Вы находитесь на странице: 1из 18

Landscape Ecology 12: 119136, 1997

1997 Kluwer Academic Publishers. Printed in the Netherlands.

Regional ecosystem simulation: A general model for simulating snow


accumulation and melt in mountainous terrain
Joseph C. Coughlan1 and Steven W. Running2
1Johnson Controls World Services, NASA AMES Operation, NASA Ames Research Center M.S. 242-4, Moffett
Field, California 94035-1000, USA; 2University of Montana, Missoula, Montana 59812, USA
Keywords: snow, model, regional simulation, forest, remote sensing
Abstract
A general snow accumulation and melt model was developed to (1) determine how accurately snow accumulation and ablation can be modeled over heterogeneous landscapes with routinely available climatologic,
topographic, and vegetation data, and (2) improve estimates of annual forest snow hydrology for point and
regional calculations of annual forest productivity. The snow model was designed to operate within the
Regional Hydroecological Simulation System (RHESSys), a GIS based modeling system to manage spatial
data for distributed computer simulations on watershed scales. One feature of the RHESSys Snow Model
(RSM) is it can use satellite derived forest leaf area index (LAI) to represent catchment forest cover; difficult
to obtain in adequate cover and resolution by any other means. The model was tested over 3 water years
(October to September) with data recorded by 10 snow telemetry stations (SNOTEL) in 5 states ranging in
meso-climate and elevations from a coastal Oregon site (1067 m) to a continental Colorado site (3261 m).
Predictions for the 10 sites were made with identical parameter values and only site climate varied for all
sites. The average difference between observed and predicted snow depletion dates for all sites and water
years was 6.2 days and 8 of the 30 simulations were within 2 days (R2 = 0.91). Radiation melt was the dominate snow ablation component at the Colorado site where sublimation was 10% (LAI = 0) to 20% (LAI = 6)
of snow loss while air temperature was the dominate component at the Oregon site with sublimation reduced
to 1% (LAI = 0) to 6% (LAI = 6) of snow loss. LAI had a greater effect determining snow depletion than site
aspect. Aspect increased in importance if the snow depletion occurred during early spring when solar insolation differences between hillslopes is greater than in the late spring.
An accurate prediction of daily snowpack water equivalent (SWE) was not a strong determinant for making
an accurate prediction of snowpack depletion date. Predicted snowpack depletion dates were more sensitive to
timing when the snowpack reached an isothermal condition. Daily estimates of SWE were most sensitive to
correctly estimating snowfall from SNOTEL data. This means that for purposes of determining the snow
depletion dates which are useful for forest ecosystem modeling, tracking SWE is less important then triggering snowmelt. Comparisons of simulations to published snow depletion dates show that RSM predicted the
relative ranking and magnitude of depletion for different combinations of forest cover, elevation, and aspect.
1. Introduction
Snow is an important component of the hydrologic
cycle for coniferous forests in western North America (Waring and Schlesinger 1985) and a significant water source for evergreen conifers which are
photosynthetically active in the spring and early
summer when melt water is available (Waring and
Franklin 1979). Forest transpiration and growth are

partially determined by the length of the effective


growing season as it can be constrained by site
water availability (Running 1984). Soil water
depletion begins soon after the snowpack disappears and continues into the summer, causing
forests to undergo increasing water stress (Running
and Nemani 1991; Knight et al. 1985; Black and
Spittelhouse 1981; Troendle and Leaf 1980). Site
water balance models, some incorporating biophys-

120
ical controls on forest transpiration, require an
accurate estimate of snowpack amount and loss in
order to accurately time soil water depletion and
the onset and severity of summer water stress
(Giles et al. 1985; Running 1984; Black and Spittelhouse 1981; Troendle and Leaf 1980).
Forest ecosystem process models predominately
use daily to monthly time steps to model the forest
hydrologic budget and its effect on canopy processes like net photosynthesis and transpiration
(McMurtrie 1992; Ryan 1993; Friend et al. 1993).
One general forest ecosystem model, FORESTBGC (Running and Coughlan 1988; Running and
Gower 1991) has demonstrated how LAI and daily
climate data can explain differences in forest carbon and water cycles without site specific model
calibration.
Process models represent snowpack as a state
variable contains a mass of water (M3/HA) which
accumulates during the winter months and releases
during the spring months. The melt water can be a
significant source of water for generating spring
runoff and resaturating the soil to provide water for
transpiration and growth (Running and Coughlan
1988). For site water balance modeling, simulations can use data describing spring conditions
such as an accumulated and isothermal snowpack
(McLeod and Running 1988). If initialized with an
isothermal snowpack, a process model like FOREST-BGC can employ a simple air temperature
index melt model. There are critical limitations,
however, to an isothermal snowpack model including obtaining the initial amount of snow and estimating when it becomes isothermal. This data can
be measured for a single site, but for landscape level simulations in mountainous terrain data are often
difficult to obtain and are highly spatially variable.
One approach to modeling regional hydrologic
budgets is to divide a heterogeneous study area into
smaller, more homogeneous sub-sites, and execute
individual model runs on each sub-site (Nemani et
al. 1993; Paniconi and Wood 1993). For mountainous terrain, the amount of snow accumulation and
the rate of snowmelt are partially dependent on
landscape features that tend to be highly variable
over short distances (U.S. Army Corps 1956). For
parameterizing a temperature index snow model
for remote sites or regions, snow amounts and the
onset of isothermal snowpack conditions must be
estimated remotely. Direct measurements of

regional snowpack depth and thermal content are


impractical, but the snowpack condition can be
simulated with available data. For predictive purposes a regional estimate of snow hydrology
should be sensitive to environmental conditions,
such as site microclimate and the effects of forest
cover, and topograpic position.
Fortunately, the physical processes governing
snow accumulation and melt are understood and
well documented for small, intensively studied
sites. Energy and mass balance snow models incorporate major physical processes and can produce
good results for a variety of site conditions and climatic scenarios (Anderson 1968; Leaf and Brink
1973a; Leaf and Brink 1973b; Solomon et al. 1976;
Price 1988; Motoyama 1990; Prevost et al. 1991;
Marks et al. 1992; Marks and Dozier 1992). For
watershed level simulations, forest cover variations
and topography were incorporated in models to
simulate snow accumulation and melt for remote
forest stands (Anderson 1973; Leaf and Brink
1973a; Leaf and Brink 1973b; Solomon et al. 1976;
Prevost et al. 1991). These models use either percent canopy closure or cover density to describe
canopy structure. Currently forest process models
use LAI as a definition of total canopy surface area
and (with specific leaf area) canopy biomass that
can be derived from satellite data.
For this study, a different approach towards
modeling snow dynamics was taken for three reasons. 1) There are new methods to sample and
parameterize LAI that are incompatible with previous canopy cover definitions used by earlier snow
models like Leaf and Brink (1973a) and Solomon
et al. (1976). 2) A general snow model must be climate driven and parameterized with generic values for sites across a wide range of climates. There
were no published results available from the previous snow accumulation models or studies demonstrating this general ability nor was such a general
design part of their objectives. 3) If the snow model is linked to an ecosystem model, climatic and
site requirements must be compatible with the data
needs of the regional ecosystem model's approach.
The data requirements and abstraction of forest and
site characteristics must be consistent with the
established ecological abstractions.
The objectives of this study are to (1) test a new
model, RHESSys Snow Model (RSM), simulating
snow accumulation and ablation for a wide variety

121
Coughlan 1988) and the climate simulator model
MTCLIM (Running et al. 1987). The RHESSys
Snow Model (RSM) was built to compute a spatially varying snowpack for regional applications of
FOREST-BGC in RHESSys. This section describes
RSM and the ecosystem process model FORESTBGC, and the climate simulator, MTCLIM, as they
pertain to the RSM.

Fig. 1. Approximate locations and elevations for the study's 10


western SNOTEL (Snow Telemetry stations) study sites operated by the U.S. Soil Conservation Service.

of climates in the western United States. (2) Test


the snow model's sensitivity to changing site composition through tests with combinations of hillslope aspects and forest LAIs. (3) Evaluate if a simple model can represent the basic dynamics of
snow accumulation and melt using climatic driven
differences between sites, and subsequently show
how these climatic driven differences in snow
dynamics are sensitive to forest and topographic
orientation.
The model was tested using data collected from
sensors on 10 U.S. Geologic Survey (USGS) SNOTEL stations; 5 station pairs located in 5 western
States (California, Colorado, Arizona, Oregon and
Montana) over the course of 3 consecutive water
years (October to September) (Fig. 1). Daily snow
water equivalent (SWE) was estimated with the
SNOTEL snow pillow sensor. The absolute difference between observed and simulated SWE calculated daily was used as a measure of error. The
snow model's sensitivity to forest LAI, aspect, and
elevation was tested at the Oregon, Colorado and
Montana stations. For Colorado (highest elevation)
and Oregon (lowest elevation) a single snow season's melt was partitioned into sublimation, radiation driven, and temperature driven melt for comparison of the model in these contrasting environments.
2. Model descriptions
At the core of the Regional Ecological Simulation
System (RHESSys) (Band et al. 1991; Band et al.
1993 and Coughlan and Dungan 1996) is the FOREST-BGC ecosystem process model (Running and

2.1. FOREST-BGC and MTCLIM


FOREST-BGC is a nonlinear process model built
to calculate carbon, nitrogen and water cycling
through a homogeneous forest ecosystem (Running
and Coughlan 1988; Running and Gower 1991). It
computes a daily water balance including precipitation, evaporation, snow accumulation, snowmelt,
soil water content, runoff, and transpiration. The
model treats ecosystem processes in a general manner and has been applied for diverse climates
including Washington, Florida, Wisconsin, Tennessee, and Montana (Running and Coughlan
1988). The climatic data required by FORESTBGC are daily maximum and minimum air temperatures, dew point temperature, precipitation, and
shortwave radiation. When remote site climate data
are not available, FOREST-BGC uses a climatic
simulator, MTCLIM, to calculate remote site
microclimate including daily shortwave radiation
(S) (Running et al. 1987).
MTCLIM was designed for mountainous catchments where microclimate is heavily influenced by
topography and LAI. MTCLIM requires only daily
maximum and minimum air temperatures and precipitation collected from a base meteorological station and can simulate daily S for any sloping surface using sun-earth geometry and daily cloud cover computed from maximum and minimum air temperatures (Bristol and Campbell 1984). Incoming S
on a sloping surface will change maximum air temperatures up to 2C relative to observed temperatues from a flat surface. Forest LAI reduces the
topographic influence on microclimate by reducing
S penetration to the forest floor (Sf) and thereby
reducing the associated air temperature adjustments.
Together, FOREST-BGC and MTCLIM provide
a framework for defining the constraints and data
requirements for building a snow model. One

122
T min + T day
Tnight = __________
2

(1)

where Tmin is the minimum air temperature (C) for


the 24 hr day and Tday is daylight average air temperature (C) which is also from Running and
Coughlan (1988). Tday is defined as
Tday = 2.12(Tmax Tave) + Tave

Fig. 2. Diagram of the RHESSys Snow Model indicating water


flows through the model and major snow accumulation and
melt processes. Listed in the box are essential climatic driving
variables and site description data.

important feature of both models is using a definition of forest structure (leaf, stem and root biomass) obtainable from remotely sensed forest LAI.
Previous stand level models (Leaf and Brink
1973a; Solomon et al. 1976) used canopy closure
or cover density from aerial photographs to sample
catchment forest cover. In the RSM, a catchment
can be parameterized using satellite derived LAI
and topography derived from digital elevation
models. Another change from previous snow models is having a separate climatic simulator,
MTCLIM, which can explicitly estimate daily S
from integration of potential inputs at hourly intervals mitigated by daily transmissivity estimated
from air temperature amplitude.

2.2. RHESSys Snow Model (RSM)


This section contains a description of the major
components and processes in RSM (Fig. 2) and
some variables used by RSM but previously
defined in FOREST-BGC (Running and Coughlan
1988) and MTCLIM (Running et al. 1987, Hungerford et al. 1990).
2.2.1. Snow accumulation
Precipitation falls as snow when the average night
air temperature is below 0.0C. Night average air
temperature is a variable computed in Running and
Coughlan (1988) as

(2)

where Tmax is the 24 hour maximum temperature


(C), Tave is the arithmetic mean of maximum and
minimum temperatures, and 2.12 is a correction for
the sine wave approximation of the diurnal air temperature amplitude that left uncorrected, would
underestimate average daylight air temperature
(Running et al. 1987).
Snowpack water equivalent (SWE) is computed
from incoming precipitation, in this case from the
SNOTEL station gauge. RSM only tracks SWE and
does not model snowpack density or the liquid
water content of the snowpack. Rain water falling
on the snowpack is directly routed into either the
soil compartment or is lost as runoff. The forest
canopy intercepts both rain and snow in proportion
to amount of LAI (Hardy and Hansen-Bristow
1990) and both rain and snow evaporate as a function of daily S. Intercepted snow is held in the
canopy for one day where it is exposed to S and
sublimates. Canopy sublimation losses are the minimum of either intercepted snow or the amount of
snow that can be evaporated by the total S for that
day. At day's end, any residual intercepted snow is
redistributed into the snowpack below as done in
Leaf and Brink (1973a).
Canopy sublimation (qsub) is computed as

Cint

____
qsub = MIN LAI .

_____
,
+

(3)

where Cint is a rain interception coefficient 1 (1.0


mm LAI1) from Running and Coughlan (1988),
is the latent heat of fusion (MJ mm1 M2), is the
latent heat of vaporization (MJ mm1 M2), and S is
the total daily incident S at forest canopy surface
(MJ M2 Day1). Snow interception is calculated as
a function of LAI using an interception coefficient
half the value of the FOREST-BGC defined rain

123
interception coefficient (1.0 mm LAI1). Although
water losses from canopy interception of snow can
be greater than of rain (Waring and Schlesinger
1985), FOREST-BGC's rain interception coefficient also includes interception losses from the litter layer. Sublimation driven by turbulent transfer
is not incorporated in this formulation.

is the sum of the arithmetic average of daily


maximum and minimum air temperatures weighted
by the duration, in fraction of 24 hours, of day and
night. The weighted temperature average is calculated as

2.2.2. Snow thermal content


The RSM uses a simple temperature sum ( ) to
approximate the snowpack's thermal content for the
purpose of determining when the snowpack is
isothermal (U.S. Army Corps 1956; Prevost et al.
1991; Hinzman and Kane 1991). An upper threshold for the is defined to mark when the snowpack switches to isothermal conditions and triggers
a temperature-driven melt process (Hinzman and
Kane 1991). A minimum threshold is also defined
to prevent the from accumulating unrealistically
large temperature deficits which cannot be compensated for during spring warming (Solomon et
al. 1976). The RSM uses one parameter to define
both the minimum value and upper threshold
value. The parameter, called the temperature increment (), defines a temperature range symmetrical
around = 0.0C. The negative value defines
the snowpack's minimum temperature deficit (i.e.,
= 30; - 30C), and the positive value
defines the isothermal threshold (i.e., = 30; =
30C). A large value creates a wide range
between the minimum temperature deficit () and
upper isothermal snow conditions (+) and, therefore more days with above freezing temperatures
are needed to reach isothermal conditions, and consequently to trigger temperature melt.
The is calculated as

with T24 as the weighted air temperature (C), fday


as the daily fraction of 24 hours in light and fnight as
the night fraction. A weighted air temperature average is necessary to adjust maximum and minimum
temperatures to their proportion of a 24 hour period. Maximum air temperatures, usually associated
with daylight hours, contribute less to the in the
short winter days than they do in the spring, when
the daylight period is longer. Day and night length
for a flat surface are calculated with 15 minute precision using yearday and latitude from Running
and Coughlan (1988).
A simple temperature index (mm C1) defines
the daily amount of melt attributed to sensible heat
as measured by air temperature (Equation 2). Temperature driven melt, qmelt, only occurs when the
is above +, regardless of how warm air temperatures may becomes prior to reaching melt
threshold. The index value falls in the lower range
of inex values reported by the U.S. Army Corps
(1956). This component will become less significant in the open where S is much larger and tends
to dominate the melt process.

t = MAX[( t-1 + T24),()]

(4)

where is the current temperature sum for determining isothermal snow conditions (C), 1 is
the temperature sum of the previous day (C), is
the temperature increment parameter (C), and T24
is the weighted sum of the daily maximum and
minimum air temperatures (C) (Equation 5). is
initialized to 0.0C with a new snowpack and is
defined for all days with a snowpack. Snowfall
events do not change nor is the calculation
dependent on the mass of water in the snowpack.

T24 = (Tmax fday) + (Tmin fnight)

(5)

2.2.3. Snow albedo


Snow albedo is an important characteristic needed
to model the S component of snow ablation. In
RSM albedo decay is a linear function of the accumulated air temperature at the snow surface. The
RSM albedo calculation uses a temperature sum
( ) which is computed similar to in Equation 4.
Both temperature indices and time-since-snowfall
are highly correlated during the melt phase (U.S.
Army Corps 1956; Baker et al. 1990).
Time-since-snowfall has commonly been used as
the predictor for albedo since it is easy to compute
and measure (Anderson 1968; Leaf and Brink
1973; Baker et al. 1990). The rational for a temperature based albedo decay is that for the entire fallwinter-spring snow season, the amount of thermal
energy is not linearly related to time, so as a work

124
around, a seasonally adjusted decay rate is calibrated empirically (Leaf and Brink 1973) to reflect the
changing albedo-time relationship. A temperature
based albedo decay need not be seasonally adjusted
because by definition it will have a slower decay
rate in the cold winter month and correspondingly
an accelerated decay rate during the warm melt
phase (U.S. Army Corps 1956; Baker et al. 1990).
Snow surface albedo is calculated as


t = MIN (t1), max . ___ (6)

where t is the current albedo (fraction), t1 is the


previous day's albedo (fraction), max is the maximum albedo (fraction), is the temperature increment (C) and is the temperature sum (C) from
Equation (4) with one modification. Since snow
albedo is refreshed to a maximum albedo after
snowfall events (Aguado 1985), the temperature
sum used to calculate albedo is reset with each
snow event. was created independently of
because cannot be reset to 0.0C with each
snowfall event without altering the isothermal
snowpack calculation, Equation 4. The representational difference between the and is that
represents the snowpack thermal content of the
entire snowpack while represents the accumulated effect of thermal energy on the snowpack surface. is reset to 0.0C with each snowfall event
to recalculate snow surface albedo while is never reset unless there is a new snowpack.
Snow albedo has a maximum value of 0.88
(fraction) and a minimum value of 0.50 (fraction)
(Aguado 1985; Anderson 1968; U.S. Army Corps
1956). Model parameter values that define the linear relationship between and albedo were selected from calibration on empirical observations made
with data collected at a ski resort approximately 60
km southwest of the MT study sites (Potts unpublished). The decay rate was set at 0.004 (fraction)
decrease in albedo (fraction) for each 1C increase
in . For example, when is 30C, the albedo is
0.76 (0.88 (30 0.004)).
Equation (6) selects the minimum albedo of
either the calculated albedo or the previous day's
albedo, to prevent snow albedo increasing with
decreasing temperatures. Computationally can

decrease with colder temperatures associated with


a cold snap, and a lowering of will, incorrectly, compute a higher snow surface albedo. Taking
the minimum of the past and computed albedo values produces a constant snow albedo during a cooling trend.
2.2.4. Snow ablation
Once snow albedo is defined, radiation driven
snow ablation is computed using the amount of
daily S incident on the snow surface and then
added to the amount lost from the temperature driven component. S passing through the forest
canopy is attenuated using Beer's Law.

LAI

C .___
ext 2.2

Sf = S . e .

(7)

where Sf is the amount of shortwave penetrating to


the forest floor, S is the total incident S on the
canopy surface, Cext is the Beer's Law extinction
coefficient (0.5), and all sided LAI is converted to
projected LAI by dividing by 2.2 (dimensionless)
(Running and Coughlan 1988). S can be calculated
by MTCLIM (Hungerford et al. 1990) or can be
directly measured.
Sf is incident on the snow surface under the
canopy, and Sf can either melt or evaporate the
snowpack depending on Tday (from Equation 2).
When Tday is below 0.0C the snow sublimates,
and when it is above 0.0C the snow melts. The
calculations for each process are nearly identical. S
driven snowmelt is

Sf
qmelt = ___

(8)

where Sf is the amount of S reaching the forest


floor (MJ M2 day1) and albedo is the snow surface albedo (fraction). (3.5E+5 MJ mm1 M2) is
the latent heat of fusion which is needed to change
snow to water. S driven sublimation is

S
f
qsub = ___
+

(9)

where the terms remain the same as defined in

87,88,89
87,88,89
85,86,87
84,85,86
86,88,89
87,88,89
88,89,90
85,86,87
88,89,90
88,89,90

30
101
30
30
30
30
30
30
30
30

0.52
1.0
1.0
1.0
1.0
1.0
0.52
1.0
0.52
1.0
1.0
1.0
1.3
1.3
1.4
1.5
1.0
1.0
1.3
1.0

Adjusted RSM coeff.

qmelt
Snow ppt.
(C) coeff.
Multiplier
(mm/C)
31
39
70
59
29
40
78
29
27
25

23
21
n/a
n/a
n/a
n/a
68
n/a
22
n/a

274
391
1112
853
279
789
523
170
208
358

94
171
463
336
125
359
206
99
76
143

11
10
6
7
10
5
15
17
13
7

33
23
15
18
23
11
38
29
36
17

24
12
n/a
n/a
n/a
n/a
27
n/a
29
n/a

Generic Adjusted Max. Mean Generic Generic Adjusted


Ave 3
Ave 3
Snow4 Snow5 Max 3,6 Mean 3,7 Mean 3,7
(mm/day) (mm/day) (mm) (mm) (%)
(%)
(%)

Outputs

1Generic

value of 30 () was used in generic results; 2Generic value of 1.0 (mm/C) was used in generic results; 3 = | observed - predicted | / n, where n = number of days
observed or predicted snow > 0; 4Max Snow is maximum observed SWE over 3 water years; 5Mean Snow is average observed SWE for days with snowpack for the 3 water
years; 6Max is as a % of Max Snow; 7Mean is as % of Mean Snow.

34.3
34.2
38.2
38.1
40.1
40.3
44.2
42.2
46.5
46.5

2225
2417
2682
2896
2658
3261
1067
1756
1426
1905

AZ1
AZ2
CA1
CA2
CO1
CO2
OR1
OR2
MT1
MT2

Baker Butte
Promontory
Sonora Pass
Virginia Lakes Ridge
Stillwater Creek
Lake Irene
Jump Off Joe
Quartz Mountain
Lubrecht Flume
NFK Elk Creek

Site data
Elevation Latitude Water years
(m)
()
OctSept

SNOTEL
Abbrev. Station ID name

Inputs

Table 1. SNOTEL Site descriptions, summary input data and RESSys Snow Model (RSM) parameters are listed for generic and adjusted simulations. Output results are summarized with predicted (snow water equivalent) SWE errors () which average results from 3 water year simulations. Adjusted simulations were run for sites with high and listed,
sites without adjusted results were filled with n/a.

125

126
Equation (8) with the exception of using the sum of
and (MJ mm1 M2). Total sublimation is the
sum of sublimation from intercepted snow held in
the canopy (Equation 3) and from snowpack on the
forest floor (Equation 9).

3. Methods
3.1. Study sites
10 study sites were used with site data collected
during 3 water years (defined as October to September). Two sites, each pair from five western
states: Arizona, California, Colorado, Montana and
Oregon, were chosen (Fig. 1) to represent some of
the more diverse snow regimes found in western
North America as described by Baker (1944). The
site data were collected using the Central Forecast
System's central computer database administered
by the Soil Conservation Service (SCS) (Soil Conservation Service 1985). The selection criteria for
the 10 sites were: (1) Station pairs located within
the same state, and as close as possible to one
another so that they have similar mesoscale climate
influences, (2) each state has a high and low elevation station, and (3) each station has 3 water years
data collected during the snow season.
All SNOTEL stations are located in mountainous areas in forest clearings of variying size and
were simulated with an LAI of 0.0. The one exception is the Baker Butte AZ SNOTEL station which
is located along an abandoned and very narrow
road that runs north to south, and the station is
heavily shaded to the east and west. An LAI of 1.0
was used to represent the snow pillow shading
(USGS SNOTEL site description).

3.2. SNOTEL data


Daily microclimatic data downloaded from the
SNOTEL computer included: date, maximum temperature, minimum temperature, cumulative precipitation and SWE measured by the snow pillow
sensor. Although desirable, it was not possible to
use climate data for the same water years on all
sites because some stations' data were incomplete
due to sporadic sensor failure. When possible, the

period from 1987 to 1990 was chosen, and all


water years are between 1985 and 1990.
Corrections to collected SNOTEL data were
required prior to using them in the climatic simulator, MTCLIM. Air temperatures that were missing
for periods from 1 to 3 days were substituted with
the previous day's air temperatures. The SWE values from the snow pillow sensor required infrequent fixes to correct negative values which had
been recorded during the end of spring melt. Negative SWE values were set to zero. It was assumed
that the pillow had no snow when the sensor stayed
at a constant value just above 0.0 SWE (USGS P.
Pharnes personal communication).
Some sites had significant discrepancies between the amount of SWE measured on the pillow
and the amount of precipitation measure by the
precipitation gauge. A constant correction coefficient was added to increase gauged precipitation
amounts to accommodate the greater snow accumulation on the pillow (U.S. Army Corps 1956;
Hinzman and Kane 1991). Gauge corrections were
station specific but constant for all 3 years and
ranged from 130% to 150% of each snowfall event.
Motoyama (1990) used similar coefficient values to
correct gauge measurements for predicting snow
accumulation in the mountains of Northern Japan.

3.3. Snow depletion dates and daily SWE


The RSM was incorporated into RESSys and connected to MTCLIM and FOREST-BGC. The models used parameter values listed in their respective
papers and with the site parameters footnoted in
Table 1 (Running and Coughlan 1988; Running et
al. 1987; Hungerford et al. 1990).
SWE was simulated for the 10 stations in Figure
1 using all 3 water years. Identical model parameters were used except for differences in the physical site characteristics such as latitude and elevation (Table 1). Additional simulations were run on
some sites that had large differences between
observed and predicted SWE when general parameters were used. These sites had a RSM parameter adjusted to a consistent value for all three years
(Table 1).

127
3.4. Forest cover and topography
Sensitivity tests were run for an Oregon and Colorado station and both Montana stations using
combinations of aspect, elevation, and LAI. Two
stations were selected that represent the most
extreme conditions: Jump Off Joe, Oregon (1067
m), a warm maritime climate with the lowest elevation of all ten stations and Lake Irene, Colorado
(3261 m), a cold continental climate and the highest elevation station (Baker 1944). Both Montana
SNOTEL stations represent an elevation gradient
since they are located in the same small catchment
(17 km2) with a 500 m elevation difference.
For each of the 3 sites, a set of simulations were
made for LAI's of 0 and 6, to represent a forest
clearing and a typical forest stand. An LAI of 6 is
half the maximum LAI observed for lodgepole
stands in the Central Rocky Mountains (Kaufmann
et al. 1982) and is the LAI value used to represent
an average forest stand. In addition to LAI, three
aspects; facing north and south with 25 slope and
a flat surface, were chosen to show the effect of
forest cover and aspect combinations. Elevation
sensitivity tests were replicated in Montana on a
dry (1990) and wet (1991) water year.

4. Results and discussion


4.1. Snow depletion dates and daily SWE
4.1.1. SWE
Model and site parameter values for all ten study
sites are in Table 1. Results for all simulations are
also summarized in Table 1 along with model prediction errors. Error is defined as the average of the
absolute difference between observed (O) and predicted SWE (P) and it is calculated daily for all
days (n) when either observed or predicted SWE is
non-zero ( (OiPi) / n). The error term () was put
to a ratio with the station's three year means SWE
and to the maximum SWE of the three years. The
error ratio expresses error relative to SWE and
accounts for sites with heavy snowfall that have
larger absolute errors than drier sites simply due to
their larger snowfall amounts (Motoyama 1990).
The error percentages reported in Table 1 are similar in both range and magnitude to those in

Motoyama (1990).
Summaries of daily SWE predictions are in
Table 1 where indicates how well predicted SWE
tracked observed SWE for the three snow seasons.
was less for higher elevation and colder stations
while warmer and lower elevation stations, like
Arizona and Oregon, experienced intermittent winter melt episodes and had a greater . In general, it
is easier to accumulate and melt snow when there
is a well defined accumulation and ablation season,
as is found in the higher elevation study sites.
Precipitation gauge corrections (Table 1) reconciled differences between snow pillow SWE and
SWE estimated from gauged precipitation for
SNOTEL stations in Colorado, California and
Montana. When a precipitation gauge clogs, precipitation can be approximated by accumulated
snow increments. To correct for the errors, a precipitation correction coefficient adjusts gauge measurements to match snow accumulation (U.S. Army
Corps 1956; Motoyama 1990; Hinzman and Kane
1991). The alternative explanation for discrepancies between pillow and precipitation measurements at the same location is snow redistribution
from the canopy to the snow pillow. Snow redistribution is well documented for coniferous forests
ranging from Colorado (Gray and Troendle 1967)
to Oregon (Berris and Harr 1987) to Alberta (Golding and Swanson 1978). This study used gauge corrections to SWE deficiencies because the coefficients were approximated proportional to elevation
and snowfall intensity which indicates gauge clogging (Hinzman and Kane 1991).
4.1.2. Snowpack depletion dates
A scatter plot (Fig. 3) of observed to predicted
snow depletion dates was made for all 30 simulations using generic RSM parameter values of =
30 and qmelt = 1.0 mm/1C (Table 1). Each station
is plotted three times, once for each water year, and
is identified by its state abbreviation and station
number given in Table 1. Located at the upper right
of the scatter plot (Fig. 3) are the snowy, high elevation sites like Colorado and California, while the
lower left are warmer and lower elevation sites of
Oregon and Arizona. A two week interval drawn as
dotted parallel lines in the scatter plot (Fig. 3) contained 18 of the 30 predicted dates. The average
error between observed and predicted dates is 6.2

128

Fig. 3. Scatter plot and regression line for observed and predicted snow depletion dates for each SNOTEL station and
water year. Stations are identified by state abbreviation and station numbers from Table 1, and the abbreviation's center letter
marks the exact date. Lines defining an error interval of one
week contain 18 of the 30 simulations.

days and 8 of the 30 points are within 2 days.


The linear regression between observed and predicted dates was forced through the origin (Y =
0.0) and has a R2 of 0.91 and the X coefficient is
1.01. Observed snow depletion dates began in midMarch and the last was in mid-June, a 90 day
range. Within the interval, predicted melt dates fall
around the regression line without noticeable bias
for either maritime or continental stations.
The observed snow depletion at Sonora Pass,
CA (CA1) (2682 m) had the most variation
between water years. Depletion dates varied by 30
days between water years 1985 and 1987 and the
station recorded a maximum seasonal SWE of
1150 mm in 1986 and 350 mm in 1987 (Fig. 4).
Within that wide range, RSM predicted snow
depletion dates that were in error by 1, 1, and 11
days respectively for water years, 1985, 1986, and
1987. Strangely, in 1987 RSM did a better job of
accumulating the snowpack (Fig. 4), but overestimated spring melt resulting in an 11 day discrepancy. The 1986 simulation underestimated daily SWE
by as much as 150 to 200 mm for some days but
the simulation had timed snowmelt and snow
depletion dates within 1 day of observed dates.
Results from Figure 3, Figure 4 and Table 1 suggest that predicting snow depletion is dependent on

Fig. 4. Sonora Pass, CA. Daily observed SNOTEL snow pillow


SWE (snow water equivalent) and snow simulations for 1985 to
1987 water years.

the timing and magnitude of the spring ablation


rate rather than accurately tracking SWE. SNOTEL
stations with a high (Table 1) have an average
depletion date difference of only one or two days
more than the average difference. For example,
Oregons two stations and the Lubrecht Flume, MT
station have a mean approaching 30%, yet their
predicted dates are an average of only 7.3 days
from observed. The average for all stations is 6.2
days, only one day shorter.

4.2. Albedo
Both snow depletion and daily SWE predictions
require an accurate estimate of S and snow albedo
(Aguado 1985). Snow surface albedo is usually
calculated as a function of time because time is
both a good predictor and easy to measure (Baker
et al. 1990). Models that use time usually divide
the snow season into a pre- and post-ablation
phase. In the pre-ablation phase albedo decay is
slow while in the post phase albedo decay is accelerated (Gray and Landine 1987; Loth and Graf
1993). Baker et al. (1990) demonstrated that an
albedo calculation with temperature index, , can
also predict albedo. Air temperatures change from
the winter to spring so a calculation need not
define a pre and post ablation phase. directly
relates albedo decay to air temperature which

129
makes the albedo calculation more adaptive
through the snow season and also between sites.

4.3. Model parameter sensitivity


RSM melt components use two parameters,
(30C) and qmelt (1 mm C1) and two temperature
sums, for albedo and for thermal status.
These parameters and sums have definitions that
are somewhat unique to RSM but their values can
be compared to previously published parameter
values.

Albedo is calculated with the parameter (30C)


which represents a decrease in albedo of 0.004
(fraction C1) for every degree increase in t24
(Equation 6). The U.S. Army Corps (1956) published a similar albedo-temperature decay rate of
approximately 0.0015 (fraction C1) but it used
maximum air temperature. RSM used the weighted
24 hour air temperature to better represent the
duration and extent of diurnal warming and cooling
cycle. RSM's larger coefficient (0.004 fraction C1)
is expected because RSM uses a lower temperature
driver, t24, while the Corps coefficient (0.0015 fraction C1) was driven with tmax.
qmelt
A qmelt of 1 mm C1 is lower than the 2.0 to 2.5
mm C1 rates reported for other temperature based
melt models (Prevost et al. 1988; U.S. Army Corps
1956); however, the cited temperature index models have only an air temperature melt component
while the RSM also has a radiation melt component. The lower RSM parameter value should be
expected to be smaller than a value used in a model
that melts exclusively with temperature.

RSM uses to approximate the snow's thermal


content and to estimate the heat at the snow surface for an albedo calculation. Both and were
derived from experiments conducted with data
from the Montanan SNOTEL stations and data
from other unpublished data sets collected in western Montana near the two SNOTEL stations (Potts
unpublished data). is a modification of a degree
sum described in U.S. Army Corps (1956) and the
weighted max-min air temperature approximates
the extent of diurnal heating and cooling making

Fig. 5. SNOTEL snow pillow SWE and simulated SWE made


with generic ( = 30C) and site adjusted ( = 10C) heat increment coefficient for the 1988 water year at Promonotory, AZ.

the computation more robust across climate types


and latitudes. The parameter value (30) defining
upper and lower thresholds for and is empirical and comes from examination of the SNOTEL
pillow and temperature records for the two adjacent
Montana SNOTEL stations (Fig. 1) with data for
water years prior to those used in this study.
RSM sensitivity to and
Site specific RSM parameter values reduced model
error (Table 1). The for some stations was
reduced with site specific parameter value adjustments listed in Table 1. In general these adjustments decreased snowmelt rates for the relatively
warmer sites. The site specific coefficient values
were coarse modifications made in discrete units
and were not determined by mathematical optimization methods or curve fitting.
For the Promontory, AZ station, was lowered
from 30C to 10C to slow snowmelt, and it
demonstrates RSM sensitivity to . The adjusted
had negligible effects on snowpack accumulation
when the station was getting progressively colder
(Fig. 5), but differences occurred in the timing of
spring melt. In Figure 5, the original simulation (
= 30C) missed the collapse of the snowpack while
the adjusted ( = 10C) simulation better mimicked
this sudden melt episode. The difference between
predicted and observed depletion dates was

130

Fig. 6. Simulated snow loss attributed to sublimation, temperature melt and radiation melt for Jump Off Joe, Oregon (1067
m) and Lake Irene, Colorado (3261 m) under a forest canopy
(LAI = 6) and in the open (LAI = 0).

reduced from 7 to 2 days and SWE error decreased


from 39 mm to 21 mm (Table 1).
The net effect of lowering to 10C is to both in
crease melt sensitivity to both S and air temperature. A lower value had three effects, each of
which demonstrate the methods by which RSM
ablates snow. First, a lower value reduced the
winter temperature deficit in and , therefore,
smaller temperature deficits are accumulated in the
winter which has less of an impact in the spring. In
the case in Figure 5, a value of 10C created a
winter and deficit of 10C instead of 30C.
Second, the snowpack became isothermal sooner,
when reached +10C instead of +30C, and
melts earlier (Equation 4). Lastly, a smaller
accelerates the albedo decay rate per increase in
site temperature because the albedo decay function
used a ratio of the to (Equation 6).
4.4. Forest cover and topography
4.4.1. Forest cover
Net radiation, including longwave, is not routinely
measured, so both air temperature and incoming S
have been used as surrogates for approximating net
radiation (U.S. Army Corps 1956; Male and
Granger 1978; Price 1988). Air temperatures used
for tracking longwave driven melt in RSM are

adjusted in the MTCLIM model according to site


location, forest LAI and its estimated S.
Snow ablation in a forest (LAI = 6) and in the
open (LAI = 0) are illustrated in Figure 6 for Oregon and Colorado climates. The example uses
water years that coincidentally had nearly identical
snowfall inputs (Fig. 6). For Lake Irene, Colorado,
S accounted for 88% of snow loss in the open and
45% in the forest. Less than 5% of total snowfall
losses were due to air temperature in the open, but
under the canopy it accounted for 33%. Excluding
canopy sublimation losses, air temperature melt
accounted for 65% of total snowpack loss under a
canopy in Oregon and 42% in Colorado. These
results follow the general patterns observed in
forests, where net radiation, long and shortwave, is
the principle snowmelt component and that air temperature is second (Male and Granger 1978; Price
1988).
Sublimation losses are difficult to measure but
estimates for Colorado were comparable to field
estimates reported by Golding and Swanson
(1986), which were 33 mm on the north slope and
58 mm on the south slope for a uniformly and partially harvested basin in Alberta, Canada. Their
results suggest that sublimation differences
between the two slopes were controlled by incident
S which was the only reported variable that varied
between the two slopes and caused sublimation differences. RSM's sublimation in Colorado was 80
mm in the forest and 40 mm for the bare snow surface. Oregon's sublimation losses (LAI = 6) were
less compared to CO but were greater in the forest,
and because RSM was designed to use S to sublimate canopy intercepted snow. Sublimation was a
minor portion of total snow loss off the open snowpack because the cloudy OR site has low S, and
with above-freezing temperatures, RSM was
designed to use S to melt snow (Equation 8) rather
than sublimation snow (Equation 9).
Snow interception and snow drop estimates may
be overestimated in relatively warm snowpack
zones (300 to 1100 m in Oregon), typified by Jump
Off Joe OR (Berris and Harr 1987). RSM evaporates canopy intercepted snow for one day and the
remaining snow drops into the snowpack and accumulates as snowpack SWE. The snow drop is modeled as snowfall because RSM does not model
snow quality other than estimating snowpack

131
isothermal conditions as . Therefore, the thermal
quality of the snow that drops off the canopy is, for
purposes of model simplicity, set to the snowpack's
condition, , although snow drop may really occur
as isothermal clumps or as melt water. A second
compensating error is in sublimation. Canopy snow
is sublimated with S and snow cover is sublimated
with S during sub-freezing conditions. This design
may generate an overestimate of sublimation from
S. On the opposing side, no turbulent transfer driven sublimation process is in RSM. Forest-BGC
and RHESSys does not have a dynamic windspeed
variable so the inclusion of a wind driven melt
process such as turbulent transfer would add to the
model's data needs and violate a design requirement. Further model verification is necessary to
determine if wind speed, a difficult variable to
obtain and highly variable in mountainous terrain,
is needed to improve model results.
Field data from Berris and Harr (1987) documented higher daylight air temperatures and melt
rates in the Oregon forest clearings than under the
adjacent canopy. Their reported air temperature differences were highest when S was prominent
showing a strong link between air temperature and
incoming S. Their microclimate and snowmelt data
indicate that a canopy reduces both thermal and
solar components of melt, as is predicted by RSM.
S melt dominates the temperature melt process in
the open simply because S is much greater in the
open. These data indicate that the major melt
process driving RSM are tracking observed microclimate conditions in the open and forest but the
data do not indicate if RSM snowmelt components
are correct except in relative magnitudes.
4.4.2. Topography and LAI
Aspect and elevation sensitivity were tested using
the same two Oregon and Colorado stations and
both Montana stations which are located in the
same small basin and also represent on an elevation
gradient of 500 m.
Aspect
Although their climates were different, the relative
melt rates between the three surfaces; N, S and flat,
were similar in both Oregon and Colorado climates. Colorado was more sensitive to aspect than
Oregon, which is a result of Colorado's clearer

Fig. 7. Jump Off Joe, Oregon's simulated snowpack for three


surfaces, flat, north and south facing, with a 25 slope, under a
forest canopy (LAI = 6) and bare surface (LAI = 0).

Fig. 8. Lake Irene, Colorado's simulated snowpack for three


surfaces, flat, north and south facing, with a 25 slope, under a
forest canopy (LAI = 6) and bare surface (LAI = 0).

skies and larger quantities of S. Large amounts of S


also explain why Colorado's the melt difference on
identical aspects with forest and bare surfaces was
approximately 18 days, 8 days longer than that
found in Oregon's climate. Simulated melt under a
canopy at Jump Off Joe, Oregon (LAI = 6) was
also less than melt in the open (LAI = 0) (Fig. 7)
for all aspects: N, S, and flat. The difference
between the open and forest was exaggerated by a
late April cold snap. The subfreezing temperatures
postponed melt under the canopy while the open
site had lost all snow prior to this cold snap. Snowpacks in the forest and open would have melted
within 10 days of each other, as determined by

132
Table 2. Observed and simulated snowpack depletion dates showing snow depletion sensitivity to elevation, LAI, and aspect. Elevation
sensitivity was replicated at the MT stations and entered as year 2 while all single year observations were placed into year 1. Locations
are all in the inter-mountain states of western U.S.: CO, MT, and NM. Observed depletion dates are from the literature and simulated
dates are made with SNOTEL station data, no climatic data were available for simulating snow depletion at literature sites. The SNOTEL depletion data is listed for observed, flat, LAI = 0 entries in CO and MT and simulation results were made for those exact sites.
Measurement Year Elevation

Observed

High

Low

2
Simulated

High
Low
High 1
Low 1

High 1
Low

Aspect
Location

CO

Flat
MT4

LAI 0
09-May
LAI 1.53.0 20-May1
LAI 56
02-Jun1
LAI 9
LAI 0
20-Apr
LAI 1.53.0
LAI 56
LAI 0
10-May
LAI 0
19-Apr
LAI 0
13-May
LAI 56
06-Jun
LAI 0
12-Apr
LAI 56
22-Apr
LAI 0
18-May
LAI 56
09-Jun
LAI 0
13-Apr
LAI 56
22-Apr

CO

North
NM3

MT4

CO

South
NM

06-May3

06-May3

17-Jun3

10-Jun3

MT4

13-May2
08-Apr2

09-Jun
26-Jun

16-May
07-Jun
18-Apr
26-Apr
23-May
10-Jun
19-Apr
25-Apr

05-Jun
23-Jun

12-May
05-Jun
10-Apr
21-May
17-May
08-Jun
10-Apr
21-Apr

1Observed LAI is 1.52.5 and 5.0 both at 2680 m in Colorado (CO) (Gray and Troendle 1982); 2N slope with LAI 5 and S slope with
LAI 3 at 2750 m site in New Mexico (NM) (Gray and Coltharp 1967); 3N and S hillslope pairs with and without LAI at 3400 m with 7
day sampling interval (Gray and Coltharp 1967); 4All MT data is from SNOTEL stations MT1 and MT2 in Table 1 and for water years
'89 (1) and '90 (2).

extending the melt curve in Figure 7 to the X-axis.


LAI was the most important factor determining
snowpack duration while aspect was secondary
(Figs. 7 and 8) at both warm, cloudy Oregon and
cold, sunny Colorado sites. The canopy reduced
accumulated SWE and prolonged the snowpack by
shielding snow from melt energy. South slopes
received the greatest amounts of S and had the
fastest melt rates while the opposing north slopes
had the least amount of S and retained their snow
longer. S on south facing and flat surfaces for
March 15th in CO is 30,000 and 25,000 MJ M2
Day1 respectively compared to only 15,000 MJ
M2 Day1 on a north surface (Buffo et al. 1976).
The small simulated melt differences between
north and south (N-S) aspects was understandable
for cloudy Oregon where melt is more temperature
driven. The 4 days difference between slopes at
Colorado, however, was less than expected unless
S loads were not very different on south and north
aspects during melt in the later spring (Fig. 8).

4.4.3. Field observations


Observed snow depletion dates from Gray and
Troendle (1982) and Gary and Coltharp (1967),
were tabulated (Table 2) to verify if RSM sensitivity to aspect, LAI and elevation is similar to field
observations. The comparison is relative and not
absolute in terms of specific dates between studies
and simulations because no climate data was available to do simulations for their sites. Gray and
Coltharp (1967) studied effects of forest cover and
aspect on seasonal SWE in northern New Mexico
at 3390 m and 2766 m elevations with weekly
SWE observations. Site vegetation included
Spruce, Douglar Fir, and grasses. Gray and Troendle (1982) reported the effect of stand density
(LAI) on SWE for Lodgepole Pine on flat surfaces
in Colorado at 3000 m. SWE was observed at
irregular time intervals during the melt phase. Both
studies reported stand densities which were converted to LAI using methods in Kaufmann et al.
(1982). Unlike the other studies, the MT sites testing elevation sensitivity are SNOTEL stations

133
which allows comparisons with simulated and
observed data. SNOTEL data was readily available
which allowed replication over two new seasons:
WY 1990 (as year 1) and 1991 (as year 2). All other observed and simulated comparisons were from
varying water years and merged (as year 1) in
Table 2.
At high elevations, both observed and simulated
hillslopes had small differences between opposing
N-S slopes (Table 2). The resolution of the
observed data is 7 days; hence the 7 day difference
for an LAI of 9 may not be 7 days. The simulated
differences for Colorado N-S slopes with LAI of 0
were 4 days apart and with LAI of 6, 3 days apart.
With a weekly sampling interval the difference
would also be 7 days.
Larger melt differences occurred between opposing LAI treatments (forested and bare) than opposing site aspect (N-S) but differences were greatest
between forested north slopes and bare south
slopes. With an observed LAI sensitivity test on a
south facing slope, a high elevation site with an
LAI of 9 prolonged the observed snowpack by 35
days (NM). A simulated LAI of 6 prolonged site
depletion by 18 days (CO). The largest melt differences between various LAI and aspect conditions
occurred between forested north and bare south
surfaces. When aspect was set to opposing N-S
conditions the melt differences widened from 35 to
42 and from 18 to 21 days respectively for LAI of
9 on a N (NM observed) and LAI of 6 (CO simulated). For flat surface, observed depletion in CO
had a 13 day difference between high elevation
sites with a LAI difference of 2.04.5 (1.53.0 versus 5.06.0).
Simulation results from both Montana SNOTEL
stations indicate that the effect of site orientation is
codependent on elevation (Table 2). The time lags
between depletion dates for bare N-S aspects
increased from 4 days at the high station to 8 days
at the low station. With no elevation effect the two
sites should have melted with the same time difference. The 4 day difference (84) is a doubling of
time for a 500 m elevation drop. Simulations made
again for the next water year also had a doubling
between the two elevations. The high site N-S difference was 5 days and the low site was 9.
To ascertain if radiation was responsible for the
melt differences between elevations, a simulated

LAI of 6 was added to shade the snow from S.


Adding the LAI to the N-S hillslopes decreased the
differences down from 9 and 5 days to 2 days for
both water years. There is no field data available to
test this result but it agrees with sensitivity of
snowmelt to S. In summary, simulated differences
in depletion dates on bare surfaces were larger at
lower elevations, but the simulated differences
were muted at both the high and low elevation stations by adding an LAI of 6.
In the field the same codependence between
aspect and elevation was observed by Gary and
Coltharp (1967). For their low elevation site in
NM, they measured a LAI difference of 2 (5 on a
north slope and 3 on a south slope), and a large 35
day melt difference. For their high elevation stand,
a much larger LAI difference of 9 (9 on north slope
and 0 on south slope) produced a 42 day difference
which is only 7 days longer despite the much larger
LAI difference. Both field sites were within 1.6 km
of one another and were located on similar gradients so the observed melt dates (measured at weekly intervals) may be attributed to LAI, S and site
elevation interacting with aspect. The higher elevation sites are becoming isothermal and melting
when solar input differences between hillslopes is
minimal. If the air temperatures warm in the early
spring then the differences in S that drive melt are
greater such that even the lesser of the two LAI
gradients (5 vs. 3) combined with increased S can
produce a large melt effect than a larger gradient (9
vs. 0) with lesser S.
Gray and Coltharp's 35 day N-S melt difference
does not agree data from flat surfaces from Gray
and Troendle (1982) which indicates S is a significant factor since it is the variable that varies
between N-S slopes. They reported only a 13 day
melt lag between sites with approximately the same
LAI differences (2.53.0). Their high LAI stand
had 25,000 to 50,000 stems HA1 indicating a
closed canopy while the low LAI stands were
thinned with large gaps in the canopy as is seen in
their published site photographs. These results suggest that the interaction of N-S surfaces and LAI on
incoming radiation has its greatest effect at low elevations and produces the greatest melt difference
between opposing hillslopes.

134
4.4.4. Aspect, elevation, and yearday
The differences in melt can be quantified and
explained by comparing S on different slopes. High
net radiation gradients found between N-S hillslopes decrease as the sun's azimuth increases in
the late spring (Buffo et al. 1976) causing a
decreasing effect of aspect on snow depletion. On
March 15 the simulated input is 15,000 MJ M2
Day1 on a north slope and 30,000 MJ M2 Day1
on a south slope. The respective S inputs on June
21st are 30,000 MJ M2 Day1 and 33,000 MJ M2
Day1 for N-S slopes. Lower elevation snowpacks
become isothermal early in the spring and experience greater net radiation gradients between hillslopes, and melt with increased sensitivity to
aspect. High elevation snowpacks are cold and
become isothermal in late spring when radiation
differences between hillslopes is less and total
incoming S is high. Both N-S hillslopes receive
near equal portions of S in the late spring therefore
aspect differences are less than earlier in the year
(Aguado 1985). Adding LAI shades the snowpack
from S and reduces the aspect effect in the early
part of the spring when they can be large, as seen in
Table 2 and Figure 8.
RSM results extrapolated to a forested watershed
would predict an earlier and higher peak discharge
in a harvested watershed due to the reduction in
LAI causing 1) faster melt in the forest clearings
and 2) overall increase in meltwater due to reductions in canopy sublimation losses. Sublimation
losses are difficult to accurately measure in the
field, however Golding and Swanson (1986)
reported 33 mm losses on the north slope and 58
mm on the south slope for a partially harvested
basin in Alberta, Canada with equal harvest treatments. The increased losses in water yield on the
south slope suggest that S contributes to the difference by means of increased sublimation. Troendle
and King (1985) observed annual discharge patterns for 30 years in experimental watersheds with
various harvest treatments and found runoff patterns supportive of RSM predictions. They found
large and early peak stream discharge after forest
cutting treatments and a gradual returning to predisturbance discharge patterns which are less sudden and occur later in the year as the forest grows
back and LAI increases.

5. Summary
Snow accumulation and melt can be simulated for
a wide range of conditions found in the western
United States using a simple snow model and a
generic set of model parameters. RSM predicted
snow depletion with a R2 of 0.91 and an error of
6.2 days from observed depletion dates for 30 simulations at 10 diverse sites in western North America. For the purpose of calculating site water balance and plant water stress, the snow depletion date
is important since melt water is a significant source
of site water and its disappearance marks the
beginning of soil water depletion (Knight et al.
1985). Tracking daily SWE accumulation and loss
is important for validation to SNOTEL data and for
estimating spring runoff. Strangely, a simulation
with a low SWE error did not necessarily indicate
an accurate depletion date prediction and vice
versa. This independence of the two properties is
significant when considering that validation of
regional snowmelt models can be made with
remotely sensed observations of depletion or by
streamflow measurements with SNOTEL point
estimates. For validation, regional simulated snow
cover should be compared to both remotely sensed
snow cover and stream flow in gauged watersheds.
RSM accumulated smaller snowpacks under a
forest canopy than in clearings and had greater net
water losses to sublimation. Snowpacks under a
canopy however, melted slower than in the open
because of reduced melt energy at the snow surface
which caused the melt to occur later into the season. LAI is the most important variable determining snow accumulation and melt for high elevation
sites. As elevation decreases, the model shows that
site aspect becomes increasingly important in
determining snow depletion dates, although still a
secondary variable.
Extrapolating RSM results to a landscape, a partially harvested watershed will have an earlier and
higher peak runoff due to faster melting in the
openings and less sublimation losses than a forested watershed. With RSM's simple parameterization
needs, snow cover simulations on a basin scale are
possible with satellite derived LAI and digital
topography from DEMs. A computer version of the
RHESSys Snow Model is available from the First
Author upon written request.

135
Acknowledgements
This research was guided by Steve W. Running,
and conversations with L.E. Band, D.L. Potts, R.R.
Nemani, J. Dungan. Assistance in understanding
the conditions at each SNOTEL site came from
many USGS personnel located in the States of AZ,
CA, CO, MT, and OR. Thanks to S. Alexander and
J. Dungan for helpful comments on this manuscript. This research was funded by a NASA grant.
References
Aguado, E. 1985. Radiation balances of melting snow covers at
an open site in the Central Sierra Nevada, California. Water
Resources Research 21(11): 10491034.
Anderson, E.A. 1968. Development and testing of snowpack
energy balance equations. Water Resources Research 4(1):
1936.
Anderson, E.A. 1973. SNOW-17 a model based on temperature
and precipitation. NOAA, publication SW-102.
Baker, D.G., D.L. Ruschy and D.B. Wall. 1990. The albedo
decay of prairie snows. American Meteorological Society
29: 179187.
Baker, F.S. 1944. Mountain climates of the Western United
States, Ecological Monographs 14(2): 223254.
Band, L.E., D.L. Peterson, S.W. Running, J. Coughlan, R. Lammers, J. Dungan and R. Nemani. 1991. Forest ecosystem
processes at watershed scale: Basis for distributed model.
Ecological Modeling 56: 171196.
Berris, S.N. and R.D. Harr. 1987. Comparative snow accumulation and melt during rainfall in forested and clear-cut plots
in the Western Cascades of Oregon. Water Resources
Research 23(1): 135142.
Black, T.A. and D.L. Spittelhouse. 1981. Modeling the water
balance for watershed management. Interior West watershed
management. Edited by D.M. Baumgartner. WSU. Pullman,
WA.
Bristol, K.L. and G.S. Campbell. 1984. On the relationship
between incoming solar radiation and daily maximum and
minimum temperature. Agriculture and Forest Meteorology
31: 159166.
Buffo, J., L. Fritschen and J. Murphy. 1972. Direct solar radiation on various slopes from 0 to 60 north latitude. Research
Paper PNW-142, USFS. Portland, OR.
Coughlan, J.C. and J.L. Dungan. 1996. Combining remote
sensing and forest ecosystem modeling: An example using
the Regional Hydrological Simulation System (RHESSys).
In: Gholz, H.L., Nakane, K. and Shinoda, H. (eds). The use
of remote sensing in the modeling of forest productivity at
scales from the stand to the globe. Kluwer Academic Publishers, Dordrecht pp. 135158.
Friend, A.D., H.H. Schugart and S.W. Running. 1993. A physiology-based gap model of forest dynamics. Ecology 74(3):
792798.
Gary, H.L. and G.B. Coltharp. 1967. Snow accumulation and

vegetation disappearance by aspect and vegetation type in


the Sante Fe Basin, New Mexico, USFS Fort Collins, CO.
RM-93.
Gary, H.L. and C.A. Troendle. 1982. Snow accumulation and
melt under various stand densities in Lodgepole Pine in
Wyoming and Colorado. Research Paper RM-17 USFS Fort
Collins, CO.
Giles, D.G., T.A. Black and D.L. Spittelhouse. 1985. Determination of growing season soil water deficits on a forested
slope using water balance and analysis. Canadian Journal
Forest Research 15: 107114.
Gray, D.M. and P.G. Landine. 1987. Albedo model for shallow
prairie snow covers. Canadian Journal of Earth Science 24:
17601768.
Golding, D.L. and R.H. Swanson. 1978. Snow accumulation
and melt in small forest openings in Alberta. Canadian Journal of Forest Research 8: 380388.
Golding, D.L. and R.H. Swanson. 1986. Snow distribution patterns in clearings and adjacent forest. Water Resource
Research 22(13): 19311940.
Hardy, J.P. and J.J. Hansen-Bristow. 1990. Temporal accumulation and ablation patterns of the seasonal snowpack in
forests representing vary stages of growth. In 58th Annual
Meeting Western Snow Conference. pp. 2334. CSU. Fort
Collins, CO.
Hungerford, R.D., R.R. Nemani, S.W. Running and J.C.
Coughlan. 1990. MTCLIM: a mountain microclimate simulation model. Research Paper INT-414. USFS Ogden, UT.
Hunt, E.R., F.C. Martin and S.W. Running. 1991. Simulating
the effects of climatic variability on stem accumulation of a
ponderosa pine stand: comparison with annual growth increment data. Tree Physiology 9: 161171.
Kaufmann, M.R., C.B. Edminster and C.A. Troendle. 1982.
Leaf area determinations for subalpine tree species in the
Central Rocky Mountains. Research Paper RM-238. USFS
Ogden, UT.
Knight, D.H., T.F. Fahey and S.W. Running. Water and Nutrient outflow from contrasting lodgepole forests in Wyoming.
Ecological Monographs 55(1): 2948.
Leaf, C.A. and G.E. Brink. 1973a. Computer simulation of
snowmelt within a Colorado subalpine watershed. Research
Paper RM-99. USFS Ogden, UT.
Leaf, C.A. and G.E. Brink. 1973b. Hydrologic simulation model of Colorado subalpine forest. Research Paper RM-107.
USFS, Ogden, UT.
McMurtrie, R. and J.J. Landsberg. 1992. Using a simulation
model to evaluate the effects of water and nutrients on the
growth and carbon partitioning of Pinus radiata. Forest
Ecology Management 5: 243260.
McLeod, S.D. and S.W. Running. 1988. Comparing site quality
indices and productivity in ponderosa pine stands of western
Montana. Canadian Journal of Forest Research 18: 346352.
Male, D.H. and R.J. Granger. 1981. Snow surface energy
exchange. Water Resources Research 17(3): 609627.
Marks, D., J. Dozier and R.E. Davis. 1992. Water Resources
Research Climate and energy exchange at the snow surface
in the alpine region of the Sierra Nevada 1. Meteorological
measurements and monitoring. Water Resources Research
28(11): 30293042.

136
Marks, D. and J. Dozier. 1992. Climate and energy exchange at
the snow surface in the alpine region of the Sierra Nevada 2.
Snow cover energy balance. Water Resources Research
28(11): 30433054.
Motoyama, H. 1990. Simulation of seasonal snowcover based
on air temperature and precipitation. Journal of Applied
Meteorology 29: 11041110.
Nemani, R.R., L.L. Pierce, L.E. Band and S.W. Running. 1993.
Forest ecosystem processes at watershed scale III: Sensitivity to remotely sensed LAI estimates. International Journal of
Remote Sensing 14(13): 25192534.
Nemani, N., S.W. Running, L.E. Band and D.L. Peterson. 1993.
Regional Hydrological Simulation System: An illustration
of the integration of ecosystem models in a GIS. In Environmental Modeling with GIS. pp. 296304. Edited by M.F.
Goodchild, B.O. Parks and L.T. Steyaert. Oxford University
Press, Oxford.
Paniconi, C. and E.F. Wood. 1993. A detailed model for simulation of catchment scale substrate hydrologic processes.
Water Resources Research 29(6): 16011620.
Price, A.G. 1988. Prediction of snowmelt rates in a deciduous
forest. Journal of Hydrology 101: 145157.
Prevost, M., R. Barry, J. Stein and A.P. Plamondon. 1991.
Snowmelt modeling in a balsam fir forest: comparisons
between an energy balance model and other simplified models. Canadian Journal of Forest Research 21: 110.
Running, S.W. and S.T. Gower. 1991. FOREST-BGC, a general
model of forest ecosystem processes for regional applications II. Dynamic Carbon allocation and nitrogen budgets.
Tree Physiology 9: 147160.
Running, S.W. and R.R. Nemani. 1991. Regional hydrologic
and carbon balance responses for forests resulting from
potential climate change. Climatic Change 19: 349368.
Running, S.W. 1984. Microclimate control of forest productivity: analysis by computer simulation of annual photosynthesis/transpiration balance in differing environments. Agricultural and Forest Meteorology 23: 267288.
Running, S.W. and J.C. Coughlan. 1988. A general model of
forest ecosystem processes for regional applications I.
Hydrologic balance, canopy gas exchange and primary production processes. Ecological Modeling 42: 125154.
Running, S.W., R.R. Nemani and R.D. Hungerford. 1987.
Extrapolating of synoptic meteorological data in mountainous terrain and its use for simulating forest evapotranspiration and photosynthesis. Canadian Journal of Forest
Research 19(6): 472483.
Ryan, M.G., E.R. Hunt, Jr., R.E. McMurtrie, G.I. Agren, J.D.
Aber, A.D. Friend, E.B. Rastetter, W.J. Pulliam, R.J. Raison
and S. Linder. (in press). Comparing models of ecosystem
function for coniferous forests. I. Model description and val-

idation. In Effects of Climate Change on Production and


Decomposition in Coniferous Forests and Grasslands, Scientific Committee on Problems in the Environment
(SCOPE) Volume. Edited by J.M. Melillo, G.A. Agren, and
A. Breymeyer. John Wiley and Sons.
Solomon, R.M., P.F. Ffolliott, M.B. Baker, Jr. and G.J.R.
Thompson. 1976. Computer simulation of snowmelt. RM174, USFS, Ogden, UT.
Troendle, C.A. and C.F. Leaf. 1980. An approach to water
resources evaluation of non-point silvicultural sources: III.
Hydrology. In An Approach to Water Resources Evaluation
of Non-Point Silvicultural Sources (a procedural handbook).
US EPA. Athens, Georgia.
Troendle, C.A. and R.M. King. 1985. The effect of timber harvest on the Fool Creek watershed, 30 years later. Water
Resources Research 21(12): 19151922.
U.S. Army Corps of Engineers. 1956. Snow Hydrology: Summary report of the snow investigations. U.S. Army Corps of
Engineers, Portland, OR.
Soil Conservation Service. 1985. SNOTEL system cooperator
users manual. USDA Soil Conservation Service, Portland,
OR.
Waring, R.H. and J.F. Franklin. 1979. Evergreen forests of the
Pacific Northwest. Science 204: 13801386.
Waring, R.H. and W.H. Schlesinger. 1985. Forest ecosystem:
Concepts and Management. Academic Press, London.

Glossary of symbols and variables used in equations and text

Cext
Cint

LAI

qmelt
qsub
S
Sf
t24
tday
tmax
tnight

albedo (%)
temperature index increment (C)
canopy light extinction coefficient (C)
canopy rain interception coefficient (mm/LAI)
latent heat of fusion (MJ/kg)
latent heat of vaporization (MJ/kg)
leaf area index
albedo temperature sum (C)
isothermal temperature sum (C)
radiation snowmelt (mm/day)
radiation sublimation (mm/day)
incoming shortwave radiation (MJ/M2/Day)
incoming shortwave radiation on the snow surface
(MJ/M2/Day)
24 hour weighted air temperature average (C)
daylight air temperature (C)
maximum 24 hour air temperature (C)
night air temperature (C)

Вам также может понравиться