

Erosion and delivery of sediment are a function of greater than the Amazon (4.2 vs. 1.1  109 t/year)
river runoff, basin morphology, tectonics, bedrock (Milliman et al., 1999).
lithology, human activity, and basin area. For exam- The dispersal and fate of sediment derived from
ple, sediment yield (load normalized for basin area— small rivers can also be different than that discharged
e.g., t/km2/year) is an inverse function of basin area from large rivers. Most large rivers have large flood
(Milliman and Syvitski, 1992). In the case of South- plains, deltas and/or estuaries in which much of the
east Asian and Oceanic rivers draining young moun- sediment load is stored for long time periods. Because
tains, watersheds smaller than 103 km2 have about an many large rivers (e.g., Amazon, Orinoco, Congo,
order of magnitude greater sediment yield than those Yangtze, etc.) discharge onto broad passive margins,
rivers having watersheds greater than 105 km2 (Fig. much of the sediment escaping the river or estuary
1). Smaller rivers have less basin area in which to mouth is deposited on the inner shelf; on a global
store flood-driven sediment (Milliman and Syvitski, scale relatively little escapes the shelf edge to the
1992), and they are also more likely to respond to deep sea (Meade, 1996). Most small mountainous
event-driven floods (Warrick and Milliman, in prepa- rivers draining to the oceans, in contrast, are concen-
ration). As the number of small rivers far exceeds the trated along active margins and they discharge sedi-
number of large rivers, collectively these smaller ment as a line source (e.g., Mertes and Warrick,
rivers are responsible for a much larger amount of 2001), rather than point-source dispersal seen in
sediment delivery to the global ocean than one might larger rivers. Moreover, discharge from many of these
intuitively imagine. The 460 mountainous rivers smaller systems periodically can discharge hyperpyc-
draining six high-standing islands in the East Indies nal concentrations of sediment to the coastal zone,
(Sumatra, Java, Borneo, Celebes, Timor and New whose transport path and depositional fate may be
Guinea), for example, collectively drain less than 1/ far different than that for hypopycnal plumes
3 the area of the Amazon River (1.8 vs. 6.3  106 (Mulder and Syvitski, 1995; Warrick and Milliman,
km2), but their estimated sediment load is four times in preparation).

2. Variability in sediment discharge

Present fluxes of water and sediment from rivers to


the coastal ocean are controlled by both natural and
anthropogenic influences. Variability in river dis-
charge reflects the influence of both long-term (cen-
tury to millennial) and short-term (annual and inter-
annual) climatic cycles. Superimposed on these are
the effects of anthropogenic change on both the
watershed and the river itself. Examples of ‘‘pristine’’
rivers, in fact, are difficult to find, most being
restricted to sparsely inhabited Arctic watersheds.
Human-induced changes to a drainage basin and
the river’s flow regime can also affect the transport
and delivery of such things as pollutants, pesticides
and nutrients. The impact on the coastal ocean from
these anthropogenic effects may be subtle but far-
reaching. For example, silicate uptake in reservoir
Fig. 1. Sediment yield (t/km2/year) for tropical wet rivers draining waters trapped behind the Danube River’s Iron Gates
young mountains in southeastern Asia and the high-standing islands
of Oceania. Note the order of magnitude increase in yield with Dam has decreased the silicate delivery and thus
decreasing basin size. Data from Milliman and Farnsworth (in shifted the dominant primary producers in NW Black
preparation). Sea coastal waters from diatoms to dinoflagellates and
K.L. Farnsworth, J.D. Milliman / Global and Planetary Change 39 (2003) 53–64 55

coccolithorphorids; one result of this shift in ecosys-


tems has been an increase in the occurrence of anoxic
conditions (Humborg et al., 1997).
The Yellow River, in northern China is a partic-
ularly stark example of natural cycles and anthropo-
genic change. This river is normally considered an
arid river, runoff only being about 40 mm/year.
During the last glaciation, in fact, the river may have
been dry, in response to the diminished SW Monsoon
and enhanced NE Monsoon (Xia et al., 1993). Over
the past 100 years, annual precipitation throughout the
Yellow River’s drainage basin has decreased by about
15% (Galler, 1999), and in the past decade discharge
has dropped by more than 50% in response to both a
particularly dry period and the corresponding increase
in water removal for irrigation and industrial con-
sumption. As a result, in 1997 the lower Yellow River
was dry for more than 300 days. In contrast to a mean
annual load of 1.1  109 t/year documented in the
1950s through late 1970s (Qian and Dai, 1980),
during the past decade the Yellow River’s sediment
load has averaged less than 0.2 109, and in 1997 it
was only 0.03  109 t (Yang et al., 1998; Galler,
1999).

2.1. Climate change

Climatic change is manifested in changes in solar


radiation, wind patterns, precipitation, evaporation,
and temperature (e.g., storage in glaciers). The hydro-
logic cycle presumably can absorb small changes in
climate, but larger and longer shifts in climate disrupt
this balance. Moreover, the long-held image of grad-
ual, long-term change has been modified in light of
ice-core and pollen data that show dramatic climate
changes occurring on time scales as short as several
decades (Dansgaard et al., 1993). Regional weather
changes can lead to quick responses in the character
(type and density) of vegetation, particularly in arid
and semi-arid watersheds, which in turn can have
large affects on the erosion and transport of sediment. Fig. 2. Annual deviations from the 50-year mean discharge of the
Because of the modulation of meteorological Susquehanna (drainage basin area = 71,000 km2), Merrimack
events through their watersheds, many large river (12,000 km2) and Passaic (1900 km2) rivers in the northeastern
systems tend to experience relatively little seasonal US. Deviations in Susquehanna flow are mostly less than 25% of
or inter-annual variability in their discharge. Smaller the 50-year mean, and only 1 year (1972, a 100-year event) ex-
ceeded 50%. The smaller Merrimack experienced 2 years in which
watersheds, in contrast, are more likely to reflect inter- the annual discharge exceeded 50% of the mean, but the much
annual variability (Fig. 2). Small basins are also more smaller Passaic experienced 10 significant deviations from its 50-
responsive to episodic events and therefore can year mean. Data from the US Geological Survey.
56 K.L. Farnsworth, J.D. Milliman / Global and Planetary Change 39 (2003) 53–64

deliver large portions of their fluvial discharge and exceed 400  106 t, whereas the flashy peak events in
sediment loads in relatively short periods of time. In small rivers generally deliver only a fraction as much
North America, for instance, peak 24-h loads of small sediment (Fig. 3).
rivers such as the Santa Clara (near Santa Barbara, The concentrated delivery of sediment during these
California) can exceed that of the Mississippi. The Eel peak events (not infrequently hyperpycnal; Warrick
River, in northern California, experienced a peak and Milliman, in preparation), however, can affect the
discharge in December 1964 that was more than an dispersal and fate of the event-derived sediment dis-
order of magnitude greater than the peak 24-h load charged by small rivers (e.g., Morehead and Syvitski,
measured on the Mississippi, even though its drainage 1999). Peak concentration of suspended sediment in
basin occupies less than 1/400th the area (Fig. 3). In the Santa Clara, Salinas and Eel rivers, for example,
fact, a search of the extensive USGS data base can reach 20 to 100 g/l (Table 1), whereas peak
indicates that only three Californian rivers (Eel, Sal- concentrations in the Mississippi rarely exceed 2 g/l.
inas and Santa Clara) have recorded daily sediment The importance of such infrequent events in small
loads greater than 5  106 t/day, with sediment yields rivers makes them both interesting and frustrating to
often exceeding 1000 t/km2/day during particularly study because the long-term sediment discharge can
large events (Table 1). It is important to note that depend on a few scattered events, in some cases no
while the Mississippi has no recorded load approach- more than a few times every decade.
ing 5  106 t/day, ‘‘floods’’ can extend for hundreds of
days (compared to the 2 –7 day length for most small- 2.2. Anthropogenic impact
river floods) (Fig. 3). As a result total flood-derived
discharge in a large river like the Mississippi can Man has proved to be an effective geologic agent in
altering the landscape and therefore river erosion and
delivery. Hooke (2000) estimates earth-moving activity
related to human actions to be about 100  109 t/year,
5-fold greater than the total fluvial delivery to the
global ocean. While the anthropogenic impact on
global fluvial sediment fluxes cannot yet be calculated
with any degree of accuracy, humans may be directly or
indirectly responsible for 80– 90% of the fluvial deliv-
ery to the coastal ocean (Milliman et al., 1987; Doug-
las, 1996; Milliman and Farnsworth, in preparation).
Yet any global pattern of sediment delivery belies
recent and ongoing changes that are occurring on
local and regional scales. Changing agricultural pat-
terns and reforestation over the past 50– 100 years
have resulted in decreased erosion throughout much
of Europe and North America. Decreased sediment
loads ‘‘augmented’’ by river-bed mining for sand and
gravel, for example, has resulted in the progressive
scour of many Italian river beds in recent years, one
Fig. 3. Daily sediment loads for the Mississippi and Eel rivers
during years of peak discharge, water years 1950 and 1965,
result being decreased sediment delivery to the coastal
respectively. The daily load of the Mississippi was consistently zone and thereby increased coastal erosion (Simeoni
greater than 105 t/day throughout the entire year, total annual load and Bondesan, 1997; Coltorti, 1997). In contrast to
being 500  106 t. In contrast, the Eel’s load exceeded 105 t for only the decreased erosion and sediment transport by rivers
46 days and its total annual load was about 150  106 t, but most of in much of the developed western world, erosion is
the load (110  106 t) was transported in 3 days. During this 3-day
period, average sediment concentration was 28 g/l, more than an increasing throughout many developing countries.
order of magnitude greater than in the Mississippi (1.2 – 2 g/l). Data Although definitive data are still lacking, increased
form the US Geological Survey. deforestation in the mountainous regions of southern
K.L. Farnsworth, J.D. Milliman / Global and Planetary Change 39 (2003) 53–64 57

Table 1
Major 24-h sediment loads at or near the mouths of US rivers, 1950 – 1990
River Basin Area Date Q (cm) Sediment Sediment load Sediment yield
(  103 km2) concentration (  106 t/day) (t/km2/day)
(mg/l)
Eel 8 12/23/64 18,300 32 51.8 6470
Eel 8 12/22/64 14,800 28 36.4 4550
Eel 8 12/22/56 12,000 23 24 3000
S. Clara 4.1 2/25/69 2610 69 18.5 4510
Eel 8 12/24/64 8810 23 17.3 2160
S. Clara 4.1 1/25/69 2100 61 12.9 3150
Eel 8 12/23/64 8140 17 12 1500
S. Clara 4.1 3/4/78 1720 108 10.7 2610
Eel 8 1/27/83 6360 17 9.4 1170
Salinas 11 2/26/69 1835 53 9.4 854
Eel 8 12/20/81 6440 15 8.7 1080
Salinas 11 3/3/83 1693 34 7.9 987
Eel 8 12/25/64 5380 6.2 7.7 962
S. Clara 4.1 2/10/78 1260 50 6.9 1680
Eel 8 12/26/64 4980 16 6.8 850
Eel 8 12/19/81 5460 14 6.5 812
S. Clara 4.1 3/1/83 1470 44 6.4 1560
Eel 8 1/16/74 7600 7.6 6.2 775
Salinas 11 1/27/69 1510 43 6.2 563
Eel 8 1/4/66 5470 13 6.1 762
Eel 8 1/5/66 7250 9.4 6 750
Eel 8 2/16/82 4860 13 5.5 687
Eel 8 12/19/56 3720 16 5.1 637

Asia and in Oceania (whose rivers, by virtue of their (storage capacity greater than 0.5 km3) globally store
relatively youthful morphology and geology, as well about 5000 km3 of water (nearly 15% of the global
as the seasonally heavy rains, have high sediment fluvial discharge). Collectively they intercept about
yields) almost certainly have lead to greater fluvial 40% of the world’s freshwater discharge and may trap
erosion. 25% of the global sediment delivery to the ocean.
The degree to which fluvial water and sediment Woodward (1995) lists five major rivers draining into
reach the coastal ocean, of course, depends upon other the Mediterranean whose total sediment discharge has
human activities, namely the construction of struc- decreased from nearly 190  106 t/year to less than
tures for flood control, water diversion, power gen- 10  106 t/year.
eration and recreation. Reservoirs and irrigation Dam construction in North America and Europe
channels can retain a large proportion of the fluvial essentially has ceased, both because most economic
sediment load, sometimes without actually affecting and suitable dam sites already have been occupied and
the amount of annual water discharge. Barrage con- because of increasing concern about weighing the
struction to facilitate irrigation along the lower Indus environmental impact against economic benefit. Data
River in the late 1940s, for example, resulted in a 75% from the International Commission on Large Dams
decrease in the river’s fluvial sediment load, but only (ICOLD) show that as of 1999 Canada had no large
a slight decrease in average annual water discharge; dams (defined as being higher than 15 m) under
subsequent construction of a series of high dams in construction, whereas China and India had 330 and
the Indus headwaters in the early 1960s subsequently 650, respectively. The recent acceleration of dam
reduced annual discharge of the river by nearly 80% construction in Africa and southern Asia can be
(Milliman et al., 1985). Vorosmarty et al. (1997) have illustrated in the following: of the 43 countries
calculated that the 633 dams with large reservoirs responding to an ICOLD questionnaire in 1999 (Aus-
58 K.L. Farnsworth, J.D. Milliman / Global and Planetary Change 39 (2003) 53–64

tralia and most countries in South America and within sediment discharge. Luckily, major events occurred in
the former Soviet Union were notable in their many of these rivers (Table 1), so that we can gather
absence), a total of 47,425 dams were listed, of which some insight as to their effect.
China accounted for more than half, a remarkable
number considering that in 1949 China had only three
large dams! Moreover, of the more than 10,000 dams 4. River response to change and episodic
higher than 30 m, China is responsible for 45%. events—the Salinas River example
Continuing dam construction throughout Africa and
southern Asia therefore may affect significantly the The rivers of California provide particularly inter-
amount of water and sediment delivery to the global esting examples of the impact of climate and anthro-
ocean. pogenic change on fluvial discharge to the coastal
ocean. Precipitation patterns throughout the state vary
greatly, from less than 100 mm/year in many of the
3. Problems in estimating sediment delivery central valleys to more than 2000 mm/year in the
northern mountains. These regional precipitation pat-
As discussed in the preceding paragraphs, calculat- terns are reflected in river runoff: the Tijuana River
ing or even estimating the sediment load of a river is along the Mexican border has an annual runoff of 10
complicated by the impact of natural forcings (cli- mm/year compared to the Mad River in the north with
mate, morphology, geology), watershed utilization a runoff of 1100 mm/year. But mean annual precip-
and water management. Because of the changing itation and runoff tell only part of the story because
climate and continually shifting anthropogenic im- California weather is also one of extremes: ‘‘extremely
pact, fluvial data must be continually updated. Rely- wet or extremely dry’’ (Mount, 1995). The generally
ing on old data becomes increasingly risky as the arid weather in many drainage basins is punctuated
character of the watershed and its climate changes. by periodic, intense bursts of precipitation, often
This continued need—in fact increasing need—for lasting only a few days. Moreover, with few excep-
river data stands in stark comparison with the decreas- tions, most Californian rivers have drainage basins
ing number of rivers that are presently monitored. less than 10,000 km2 in area, so that the rivers tend
Vorosmarty et al. (2001) estimate that hundreds of to be more responsive to both natural and human-
hydrologic stations are closed each year, and the data induced change.
from many of those rivers that continued to be Southern and central Californian weather patterns
monitored are difficult to obtain. The various existing are strongly controlled by the El Niño/Southern
global data banks, for instance, have few data from Oscillation (ENSO). Warmer sea-surface temperatures
African rivers more recent than 1984, and French river in the NW Pacific during El Niño periods result in the
data since 1980 are almost impossible to obtain. northward movement of moisture-laden air, often
While US fluvial data are ever easier to access over resulting in heavy rains, although they tend to be
the US Geological Survey’s web page, the number of episodic. North of San Francisco Bay the ENSO
new data decrease each year. Lanfear and Hirsch influence is diminished and more weather patterns
(1999) estimate that since the early 1990s more than roll in from Alaska and Canada and the adjacent
100 US hydrologic stations annually have been northern ocean.
closed. And the story for water quality is even Discharging into Monterey Bay, about 100 km
grimmer: only the Mississippi, Colorado, Columbia south of San Francisco Bay, the Salinas River is the
and Rio Grande (and their tributaries) are monitored northernmost Californian river that responds strongly
regularly, even though the Colorado and Rio Grande to the ENSO signal. With a watershed area of 11,000
presently effectively discharge little water and no km2, the basin is 240 km long and has an average
sediment to the coastal ocean. This contrasts strongly width of 65 km. The river flows almost entirely within
with the 20-year period from the mid-1960s to the a northwest-trending structural trough defined by two
mid-1980s when essentially all of the large US rivers parts of the NW –SE-trending Coast Range Province,
and many of the smaller rivers were monitored for the Sierra de Salinas (maximum elevations 900– 1500
K.L. Farnsworth, J.D. Milliman / Global and Planetary Change 39 (2003) 53–64 59

m) in the southwest and the Gabilan Range (750 – heavy annual precipitation (1000 mm/year) along the
1200 m) in the northeast (Fig. 4a). The mountains southeastern boundary of the watershed, and a rain
bounding the Salinas basin were formed by late shadow ( < 300 mm/year) in the valley and northern
Cenozoic uplift and deformation and are underlain mountains (Fig. 4b). Because precipitation in the
by marine sedimentary, intrusive igneous and meta- Salinas watershed is influenced by the passing of
morphic rocks. Generally older consolidated rocks are mid-latitude Pacific storms rather than northern
exposed in the mountains, whereas a cover of recent storms, it is relatively warm, falling mostly as rain
alluvium and floodplain deposits characterizes the rather than snow.
valleys and lowlands. The Salinas River has a limited base flow, with
Possessing a Mediterranean climate, nearly 90% of many years reporting runoff less than 100 mm/year
the annual precipitation in the Salinas basin falls (Fig. 5). The Nacimiento River, San Antonio River
between November and April, resulting in highly and Arroyo Seco drain the wetter western coastal
seasonal delivery of both water and sediment to the ranges, while the San Lorenzo and Estrella are the
coastal ocean. Orographically induced rainfall causes main tributaries draining the drier eastern range.

Fig. 4. Salinas River drainage basin characteristics. Elevation (a), annual precipitation (b) and land cover (c) (http://www.biogeog.ucsb.edu/
projects/gap/gap_ home.html). The western mountain ranges receive much greater levels of precipitation than the central valley, although the
rich river-derived valley alluvium is the major area for agriculture within the watershed.
60 K.L. Farnsworth, J.D. Milliman / Global and Planetary Change 39 (2003) 53–64

Fig. 5. Annual runoff for four stations within the Salinas River Basin. The major source of water to the Salinas River is from the western
mountain ranges, specifically the unrestricted Arroyo Seco.

Although equal in area to the San Lorenzo, the Arroyo absorb most of the discharge, but larger storms during
Seco has a 15-fold greater discharge (Fig. 5). that year filled the reservoir and water had to be
Three large dams along tributaries have changed released (Fig. 6).
the natural variability and timing of flow. The Salinas The number of no/low-flow days ( < 1 ft3/s) has
Dam (Lake Santa Margarita) was constructed by the changed dramatically over the past 70 years. Between
US Army Corps of Engineers in 1941. Monterey 1930 and 1960, the number of no/low-flow days was
County built the Nacimiento and San Antonio dams generally inversely proportional to the amount of
on the Nacimiento and San Antonio rivers in 1956 and precipitation (Fig. 7). But by 1960, the incremental
1965, respectively, both with storage capacities of release of water from the Nacimiento and San Antonio
approximately 0.4 km3. Water is released from these dams regulated discharge of reservoir water, thereby
dams for irrigation and groundwater recharge during modulating down-stream flow and drastically reduc-
the dry summers (USGS, 2000). The amount of water ing the number of no/low-flow days. During the past
released by the Nacimiento Dam during normal years decade, however, the pattern has changed again: even
is similar to that entering the reservoir. However, though annual precipitation throughout much of the
during extreme years, the hydrograph is highly altered 1990s was high, due to a number of severe El Niño
by the presence of the dam. During small precipitation years, the number of low-flow days at the Spreckels
events in 1969, for instance, the reservoir was able to gaging station increased, presumably reflecting the
K.L. Farnsworth, J.D. Milliman / Global and Planetary Change 39 (2003) 53–64 61

Fig. 6. River discharge data from the USGS gaging stations above
and below Nacimiento Dam. Gaging stations used were USGS
11148800 Nacimiento River near Bryson and 11149400 Nacimiento
River near Bradley. There is a significant alteration to the natural Fig. 7. Annual number of no- to low-flow days at the Spreckels
hydrograph of the Nacimiento River, which would naturally be a gaging station near the mouth of the Salinas River. The inverse
large source of water and sediment to the Salinas River. relationship between precipitation and the number of days with
essentially no flow is seen until the 1990s, when water use within
the drainage basin changed.
increased removal of upstream water for agricultural
use.
During the 70 years of monitoring, the Salinas has
delivered approximately 2.3  108 t of sediment to
Monterey Bay. Mean annual water discharge is 4 km3/
year (120 m3/s) and mean annual sediment load is 3.3
t/year, but in only 18 of those years did the annual
load actually approach or exceed the 70-year mean
(Fig. 8). In 31 of the 70 years, the river delivered less
than 0.1 106 t/year, whereas in 6 years the load
exceeded 15  106 t/year (Fig. 8). Total high-dis-
charge time, however, was less than 5 weeks. Four
of the six high-load years occurred during El Niño
years, and two during ‘‘Nada Niño’’ years; no sig-
nificant delivery events occurred during La Niña years
(Fig. 9).
The Salinas Valley has been called the ‘‘Salad
Bowl of America’’ due to the large amounts of intense
agriculture within its watershed. Land-use throughout
the basin, however, is varied, with agriculture domi-
Fig. 8. Calculated annual load of the Salinas River at Spreckels,
nating the valley floor, rangeland in the hills and using a rating curve developed from 10 years of suspended sediment
forested regions on the steep slopes (Fig. 4c) (USGS, measurements. The mean annual load is indicated, but due to the
2000). In recent years land use within the watershed episodic nature of this river, is rarely approached.
62 K.L. Farnsworth, J.D. Milliman / Global and Planetary Change 39 (2003) 53–64

Monterey County Water Resources Agency, in large


part to prevent flood damage to surrounding areas and
to prevent heavy silt loads from entering the Moss
Landing Harbor. Thus sediment input to the bay from
the Salinas River occurs only when the sand bar has
been breached during high flow periods. During these
high-flow events the amount of sediment in the river
water may exceed hyperpycnal concentrations. The
1978 measured concentrations exceeded 24 g/l, sug-
gesting that during major events, such as those during
the 1969 and 1983 floods, sediment concentrations
may have reached hyperpycnal levels. The fate of
sediment derived from episodic hyperpycnal events in
the Salinas River at present is difficult to ascertain.
Does it accumulate off the river mouth or does most of
it escape offshore; if the latter, how does it mobilize
and where does it go? The effect of such episodic
Fig. 9. The influence of the El Niño Southern Oscillation (ENSO) is sediment discharges to the shoreline, benthos, water
seen in the relationship between the Southern Oscillation Index column and the coastal zone must be immense.
(SOI) and the annual load of the river. There have been no high load
(high flow) years during La Niña years, and the largest events have
been during El Niño years. Note, however, that the magnitude of the
annual load shows little quantitative relation to the SOI.

has changed with the marked increase in viticulture;


we assume this change in land use also has affected
erosion rates, although suspended sediment in the
river has not been sampled since 1979. Water removal
from the system is mainly for agricultural use and
some years have less water passing Spreckels than in
the upper reaches of the river (Fig. 10). In 1993, for
instance, Spreckels (location of the most downstream
gauging station) monitored about 30% less discharge
than the Salinas tributaries. Even this, however, does
not indicate the entire magnitude of water withdrawal,
since agricultural activity is also intense downstream
from Spreckels.
Rivers are assumed to be the dominant source for
California beach sand (Hicks and Inman, 1987).
While event delivery to coastal ocean is periodic,
Fig. 10. Water use within the basin has created a negative water
littoral drift is continuous. The low- to no-flows in balance within the river itself. The withdrawal of water along the
the summer, combined with the wave-dominated course of the river has caused the volume of water passing a station
coastline, result in sand bar extension across the river to be less than that of the station a few kilometers upstream. This is
mouth, changing the Salinas estuary into a closed seen in the gaging stations at Bradley, Chular, and Spreckels, in the
lagoon. Therefore, for much of the year the Salinas lower Salinas basin (see Fig. 4 for locations). Although the volume
of water passing Chular should be sum of water flowing past
does not discharge to Monterey Bay. Breaching of the Bradley and the San Lorenzo and Arroyo Seco tributaries, in 3 of
sand bar occurs during flood events, and more the 4 years shown in this figure the volume decreases downstream,
recently has been ‘‘helped’’ by bulldozers from the the result of water withdrawal for agricultural use.
K.L. Farnsworth, J.D. Milliman / Global and Planetary Change 39 (2003) 53–64 63

5. Discussion ditions will be made increasingly difficult as we rely


on increasingly antiquated data. To understand the
The Salinas River is not that different from many impact of future change, be it in terms of climate or
other small rivers throughout the world, which peri- land-use, we must first understand natural variability
odically can discharge large volumes of water and of fluvial systems. Mean annual trends derived from
sediment to the coastal ocean but whose fluvial and an increasingly old database will not be adequate.
land-use character continue to change. Unfortunately,
as changes in land-use and climate continue it will be
increasingly difficult to rely on previously collected Acknowledgements
data to characterize or quantify the amount of sedi-
ment discharged to the coastal ocean. Because the Research funds for this and related papers have
Salinas River was monitored for suspended sediment come from a series of grants from the Marine Geology
by the USGS only until the late 1970s, rating curves and Geophysics program of the National Science
based on these data cannot adequately indicate sedi- Foundation. We thank our colleagues, particularly
ment loads prior to dam construction in the 1950s and Jonathan Warrick, Des Walling and James Syvitski for
1960s or changing land and water use in the past 20 continuing dialog as well as comments on a previous
years. The Nacimiento and San Antonio dams almost draft of this paper. Many thanks also to Mark
certainly reduced sediment delivery from the Salinas. Morehead and an anonymous reviewer for their
Therefore, calculations based on a rating curve from constructive comments. This paper is Contribution
the 1970s will underestimate the historical sediment No. 2529 of the Virginia Institute of Marine Science,
load of the river prior to 1965. The increased recent The College of William and Mary.
agricultural use of the lower Salinas watershed has
lead to increased water withdrawal from the river for
irrigation, thereby altering the river’s hydrograph (Fig. References
7). Present-day sediment flux therefore is probably
Coltorti, M., 1997. Human impact in the Holocene fluvial and
less than would be calculated using the 1970s-based coastal evolution of the Marche region, Central Italy. Catena
rating curve. Nevertheless, the lack of more recent 30, 311 – 335.
data means that we must continue to use this anti- Dansgaard, W., Johnsen, S.J., Clausen, H.B., Dahl-Jensen, D., Gun-
quated rating curve: there is nothing else, nor will destrup, N.S., Hammer, C.U., Hvidberg, C.S., Steffensen, J.P.,
Sveinbjornsdottir, A.E., Jouzel, J., Bond, G., 1993. Evidence for
there be in the foreseeable future.
general instability of past climate from a 250-kyr ice-core re-
Small mountainous rivers draining to the coastal cord. Nature 364 (6434), 218 – 220.
ocean play a vital role in the global sediment budget, Douglas, I., 1996. The impact of land-use changes, especially log-
but the problems associated with using mean values ging, shifting cultivation, mining and urbanization on sediment
derived from short records need to be addressed and yields in humid tropical Southeast Asia: a review with special
better understood to be able to more accurately reference to Borneo. IAHS Publication 236, 463 – 471.
Galler, J.J., 1999. Medium- and long-term changes in fluvial dis-
estimate the global sediment budget, as well as to charge to the sea: the Yellow River case study. MSc thesis,
study the role of these small rivers locally in control of College of William and Mary, 78 pp.
coastal sedimentation and erosion. It is particularly Griffiths, G.A., Glasby, G.P., 1985. Input of river-derived sediment
important that episodic events are closely monitored to the New Zealand Continental Shelf: I. Mass. Estuarine,
on these smaller rivers. Both because these often Coastal and Shelf Science 21, 773 – 787.
Hicks, D.M., Inman, D.L., 1987. Sand dispersion from an Ephem-
represent the major (perhaps only) period of signifi- eral River Delta on the Central California Coast. Marine Geol-
cant sediment discharge, but also because suspended ogy 77, 305 – 318.
sediment concentration (at least in some rivers) may Holeman, J.N., 1967. The sediment yield of major rivers of the
reach or exceed hyperpycnal levels, suggesting that world. Water Resources Research 4 (4), 737 – 747.
they may move offshore via density flows, thereby Hooke, R.LeB., 2000. On the history of humans as geomorphic
agents. Geology 28, 843 – 846.
avoiding the coastal circulation cells (Warrick and Humborg, C., Ittekkot, V., Coclasu, A., Bodungen, B., 1997. Effect
Milliman, submitted for publication). Our ability to of Danube River dam on Black Sea biogeochemistry and eco-
understand present conditions and predict future con- system structure. Nature 386, 385 – 388.
64 K.L. Farnsworth, J.D. Milliman / Global and Planetary Change 39 (2003) 53–64

Lanfear, K.J., Hirsch, R.M., 1999. USGS study reveals a decline in Mount, J.F., 1995. California Rivers and Streams. Univ. California
long-record streamgages. EOS 80 (50), 605 – 607. Press, Berkeley. 359 pp.
Li, Y.H., 1976. Denudation of Taiwan island since the Pliocene Mulder, T., Syvitski, J.P.M., 1995. Climatic and morphologic rela-
epoch. Geology 4, 105 – 107. tionships of rivers: implications of sea-level fluctuations on river
Ludwig, W., Probst, J.-L., 1998. River sediment discharge to the loads. Journal of Geology 104, 509 – 523.
oceans; present-day controls and global budgets. American Qian, N., Dai, D.Z., 1980. The problems of river sedimentation and
Journal of Science 298 (4), 265 – 295. the present status of its research in China. Proceedings of the
Meade, R.H., 1996. River-sediment inputs to major deltas. In: Milli- International Symposium on River Sedimentation, 24 – 29
man, J.D., Haq, B.U. (Eds.), Sea-Level Rise and Coastal Sub- March 1980, Beijing. The Chinese Society of Hydraulic Engi-
sidence. Kluwer Academic Publishing, Dordrecht, pp. 63 – 85. neering, vol. 1. Guanghua Press (UNESCO), pp. 19 – 39.
Mertes, L.A.K., Warrick, J.A., 2001. Measuring flood output from Simeoni, U., Bondesan, M., 1997. The role and responsibility of man
110 coastal watersheds in California with field measurements in the evolution of the Italian Adriatic coast. Bulletin de l’Institut
and SeaWiFS. Geology 29 (7), 659 – 662. Oceanographique, Monaco Special Volume 18, 111 – 132.
Milliman, J.D., Farnsworth, K.L., in preparation. River runoff, ero- USGS, 2000. Suspended sediment database of daily values and
sion and delivery to the coastal ocean: a global analysis. Oxford ancillary data, http://www.waterdata.usgs.gov/ca/nwis/sw and
University. http://www.webserver.cr.usgs.gov/sediment.
Milliman, J.D., Meade, R.H., 1983. Worldwide delivery of river Vorosmarty, C.J., Meybeck, M., Fekete, B., Sharma, K., 1997. The
sediment to the oceans. Journal of Geology 91 (1), 1 – 21. potential impact of neo-castorization of sediment transport by
Milliman, J.D., Syvitski, J.P.M., 1992. Geomorphic/tectonic control the global network of rivers. IAHS Publication 245, 261 – 273.
of sediment transport to the ocean: the importance of small Vorosmarty, C., et al., 2001. Global water data: a newly endangered
mountainous rivers. Journal of Geology 100, 525 – 544. species. EOS 82, 54 – 58.
Milliman, J.D., Shen, H.-T., Yang, Z.-S., Meade, R.H., 1985. Trans- Warrick, J.A., Milliman, J.D., in preparation. Event-driven hyper-
port and deposition of river sediment in the Changjiang Estuary pycnal discharge from southern California rivers.
and adjacent continental shelf. Continental Shelf Research 4 Woodward, J.C., 1995. Patterns of erosion and suspended sediment
(1 – 2), 37 – 44. yield in Mediterranean river basins. In: Foster, I.D.L., Gurnell,
Milliman, J.D., Yun-Shan, Q., Mei-E, R., Saito, Y., 1987. Man’s A.M., Webb, B.W. (Eds.), Sediment and Water Quality in River
influence on the erosion and transport of sediment by Asian Catchments. Wiley, Chichester, pp. 365 – 389.
rivers; the Yellow River (Huanghe) example. Journal of Geol- Xia, D.X., Wu, S.Y., Yu, Z., 1993. Changes of the Yellow River
ogy 95 (6), 751 – 762. since the last glacial age. Marine Geology & Quaternary Geol-
Milliman, J.D., Farnsworth, K.L., Albertin, C.S., 1999. Flux and ogy 13, 83 – 88.
fate of fluvial sediments leaving large islands in the East Indies. Yang, Z.-S., et al., 1998. Yellow River’s water and sediment dis-
Journal of Sea Research 41, 97 – 107. charge decreasing steadily. EOS, 79, 589, 592.
Morehead, M.D., Syvitski, J.P.M., 1999. River-plume sedimentation
modeling for sequence stratigraphy; application to the Eel mar-
gin, Northern California. Marine Geology 154 (1 – 4), 29 – 41.

You might also like