Sie sind auf Seite 1von 43

Journal of Hydrologic Engineering.

Submitted March 28, 2012; accepted December 21, 2012;


posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

Title:

Impact of SWMM Catchment Discretization: A Case Study in Syracuse, NY

Authors:
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

Ning Sun a, Myrna Hall b, Bongghi Hong c, LianJun Zhang d

t
ip
a
The Graduate Program in Environmental Science, State University of New York,

d cr
College of Environmental Science and Forestry (SUNY-ESF), 1 Forestry Drive,

te s
Syracuse, NY 13210-2778, Email: nsun@esf.edu

di nu
b
Department of Environmental Studies, SUNY-ESF, 1 Forestry Drive, Syracuse, NY

13210-2778, Email: mhhall@esf.edu.


ye a
op M

c
Department of Ecology and Evolutionary Biology, Cornell University, 103 Little

Rice, Ithaca, NY, 14850, Email: bh43@cornell.edu.


C ted

d
Department of Forest and Natural Resources Management, SUNY-ESF, 1 Forestry

Drive, Syracuse, NY 13210-2778, Email: lizhang@esf.edu.


ot p

Corresponding Author:
N ce

Myrna Hall, Email: mhhall@esf.edu.


Ac

Key Words:

SWMM, Discretization, Parameterization, GLUE, Uncertainty, Calibration,

Validation

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

Abstract

This study examined how the level of catchment discretization influenced the

model parameterization and output uncertainty of the Storm Water Management

Model (SWMM) 5.0. We developed two catchment delineations for a highly


Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

urbanized sewershed in Syracuse, NY: 1) the macro-scale model containing a

t
ip
minimum required number of subcatchments to retain the original sewer network

d cr
properties, and 2) the micro-scale model in which each subcatchment was defined for

te s
a unique soil and land use combination. For both scales, we calibrated the model

di nu
parameters and quantified the uncertainty of model outputs using the Generalized

Likelihood Uncertainty Estimation (GLUE) methodology. Then we applied calibrated


ye a
op M

posterior parameter sets at micro and macro scales individually to a second

sewershed, which was also delineated at both micro and macro scales, to test observed
C ted

versus simulated flows. The results indicated that the catchment disaggregation level

had a great impact on both parameterization and simulation results, and that the
ot p
N ce

majority of the parameters were sensitive to the modeling scales. Overall, the

posterior parameters calibrated based on the micro delineation resulted in a higher


Ac

degree of reduction in output uncertainties for both calibrated and validated

sewersheds. Hence we argue that the calibrated parameters obtained, based upon the

macro delineation, would result in reduced confidence in simulated runoff for another

site unique in its characteristics, while the posterior parameters derived from the

micro delineation could provide a higher confidence level in terms of parameter

transferability for modeling other, particularly ungaged sites.

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

1. Introduction

1.1 Combined Sewer Overflows

Urban sprawl over the last three decades has induced an expansion of impervious

areas accompanied by more frequent occurrences of sewer overflow events in cities


Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

where sanitary and storm sewers share a main sewer line. Instead of letting rain water

t
ip
run its pre-developed natural courses, the water flowing off city surfaces is connected

d cr
directly to the combined storm sewer system through pipes and gutters in most urban

te s
settings. Combined sewer overflows (CSOs) occur when stormwater runoff exceeds

di nu
infrastructure capacity. This normally is caused by large rainfall events. The CSO is

released into neighboring surface waters, and the resulting effluent pollutes receiving
ye a
op M

waters and poses a considerable threat to aquatic life, human health and groundwater

quality (Grigg, 1996). A major challenge of urban stormwater management is to


C ted

design effective and efficient stormwater control facilities and hydraulic structures

that can handle peak flows and hence reduce CSO events. In practice, the design
ot p
N ce

parameters of these facilities are generally derived from the actual hydrograph

developed based on the measurements of historical storm events. Flow records,


Ac

however, are often not available for estimating these design parameters because of

sparsely installed flow meters that are used to measure stormwater runoff from urban

sewersheds. To overcome this problem, hydrologic models are often used to simulate

storm hydrographs and provide discharge estimates at ungaged sites. On the other side,

hydrologic models cannot be calibrated or validated without flow measurements,

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

however, which indicates significantly overlooked uncertainty associated with applied

model parameters and flow predictions.

1.2 Issues with Model Scales

More than a quarter century ago, Huber et al. (1975) presented the possible
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

catchment discretization methods, a “fine” versus “coarse” discretization, for the

t
ip
Northwood section of Baltimore, Maryland. Zaghloul (1981) suggested using a finite

d cr
set of area- or length-weighted parameters to represent the spatially varying physical

te s
properties of continuous spaces. This is also known as the spatially lumped-parameter

di nu
approach to hydrological modeling. Since then, the majority of hydrological models

used in practice today are lumped parameter models (Butts et al., 2004). This method
ye a
op M

is widely used because it requires fewer input data and is more computationally

efficient than a model based on spatial units of a fine resolution. However, the lumped
C ted

hydrological modeling approach does not explicitly account for the heterogeneity of

surface properties (e.g., variation of land covers, soils, slopes and etc.) that influence
ot p
N ce

how water moves through the landscape. It is also very difficult to represent the

complexity of flow paths mathematically at a coarse scale. Because physical


Ac

hydrological processes are characterized by strong nonlinearity, Moore and Gallant

(1991) argued that it may well be impossible to use the lumped value of spatially

variable parameters to represent the mean hydrologic response related to catchment

characteristics. Beven (1989) suggested that future development in physically based

(hydrological) modeling must take into account the lumping effect of subgrid scale

processes. An alternative to the lumped-parameter model is the distributed-parameter

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

model, which ideally considers all spatial variability of the landscape by solving the

governing equation within each spatially-defined landscape unit (Baker, 1989; Chow

et al., 1988). It is important to note that any so called “lumped” model is rarely

entirely lumped, and so may be considered a partially “distributed” model. The model
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

scale, lumped or distributed, is relative to the size of the study area, scale of the input

t
ip
dataset, model complexity, and etc. The constraints of distributed-parameter models

d cr
result from the model complexity in terms of model structure and requirements for

te s
inputs. Consequently, the choice of modeling scale has always been a dilemma. Few

di nu
past studies examined the impact of modeling scale on distributed hydrologic model

performance. Mamillapalli et al. (1996) found that, using the Soil and Water
ye a
op M

Assessment Tool (SWAT), simulated streamflow predictions varied with the number

of subwatersheds used to divide the watershed. Discretization of Hubbard Brook


C ted

Watershed 6, New Hampshire into 208 cells led to an improvement in streamflow and

nutrient flux estimation simulated with Simple Nitrogen Cycle (SINIC) model (Hong
ot p
N ce

et al., 2006). Barco et al. (2008) applied the SWMM model coupled with GIS

procedures to simulate runoff for a large urban catchment in Southern California, and
Ac

employed the ‘complex method’ in multi-objective optimization to estimate four

SWMM runoff parameters. The results demonstrated that integrating GIS and a

stormwater model with a constrained optimization technique can be applied to large

watersheds. The results from Guo and Urbonas (2008), and Dankenbring and Mays

(2009) demonstrated that successive catchment discretization in the Storm Water

Management Model (SWMM) caused increasing predicted peak flows. Ghosh and

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

Hellweger (2012) investigated the effects of spatial resolution on SWMM model

predictions in an urban catchment using the Artificial Network Generator (ANGel).

However, these studies were conducted in a deterministic fashion without taking the

uncertainty issues into consideration. Uncertainties or errors associated with model


Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

structure and model parameterization can propagate to the model outputs. Thus it is

t
ip
desirable and necessary to integrate such uncertainties into the modeling processes in

d cr
order to quantify the overall uncertainty in model outputs. The Generalized

te s
Likelihood Uncertainty Estimation (GLUE) methodology, developed by Beven and

di nu
Binley (1992), is a widely adopted approach to evaluate model uncertainties

particularly for nonlinear hydrologic models. Recent research that demonstrated


ye a
op M

applications of the GLUE procedure in rainfall-runoff models can be found in Beven

and Binley (1992), Lamb et al. (1998), Blazkova et al. (2002), Hong et al. (2005),
C ted

Choi and Beven (2007), Fang and Ball (2007), and Xiong and O’Connor (2008).

The research reported here focused on three questions: 1) how would model
ot p
N ce

outputs and parameter values respond to changing model scale? 2) Is there any scale

dependence of the control parameters? 3) Do the parameters calibrated from finer


Ac

catchment delineation in one catchment lead to better model performance in another

catchment? To answer these questions, we developed two catchment delineations with

different discretization levels, i.e., micro and macro scale. For both catchment

delineations, we ran the SWMM model and quantified the uncertainties of model

predictions using the GLUE methodology. We had the measured hydrographs from

sewershed outlets available to us, which allowed us to calibrate the model parameters

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

and test parameter transferability through model validation. The ultimate goal of this

study was to investigate urban runoff modeling techniques and provide better

estimates of sewer hydrographs, particularly for ungaged sites.

2. Methods
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

2.1 SWMM Parameterization

t
ip
SWMM is a physically based, spatially distributed model for simulating all

d cr
aspects of hydrologic and water quality cycles primarily within urban areas (Huber et

te s
al., 1988; Rossman, 2010). SWMM treats each catchment as a nonlinear reservoir and

di nu
employs the combined continuity equation (1) and Manning’s equation (2) on each

subcatchment (Huber et al., 1988). Here a subcatchment is defined as an area of land


ye a
op M

containing its own fraction of pervious and impervious surfaces whose runoff drains

to an outlet point, which could be either a storm drain or another subcatchment


C ted

(Rossman, 2010).
ot p

(1)
N ce

where V is the volume of water on the subcatchment, dt is the computational time


Ac

step, A denotes the area of the subcatchment, ie is the effective inflow, and Q is the

subcatchment outflow, derived as follows:

(2)

where W denotes the subcatchment width, k is a conversion constant equal to 1.486

for U.S. metric units or 1.0 for SI units, n is the Manning’s roughness coefficient, DS

is the surface depression storage, and S is the subcatchment slope.

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

The input parameters of physical process-based models such as SWMM often

correspond to the interpretable physical properties of the site. As geographic

information systems (GIS) has been increasingly linked with SWMM for data pre-

processing and model parameterization, the majority of the SWMM control


Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

parameters can be extracted from spatially defined GIS layers, e.g. catchment area

t
ip
and slope. The principal measured SWMM parameters are presented in Table 1.

d cr
Ideally, every SWMM parameter would be developed based on actual measurements

te s
of catchment characteristics. In practice, however, some parameters with a high

di nu
degree of spatial variability cannot be directly measured or extracted from GIS layers,

and, hereafter will be referred to as inferred parameters. Previous studies either


ye a
op M

developed empirical relationships from which these parameters could be inferred

based on obtainable measurements, or established the ranges for the parameter values
C ted

through statistical approaches, or lab or field experiments. The inferred parameters

are rarely well-defined due to model scale issues and the spatially and temporally
ot p
N ce

varying meteorological and geographical conditions under which the models were

built. In this study, based upon the actual sewer hydrographs, we aimed to refine the
Ac

following inferred overland flow parameters including Manning’s n and depression

storage (DS) for each surface cover, the coefficient for subcatchment width (KWidth),

and Manning's n for closed conduit (nConduit). Manning’s n, often called the surface

roughness coefficient, determines the overland runoff rate from a subcatchment. A

greater n value results in a lower runoff rate and prolongs the flow duration. DS is the

volume of water that can be held in natural depressions in the land surface (Horton,

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

1935). In other words, DS represents the depth of water that has to be exceeded on the

subcatchment before runoff occurs. A relatively small value of DS indicates less

rainfall retention capacity in a subcatchment, and subsequently leads to higher peak

flow rates and earlier occurrence of runoff. Manning’s n and DS are difficult to
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

measure because of the complexity and variability in surface covers in an urban

t
ip
context. Subcatchment width (W), conceptually, represents the width of the

d cr
downstream side of the idealized sloping rectangular subcatchment (Huber et al.,

te s
1988). A smaller W is equivalent to a longer flow path, leading to the attenuated peak

di nu
flow occurrence. Since real-world subcatchments are mostly irregular in shape with

drainage channels off the center, it can be very difficult to determine the actual
ye a
op M

subcatchment width. We therefore adopted the equation used in the InfoSWMM User

Manual (MWH Soft Inc., 2005):


C ted

(3)

where W is computed as a ratio (KWidth) of the square root of subcatchment area. We


ot p
N ce

chose to use so the estimated flow width was less sensitive to the actual

subcatchment width (particularly important when the subcatchment has a very large
Ac

width), as opposed to the method suggested in the SWMM applications manual

(Gironás et al., 2009): W = A/ L, where L denotes a conceptual constant maximum

overland flow length. The parameter nConduit is an indicator of the smoothness of

interior pipe walls, and determines the volume of water a pipe can convey (i.e., its

hydraulic capacity). The lower nConduit the smoother the wall, hence the greater is the

hydraulic capacity of the pipe. Although nConduit can be measured directly, it seldom

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

is, due to the fact that numerous studies have been done to establish a range of

Manning’s n values for various types of pipes. Nevertheless, the selection of the

nConduit value is a rather arbitrary process.

In this study, the kinematic wave method (Lighthill and Whitham, 1955) was
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

chosen for the channel flow routing. The runoff from subcatchments was computed at

t
ip
a 5-minute time step, while the computational time used for flow routings was set to

d cr
30 seconds. Compared to the diffusive wave method, the kinematic wave method

te s
substitutes the simple stage-discharge relationship for the momentum equation by

di nu
assuming that the bed slope is approximately equal to the friction slope and stage-

discharge (MacArthur and DeVries, 1993). As urban conveyance facilities, such as


ye a
op M

pipes, are usually designed for partially full flow conditions, the kinematic wave

method is considered appropriate for simulating the movement of stormwater over


C ted

small urban basins (Overton and Meadows. 1976). It is also important to bear in mind

that the diffusion wave equation is more suitable than the kinematic wave method for
ot p
N ce

the milder slopes (0.001–0.0001) (Kazezyılmaz-Alhan et al., 2005; Kazezyılmaz-

Alhan and Medina, 2007).


Ac

2.2 Multi-scale Catchment Delineations

We developed two catchment delineations, one at the micro and one at the macro

scale, to test how much the model outputs are influenced by the aggregation level, or

scale, of input data. The two scales differ in the aggregation level of subcatchments

and the flow routing element settings. The routing element of each subcatchment is

comprised of two main components: outlet and internal subarea routing. The outlet

10

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

(either another subcatchment or a storm drain) receives runoff from a subcatchment,

and is represented by a user-specified flow direction. If a subcatchment contains both

impervious and pervious subareas, internal subarea routing specifies the runoff

direction within the subcatchment, i.e., impervious to pervious subarea or pervious to


Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

impervious subarea.

t
ip
2.2.1 The Micro-scale Delineation

d cr
The micro-scale model aims to reduce heterogeneity within each subcatchment of

te s
the sewershed in order to better replicate the heterogeneity across the sewershed and

di nu
thus reduce model uncertainty. Each subcatchment has a unique soil and land cover

combination and is treated as a distinct entity or hydrologic response unit (HRU) as


ye a
op M

defined by Leavesley and Stannard (1990). The hydrologic response is simulated

within each subcatchment. Due to the homogeneity of each subcatchment, the micro
C ted

delineation can represent better the actual hydrologic behavior of different surface

covers. Slope variance within the subcatchment, however, could not be captured due
ot p
N ce

to the scale of the input digital elevation model (DEM), hence we used the average

slope for each subcatchment.


Ac

2.2.2 The Macro-scale Delineation

The subcatchment boundaries for the macro-scale delineation were developed by

overlaying the surface cover map with the sewer network map, and the sewer network

was preserved with the original level of detail. The majority of subcatchments

contain more than one surface cover type. The subarea routing direction within the

subcatchment was manually determined based on landscape slope and connectivity,

11

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

and the percent of runoff routed between subareas was assumed 100% for all

subcatchments. Within each subcatchment, all inferred parameters of interest were

calculated from an area-weighted average of existing surface conditions defined by

the micro-scale model.


Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

2.3 Study Site and Input Data

t
ip
Two urban sewersheds located in the City of Syracuse, NY were used in this

d cr
study, CSO 027 for calibration and CSO 044 for validation (Figure 1). CSO 027,

te s
46.76 hectares in size, is comprised of 68.4% impervious surfaces, 18% tree cover

di nu
and 13.6% lawns as viewed from aerial photography (Table 2). In CSO 044, 52.2% of

the total 17.75 hectares is covered by impervious surfaces, 21.3% is covered by trees,
ye a
op M

and the rest 26.5% is lawns. We defined impervious surfaces as all paved roads and

streets, sidewalks, rooftops, and parking lots, and assumed an identical hydrological
C ted

response from all impervious surfaces. Lawns and trees account for all pervious

surfaces in both sites. No adjustment is made for tree canopy over impervious
ot p
N ce

surfaces.

CSO 027 and CSO 044 were both delineated on micro and macro scales. The
Ac

macro-scale CSO 027 consisted of 70 subcatchments while the micro-scale

delineation contained 292 subcatchments (Figure 2). The macro-scale CSO 044

contained 19 subcatchments and the micro-scale CSO 044 was partitioned into 478

homogeneous subcatchments (Figure 3). The GIS layers used to prepare SWMM

input parameters included the Digital Elevation Model (DEM), the soil map, the

sewershed surface cover map that we digitized from the sewershed orthoimages, and
12

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

the sewer network map (all described in Table 3). We employed the SCS Curve

Number Method (USDA SCS, 1986) for infiltration estimations. The hydrologic soil

group (HSG) map derived from the 30-m soil map (NRCS, 2010) was overlaid with

the surface cover map to determine the curve number for each surface cover
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

according to the table “Runoff Curve Numbers for Urban Area” (USDA SCS, 1986).

t
ip
This method was selected because it is conceptually simple, has less intensive

d cr
parameterization requirements, yet provides reliable estimates of infiltration rates

te s
(Bales and Betson, 1982; Xiong and Melching, 2005). Additionally, the curve number

di nu
method can easily incorporate changing land uses and covers in hydrologic modeling.

The 5-min rainfall data, daily temperature and monthly evaporation data from a
ye a
op M

nearby weather station provided the model meteorological inputs. The selected storm

events used in model calibration and validation are presented in Table 4. The
C ted

monitored 5-minute combined sanitary-storm flows from August through October,

2009 at the two sewersheds, CSO 027 and CSO 044, were provided by Onondaga
ot p
N ce

County consultants, CH2MHill. The data were used for model calibration and

validation. We developed a baseline hydrograph at each sewershed outlet, which


Ac

represents the sanitary flow contribution only. To do this we adjusted the sanitary

hydrograph based on evaluation of the empirical time patterns (monthly, daily,

hourly) of human water use. The adjustment factors were derived from the collected

5-minute sanitary inflow data during dry weather conditions. A full description of the

method can be found in Rossman (2010).

13

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

2.4 Model Calibration and Validation Procedure

We applied GLUE in model calibration and validation to incorporate the sources

of uncertainties for inputs and outputs. GLUE uses a likelihood function to measure

the goodness-of-fit between simulated and observed data and produces credible limits
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

for model input parameters and simulated outputs for a given uncertainty level

t
ip
(Beven, 2006). A full mathematical description can be found in Beven and Binley

d cr
(1992), Lamb et al. (1998), Hong et al. (2005). The steps for the calibration and

te s
validation procedures are briefly described as follows.

di nu
A uniform prior distribution was assumed for each parameter over a sufficiently

wide range of parameter values (Table 5). We ran SWMM for the micro/macro CSO
ye a
op M

027 delineation using 25,000 sample parameter sets drawn from the prior distribution

using a Monte Carlo technique. From this the prior output distributions were
C ted

generated. We then computed a “goodness-of-fit” measure P (i.e., likelihood) for each

of the 25,000 simulations against total flows measured from each of the selected rain
ot p
N ce

events using the Gaussian distribution function. The selection of rainfall events was

based on the observed continuous hydrographs. The lag time between the selected events
Ac

had to be long enough so that the events could be considered independent from one another

(Table 4). In each instance the runoff from a previous event had returned to base flow,

meaning that the contribution of the direct runoff generated from the previous storm events to

the hydrograph was negligible when the following storm event started. The likelihood

function was then updated using Bayes’ theorem (Box and Tiao, 1992) when

observed peak flow rates were added as a new dataset:

14

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

(4)

where D denotes observed data (DQ for observed total flow volume and DP for
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

observed peak flow rate); θ denotes model parameter sets; θQ is simulated total flow

t
ip
volume and θP is simulated peak flow rate; SQ and SP is the standard deviation of

d cr
observed total flows and peak flow rates, respectively; i is numbered rainfall event

te s
and n is the total number of events. Note that the underlying assumption of applying

di nu
the Gaussian distribution function (Equation 4) is that the total flow and peak flow

data are independent of each other. The Gaussian likelihood function was used here in
ye a
op M

order to approximate the true form of posterior probability distributions under the

implicit assumption of error distribution and data independence. Based on the


C ted

likelihood values obtained from Equation (4), we applied the sampling/importance

resampling (SIR) algorithm (Rubin, 1987, 1988) to generate the posterior probability
ot p
N ce

distributions for each inferred parameter of interest and the uncertainty bounds for

runoff predictions. Subsequently, we compared the posterior parameter distribution


Ac

and flow uncertainty bounds for the micro- and macro- CSO 027 to test how

parameter values and model outputs would respond to the changing model scale.

To examine whether or not there is scale dependence of any inferred parameters

of interest, we applied the first-order sensitivity method (Saltelli, 2002) at both model

scales. This method quantifies the interactions between input parameters and

identifies the relative influence of each input parameter on the model output

15

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

uncertainty (Saltelli et al., 2000). The fundamental equation used to compute the

fractional contribution (IX) of each model input parameter θX to the total variance of

model outputs (Y) is:

(5)
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

t
The sum of IX closer to one indicates less interaction among the parameters. A higher

ip
IX value represents higher sensitivity, indicating greater influence on the model output

d cr
from changes in an individual parameter (Saltelli, 2004).

te s
Finally, we conducted two same-scale and two cross-scale model validation runs

di nu
to identify whether finer scale catchment parameterization produces better model
ye a
performance when transported to another catchment. The model validation was based
op M

on the assumption that DS and Manning’s n for impervious surfaces, tree and lawns,
C ted

KWidth and nConduit were the same from one catchment to the other. Two same-scale

validations include: the posterior parameters derived from the micro CSO 027 were
ot p

applied on the micro delineations of the second watershed CSO 044 to simulate the
N ce

flows, and the parameters from the macro CSO 027 applied on the macro CSO 044.
Ac

Conversely, parameters from the macro-scale CSO 027 applied on the micro CSO 044

and parameters derived from the micro-scale CSO 027 applied on the macro CSO 044

comprise the cross-scale model validations. The simulation results were validated

against the measurements taken at the CSO 044 outlet under storm events of varying

magnitude.

16

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

3. Result and Discussion

3.1 Parameter Calibration for CSO 027

We obtained over 1300 posterior parameter sets with unique values for both

delineations using the SIR approach. This number was considered as a sufficient
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

sample of parameter sets to provide a reasonable estimate of the posterior distribution

t
ip
(Bates et al, 2000; Lamb et al. 1998; Xiong and O’Connor, 2008). The posterior

d cr
parameter distributions generated from the macro CSO 027 delineation were

te s
compared with those from the micro delineation in Figure 4. Overall, a greater degree

di nu
of updating from prior to posterior parameter distributions was observed in the micro

delineation. In particular, we observed a greater change of the distribution shapes


ye a
op M

from the uniform prior to the posterior for DSIS, nIS, DSTree and nTree in the micro

delineation compared to the macro one, indicating that observations of flow data
C ted

better constrained the likely values of these parameters at the micro delineation. Since

the same likelihood measure and posterior sampling approach conditioned on same
ot p
N ce

sets of runoff observations were applied consistently to calculate the posterior

parameters for CSO 027 at micro and macro scales, we argue that the difference
Ac

between estimated posterior parameters from the micro and macro scale models is due

primarily to the modeling scale. The posterior distribution for KWidth and nConduit

showed slight differences between the micro and macro delineation, indicating that

the scale change did not have considerable impact on these two parameters. This may

be because the original properties of the pipes were retained on both scales without

spatial aggregation and the subcatchment width was dependent on the corresponding

17

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

subcatchment area that was measured based on the GIS map. In addition, DSLawn and

nLawn at both scales showed little update from the prior distribution.

The first-order sensitivity analysis suggested that: 1) at both micro and macro

scales, DSIS was the most important parameter, followed by nIS in determining the
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

total flow (Figure 5). Relatively speaking, DSIS and nIS were more important in

t
ip
determining the total flow at micro scale than at macro scale; 2) the peak flow showed

d cr
greatest sensitivity to nConduit followed by DSIS at both scales; 3) the total flow showed

te s
very little first-order sensitivity (Ix < 4%) to DSTree and nTree at both scales, while

di nu
DSTree and nTree each accounted for a significant percentage (Ix > 9%) of the total

variation in the peak flow in the micro model. However, the peak flow of the macro
ye a
op M

model showed less response to these two parameters; and 4) the peak flow and total

flow at both scales exhibited little sensitivity to DSLawn or nLawn, suggesting that the
C ted

variability in these lawn parameters had almost no effect on the variability in total and

peak flow predictions. For the micro-scale model, the sum of IX of all parameters was
ot p
N ce

0.92 for the total flow and 0.88 for the peak flow. For the macro-scale model, the sum

of IX was 0.82 for the total flow and 0.89 for the peak flow. The IX value of 0.82 for
Ac

the total flow suggested that approximately 18% of the variability in the total flow

data was left unexplained by considering the first-order effects of the model input

parameters. With techniques described in Ratto et al. (2001), we also examined the

interaction structures among SWMM parameters conditioned on the observed total

flow and peak flow for the micro and macro scale. The results (not shown here) did

not show a significant level of interaction between any parameters. Overall, the

18

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

parameter sensitivity results were consistent with the findings shown in Figure 4,

where the posterior distribution of the parameters with higher first-order sensitivity

values was greatly updated from the prior distributions.

3.2 Flow Uncertainty for CSO 027


Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

Three sets of flow uncertainty bounds, at a significance level α = 90% against

t
ip
observations for 6 representative events of different magnitudes, for CSO 027 were

d cr
estimated from the prior parameters (labeled “Prior”), and the posterior parameters for

te s
the macro-scale (labeled “Macro”), and the micro-scale delineation (labeled “Micro”)

di nu
(Figure 6). The “Micro” and “Macro” flow uncertainty bounds were substantially

refined compared to the “Prior” uncertainty bounds (Figure 6A-E), indicating that the
ye a
op M

combined peak flow and total flow were informative in reducing model output

uncertainty at both scales. Comparing the “Micro” and “Macro” uncertainty bounds,
C ted

the “Micro” bounds were better constrained towards the observations while the wider

“Macro” bounds indicated a greater uncertainty. The discretization level, however,


ot p
N ce

did not show any noticeable influence on the timing of peak flow occurrence. The

results suggested that, given the same observed data, the macro-scale model was less
Ac

informative in updating the “Prior” bounds to the posterior uncertainty bounds than

the macro-scale model. The only exception is shown in Figure 6F, in which both

“Micro” and “Macro” flow bounds fail to envelope the observations of the small rain

event (< 1.94 mm/hour), implying that the SWMM model has limitations in

simulating small rain events.

19

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

3.3 Model Validation for CSO 044

For validation purposes, the posterior distributions of the parameters calibrated

based upon the micro- and macro-scale CSO 027 were first applied to the micro- and

macro-scale CSO 044 to generate the flow rates, respectively (i.e., same-scale
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

validation). Figure 7 shows the flow uncertainty bounds at α = 90%. Overall, the

t
ip
simulation uncertainty bounds estimated from the posterior parameters of the macro

d cr
CSO 027 delineation (labeled “Macro”) were wider than those computed from the

te s
posterior parameters of the micro CSO 027 delineation. The results indicated that the

di nu
employment of the posterior parameters calibrated from the macro-scale model led to

acceptable simulation results with the observations falling within the uncertainty
ye a
op M

bounds. However, the posterior parameters derived from the micro-scale model

reduced the simulation uncertainty to a greater degree on a catchment of the same


C ted

modeling scale. We then applied the posterior parameters derived from the micro- and

macro-scale CSO 027 on the macro- and micro-scale CSO 044 (i.e., cross-scale
ot p
N ce

validation). We found that all uncertainty bounds generated from the posterior

parameters calibrated based upon the macro-scale CSO 027 led to a much greater
Ac

degree of uncertainty represented by wider flow bounds, while the flow uncertainty

bounds estimated from the posterior parameters of the micro CSO 027 narrowed

down the uncertainties of flow predictions at the macro-scale CSO 044 (Figure 8). In

order to determine why the posterior parameters derived from the micro-scale CSO

027 outperformed the ones from the macro-scale CSO 027, we computed the 25% and

75% credible interval and the mode value (as an alternative to the optimal parameter

20

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

value) for each posterior parameter distribution derived from micro and macro CSO

027 (Table 6). It indicated that the credible interval of DSIS and nTree generated from

micro-scale CSO 027 was much smaller than those from macro-scale CSO 027. As

shown in the sensitivity analysis (Figure 5), DSIS showed greatest influence on the
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

simulated peak flow rates. In this case, it may be the wider range of DSIS value

t
ip
derived from the macro CSO 027 that contributed largely to the greater uncertainty in

d cr
flow predictions. Hence we argue that the calibrated parameters obtained based upon

te s
the macro delineation might not be adequate to be applied to simulate another

di nu
sewershed delineated at a different scale, while the posterior parameters derived from

the micro delineation might be able to accommodate the variation as a result of


ye a
op M

different catchment aggregation level to a much better degree and thus provide more

reliable predictions.
C ted

4. Conclusions

In this study, we examined how the level of catchment discretization influenced


ot p
N ce

SWMM modeling uncertainty in the context of urbanized sewersheds. Based on our

findings we concluded that the manner in which continuous geographic spaces were
Ac

disaggregated into discrete spatial units had a significant impact on both the model

parameterization and simulation results. The results showed that the parameter values

were quite sensitive to the modeling scales. Furthermore, both model calibration and

validation results showed that the posterior parameter sets calibrated based upon the

micro delineation reduced the uncertainty of flow predictions to a greater degree

compared to those from the macro delineation. Accordingly, we argue that the

21

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

calibrated parameter sets given the macro-scale delineation may be sufficient to

provide acceptable simulations for this specific site at the matching scale. However, it

might not be considered reliable to be used for simulating another site unique in its

characteristics (e.g., surface cover, spatial configuration, etc.). As the model


Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

framework at micro scale takes care of spatial connectivity between the homogeneous

t
ip
subcatchments via flow routing component, the posterior parameter distributions

d cr
derived from the micro delineation might better accommodate the variation in

te s
catchment characteristics, and provide higher confidence level in terms of parameter

di nu
transferability for modeling other sites. These findings are important for future

development of hydrological models. Hydrological modeling, especially the small-


ye a
op M

scale modeling has always been hindered by availability of runoff measurements, in

which case model parameters cannot be calibrated for better model fit or validated to
C ted

test its acceptance. In this instance, we recommend using the posterior parameter

distributions calibrated based upon a micro-scale model of a gaged site to model other
ot p
N ce

ungaged sites delineated on the same fine scale, although such parameters derived

from small-scale model may lose physical significance at larger scales. Furthermore,
Ac

the micro-scale model can easily incorporate changes in surface covers at a small

scale. This is significant for assessment of the potential runoff control benefits given

small-scale landscape modification or on-site green infrastructure (GI)

implementation, e.g., planting more trees and replacing impervious roofs with green

roofs. Since onsite GI control measures are commonly implemented in a

disaggregated and distributed network, a micro-scale model is needed to capture the

22

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

net utility of decentralized GI facilities at the parcel scale while the lumped models

are inadequate to simulate such decentralized GI measures.

Acknowledgements

We gratefully acknowledge the National Science Foundation Award BSC-


Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

0948952 for an Urban Long Term Research Area Exploratory project (ULTRA-EX)

t
ip
that supported and inspired this research. We also want to thank the Onondaga County

d cr
Water and Environment Program and CH2MHILL, Syracuse for the provision of

te s
monitored flow data.

di nu
ye a
op M
C ted
ot p
N ce
Ac

23

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

References

Baker, W. L. (1989). A review of models of landscape change. Landscape Ecol., 2(2),

111-133.
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

Bales, J., Betson, R. (1982). The curve number as a hydrologic index, 371-386, ed. V.

t
P. Singh. Rainfall-Runoff Relationships. Water Resources Publications,

ip
Highlands Ranch, CO.

d cr
te s
Barco, J., Wong, K.M., Stenstrom, M. K. (2008). Automatic calibration of the

di nu
U.S.EPA SWMM model for a large urban catchment. J. Hydr. Engrg., 134(4),

466-474.
ye a
op M

Bates, S. C., Raftery, A. E., Cullen, A. C. (2000). Bayesian uncertainty assessment in

deterministic models for environmental risk assessment. NRCSE Tech. Rep.


C ted

Ser. 58, 17 pp., U.S. Environ. Prot. Agency, Washington, D. C.

Beven, K. (1989). Changing ideas in hydrology--the case of physically-based models.


ot p

Journal of Hydrology, 105(1-2), 157-172.


N ce

Beven, K. (2006). A manifesto for the equifinality thesis. Journal of Hydrology,


Ac

320(1-2), 18-36.

Beven, K. and Binley, A. (1992). The future of distributed models: model calibration

and uncertainty prediction. Hydrol. Process., 6(3), 279-298.

Blazkova, S., Beven, K. J., Kulasova, A., 2002. On constraining TOPMODEL

hydrograph simulations using partial saturated area information. Hydrol.

Process., 16(2), 441-458.

24

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

Box, G.E.P. and Tiao, G.C. (1992). Bayesian inference in statistical analysis. Wiley

Online Library.

Butts, M.B., Payne, J.T., Kristensen, M., Madsen, H. (2004). An evaluation of the

impact of model structure on hydrological modelling uncertainty for


Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

streamflow simulation. J. Hydrol., 298(1-4), 242-266.

t
ip
Choi, H. T., Beven, K. (2007). Multi-period and multi-criteria model conditioning to

d cr
reduce prediction uncertainty in an application of TOPMODEL within the

te s
GLUE framework. J. Hydrol, 332(3-4), 316-336.

di nu
Chow, V. T., Maidment, D. R., Mays, L. W. (1988). Applied hydrology. McGraw-

Hill Book Company.


ye a
op M

Dankenbring, S. and Mays, D. (2009). Catchment discretization in the Colorado urban

hydrograph procedure: a case study in the East Toll Gate creek watershed.
C ted

Arapahoe County, Colorado.

Fang, T., Ball, J. E., 2007. Evaluation of spatially variable control parameters in a
ot p
N ce

complex catchment modeling system: a genetic algorithm application. J.

Hydroinform, 9(3), 163-173.


Ac

Ghosh, I., Hellweger, F. L. (2011). Effects of Spatial Resolution in Urban Hydrologic

Simulations. J. Hydrol. Eng., 17(1), 129-137.

Gironás, J., Roesner, L. A., Davis, J., Rossman, L. A. (2009). Storm Water

Management Model Applications Manual, National Risk Management

Research Laboratory, Office of Research and Development, US

Environmental Protection Agency, Cincinnati, OH.

25

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

Grigg, N. S. (1996). Water resources management: principles, regulations, and cases.

McGraw-Hill Professional, 311-314.

Guo, J.C.Y. and Urbonas, B. (2008). Consistency between CUHP and rational

methods. Urban Drainage and Flood Control District, Denver, Colorado.


Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

Hong, B., Strawderman, R.L., Swaney, D.P., Weinstein, D.A. (2005). Bayesian

t
ip
estimation of input parameters of a nitrogen cycle model applied to a

d cr
forested reference watershed, Hubbard Brook watershed six. Water Resour.

te s
Res., 41(3), W03007.

di nu
Hong B, Swaney, D.P., Weinstein, D.A. (2006). Simulating spatial nitrogen dynamics

in a forested reference watershed. Hubbard Brook Watershed 6, New


ye a
op M

Hampshire, USA. Landscape Ecology 21 (2): 195-211.

Horton, R E. (1935). Surface runoff phenomena. Edwards Brothers, Inc..


C ted

Huber, W. C., Heaney, J. P., Medina, M. A., Peltz, W. A., Sheikh, H., Smith, G. F.

(1975). Storm Water Management Model User's Manual, Version II. EPA-
ot p
N ce

670/2-75-017, 45-48.Environmental Protection Agency, United States,

Huber, W.C., Dickinson, R.E., Barnwell Jr, T.O. (1988). Storm water management
Ac

model; version 4. Environmental Protection Agency, United States.

Huber, W. C., Dickinson, R. E. (1992). Storm water management model user’s

manual, version 4. Environmental Protection Agency, Georgia.

Kazezyılmaz-Alhan, C. M., Medina, M. A., Rao, P. (2005). On Numerical Modeling

of Overland Flow. Applied Mathematics and Computation, 166(3), 724-740.

26

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

Kazezyılmaz-Alhan, C. M., Medina M. A. (2007). Kinematic and Diffusion Waves:

Analytical and Numerical Solutions to Overland and Channel Flow. J.

Hydraul. Eng., 133(2), 217-228.

Lamb, R., Beven, K., Myrabų, S. (1998). Use of spatially distributed water table
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

observations to constrain uncertainty in a rainfall-runoff model. Adv. Water

t
ip
Resour., 22(4), 305-317.

d cr
Leavesley, G.H. and Stannard, L.G. (1990). Application of remotely sensed data in a

te s
distributed-parameter watershed model. Proc., Workshop on Applications of

di nu
Remote Sensing in Hydrol, 47–68.

Lighthill, M., and Whitham, G. (1955). On Kinematic Waves. I. Flood Movement in


ye a
op M

Long Rivers. Proc. Roy. Soc. A, 229(1178), 281-316.

MacArthur, R., and DeVries, J. J. (1993). Introduction and Application of Kinematic


C ted

Wave Routing Techniques using HEC-1. US Army Corps of Engineers,

Institute for Water Resources, Hydrologic Engineering Center.


ot p
N ce

Mamillapalli, S., Srinivasan, R., Arnold, J.G., Engel, B.A. (1996). Effect of spatial

variability on basin scale modeling. Proc., Proceedings, Third International


Ac

Conference/Workshop on Integrating GIS and Environmental Modeling,

Santa Fe, New Mexico.

Moore, I.D., and Gallant, J.C. (1991). Overview of hydrologic and water quality

modeling. Modeling the Fate of Chemicals in the Environment, 1-8.

MWH Soft Inc. (2005), InfoSWMM User Manual. MWH Soft Inc., Pasadena,

California.

27

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

Natural Resources Conservation Service (NRCS) (2010). Soil Survey Geographic

(SSURGO) database for Onondaga County, New York. U.S. Department of

Agriculture, NRCS, Fort Worth, Texas.

Overton, D. E., and Meadows, M. E. (1976). Stormwater Modeling, 358P. Academic


Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

Press, New York, N.Y.

t
ip
Ratto, M., Tarantola, S., Saltelli, A. (2001). Sensitivity analysis in model calibration:

d cr
GSA-GLUE approach. Comput. Phys. Commun., 136(3), 212-224.

te s
Rossman, L.A. (2010). Storm water management model user's manual, version 5.0.

di nu
National Risk Management Research Laboratory, Office of Research and

Development, US Environmental Protection Agency.


ye a
op M

Rubin, D.B. (1987). The calculation of posterior distributions by data augmentation:

comment: a noniterative sampling/importance resampling alternative to the


C ted

data augmentation algorithm for creating a few imputations when fractions

of missing information are modest: the SIR algorithm. Journal of the


ot p
N ce

American Statistical Association, 82(398), 543-546.

Rubin, D.B. (1988). Using the SIR algorithm to simulate posterior distributions.
Ac

Bayesian Statistics, 3, 395-402.

Saltelli, A. (2002). Making best use of model evaluations to compute sensitivity

indices. Comput. Phys. Commun., 145(2), 280-297.

Saltelli, A. (2004). Sensitivity analysis in practice: a guide to assessing scientific

models. John Wiley & Sons Inc.

28

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

Saltelli, A., Tarantola, S., Campolongo, F. (2000). Sensitivity analysis as an

ingredient of modeling. Statistical Science, 377-395.

Tsihrintzis, V. A., and Hamid, R. (1998). Runoff quality prediction from small urban

catchment using SWMM. Hydrol. Process., 12(2), 311-329.


Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

US Department of Agricluture Soil Conservation Service (USDA SCS) (1986). Urban

t
ip
hydrology for small watersheds. Technical release 55, 2nd ed., NTIS PB87-

d cr
101580. USDA SCS, Springfield, Virginia.

te s
Xiong, L., O’Connor, K. M. (2008). An empirical method to improve the prediction

di nu
limits of the GLUE methodology in rainfall–runoff modeling. Journal of

Hydrology, 349(1-2), 115-124.


ye a
op M

Xiong, Y., and Melching, C. S. (2005). "Comparison of Kinematic-Wave and

Nonlinear Reservoir Routing of Urban Watershed Runoff." J. Hydrol. Eng.,


C ted

10(1), 39-49.

Zaghloul, N.A. (1981). SWMM model and level of discretization. Journal of the
ot p
N ce

Hydraulics Division, 107(11), 1535-1545.


Ac

29

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Figure 1

Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

Accepted Manuscript
Not Copyedited

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Figure 2

Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

Accepted Manuscript
Not Copyedited

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Figure 3

Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

Accepted Manuscript
Not Copyedited
Copyright 2012 by the American Society of Civil Engineers
J. Hydrol. Eng.
Figure 4

Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

Accepted Manuscript
Not Copyedited
Copyright 2012 by the American Society of Civil Engineers
J. Hydrol. Eng.
Figure 5

Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

Accepted Manuscript
Not Copyedited

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Figure 6

Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

Accepted Manuscript
Not Copyedited

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Figure 7

Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

Accepted Manuscript
Not Copyedited

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Figure 8

Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

Accepted Manuscript
Not Copyedited

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

Table 1. The principal measured SWMM parameters

Object Measured Parameters Sources


Area, precentage of imperviousness, GIS layers including the
Subcatch-
slope, curve number of different land use sewer network map, soil map,
ment
type DEM, and orthoimagery
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

Length, shape, maximum depth of the


Conduit
conduit's cross section Syracuse city sewer system
evaluation survey
Junction Invert elevation

Accepted Manuscript
Not Copyedited

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

Table 2. Summary of surface characteristics in the studied sewersheds

Percentage of Land Cover


Total Area
Site Slope
(hectare) Impervious
Tree Lawn
Surface

CSO27 46.76 1.36% 68.4% 18.0% 13.6%


Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

CSO44 17.75 0.97% 52.2% 21.3% 26.5%

Accepted Manuscript
Not Copyedited

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

Table 3. Input datasets used for SWMM parameterization

Data Description Source

The U.S. Geological Survey


Digital Elevation The DEM published in 2008 is at a 10-m resolution,
(USGS)
Model (DEM) and was used to derive the 10-m slope raster file.
(http://www.usgs.gov/)
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

The Natural Resources


The data set consists of the georeferenced 30-m soil
Conservation Service (NRCS)
map developed based on the surveys conducted
Soil Survey Geographic
NRCS Soil Map during 2004 – 2010. We used it to generate the
(SSURGO) Database
hydrologic soil group (HSG) map, which were then
(http://soils.usda.gov/survey/ge
used to compute the curve numbers.
ography/ssurgo/)
The orthoimagery taken in 2009 is a 1-m resolution
NYS GIS Clearinghouse
remotely-sensed digital picture in a raster data
Orthoimagery (http://www.nysgis.state.ny.us/
format, bases on which the sewershed land cover
gateway/mg/)
classification was created.

The Sewer The map includes georeferenced sewerlines, Onondaga County consultants,
Network Map junctions and sewershed outlets in a vector format. CH2M Hill, Syracuse

A SUNY ESF meteorological


station (LAT. 39°2'50'' LONG.
The data set includes five minute interval rainfall
76°9'23'') installed at and
Meteorological data, daily maximum and minimum air
maintained by the Museum of
Data temperatures, and monthly evaporation rate for
Science and Technology
August – October, 2009.
(MoST*) at Armory Square,
Syracuse
The data set contains the monitored 15-min interval
combined sanitary-storm flows during August 13th –
Monitored Flow Onondaga County consultants,
September 10th, and 5-min interval combined
Data CH2M Hill, Syracuse
sanitary-storm flows spanning the period from
September 10th – October 29th, 2009.

Accepted Manuscript
Not Copyedited

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

Table 4. Characteristics of rainfall events

Lag Time Date Total Ave.


Event Duration
Magnitude from Last (Year Depth intensity Purpose*
no. (hours)
Rain (hours) 2009) (mm) (mm/hr)
1 9.6 10/5 1.08 1 0.92 C
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

Light 2 87.4 9/22 0.67 1.3 1.94 C&V


3 8.4 10/1 0.67 1.3 1.94 C
4 26.6 11/6 0.42 1 2.38 C
5 40.1 11/5 0.75 2 2.67 C
6 57.5 8/21 0.58 1.78 2.71 V
7 73.1 9/26 2.75 8.89 3.23 V
Meduim 8 17.6 9/29 1.17 4.3 3.68 C
9 11.9 10/6 1.67 9.14 5.47 V
10 9.9 10/24 1.42 9.4 6.62 C&V
11 5.0 10/7 0.83 6.4 7.71 C&V
12 81.5 10/28 3.33 27.69 8.32 C&V
13 200.9 8/18 0.25 5.1 20.40 C
Heavy 14 15.3 8/29 1.08 24.6 22.78 C&V
15 35.0 9/23 0.25 8.1 32.40 C
* C = Calibration, V= Validation

Accepted Manuscript
Not Copyedited

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

Table 5. The prior distribution of studied SWMM inferred parameters

Inferred Parameters Description Prior Distributoin Typical values from literature

Manning's n for overland


0.01~0.015 (Huber et al.,
nIS flow over impervious Uniform (0, 0.2)
1988)
surfaces
Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

Manning's
n 0.02~0.8 for pervious surfaces
nTree Manning's n for trees Uniform (0, 1)
(Huber et al., 1988; Huber
and Dickinson, 1992)
nLawn Manning's n for lawn Uniform (0, 0.5)
Depth of depression 0.012~0.1 (Huber et al., 1988;
DSIS storage of impervious Uniform (0, 0.5) Huber and Dickinson, 1992;
surfaces Tsihrintzis and Hamid, 1998)
Depression
Storage Depth of depression 0.1~0.2 for pervious surfaces
DSTree Uniform (0, 0.8)
(inch) storage of trees (Huber et al., 1988; Huber
Depth of depression and Dickinson, 1992;
DSLawn Uniform (0, 0.5) Tsihrintzis and Hamid, 1998)
storage of lawns
Fraction of the square
0.2~0.5 (MWH Soft Inc.,
KWidth root of subcatchment Uniform (0, 1)
2005)
width
0.011~0.017 (Huber et al.,
Manning's roughness
nConduit Uniform (0, 0.1) 1988; Huber and Dickinson,
coefficient for conduits
1992)

Accepted Manuscript
Not Copyedited

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.
Journal of Hydrologic Engineering. Submitted March 28, 2012; accepted December 21, 2012;
posted ahead of print December 26, 2012. doi:10.1061/(ASCE)HE.1943-5584.0000777

Table 6. The 25% and 75% credible interval and the mode value of the posterior parameter distribution

Micro CSO 027 Macro CSO 027


Calibrated
Parameter
Mode 25% CI 75% CI Mode 25% CI 75% CI

DSIS 0.134 0.117 0.156 0.084 0.036 0.293


Downloaded from ascelibrary.org by UNIVERSIDADE DE BRASILIA on 05/31/13. Copyright ASCE. For personal use only; all rights reserved.

nIS 0.048 0.042 0.073 0.094 0.078 0.124


DSTree 0.451 0.403 0.525 0.224 0.202 0.479
nTree 0.491 0.431 0.564 0.370 0.278 0.619
DSLawn 0.251 0.146 0.279 0.188 0.168 0.349
nLawn 0.226 0.188 0.329 0.314 0.181 0.327
KWidth 0.698 0.447 0.708 0.452 0.392 0.744
nConduit 0.020 0.016 0.032 0.019 0.025 0.040

Accepted Manuscript
Not Copyedited

Copyright 2012 by the American Society of Civil Engineers


J. Hydrol. Eng.

Das könnte Ihnen auch gefallen