You are on page 1of 55

9th International Conference on Urban Drainage Modelling

Belgrade 2012

Modelling Internal Boundary Conditions of a Sewer


Network
Nuno Melo1 , Jorge Leandro2, James Shucksmith3, Matteo Rubinato3,
Slobodan Djordjevic4, Adrian J. Saul3, Helena Ramos5, Joo L. M. P. de
Lima2
1

UDI Research Unit for Inland Development, Polytechnic Institute of Guarda, Portugal, nuno_melo@ipg.pt
IMAR Institute of Marine Research, University of Coimbra, Portugal, leandro@dec.uc.pt, plima@dec.uc.pt
3
University of Sheffield, United Kingdom, j.shucksmith@sheffield.ac.uk, m.rubinato@sheffield.ac.uk,
a.j.saul@sheffield.ac.uk
4
University of Exeter, United Kingdom, S.Djordjevic@exeter.ac.uk
5
IST Instituto Superior Tcnico, Technical University of Lisbon, Portugal, hr@civil.ist.utl.pt
2

ABSTRACT
Due to the increased frequency of rainfall events caused by climate change,
flooding in urban areas are becoming increasingly frequent. Thus the accurate
modelling of drainage systems is a fundamental tool to enable operators to
minimize flooding. In this paper we compare the experimental data obtained from a
facility in the University of Sheffield with the numerical results obtained with two
one-dimensional numerical models (1D), SIPSON and SWMM. The experimental
facility is a scaled model of an urban drainage system located in the north of
England. The inputs of the scaled model were taken from two rainfall events that
occurred on the 12th December 2008 and on the 17th January 2009 (data measured
by a rain gauge installed in the basin). It was found that SIPSON internal boundary
conditions at the manhole level represented well the head losses of the flow inside
the manhole, thus the model was able to reproduce fairly well the water depths
along the drainage system.

KEYWORDS
Numerical modelling, urban flooding, urban drainage, internal boundary conditions

INTRODUCTION

The risk of flooding in urban areas is directly related to the capacity of the sewer system to convey or
hold the excess runoff generated by a particular rainfall event. Accurate modelling of these systems is
a fundamental tool for real time management, enabling operators to take advantage of the systems full
capacity and minimize flooding.

Some national regulations for preliminary design of stormwater drainage systems is based on the
concept of uniform flow with a filling ratio of approximately 85%, for which the discharge is equal to
the full pipe discharge. This concept was tested in the past and is still the basis of the present
guidelines, although this is only strictly true for subcritical flow (Gargano and Hager, 2002). The free
surface of supercritical flow is dominated by shock waves due to flow perturbations. For example, in a
manhole connected to both up-and downstream sewers, there are changes of cross-sectional shape
from the circular to the U-shaped profiles of equal diameter, which generates shock waves within the
flow (Gargano and Hager, 2002).
Manholes allow the aeration of the drainage system and its inspection and maintenance. Regulations
request that they should be placed whenever sudden changes along the system occurs, e.g. in terms of
diameter, bottom slope, discharge addition or reduction, or changes in boundary roughness. Typically,
the spacing of manholes in meters should be equal to the sewer diameter in centimetres, and not
exceed 100m (Del Giudice et al., 2000; Hager, 1999).
Free surface flow, as occurs in partially filled sewers, depend essentially on the Froude number. For
subcritical flow, disturbances propagates also in the upstream direction, such that these flows must be
computed against the flow direction. In contrast, the computational and the flow directions are
identical for supercritical flows, for which the average-flow velocity V is larger than the wave celerity
c (Hager and Gisonni, 2005). Hager (1999) proposed a simplification to calculate the Froude number F
for circular sewers in terms of discharge Q, gravitational acceleration g, sewer diameter D and flow
depth h for sewer filling y=h/D between 20% and 95% (Equation 1).

F Q / gDh4

12

(1)

This is similar to the expression for the rectangular channel, yet with a larger effect of flow depth, or
relative sewer filling y. For complete sewer filling, i.e. the transition from free surface to pressurized
pipe flow (y=1), (1) degenerates to the so-called pipe (subscript D) Froude number FD Q /( gD 5 )1 2
(Hager and Gisonni, 2005).
Loss coefficients at the manholes can be split into inlet and outlet coefficients being the global
coefficient of the manhole obtained by Out In (Merlein, 2000). Losses at sewer junctions
depend on flow rate, junction geometry, and the change in pipe diameter between the inflow and
outflow lines (Wang et al., 1998).
According to Leandro et al. (2009) the head losses investigation at sewer junctions have been made
through experimental studies for a better understand of the hydraulic conditions for both subcritical
and supercritical flows. However most of these studies remain purely experimental and often they do
not translate back into commercial urban flood models.
The aim of this work is to compare the observed flow in a scale model of an urban drainage system,
with the results obtained using two one-dimensional (1D) numerical models, SIPSON and SWMM, in
order to validate the internal boundary conditions. The calibration of the models is done using the
experimental data of the two storm events occurred on 12th December 2008 and 17th January 2009
(data measured by a rain gauge installed in the basin), thus validating the internal boundary conditions.
The facility is a scale model of an urban drainage system located in the North of England, UK, being
this composed of three inlet pipes fed from a header tank, and six manholes (diameter of 240 mm),
connected to each other by five circular pipes of two sizes (diameters of 75 and 100 mm).

2
2.1

METHODOLOGY
Physical model

The physical model used in this study represents a section of an urban drainage system located in the
North of England, UK, which can record water depths and flow in real time. The facility was
constructed in the laboratorial facilities of the Department of Civil and Structural Engineering of the
University of Sheffield. The installation is composed by three inlet conduits (A, B and C) 75mm,
six manholes and two CSO (Combined Sewer Overflow), as shown in Figure 1. The six manholes and
the two CSO are connected to each other at the bottom level by five pipes 75mm, and the last
manhole (M6) is connected to the Frontal and Lateral CSO by two pipes of 100mm. The first leaves
the M6 at the bottom level and the second 90mm above the bottom level. The bottom level of the
manholes is equal for all. Every manhole has an internal diameter of 240 mm. This facility is supplied
by a header tank of constant level, which in turn is supplied by the water that is recirculated from de
CSO tanks.

Figure 1. Scheme of the facility (M stands for manhole).


In each inlet of the facility the flow is controlled by calibrated butterfly valves, remotely controlled
from a computer. The flow in each pipe is varied independently in real time, thereby enabling the
simulation of a variety of precipitation events. Through the system, non-intrusive pressure transducers
are installed, which enable the acquisition of pressure data in real time, and six transducers are located
within the manholes, which allow determining the flow depth at these sites by depth versus pressure
relationships. The installation has also three flow meters, one at each inlet pipe system.
2.2

Hydraulic Numerical Models

To model the system two one-dimensional models, SIPSON and SWMM were used.
SIPSON is a 1D/1D integrated hydraulic model developed by Djordjevic (2001) at the University of
Belgrade. The acronym SIPSON stands for Simulation of Interaction between Pipe flow and
Overland flow in Networks. SIPSON, besides being a hydraulic model, also incorporates a
3

hydrologic model (rainfall runoff), called BEMUS, which is used for calculating the surface runoff
input to the hydraulic model. A GIS interface, named 3DNet, works as the platform for management
and editing of data and visualization of the SIPSON results. The hydraulic model is based on the
Preissman finite difference method and the conjugate gradient method, solving simultaneously the
continuity equations for network nodes, the complete St. Venant equations for the 1D networks and
the links equations (Djordjevic et al., 2005).
The modelling of head losses in manholes (in SIPSON) is done by choosing for each manhole, one of
five options (high head loss, normal head loss, special type 1, special type 2 and special type 3)
ordered by degree of head loss.
Storm Water Management Model (SWMM) is a dynamic rainfall-runoff model. The component of
runoff operates on a collection of sub-catchment areas that receive precipitation and generate runoff.
The routing of the SWMM runoff is done through the system of channels, pipes and devices. The flow
routing in this case is calculated, using the complete one-dimensional Saint Venant flow equations
(Dynamic Wave Routing) (Rossman, 2010). This routing method can account for channel storage,
backwater, entrance/exit losses, flow reversal, and pressurized flow (Rossman, 2010).
The modelling of the head losses in manholes (in SWMM) is done introducing in the pipes, local loss
coefficients at entry and exit of this.
2.3

Data Analysis

Modelling using SIPSON and SWMM was based on the geometric characteristics of the installation,
pipe materials and inflow hydrographs at each inlet pipes of the system (branches A, B and C).
Calibration was based on the results obtained experimentally and the reported storm events (12th
December 2008 and 17th January 2009) (Figure 2), acting on the roughness of the pipes and head
losses in the manholes to minimize difference in the manholes depths. The final Manning roughness
coefficient of the pipes was 0.01 m-1/3s.

a)

b)

Figure 2. Inflow hydrographs at each inlet pipes of the system, a) event of 12th December 2008 and b)
event of 17th January 2009.

RESULTS AND DISCUSION

Figures 3 to 8 show the variations of the water depth in the manholes over the simulated events. In
each graph the results obtained experimentally in the scale model and the results obtained by the
modelling made on SIPSON and SWMM are compared. Two cases are analysed, first considering
only the continuous head losses and a second case considering the continuous head losses and the local
head losses at the manholes.
It was found that the flow Froude number is always less than unity irrespectively of the event (and
regardless of local head losses at the manholes), indicating that we are in the presence of subcritical
flow. According to Zhao et al. (2006) subcritical flow in sewer junctions has relatively small energy
losses and may be described as an open-channel ow junction. In SIPSON the head losses in the
manholes was Special Type 3 (least energy loss). In SWMM for the inlet pipe it was set equivalent to
the passage in sharp edge of a pipe to a reservoir (K =1), for the outlet pipe it was set equivalent to the
sharp edge passage from a reservoir to a conduit (K= 0.5) (Quintela, 1981).
Regarding the water depths found in manholes 1, 2 and 3 we verified that for the event of 12th
December 2008, the software that best reproduced the experimental data was SIPSON by neglecting
the head losses in the manholes, including the reproduction of the maximum peak recorded in the
manholes. It was verified that SWMM had a higher peak damping. Nonetheless and in terms of flow
regime near to steady-state the results obtained by both models were approximately equal. For the
event of the 17th of January 2009, the conclusions are approximately the same. Nonetheless in this case
both SIPSON and SWMM overshoot the time of peak obtained when compared with the experimental
data.
Manholes 4 and 5 are typical of a situation of confluence of two pipes in a single manhole, one aligned
with the outlet conduit (main flow direction) and the other with an angle of 45 degrees with the main
direction of flow. For the two simulated events, it was verified that for the situation where the head
losses in manholes are no considered:

In manhole 4 there was a significant gap between the simulated and experimental depths. This
could be justified by the fact that the flow of the pipe that enter the manhole with an angle of
45 to be greater than 1/3 of the flow that circulate in the main flow direction; it may cause
higher turbulence inside the manhole and a consequent increase in the water level, a situation
that is not reproduced by the models.

In manhole 5 the flow depths resulting from SIPSON fit relatively well to the experimental
data. Contrarily to the previous case, here, the incoming lateral flow is smaller than 25% of
the flow that circulates in the main flow direction.

a)

b)
th

Figure 3. Variation of water depth at manhole 1, a) event of 12 December 2008 and b) event of 17th
January 2009.
5

a)

b)
th

Figure 4. Variation of water depth at manhole 2, a) event of 12 December 2008 and b) event of 17th
January 2009.

a)

b)
th

Figure 5. Variation of water depth at manhole 3, a) event of 12 December 2008 and b) event of 17th
January 2009.

a)

b)
th

Figure 6. Variation of water depth at manhole 4, a) event of 12 December 2008 and b) event of 17th
January 2009.

a)

b)
th

Figure 7. Variation of water depth at manhole 5, a) event of 12 December 2008 and b) event of 17th
January 2009.

a)

b)
th

Figure 8. Variation of water depth at manhole 6, a) event of 12 December 2008 and b) event of 17th
January 2009.
In the case of manhole 6 (Figure 8), for both simulated events, the water depths obtained by the
numerical models are higher than the ones obtained by the experimental data. A possible justification
is the position of the pressure sensor that is next to the downstream pipe. This is the only manhole that
the downstream pipe has a larger diameter than 75mm, i.e. 100mm.
In Figures 9 to 12 the flow rate variations obtained using the models SIPSON and SWMM for the
different simulation conditions are presented (with and without head losses in manholes).
Comparing the flow rates obtained with both rainfall events, it was verified that the values resulting
from the modelling on the SIPSON are larger than those obtained by SWMM.
Comparing the flow rates obtained for each model for the situation without head losses in manholes
and with head losses in manholes, SIPSON results did not show significant differences, since the loss
of energy introduced in manholes was very small, thus only a small influence on the flow variation
occurred. On the other hand SWMM results of the level of the flow in the facility are increased, which
caused a diminished of the flow peaks due to the effect of storage observed in the system.

a)

b)

Figure 9. Variation of flow rate at manholes, for the situation without considering head losses in
manholes for the event of 12th December 2008, a) SIPSON results and b) SWMM results.

a)

b)

Figure 10. Variation of flow rate at manholes, for the situation considering head losses in manholes for
the event of 12th December 2008, a) SIPSON results and b) SWMM results.

a)

b)

Figure 11. Variation of flow rate at manholes, for the situation without considering head losses in
manholes for the event of 17th January 2009, a) SIPSON results and b) SWMM results.

a)

b)

Figure 12. Variation of flow rate at manholes, for the situation considering head losses in manholes for
the event of 17th January 2009, a) SIPSON results and b) SWMM results.
In SWMM is not possible to define the size of the junctions (diameter), which does not allow that the
effects of storage on these components to be taken, unless junctions are changed to Reservoirs; and
even so only if the reservoir diameter is larger than 1.2m, which is much larger than the 0.24m of the
manholes of the facility (Figure 13).

Figure 13. Comparison of the variation of water depth at manhole 6 for the event of 12th December
2008, for two different simulations (with junctions or reservoirs to modelling the manholes in
SWMM).
It was found that SIPSON calculated the flow depth inside manholes taking into consideration the
depth increase due to transfer of kinetic energy to potential energy. This was verified by looking at the
water depth in the pipes upstream and downstream of the manholes and verifying that the depth of the
flow inside the manhole was greater than the depth of the flow in the pipe upstream or downstream.
A similar analysis was done to the SWMM results and it was found that the depth of the flow within
the manhole was an average between the depths of the flow in conduits that are upstream and
downstream of the manhole.

CONCLUSION

In this paper we compared the experimental data obtained from a facility in the University of Sheffield
with the numerical results obtained with two one-dimensional numerical models (1D), SIPSON and
SWMM.
When modelling the experimental facility it was found that SIPSON reproduced fairly well the water
depths in the manholes of the drainage system. In SWMM the results are less representative, in part,
due to the limitations to define the size of the junctions (diameter).

ACKNOWLEDGEMENT

This research was supported by projects PTDC/ECM/105446/2008 and PTDC/AACAMB/101197/2008, funded by the Portuguese Foundation for Science and Technology (FCT) and by
the Operational Programme Thematic Factors of Competitiveness (COMPETE), shared by the
European Regional Development Fund (ERDF).
The first author is also grateful to UDI Research Unit for Inland Development, Polytechnic Institute
of Guarda, Guarda (Portugal), by the support to this work.

REFERENCES

Djordjevic S. (2001). A mathematical model of the interaction between surface and buried pipe ow in
urban runoff and drainage. PhD thesis, University of Belgrade, Belgrade, Serbia.
Djordjevic S., Prodanovic D., Maksimovic C., Ivetic M. and Savic D. (2005). SIPSON - simulation of
interaction between pipe flow and surface overland flow in networks. Water Science and
Technology, 52(5), 275-283.
Gargano R. and Hager W. H. (2002). Supercritical flow across sewer manholes. Journal of Hydraulic
Engineering, 128(11), 1014-1017.
Del Giudice G., Gisonni C. and Hager W. H. (2000). Supercritical flow in bend manhole. Journal of
Irrigation and Drainage Engineering, 126, 48-56.
Hager W. H. (1999). Wastewater Hydraulics: Theory and Practice. New York: Springer, Berlin.
Hager W. H. and Gisonni C. (2005). Supercritical flow in sewer manholes. Journal of Hydraulic
Research, 43(6), 660-667.
Leandro J., Abreu J. M. and de Lima J. L. M. P. (2009). Laboratory set-up to validate a dual drainage
concept numerical model. In 8th International Conference on Urban Drainage Modelling, Tokyo,
Japan.
Merlein J. (2000). Flow in submerged sewers with manholes. Urban Water Journal, 2(3), 251-255.
Quintela A. C. (1981). Hidrulica (Hydraulics). 1st ed, Fundao Calouste Gulbenkian, Lisboa.
Rossman, L. A. (2010). Storm Water Management Model - User's manual (Version 5.0). Cincinnati,
Environmental Protection Agency.
Wang K. H., Cleveland T. G., Towsley C. and Umrigar D. (1998). Head loss at manholes in
surcharged sewer systems. Journal of the American Water Resources Association, 34(6), 13911400.
Zhao C., Zhu D. and Rajaratnam N. (2006). Experimental study of surcharged flow at combining
sewer junctions. Journal of Hydraulic Engineering 132(12), 1259-1271.
10

9th International Conference on Urban Drainage Modelling


Belgrade 2012

Automated Pipe-sizing of Storm Sewer or Combined


Sewer Systems Based on Hydrodynamic Modelling
Kegong Diao1, Michael Mair2, Michael Mderl3, Manfred Kleidorfer4, Robert
Sitzenfrei5, Christian Urich6, Wolfgang Rauch7
1

University of Innsbruck, Innsbruck, dkg630630@gmail.com


University of Innsbruck, Innsbruck, Michael.Mair@uibk.ac.at
3
University of Innsbruck, Innsbruck, Manfred.Kleidorfer@uibk.ac.at
4
University of Innsbruck, Innsbruck, Robert.Sitzenfrei@uibk.ac.at
5
University of Innsbruck, Innsbruck, Michael.Moederl@uibk.ac.at
6
University of Innsbruck, Innsbruck, Christian.Urich@uibk.ac.at
7
University of Innsbruck, Innsbruck, Wolfgang.Rauch@uibk.ac.at
2

ABSTRACT
This paper introduces a method for automated pipe-sizing of storm sewer or
combined sewer systems based on hydrodynamic modelling. The methodology
includes three steps. Initially, graph theoretical description of network topology
(e.g. Strahler number) is utilized for classification of the studied sewer networks
topology. Then, the network is decomposed hierarchically into a number of
subsystems based on the network topology. Finally, the pipe sizing is carried out
subsystem by subsystem with no flooding in the whole system as the objective. To
verify the results of the method, the algorithm is tested on a real world sewer
network, and then the solution is compared with the global optimal solution. As
proved by the case study, the author-designed method could guarantee a nearoptimal solution that is very close to the global optimal solution, while requires
dramatically less computational effort than global optimization method. Compared
with evolutionary methods, the method has its own advantages, since it does not
require any parameter for configuration and execution control, and could produce
unique solutions as long as the design principles are fixed.

KEYWORDS
Automated pipe-sizing; combined sewer system; hydrodynamic modelling; storm sewer system;
SWMM

INTRODUCTION

Regarding flood protection-based sewer network design, the task is to minimize the construction costs
whilst ensuring no flooding. The design problem was mostly handled as a pipe sizing and slope design
problem for sewer networks with a fixed plan layout. There are two major kinds of methodologies for
1

solving this problem. The first kind of methodologies is based on estimating a fixed design discharge
for each pipe. The developed models mainly resort to the application of dynamic programming (Mays
and Wenzel, 1976; Walters and Templeman, 1979; Yen et al., 1984; Kulkarni and Khanna, 1985),
linear programming (Deininger, 1966; Dajani and Hasit, 1974; Elimam et al., 1989), and non-linear
programming (Holland, 1966; Price, 1978; Gidley, 1986). But, as commonly known, the system
capacity can accommodate a considerable surcharge before surface flooding occurs. Hence, these
approaches may result in a serious over dimensioning of the system capacity. To deal with this
problem, the second kind of methodologies could be used to achieve good system performance (e.g.,
no flood occurrence) based on assessing system performance as a whole under a predefined design
storms. As the optimal pipe sizing is an NP-hard (non-deterministic polynomial-time hard) problem
(Yates et al. 1984), approximation methods are required to solve the optimization problem. In this
regard, the evolutionary methods have performed well. For instance, Savic and Walters (1997) have
successfully applied the GA method (Goldberg, 1989) for this task. Other techniques, such as the ant
colony optimisation method (Zecchin et al., 2006), the particle swarm optimisation method (Izquierdo
et al., 2008) and the cellular automata (Guo et al., 2007) have also been applied successfully.
Although the evolutionary methods are proved to be efficient and robust in finding near optimal
solutions, they are suffering from several drawbacks. Firstly, this kind of algorithms usually requires
the users to establish several parameters related to configuration and execution of the algorithm.
Nevertheless, no general rule is available for the determination of these parameters and pipe-sizing
cannot be linked to existing design standards which have to be regarded. Hence, a large number of
trial-and-error tests are unavoidable to find appropriate parameter values. Secondly, the methods
(especially, GAs) always entail a high computational cost in order to achieve a sound level of good
solutions. Thirdly, the methods are inherently stochastic even though it finds out the solution with real
minimization of costs for each run, i.e. different solutions would be produced after different
implementations.
Given the limitations of currently available methods, a novel automated pipe-sizing method is
developed in this study. Compared with evolutionary methods, the method introduced in this paper has
the following advantages. First of all, there is no need of parameters for configuration and execution
control. So, no large amounts of pre-runs are necessary. Secondly, the method is deterministic, i.e. the
optimization results are unique as long as the design principles (described e.g. by legal regulations) are
fixed. The method simply imports a model input file (e.g. SWMM file, Rossman, 2010) of the studied
network for optimization. The optimization process could start with the sizes of all pipes being set to
the minimum required value. It could also run based on a pre-designed layout and then further
optimize the pipe-sizing to improve the system performance. Admittedly, the implementation of this
method could be computational costly as well but it is possible to improve the efficiency using threadsafe version or a parallel version of SWMM (Burger and Rauch, 2012).

METHODOLOGY

With surcharging, the intrinsic storage capacity of the system (before surface flooding occurs) can be
increased dramatically beyond (e.g. even doubled) the design capacity (Butler and Davies, 2000). The
method introduced in this paper is hence applying automated pipe-sizing based on hydrodynamic
modelling so that the storage capacity of pipes and manholes could be utilized in a near-optimal way.
To optimize the storage capacity, the key issue is to determine where are the proper locations to store
excess water. If surcharges are allowed, network discharge could flow reversely against the slope due
to the backwater effect and thus be stored in upstream pipes. Therefore, the basic principle of the
2

method developed in this study is to maximize the storage capacity of upstream pipes so that the sizes
of downstream pipes can be minimized. Since the sizes of downstream pipes are commonly much
larger (and more costly) than that of upstream pipes, the method might be able to reach near-optimal
solution with nearly minimized cost by always avoiding increasing the size of downstream pipes.
The method includes three steps: 1) Sewer branch order; 2) Network decomposition; 3) Pipe sizing.
2.1

Sewer branch order

The Sewer branch order is used to describe the network topology (Sitzenfrei et al., 2012; Urich et
al., 2010). In this way, the Strahler numbers are assigned to each pipe (Strahler, 1952; Urich et al.,
2010). An example of Strahler numbers determination is given in Figure 1.

Figure 1. An example of Strahler numbers assignment for a small network. The meaning of all the
symbols remains unchanged for all the rest figures unless otherwise specified.
2.2

Network decomposition

Based on the Strahler numbers, the system is hierarchically decomposed into a number of subsystems
labelled as PSN(i). The superscript SN refers to the level of decomposition. The index i refers to the ID
of each subsystem at the same level. Except the top level, all subsystems at each level are comprised
of pipes with Strahler numbers being equal to or smaller than the current level (SN) and a downstream
pipe at the higher level connected to them. For the top level, the corresponding PSN is the whole
drainage system and the downstream pipe is the outlet pipe of the system. The sizes of those
downstream pipes are set to be infinite (e.g. >100m). Thus, flooding appears only when the capacity
of the corresponding subsystem is not sufficient, since the back water effects from the downstream of
PSNs have been eliminated. As an instance, Figure 2 illustrates the decomposition of the small network
shown in Figure 1. For combined sewer systems, notice that the system would be decomposed
according to the locations of CSO (Combined Sewer Overflow) structures first, and then be further
divided into PSNs.

Figure 2. Decomposition of the small network shown in Figure 1.


3

2.3

Pipe sizing

The automated pipe sizing procedure is illustrated as a flowchart in Figure 4. As it can be seen from
the flowchart, the procedure optimizes the capacity of each subsystem step by step following the
hierarchical structure of the studied network. Take the network in Figure 2 as an example, the method
starts from analyzing subsystems at the first level of decomposition (PSN=1) till there are no flooding in
all the subsystems (PSN=1(1),..., PSN=1(5)). Subsequently, the same process would be repeated for the
two subsystems PSN=2(1) and PSN=2(2) at the second level, and then the PSN=3. Applying this strategy
ensures the storage capacity of the upstream subsystems to be maximized. As discussed above, no
flooding in subsystems at the current level of decomposition is the prerequisite for the next step. For
this reason, if pipes in the PSN=1(1) are enlarged further during the optimization for PSN=2(1), this means
the capacity of PSN=1(1) is further increased to not only deal with the local flooding in its served region
but also to accommodate the excess flows from downstream due to the back water effect.
In the application, the so called infinite size should be defined with care. On the one hand, it must be
large enough to eliminate the backwater effects from the downstream PSN (higher level) to the
upstream PSN and the interactions between PSNs. On the other hand, it can also not be too large as to
avoid computation errors. However, the value of infinite size can be determined based on a few
trial-and-error tests, using a hydrodynamic model.
J and C in Figure 3 and 4 refers to junctions and pipes respectively. Regarding the JF and CF, JF is the
most upstream flooding node in the studied PSN. CuF is the most upstream pipe with JF being one end.
CF is the first downstream pipe of CuF, and CdFs are downstream pipes of CF (Figure 3B).
A JS node refers to a node with the following two principles being satisfied. As shown in Figure 3(A),
first the upstream pipe (Cus) connected to a JS node has its capacity (Ca = actual Depth/Max. Depth)
being equal to 1; second the downstream pipe (Cds) connected to that node has a larger size (Max.
Depth) than Cus. This is to ensure that only pipes without enough capacity are enlarged, and the
upstream pipes are always preferable. All of the variables above are defined just for facilitating the
implementation of the method.

Figure 3. Definition of JS, JF, and CF.

Figure 4. The flowchart of the automated pipe-sizing procedure.

RESULTS AND DISCUSSION

A case study of a real-world storm sewer network (Example USER1, Rossman, 2006) is carried out to
test the method. The case study network serves for a 175 hectare drainage area, divided into 58
subcatchments. The network layout is given in Figure 5(A), in which there are 59 circular pipes
connected to 59 junctions and to a single outfall. The elevation profile of the trunks drops almost 19
meters over a distance of 2.5 km (see Figure 5(B)). Figure 5(C) describes the design storm used for the
simulation. The system was solved using the software SWMM with a 5 second flow routing time step
for a 7 hours duration with a 1 minute reporting time step.

Figure 5. (A) Schematic of the case study drainage network. (B) Elevation profile of the main stem of
the case study drainage network. (C) Rainfall hyetograph for the design storm used for the case study
drainage network. The figures are cited from Rossman (2006).
As for the case study, the method is implemented based on using the default pipe sizes specified in the
model as initial estimates. For a new layout, however, the initial pipe sizes could also be determined
by using the time area method. For the optimization of each subsystem, all pipes except the one with
infinite size are selected as decision variables. The values available for the decision variables are listed
in Table 1. Two design principles are imposed in this investigation, one allowing surcharge and the
other not. This is to confirm whether the method works for both cases and can determine logic correct
solutions. A comparison between the two alternative solutions could be found in Figure 7.

Figure 6. Decomposition of the case study drainage network.

Table 1. Price list of reinforced concrete pipes (Cited from Con Cast Pipe Pricelist, 2012)
Size
(mm)
300
375
450
525
600
675
750
825
900
975
1050
1200

Unit mass
(kg/m)
225
306
381
470
578
691
780
912
1039
1195
1277
1561

Class 50-D
($/m)
65.9
81.4
83.9
91.5
131.5
201.6
265.7
308.3
369.8
405.7
464.6
582.3

Size
(mm)
1350
1500
1650
1800
1950
2100
2250
2400
2550
2700
3000

Unit mass
(kg/m)
1939
2123
2500
2865
3324
3807
4311
4869
5179
5752
7043

Class 50-D
($/m)
713.3
872.4
1,044.70
1,262.40
1,464.10
1,680.00
1,909.30
2,234.70
2,516.90
2,793.30
3,420.60

Figure 7. Comparison between alternative design strategies: (A) With surcharging (B) Without
surcharging.
The further verification of the method is to compare its solution with the global optimal solution. In
this study, a brute force method is applied to the case study network to get the global optimal solution
(simply testing all possible options in the parameter space). The objective is also to ensure no flooding
occurs during the whole simulation period. Surcharges are allowed in this case according to the design
principle. For the sake of saving computational cost, however, only the group of pipes resized by the
method developed in this study are chosen as decision variables. This simplification is introduced
because the other pipes have enough capacity and there is no necessity to enlarge them. Also, there is
no de-sizing procedure involved in the current algorithm. The range for all decision variables is from
0.3 m to 1.2 m with the same stepsize specified in Table 1. Comparison between the authors solution
and the global optimal solution is provided in Table 2.
Table 2. Comparison between the global optimal solution and authors solution
PipeID
25
26
29
30
31
41
42
43
44
45
46

Initial Max.
Depths (m)
0.600
0.600
0.400
0.450
0.500
0.300
0.230
0.300
0.300
0.500
0.500

Global
optimal
0.675
0.675
0.450
0.600
0.600
0.525
0.525
0.525
0.450
0.600
0.825

Authors'
method
0.675
0.675
0.450
0.600
0.600
0.825
0.750
0.750
0.600
0.750
0.825

PipeID
47
48
49
64
65
66
68
69
70
71
72
8

Initial Max.
Depths (m)
0.500
0.600
0.600
0.450
0.450
0.450
0.450
0.300
0.300
0.300
0.300

Global
optimal
0.825
0.900
0.900
0.750
0.750
0.750
0.750
0.675
0.675
0.600
0.600

Authors'
method
0.825
0.825
0.825
0.750
0.750
0.750
0.750
0.750
0.675
0.675
0.450

As shown in Table 2, the differences between the two solutions are not significant for 85% of resized
pipes in the network. In terms of the cost, the authors solution is about 2% higher than the optimal
solution. Regarding the computational expense, the author-designed method takes about 5 min when it
was executed on a desktop computer configured with Intel(R) Core(TM) i5-2400 CPU @ 3.10GHz
3.10 GHz, and 4.00 GB RAM. In the same environment, however, the brute force method consumes
11 hours. As proved by this case study, the method developed in this research might guarantee a nearoptimal solution that is very close to the global optimal solution, while requiring dramatically less
computational effort than global optimization method.
At present the methodology has some shortcomings that will be addressed in subsequent studies.
Firstly, the qualities of the results highly depend on the computation accuracy of hydrodynamic
simulation. Although it is rarely difficult to limit the errors to a rather small quantity, a slight
difference in the accuracy could also lead to tremendous oversize problems for some pipes. In this
case, three pipes (41, 42, and 43) are considerably oversized (more than 0.2 m larger than necessary)
by the authors method for instance. Two factors are attributing to this problem. On the one hand, the
usage of pipes with infinite size causes an increase in flow routing errors. On the other hand, the
computation error of the SWMM engine may also be a reason. The authors witnessed that enlarge of a
pipe using a reasonable increment may consequently cause flooding at upstream nodes, which might
not be logic.
Secondly, the initial estimates for the pipe sizes have considerable effects on the final solution. One
solution for this problem is to use network layouts with well-defined pipe sizes according to time area
method (Urich et al., 2010) and engineering guidelines. Another solution is to simply use the
minimum allowed pipe sizes as initial estimates, e.g. 300 mm. Further work is essential for addressing
this problem in more details.
Thirdly, oversizing of pipes at some places is unavoidable. Rules or constraints taking into account
economic factors could be imposed on the mechanism of the method. However, a good point is that
the economic factor is somehow considered implicitly by this method. Since the method could be
understood as to address a certain amount of storage capacity in the most proper location in a sewer
system with surcharge being accepted, the increase on the size for a long pipe would be definitely
smaller than that for a short one as the storage capacity required is the same.

CONCLUSIONS

A method for flood protection-based sewer network design is introduced in this paper. The method is
developed for optimal pipe sizing for both storm sewer network and combined sewer network. On the
basis of system decomposition according to Sewer branch order, the automated pipe-sizing for a
studied network could be executed to optimize the capacity of each subsystem in the network step by
step following the hierarchical structure of the network.
The reliability of the method is examined through a case study using a real world drainage system. The
case study network includes 59 junctions, 59 circular pipes and one outfall. To implement the method,
the system was decomposed into 17 subsystems belonging to three levels respectively. The default
pipe sizes specified in the model are utilized as initial estimates. For the optimization of each
subsystem, all pipes are selected as decision variables. Commercial pipe sizes (Table 1) are used as the
available values for decision variables.

The solution for the case study was then compared with the global optimal solution achieved using the
brute force method. The differences between the two solutions are not significant for 85% of resized
pipes in the network. Only three pipes are oversized by more than 40% in the authors method.
However, the loss of accuracy is compensated by the reduced computational expense, since the authordesigned method takes less than 1% than the brute force method. Compared with evolutionary
methods, the method also has its considerable advantages, since it does not require any parameter for
configuration and execution control, and could produce unique solutions as long as the design
principles are fixed.

REFERENCES

Butler, D. and Davies, J. (2000). Urban drainage. E & FN Spon, London, UK.
Burger, G. and Rauch, W. (2012). Parallel Computing in Urban Drainage Modeling: A Parallel
Version of EPA SWMM. 9th International Conference on Urban Drainage Modelling.
Con Cast Pipe website. (2012). http://www.concastpipe.com/pricing/CC_2012Pricelist.pdf (accessed
19 March 2012)
Dajani, J. S. and Hasit, Y. (1974). Capital cost minimization of drainage networks. J. Environ. Eng.ASCE, 100(2), 325-337.
Deininger, R. A. (1966). Computer aided design of waste collection and treatment systems. Proc. 2nd
Annual Conf. of American Water Resources, Chicago, USA, 247-258.
Elimam, A. A., Charalambous, C., and Ghobrial, F. H. (1989). Optimum design of large sewer
networks. Journal of Environmental Engineering, 115(6), 1171-1190.
Gidley, J. S. (1986). Optimal design of sanitary sewers. Proc. 4th ASCE Conf. on Computing in Civil
Engineering, Boston, USA, 162-177.
Goldberg, D. E., (1989). Genetic algorithms in search, optimization and machine learning. MA:
Kluwer Academic Publishers, Boston.
Guo, Y. G., Walters, G. A., Khu, S. T., and Keedwell, E. C. (2007). A novel cellular automata based
approach to storm sewer design. Engineering Optimization, 39 (3), 345364.
Guo Y. G., Walters G. A., and Savic D. A. (2008). Optimal design of storm sewer networks: Past,
Present and Future. 11th International Conference on Urban Drainage, Edinburgh, Scotland,
UK, 2008.
Holland, M. E. (1966). Computer models of wastewater collection systems. PhD dissertation, Harvard
University, Cambridge, Massachusetts, USA.
Izquierdo, H., Montalvo, I., Perez, R., and Fuertes, V. S. (2008). Design optimization of wastewater
collection networks by PSO. Computer and Mathematics with Applications, 56(3), 777-784.
Kulkarni, V. S. and Khanna, P. (1985). Pumped wastewater collection systems optimization. Journal
of Environmental Engineering, 111(5), 589-601.
Mays, L. W. and Wenzel, H. G. (1976). Optimal design of multi-level branching sewer systems. Water
Resour.Res., 12(5), 913-917.
Price, R. K. (1978). Design of storm water sewers for minimum construction cost. Proc. 1st Int. Conf.
on Urban Strom Drainage, Southampton, UK, 636-647.
Rossman, L. A. (2006). STORM WATER MANAGEMENT MODEL QUALITY ASSURANCE
REPORT: Dynamic Wave Flow Routing. Water Supply and Water Resources Division National
Risk Management Research Laboratory Cincinnati, OH 45268, USA.

10

Rossman, L. A. (2010). Storm Water Management Model users manual (version 5.0). U.S.
Environment Protection Agency, Cincinnati, USA.
Savic, D. A. and Walters, G. A. (1997). Genetic algorithms for least cost design of water distribution
networks. Journal of Water Resources Planning and Management, 123(2), 67-77.
Sitzenfrei, R., Urich, C., Mderl, M. and Rauch, W. (2012). Assessing the efficiency of different CSO
positions based on network graph characteristics. 9th International Conference on Urban
Drainage Modelling, Belgrade 2012.
Strahler, A. N. (1952). Dynamic basis of geomorphology. Geol. Soc. Am. Bull. 63, 923-938.
Urich C., Sitzenfrei R., Moderl M. and Rauch W. (2010). An agent-based approach for generating
virtual sewer systems. Water Sci Technol, 62 (5), 1090-7.
Walters, G. A. and Templeman, A. B. (1979). Non-optimal dynamic programming algorithms in the
design of minimum cost drainage systems. Eng. Optimiz., 4, 139-148.
Yates, D. F., Templeman, A. B., and Boffey, T. B. (1984). The computational complexity
of the
problem of determining least capital cost designs for water supply networks. Engineering
Optimization, 7(2), 142-155.
Yen, B. C., Cheng, S. T., Jun, R. I., Voorhees, M. L., Wenzel, Jr, H.G., and Mays, L. I. (1984). Least
Cost Sewer System Design Model. Users Guide. Illinois Austin, TX.
Zecchin, A. C., Simpson, A. R., Maier, H. R., Leonard, M., Roberts, A. J., and Berrisford, M. J.
(2006). Application of two ant colony optimization algorithms to water distribution system
optimization. Mathematical and Computer Modelling, 44 (5-6), 451-468.

11

9th International Conference on Urban Drainage Modelling


Belgrade 2012

Parallel Computing in Urban Drainage Modeling:


A Parallel Version of EPA SWMM
Gregor Burger1, Wolfgang Rauch2
1
2

University of Innsbruck, Austria, gregor.burger@uibk.ac.at


University of Innsbruck, Austria, wolfgang.rauch@uibk.ac.at

ABSTRACT
The hydrodynamic rain-fall run-off simulation model SWMM is state of the art in
research and practice. In order to reduce the burden of long simulation runs and use
the extra power of modern multi-core computers a parallel version of SWMM is
presented. The challenge was to modify the software in such a minimal way that
the changes may find its way into the several commercial and non-commercial
tools that depend on SWMM for its calculations. A pragmatic approach to identify
and enhance the most runtime intense parts of the software was chosen in order to
keep the code changes as low as possible. The enhanced software was then
benchmarked on four different input scenarios ranging from a very small village to
a medium sized city. In the investigated sewer systems a speedup of six to ten times
on a twelve core system was realised, thus decreasing the execution time to an
acceptable level even for tedious system analysis.

KEYWORDS
Multi-Core, OpenMP, Parallel Computing, SWMM, Urban Drainage Modelling

INTRODUCTION

The Storm Water Management Model (SWMM) is a dynamic rainfall-runoff simulation model used
for single event or long-term simulation of runoff quantity and quality from urbanized areas (Huber
1995). SWMM was developed by the US EPA and therefore the code it is built upon is in possession
of the public domain. The fact that a code of such a high quality model is available for everyone to use
and alter, made the tool very popular in research and in the applied civil engineer field.
Nowadays numerical models like SWMM are not just used in run-once scenarios but in sensitivity
analyses (C.B.S. Dotto et al. 2011; Mair et al. in press), uncertainty analyses (Kleidorfer, Deletic, et al.
2009; Kleidorfer, Mderl, et al. 2009; Cintia B.S. Dotto et al. 2012) and vulnerability analyses
(Mderl et al. 2009). All of those requiring a multitude of execution runs. In fact, a comparison of
different uncertainty assessment techniques (GLUE, Bayesian methods, etc) revealed that the number
of simulation runs, depending on which technique used, ranged in that particular example from 1600

to over 30000 runs, but the mean is around 3500 runs, until the uncertainty of the different system
parameters is devised (Cintia B.S. Dotto et al. 2012).
Given the fact that these analyses are essential for a deeper understanding of the underlying model and
input system there is an imminent need for action. For researchers applying these analyses a way to
reduce the time spent for computer simulation is to restrict the system parameters or to reduce the
size/resolution of the input system. Both solutions are suboptimal and may degrade the quality of the
results which in the best case can cause additional simulation runs or in the worst case a
publication/assessment with imprecise or wrong results.
A better way to reduce the run-time of the simulations is to improve the performance of the underlying
simulation codes. One step in this direction has already happened. CITYDRAIN (Achleitner, Moderl,
and Rauch 2007), a hydrological model developed at the University of Innsbruck, was redesigned and
rewritten into CityDrain3. CityDrain3 was designed from the ground up to have native performance
instead of interpreted MATLAB and use all the extra parallel computing power that is available in
modern multi-core computers (Burger et al. 2010). In this paper we will outline the efforts of bringing
one of the most prominent hydrodynamic codes, namely SWMM, into the multi-core era. In the course
of this manuscript we describe what parallel computing is, what challenges there are, why it is
essential to do it nowadays, how a parallel version of SWMM looks like and how much performance
there is to be expected from a parallel SWMM.
Beside the apparent reason of an immediate performance gain that is due to parallel computing there is
a not so obvious second reason for parallelizing simulation software and software in general. Semi
conductor researchers are hitting hard physical limits that prevent them to scale in single thread
performance. They cannot reduce the chip size indefinitely neither can they increase the speed of light.
In order to keep customers satisfied and stay true to Moores law, which states that the transistor count
doubles roughly every two years, they needed to make radical changes (Moore 1998). For chipmakers,
the way out of this misery was to pack several cores onto one die. The multi-core era was born and
now they can again - double the number of cores every two years (Olukotun et al. 1996).
Although multi-cores solve the problem for the hardware industry, the real problem was shifted onto
the software developers. Every application out there must be altered in order to take advantage of the
extra cores that are now built into every CPU. Software is not getting faster anymore on modern CPUs
as it was the case before the multi-core era. In a very famous paper called the free lunch is over Herb
Sutter describes it in an even more drastic way and states that the performance of software that does
not take advantage of parallel computing is not only stagnating but will degrade on future many-core
devices (Sutter 2005). Many-core devices and hybrid solutions is the future that software developers
will face. In a recent article the same author describes the complexity and diversity of the modern and
future parallel computing landscape (Sutter 2011).

METHODOLOGY

The architectural landscape that parallel computing is applied to, can be quite widespread nowadays;
ranging from dual-core phones up to the overly hyped cloud-computing architecture (Sutter 2011).
Beside all the hype of cloud computing, urban drainage simulations are used by engineers,
infrastructure planners and researchers in the field of urban drainage for which the desktop computer is
still the preferred computing platform. A state of the art, off-the shelf desktop computers hosts a CPU
with up to six or eight cores. Therefore, we focused on this parallel computing (PC) architecture called
multi-core computing (Olukotun et al. 1996). The advantage of focusing on the multi-core architecture

is that it is readily available and that the porting overhead is much lower compared to other parallel
computing architectures like GPUs or cloud based systems.
SWMM has a very mature code base that is tested and verified. The code is heavily used in its standalone version and included in commercial applications. We wanted to have as little impact as possible
on the code structure. The idea was that such minimal code changes would result in a higher
confidence and understanding of the changes so that adoption of the code or even inclusion in the main
SWMM code can happen. This minimal impact requirement and the fact that a mature code needs to
be portable drove us to decide on OpenMP as a base parallel application programming interface (API).
OpenMP is a compiler extension that allows parallelizing of new and old code. In the best case the
already existing code only needs to be decorated with parallel instructions and synchronization
primitives. Therefore, OpenMP allows us to have the minimum code changes as described in the
previous paragraph. Another reason for choosing OpenMP was that it is supported by most compilers
and it has matured into a well understood and standardized parallel computing API for which a lot of
performance and testing tools are available.

Figure 1 Profile of the sequential SWMM code.


The first step for a parallel SWMM was to search for performance sensitive spots. It makes no sense to
parallelize a code that does only take five percent of the overall SWMM runtime. After intensive
profiling the critical parts for performance where identified. As expected, the routing of the conduits
and sewers are the most time intensive tasks. A profile as shown in Figure 1 reveals exactly at which
line the most time is spent. For someone not familiar with a code base, as it was the case in this work,
this is very helpful. The profile shows that the function getConduitFlow is responsible for 65% of the
overall run-time. This function is called by the execRoutingStep function that iterates over all conduits
and calls find/getConduitFlow for each link.
In the first step we identified parts of the code that have the potential to run parallel. The code is
already in a structure that allows easy parallelisation. In execRoutingStep a loop iterates over an array
of Link structures. Each link is then argument to the getConduitFlow function that calculates the new
flow for each Link. The most challenging part in this step was to check if each Link can be updated
separately. (Rossman 2006) states that SWMM solves the Saint Venant equations with a finite
difference scheme. In the first step the flow for each link is being calculated. After that the hydraulic
3

head (elevation head plus any possible pressure head) for each node is updated. The model of SWMM
is that links are conduits and nodes are junctions in the graph describing the wastewater system. This is
already a first hint that a parallelization at the level of each conduit and link is possible.
The ultimate proof, though, is the code. This means that each and every line of code that may be
executed in the course of the call of findConduitFlow needs to be checked for race-conditions and
critical sections. As one can imagine reading and understanding other people code and then checking
for parallel programming errors is a very tedious, time consuming task that needs full attention. An
error in this part can lead to undefined behaviour, data corruption, wrong results or even crashes of the
program. The problem with these errors is that they may not be triggered for a long time and are hard
to reproduce. Besides fixing the introduced race-conditions, the code also contained non reentrant
functions that needed to be fixed so that multiple threads do not interfere with global variables in a
multi-threaded simulation.
The same procedure described in the previous section for findConduitFlow has been applied to
findNonConduitFlow, initNodeState, link_setOutFallDepth and setNodeDepth by means of several
benchmarking, implementation and testing cycles.

RESULTS

The benchmarks were performed on a Dual Xeon System. Each processor featured six hyper-threaded
cores resulting in up to 24 virtual threads for the whole system. Each core runs with a clock frequency
of 2.67 GHz. The cores of each CPU package share a 12Mb L3 cache, enhancing further the parallel
performance of the CPU. The system was equipped with 24 GB of main memory. The system was
operated by Linux with a kernel at version 3.2.
Each input system was run with one thread as a base measurement, followed by runs with increasing
thread count up to the 24 maximum threads the system features. A step size of two was chosen
because an uneven thread number is not ideal in a hyper-threaded system. The average run-time of
four runs was then taken as the resulting run-time (left images). The speedup refers to how much a
parallel algorithm is faster than a corresponding sequential algorithm (right images).
The four input systems are in an increasing number of nodes and links, i.e. they are getting bigger. The
general trend that the benchmarks show is that the bigger a system the better the parallel version of
SWMM scales. A reason for this is that there is a sequential part in the routing algorithm that cannot
be parallelized. This sequential part seems to be more or less independent of the system size. Amdahls
law states that the bigger the sequential part of a parallel algorithm is, the worse is the scaling of the
speedup (Amdahl 1967; Hill and Marty 2008).
Table 1 Number of Elements per Input System
Input System
Nodes
Links

Sub catchments

Population

Artificial

50

49

42

Unknown

Village

1709

1722

440

10760

Small Town

1254

1274

3062

12695

Town

5485

5834

4498

120147

In order to show at what system size it is feasible to use the parallel SWMM code we used sewer
systems of different extension. This first benchmark system is an artificially generated system having
an alpine character. The system was generated using the Case Study Generator (Moderl, Butler, and
Rauch 2009). The next three benchmarks systems, listed in Table 1, are in order of population. All
four systems have an alpine character, which means they are more or less in a tree structure with only
few loops. Nevertheless as the dynamic wave routing was optimized and used in the benchmark
systems loops do not necessarily have an influence on the performance. The village module represents
a small village in a rural area, the small town is in a suburban area and the town is a regional capital
city. Beside the dynamic wave routing, Horton infiltration was used in the SWMM options, allow
ponding was disabled and skip dry weather periods was disabled in the input systems.
3.1

Artificial

As mentioned previously the artificial system was chosen to see how the optimizations perform when
a small input system is used. In a small system the parallel overhead and the sequential part of the
parallel algorithm are high compared to the part of the algorithm that runs in parallel. Figure 2 shows
how this influences the overall runtime and speedup. Up until eight threads the runtime decreases but
then rises again. Although the speedup is not good there are no signs of a slowdown, which means that
there is no disadvantage of using the parallel version of SWMM.

Figure 2 Runtime and Speedup for the Artificial input system.


3.2

Village

The village system is the first one having enough runtime to cover the overhead and sequential parts of
the routing algorithm. Up until twelve threads the parallel SWMM version shows a very good speedup and the code is 8.3 times faster. At 24 threads the minimum runtime is reached and a maximum
speed-up of 9.5 times. This means that a parallel version of SWMM is almost 10 times faster at 24
threads than the standard SWMM. The speed-up curve in Figure 3 has a slight kink at twelve threads.
This kink is because up until twelve threads the Linux scheduler can schedule all threads on real
hardware threads, beginning with the 13th thread the virtual hyper-threads must be used, which causes
an additional overhead in the CPU if an application is floating point intensive. The overhead stems
from the fact a real- and a hyper-thread share the floating point unit. Because of this additional overhead the speedup-up curve is shallower beginning with the 13th thread.

Figure 3 Runtime and Speedup for the Village input system.


3.3

Small Town

The small town input system shows a slight worse speed-up than the village. Although the system has
a higher population the modeller chose a lower resolution for links and nodes but a higher resolution
for sub-catchments. As stated in the previous sections the primary target of parallelization was the
routing of the channels so therefore the scaling is better when more pipes and less sub-catchments are
in the system. Nevertheless parallel SWMM version reduces the simulation time 6 times from 36
seconds town to 6 seconds.

Figure 4 Runtime and Speedup for the Small Town input system.
3.4

Town

The town input system is the biggest and most detailed system. With around 5000 links and nodes the
sequential runtime is over 3.5 minutes. With the parallel SWMM code this runtime is reduced to 26
seconds resulting in a speedup of 9.3 times. At the University Innsbruck this sewer system was heavily
used and analysed including vulnerability and sensibility analyses. These analyses require several runs,
up to the hundreds, of a slightly modified input system. One can imagine what a ten time reduction in
run-time could mean to such analyses, heavily bound on the run-time of a single simulation (e.g. longterm simulations could be feasible).

Figure 5 Runtime and Speedup for the Town input system.

CONCLUSION AND DISCUSSION

Take for example the uncertainty assessment described in the introduction where the analysis took
around 3500 iterations to evaluate the uncertainty of five system parameters for the given case. When a
single simulation run takes around five minutes, researchers must wait twelve days until she/he can
finally interpret the results. The processing unit running this simulation is blocked for twelve days.
And five minutes is not even a long run for an urban drainage simulation. Applying the techniques
described in this paper one can reduce the runtime of this uncertainty simulation down to 29 hours.
Although it should not be, but such delays do affect the decisions of researchers whether they run such
analysis or just skip them because they cannot sacrifice twelve days of their precious time. A parallel
version of SWMM also opens ways to longer and more detailed simulations. And in case of a certain
pressure more money for some extra cores can speed up the simulation additionally. Without a parallel
version of SWMM the performance of hydrodynamic simulations will stagnate or may even degrade
on future computing technologies.
In this paper we outlined the fundamentals of parallel coding for the well known hydrodynamic
software SWMM. We demonstrated that a rather small part of the code is decisive for the execution
time. The change of this part of the code into a parallel version resulted into a significant speedup in
the execution. The speedup is not linear but increases both with the complexity of the system (the
more pipes the better) and the number of threads. In the investigated real world systems the speedup
amounted to 6 to 10 times on a PC with 12 threads.

ACKNOWLEDGMENT

This work has been financially funded in the course of the PaCoWaDi project by the
Bundesministerium fr Verkehr, Innovation und Technologie in the program FIT-IT/ ModSim (FFG
projectnumber 2059687).

REFERENCES

Achleitner, S., M. Moderl, and W. Rauch. 2007. CITY DRAIN\copyright-An Open Source
Approach for Simulation of Integrated Urban Drainage Systems. Environmental Modelling &
Software 22 (8): 11841195.
Amdahl, G.M. 1967. Validity of the Single Processor Approach to Achieving Large Scale Computing
Capabilities. In Proceedings of the April 18-20, 1967, Spring Joint Computer Conference,
483485.
Burger, G., S. Fach, H. Kinzel, and W. Rauch. 2010. Parallel Computing in Conceptual Sewer
Simulations. Water Science and Technology: a Journal of the International Association on
Water Pollution Research 61 (2): 283.
Dotto, C.B.S., M. Kleidorfer, A. Deletic, W. Rauch, D.T. McCarthy, and T.D. Fletcher. 2011.
Performance and Sensitivity Analysis of Stormwater Models Using a Bayesian Approach and
Long-term High Resolution Data. Environmental Modelling & Software 26 (10) (October):
12251239. doi:10.1016/j.envsoft.2011.03.013.
Dotto, Cintia B.S., Giorgio Mannina, Manfred Kleidorfer, Luca Vezzaro, Malte Henrichs, David T.
McCarthy, Gabriele Freni, Wolfgang Rauch, and Ana Deletic. 2012. Comparison of Different
Uncertainty Techniques in Urban Stormwater Quantity and Quality Modelling. Water
Research 46 (8) (May 15): 25452558. doi:10.1016/j.watres.2012.02.009.
Hill, M.D., and M.R. Marty. 2008. Amdahls Law in the Multicore Era. Computer 41 (7): 3338.
Huber, WC. 1995. EPA Storm Water Management ModelSWMM. Computer Models of Watershed
Hydrology 1: 783808.
Kleidorfer, M, M Mderl, R Sitzenfrei, C Urich, and W Rauch. 2009. A Case Independent Approach
on the Impact of Climate Change Effects on Combined Sewer System Performance. Water
Science and Technology: A Journal of the International Association on Water Pollution
Research 60 (6): 15551564. doi:10.2166/wst.2009.520.
Kleidorfer, M., A. Deletic, T. D. Fletcher, and W. Rauch. 2009. Impact of Input Data Uncertainties
on Urban Stormwater Model Parameters. Water Science & Technology 60 (6) (September):
1545. doi:10.2166/wst.2009.493.
Mair, M., R. Sitzenfrei, M. Kleidorfer, M. Mderl, and W. Rauch. in press. GIS-based Applications
of Sensitivity Analysis for Sewer Models. Water Science & Technology.
https://mail.google.com/mail/?shva=1#search/manfred.kleidorfer@uibk.ac.at/13567075afcc7ffe.
Moderl, M., D. Butler, and W. Rauch. 2009. A Stochastic Approach for Automatic Generation of
Urban Drainage Systems. Water Science & Technology 59 (6): 11371143.
Mderl, M., M. Kleidorfer, R. Sitzenfrei, and W. Rauch. 2009. Identifying Weak Points of Urban
Drainage Systems by Means of VulNetUD. Water Science & Technology 60 (10) (November):
2507. doi:10.2166/wst.2009.664.
Moore, G. E. 1998. Cramming More Components Onto Integrated Circuits. Proceedings of the IEEE
86 (1) (January): 8285. doi:10.1109/JPROC.1998.658762.
Olukotun, Kunle, Basem A. Nayfeh, Lance Hammond, Ken Wilson, and Kunyung Chang. 1996. The
Case for a Single-chip Multiprocessor. SIGPLAN Not. 31 (9) (August): 211.
doi:10.1145/248209.237140.
Rossman, L.A. 2006. Storm Water Management Model, Quality Assurance Report: Dynamic Wave
Flow Routing. Cincinnati,OH: US Environmental Protection Agency, Office of Research and
Development, National Research Management Research Laboratory.

Sutter. 2005. The Free Lunch Is over: A Fundamental Turn Toward Concurrency in Software. Dr.
Dobbs Journal 30 (3): 202210.
Sutter. 2011. Welcome to the Jungle. Sutters Mill. http://herbsutter.com/2011/12/29/welcome-to-thejungle/.

9th International Conference on Urban Drainage Modelling


Belgrade 2012

Modelling of percolation rate of stormwater from


underground infiltration systems
Ewa Burszta-Adamiak1, Janusz Lomotowski2
1
2

Wroclaw University of Environmental and Life Sciences, Poland, ewa.burszta-adamiak@up.wroc.pl


Wroclaw University of Environmental and Life Sciences, Poland, janusz.lomotowski@up.wroc.pl

ABSTRACT
Underground or surface stormwater storage tank systems with infiltration of water
into the ground constitute the basic elements used in Sustainable Urban Drainage
Systems (SUDS). So far, the methods of designing such facilities have not taken
into account the phenomenon of ground clogging during the infiltration of
stormwater. Sealing of the top layer of the filter bed influences the infiltration rate
of water into the ground.
This study presents an original, mathematical model describing the changes in the
infiltration rate in the phases of filling and emptying of storage and infiltration tank
systems, which enables to determine the degree of clogging of the top layer of the
ground. The input data for modeling were obtained from studies conducted on
experimental sites on objects constructed in semi-technological scale.
The tests have proved that the developed model is useful on the stage of designing
stormwater infiltration facilities and that it helps to control the degree of clogging
of absorptive surfaces during the exploitation of such facilities.

KEYWORDS
clogging; hydraulic resistance; modelling; storm water management; underground infiltration system

NOMENCLATURE
F , bottom surface of infiltration facility (cm2)
H 0 water level in the infiltration module at the end of the filling phase (cm)

H (t ) water level in the infiltration module at time t (cm)

H s water level above ground surface (cm)


h f negative pressure head at wetting front (cm)

I (t ) accumulated water infiltration into the ground (cm)


K f wetting zone hydraulic conductivity (cm min-1),

Q infiltration flow (cm3 min-1),

QF variable parameter used in model (10) (cm min-1),


q(t ) infiltration rate at time t (cm min-1),
R substituted hydraulic resistance (min)
t time (min)
Z (t ) depth of wetting front (cm)

INTRODUCTION

Stormwater management is becoming aimed at local infiltration or retention. Such tasks can be
realised with use of underground or surface storage and infiltration systems. In spite of a growing
interest in the application such facilities, they are still designed and exploited without taking into
consideration the process of clogging during usage. Assuming, for calculation purposes, the lowest
value of the filtration coefficient from the collection of results obtained during geotechnical studies,
does not correspond to the dynamic changes in the filtration coefficient caused by the clogging
process. The assumption that the filtration coefficient remains fixed throughout the exploitation period
is not only incorrect, but also harmful, as with time the need occurs to modernize existing elements of
the system and to invest additional funds. Numerous studies (Burszta-Adamiak, 2005; BursztaAdamiak and omotowski, 2005b; Mallin et al., 2009; Marla and Lee-Hyung, 2010; Rupak et al.,
2010; Zhuanxi et al., 2012) have shown that water flowing into the facilities contains a significant
amount of pollutants, which deteriorate the infiltration parameters of the infiltration modules. With
time of exploitation of infiltration modules, layers of sediments start to form on the bottom and side
walls of such reservoirs, and soil pores are being sealed. This process is known as ground clogging.
We distinguish between physical, biological and chemical clogging. In fact, the process is a very
complex one, being a resultant of specific individual processes. Physical clogging during the
infiltration of stormwater is caused by additives that remain in a suspended state. Rain washes out of
the air the molecules remaining in gaseous state, aerosols and dusts, of natural or anthropogenic origin.
Physical clogging can also be caused by bubbles of gas exuded from water or from the soil. The
development of biofilm in the sediment zone and in the layer adjacent to soil contributes to biological
clogging. Chemical clogging takes place when suspensions or insoluble minerals are deposited on
grains of soil. Chemical clogging is mainly caused by calcium carbonate and insoluble ferrite
compounds deposited from the water (Hua et al., 2010; Nivala et al., 2012; Skolasiska, 2006;
Vigneswaran and Suazo, 1987).
The phenomenon of clogging occurs on infiltration water intakes, during the filtration of water
through rapid and slow filters, in the course of exploitation of underground water intake facilities
(clogging of well filters and drains), trickling filters, sand filters and subsurface wastewater infiltration
systems (Oe et al., 1996; Rinck-Pfeiffer et al., 2000; Hiscock and Grischek, 2002, Lloyd et al., 2009;
Mays and Asce, 2010; Pedretti et al., 2012). Regardless of the type of the given infiltration system,
clogging is an undesirable phenomenon (Bouwer, 2002; Bouwer et al., 2001; Gautier et al., 1999). The
thickness of the clogging layer can range from several millimetres to several centimetres or even
decimetres, for larger amounts of accumulated sediments (Bouwer, 2002). The phenomenon of
clogging develops at various rates. Usually, in the first year of exploitation of infiltration facilities, no
significant decrease in water infiltration to the ground is observed, although in the subsequent years of
usage such decrease can reach even up to 50% of the initial permeability( Balades et al., 1995). On the
other hand, Geiger and Dreiseitl (1999) disagree, claiming that the phenomenon of clogging of
stormwater infiltration facilities is the most intense during the start-up phase. During that time the
adjacent area is usually not yet overgrown by plants, but heavily polluted with fine dust that appears as

a result of construction works. Erosion causes soil particles and other pollutants to enter the filtration
zone together with stormwater, causing rapid mechanical sealing.
The basic technological parameter that is taken into account during the process of designing artificial
infiltration systems is the infiltration rate. The first model of infiltration of water into the ground was
developed by Green and Ampt in the year 1911 (Ravi, 1998; Williams et al., 1998; Ying et al., 2010).
The model is based on the assumption that the water penetrates into the ground according to Darcy
law, whereas the infiltration rate is determined by head loss in the saturated and wetted zones Z:

q( t )

H Z (t ) h f
dI (t )
K f s
dt
Z (t )

Assuming that:

H s h f

Z (t ) / K f

Z(t)=Zconst

and introducing a parameter R defined by the ratio:


R=Zconst/Kf.
equation (1) will has the form:

q(t )

H s hf
R

Kf

Kf

(1)

(2)
(3)
(4)

The Green-Ampt model offers a good description of measurement results in cases characterised by
stable inflow conditions, i.e. when a fixed layer of water remains on the ground surface. For variable
inflow conditions (which are typical for stormwater infiltration facilities due to the random nature of
rainfall) the error in transient infiltration rate prognosis and accumulated infiltration of water into the
ground increases.
A well-known model that describes the mechanical clogging of filter beds while taking into account
the changes in the concentration of suspended solids in liquid during the flow through porous media is
the Iwasaki equation, developed in 1937 (Iwasaki, 1937 in: Tesaik ,1980). Some examples of models
describing chemical and biological clogging can be found, among others, in the works of Vandevivere
et al. (1995), Teylor and Jaffe (1990), Taylor et al. (1990), Prez-Paricio (2001) and Seki and
Miyazaki (2001).
In spite of the fact that numerous models have been developed that allow for a better understanding of
the nature of phenomena occurring in porous media, the process of clogging in stormwater infiltration
facilities is still typically evaluated by means of an evaluation of the changes in the infiltration
intensity during a given test period (Raimbault et al., 1999). The intensity of infiltration decreases
gradually, until, with time, a layer of low-permeable soil is created that does not meet design
requirements (Balades, 1995). This can be illustrated by the equation used for the hydraulic evaluation
of the functioning of clogged infiltration systems, which was developed by Bouwer (1969). Numerous
variations of the Bouwer model can be found, among others, in the works of Dechesne et al. (2004)
where it has been applied for the purposes of evaluation of infiltration basins with a clogged sediment
layer on the bottom and by Gautier et al. (1999), who tested and then described the process of
infiltration of water through absorptive surfaces, dividing the flow into infiltration through the clogged
bottom and banks of the basin.
The models used to evaluate the phenomenon of clogging in stormwater infiltration facilities presented
in literature are developed basing on the results of tests performed on surface infiltration systems. Due
to the fact that the area designed for the construction of infiltration systems is often limited, it is quite
often required to use underground infiltration systems, e.g. in form of infiltration module systems or
infiltration trenches, etc. These facilities function properly when there is a need to absorb a larger
amount of water than the infiltration and retention capabilities of the adjacent land allow to absorb in a
3

specific period of time. Underground storage facilities first store the collected inflow of stormwater
and then enable free infiltration of water into the ground. The evaluation of the progress of the
clogging phenomenon in underground storage facilities is difficult due to the fact that the layer of
suspended solids and/or biomass is deposited on the infiltration surface located below land surface,
thus in this type of facilities it is difficult or even impossible to perform any declogging activities that
are traditionally performed in surface infiltration systems.
Insufficient information concerning the hydraulic fundamentals of designing underground stormwater
infiltration facilities lead to experiments. The aim of the analyses was to develop simple models
describing the changes in the infiltration rate in the phase of filling and emptying of retention and
infiltration reservoirs and to test their usability in the evaluation of the progress of clogging.

2
2.1

METHODS
Characteristics of the model

The water volume balance equation for infiltration module systems with an absorptive surface F which
is filled at a constant rate Q, without taking into account the impact of infiltration through side walls
can be formulated as follows:

F dH (t ) Q dt F q(t ) dt

(5)

where the left side of the equation describes the increase in water volume in the infiltration module
and the right side describes the volume of water flowing into the reservoir, less the amount of water
infiltrating into the ground. This is illustrated in Figure 1

Figure 1. Sample drawing illustrating momentary water volume balance during the process of water
infiltration from the module.
Assuming that:

q( t )

H (t )
Kf
R

and introducing a parameter defined by the ratio:


4

(6)

QF

Q
F

(7)

it is possible to determine the equation describing the changes in water surface level in an infiltration
module for water flowing in at a constant rate:

t
t
H (t ) QF R 1 exp K f R 1 exp
R
R

(8)

In the case if water will not be flowing into the module, a decrease in the water level due to filtration
will be observed. The function of change of the water surface level in the module in the emptying
phase will be described by the following equation:

t
t
H (t ) H 0 exp K f R 1 exp
R
R

(9)

During the infiltration of water containing suspension a significant decrease of the filtration coefficient
of the top layer of soil is observed. In the case if the value of the product of the Kf R constants is close
to 0, equations (8) and (9) will have the respective forms:

t
H (t ) QF R 1 exp
R

t
H (t ) H 0 exp
R

2.2

( 10 )

( 11 )

Description of the experimental site and methodology

The experimental sites were constructed with use of prefabricated openwork modules. The side walls
of those modules contain apertures that enable the infiltration of inflowing water into the ground.
Modules were wrapped in 1.6 mm thick geotextile made from polypropylene, characterised by
perpendicular water permeability of 0.0026 m/s. The dimensions of the infiltration modules are

500x1000x400 mm (length x height x width). These are systems prepared for the management
and infiltration of stormwater, commonly used in engineering practice.

Infiltration modules were placed in the ground according to the guidelines provided by the
manufacturer. Prior to the beginning of the experiments, geotechnical tests were conducted in the site
where the modules were located. Lithological profiles of the soil on the site of experiments are
presented in Table 1. Below the bottom of experimental site no.1 there was a deposit of sandy clay,
covered by medium sand reaching up to the surface of the soil. Experimental site no 2 was located on
cohesive clay, covered by sandy clay up to the depth of 0.4 m below land surface. Above that depth, it
was covered with a deposit of fine sand.

Table 1. Lithological profiles of the soil on the site of experiments.

In order to prepare the clogging agent (dispersion suspension), kaolin clay was used on both
experimental sites. The kaolin clay suspension used for the tests was determined basing on two years
studies of the grain size distribution and concentration of suspensions present in rainfall, snowfall and
roof runoff. These study was conducted with use of a laser particle sizer Mastersizer 2000
manufactured by Instruments Ltd. Samples were collected both on the site where later studies on
infiltration modules were conducted, and in several other Polish cities. The preliminary studies show
that the average respective concentrations of suspensions in samples of rainfall, snowfall and roof
runoff were 0.0075% vol., 0.0082% vol. and 0.012% vol. By referring the obtained results to the
concentration of kaolin clay suspension and the volume of suspension introduced into the test sites, it
can be stated that one year of conducted analysis corresponds to approximately five years of
exploitation in real conditions. The use of higher concentrations of kaolin clay suspension resulted
from the need to intensify the studies of the clogging process, which is much slower in real conditions.
The particle size of the clogging agent fell within the range from 0.25 to100 m. These particles
accounted for approximately 60 % of the pollutants present in rainwater. Sample particle size
distribution of the pollutants present in stormwater and in kaolin clay is presented in Figure 2. The
selection of a model suspension characterised by a particle size composition similar to that of
stormwater enabled us to model the processes occurring in nature in a more precise way. Modules
were each time filled with 60 dm3 of suspension of the concentration of 2.5 g/dm3. Suspension used
for the tests was prepared on the basis of tap water, after several days of soaking in order to eliminate
the process of expansion of minerals. The filling lasted for 8-10 minutes, and the infiltration timefrom 0.5 hours at the beginning of the test period to 9 hours after a year of exploitation. The tests on

the sites were conducted in the period from 18.06.2003 to 14.06.2004. During that time the
storage and infiltration modules were filled on the average once a week. During the tests the
duration of the experiment was measured, along with the changes in the level of water with kaolin clay
suspension in the modules in the filling and infiltration stages.
6

Percent finer by weight, %

100
80
60

40
20
0
0,1

10

100

1000

10000

Particle Size, mm
Sample 1
Sample 4

Sample 2
Kaolin clay

Sample 3

Figure 2. Particle size distribution in suspensions present in rainwater collected at the location of the
measurement sites and in kaolin clay used for the tests.

RESULTS AND DISCUSSION

Due to a large amount of factors influencing clogging and to their significant variability in time, it is
practically impossible to model the processes occurring in nature. The construction of a simplified
experimental model, together with the application of a dispersion suspension characterised by an
adequate concentration and size of particles enables us to model processes that would take even
several or over ten years in objects functioning in the real world, in short periods of time.
The impact of infiltration through side walls was omitted in calculations, as the scanning photos of the
sediment deposited on the surface and inside the geotextiles, collected from the bottom and walls of
the sites after the tests were ended, show that the flow of suspension occurred mainly through the
bottom, which is proved by a larger amount of sediments deposited in this part of the sites (Figure 3).
This is mainly a result of the mechanisms of the sedimentation process, which is one of the specific
processes occurring during the infiltration of polluted waters. In experimental site no.1, as much as
88.4% of the total mass of kaolin clay sediments deposited in the module clogged the bottom, while
only 11.6 % was deposited on the walls. A similar situation was observed in experimental site no. 2,
where 90.1% of the total mass of kaolin clay was deposited on the bottom in form of sediment, while
9.9 % was found in the geotextile covering the walls.

Figure 3. Scanning photos of kaolin clay sediment deposited in the geotextile on the bottom of the
module (left image) and on the side walls of the site (right image) magnified 1000 times
7

3.1

Changes in the water level during filling and infiltration

Sample diagrams showing the measured change in water levels in the infiltration module systems
during filling and infiltration as well as regression functions described by the general models (equation
10 and 11) calculated with use of STATISTICA 10.0 PL software are shown, respectively, on Figures
4, 5 and 6, 7.
25

35
30

20
Water level, cm

Water level, cm

25
15

10

20
15
10

5
5
0

10

Time, min

10

12

Time, min

30

30

25

25

20

20

Water level, cm

Water level, cm

Figure 4. Comparison of the changes in water levels during the filling of experimental site no.1,
measured on the 15.09.03 (left) and the 10.12.03 (right) with a regression function described by the
general model (10)

15
10
5
0

15
10
5

10

Time, min

10

Time, min

Figure 5. Comparison of the changes in water levels during the filling of experimental site no. 2,
measured on the 11.07.03 (left) and the 18.08.03 (right) with a regression function described by the
general model (10)
25

20
Water level, cm

Water level, cm

20

15

10

15

10

10

15

20

25

30

0
0

Time, min

10

15

20

25

30

35

40

Time, min

Figure 6. Comparison of the changes in water levels during the infiltration in experimental site no.1,
measured on the 30.07.03 (left) and the 11.08.03 (right) with a regression function described by the
general model (11)

30
25
25
Water level, cm

Water level, cm

20
15
10
5
0

20
15
10
5

20 40 60 80 100 120 140 160 180 200 220 240

Time, min

20

40

60

80 100 120 140 160 180 200 220


Time, min

Figure 7. Comparison of the changes in water levels during the infiltration in experimental site no. 2,
measured on the 09.07.03 (left) and the 14.07.03 (right) with a regression function described by the
general model (11)
The calculated regression functions describe the water level fluctuations in time in tested underground
facilities very well, as is proved by high values of determination coefficients obtained for these
functions, falling within the range from 0.858 to 0.999 for modelling with use of equation (10) and
from 0.845 to 0.992 when model (11) was used to describe the process of water infiltration. During the
evaluation of the measured water layer fluctuations during the emptying of the test infiltration
modules, an acceleration of the transient infiltration rate was noted at the end of the infiltration phase.
In the analysed test sites the change in the gradient of transient infiltration rates occurred at water level
between 4-6 cm. This phenomenon was noted for all measurements. It could be explained by an
increase in the suction power of the soil located below the geotextile. At high transient infiltration
rates the thickness of fully saturated zone stabilises. When the transient infiltration rate decreases, the
thickness of the fully saturated zone starts to decrease, as more water flows out than it flows in from
the land surface direction in the zone that has not been fully saturated with water. This causes a
decrease in soil moisture and an increase of suction pressure below the surface of sediments. The
decrease in the thickness of the saturated layer combined with an increase in the suction pressure lead
to an acceleration of the transient rate, as described by the Green-Ampt equation (1). The work of
Burszta-Adamiak and omotowski (2005a) presents a description of segment regression, which
provides a better description of the relation between water level in the modules and the duration of the
infiltration process in the final phase of infiltration. However, due to the complexity of the required
parameter calculation for that function, its practical application for engineering purposes is limited,
and thus it has not been presented in this study.
3.2

Calculation of hydraulic resistance

The models described by equations (10) and (11) allow to determine hydraulic resistance of a soil
layer during the infiltration of water in the phase of filling and emptying of infiltration modules.
The relations between the volume of suspension introduced with infiltrating water and the hydraulic
resistance determined in the phase of filling and emptying of module infiltration systems are presented
in Figures 8 and 9.

Suspended hydraulic resistance R, min

90
80
R=0.0201Vz+19.553

70
60
50
40
30
20
10
0

R=0.0082Vz+2.520

500

1000

1500

2000

Suspension volume Vz, cm

2500

3000

Suspended hydraulic resistance R, min

Figure 8. Comparison of hydraulic resistance of soil during the infiltration of water in the filling phase
on experimental site no.1 (white points) and experimental site no. 2 (black points)
350
R=0.0664Vz+124.46

300
250
200
150
100

R=0.0107Vz+4.113

50
0

500

1000

1500

2000

2500

3000

Suspension volume Vz, cm3

Figure 9. Comparison of hydraulic resistance of soil during the infiltration of water in the infiltration
phase on experimental site no.1 (white points) and experimental site no.2 (black points)
The values of resistance determined for infiltration module no.1, filled with a suspension of kaolin
clay and placed in sandy clay were not characterised by such rapid increase with the duration of the
experiment. In contrast, on experimental site no.2, filled by the same suspension but embedded in
cohesive clay, significant differences were noted between the resistance values calculated for the
filling and emptying phases. For both sites, the values of hydraulic resistance calculated for the
emptying phase were higher than those in the filling phase. The difference between these values
increased along with the increase in the volume of suspension introduced with infiltrating water,
which can be explained by a gradual clogging of the geotextile and of the superficial layer of soil.
Lower values of the calculated hydraulic resistance in the filling phase, as compared to those in the
emptying phase, are caused by capillary suction of the soil. As the top layer of soil becomes saturated
with water, hydraulic conditions stabilise, which in turn influences the changes in transient rate of
infiltration of water into the ground. This means that the calculated values of hydraulic resistance
depend on hydraulic and soil-related conditions. When comparing the values of substitute hydraulic
resistance obtained for experimental site no.1 with those measured on experimental site no.2, one can
easily note that these values are significantly lower for the module embedded in sandy clay. This leads
10

to the conclusion that the type of soil has a significant influence on the values of hydraulic resistance
occurring during the infiltration of water into the ground. Cohesive clay caused a rapid increase of the
resistance both in the filling and in the infiltration phase.

CONCLUSIONS

Conducted tests and analyses lead to the following conclusions:


1. Complex models of soil clogging have a limited practical value for the purposes of designing of
stormwater infiltration facilities. This results from a large variability of the conditions in which the
infiltration process takes place (variable amount of stormwater introduced to the facility in time,
varied concentration of suspended solids in stormwater, chemical and biological composition,
temperatures etc.) and from the complexity of the formal description of the models.
2. For engineering purposes, simple models of changes in the infiltration rate in the phase of filling
and emptying of infiltration modules, as described by equations (10) and (11) can be applied. The
practical value of the developed models for the designing of stormwater infiltration systems translates
mainly in the practical possibility to evaluate the degree of clogging and the need to take declogging
actions. This is particularly important for the purpose of designing underground stormwater infiltration
facilities, as all works related to the modernization of such systems, including declogging, are costly
and difficult.
3. Substitute hydraulic resistance R of the soil in the phase of filling of storage and infiltration
modules is significantly lower than that in the emptying phase. The difference between these values
increased along with the increase in the volume of suspension introduced with infiltrating water,
which can be explained by a gradual clogging of the geotextile and of the superficial layer of soil.
Lower values of the calculated hydraulic resistance in the filling phase are also caused by capillary
suction of the soil. This means that the calculated values of hydraulic resistance depend on hydraulic
and soil-related conditions.
4. The degree and rate of clogging of the ground in the course of long-term exploitation of stormwater
storage and infiltration facilities can be determined basing on the measured values with use of
substitute hydraulic resistance R in the emptying phase.

REFERENCES

Balades J-D., Legret M., Madiec H. (1995). Permeable pavements:Pollution management tools. Water
Science and Technology 32(1), 49-56
Bouwer H. (2002). Artificial recharge of groundwater: hydrogeology and engineering. Hydrogeology
Journal ,10,121-142
Bouwer H., Ludke J., Rice RC. (2001). Sealing pond bottoms with muddy water. Journal Ecological
Engineering 18 (2), 233-238
Bouwer H.(1969). Theory of seepage from open chanels. Advances in Hydrosciences, 5, , 121-170
Burszta-Adamiak E. (2005). Badania nad zastosowaniem geowknin do przeciwdziaania kolmatacji
w procesie infiltracji. (Research on nonwoven geotextile use in order to counteract clogging
process during infiltration into a ground). Ph.D Thesis, University of Technology, Wroclaw
Burszta-Adamiak E. omotowski J. (2005a). Prognozowanie zmian poziomu wody w czasie infiltracji
z podziemnych zbiornikw chonnych. (Forecasting water level changes during infiltration from
11

subsurface stormwater facilities). VII International scientifically-technical conference


Computer in environmental protection, Gniezno, Poland, 31-37
Burszta-Adamiak E., omotowski J. (2005b). Badania skadu granulometrycznego wd opadowych i
powierzchniowych z zastosowaniem granulometru laserowego. (Grain composition studies of
rainwater and the surface water by means of laser granulometer). Monograph of Environmental
Engineering Association PAN, II Environmental Engineering Congress, Vol.33 , 331-338
Dechesne M., Barraud S., Bardin J. P.(2004). Indicators for hydraulic and pollution retention
assessment of stormwater infiltration basins. Journal of Environmental Management, 77, 371380
Gautier A., Barraud S., Bardin J.P.1999. An approach to the characterization and modeling of
clogging in storm water infiltration facilities. Proc. the Eighth International Conference on
Urban Storm Drainage. August 30 . September 3, Sydney, Australia. Edited by IB Joliffe and JE
Ball. The Institution of Engineers Australia, The International Association for Hydraulic
Research, and The International Association on Water Quality, 1007-1015.
Geiger W., Dreiseitl H. (1999). Nowe sposoby odprowadzania wd deszczowych; Poradnik
retencjonowania i infiltracji wd deszczowych do gruntu na terenach zabudowanych. (New
ways to stormwater drainage. Guide of stormwater retention and infiltration into the ground in
urban areas) Oficyna Wydawnicza Projprzem-EKO, Bydgoszcz.
Hiscock K.M., Grischek T.(2002). Attenuation of groundwater pollution by bank filtration, Journal of
hydrology, 266 (3-4), 139-144
Hua, G. F.; Zhu, W.; Zhang, Y. H. (2010). A conceptual approach based on suspended solids to
estimate clogging time in constructed wetlands. Journal of Environmental Science and Health,
Part A: Toxic/Hazardous Substances and Environmental Engineering, 45 (12), 1519-1525.
Iwasaki T. (1937). Some notes on sand filtration. Journal of the American Water Works Association
(AWWA), 29, 1591-1602
Lloyd H. C. Chua, Melvin C. M. Leong, Edmond Y. M. Lo, Martin Reinhard, Alexander P. Robertson,
T. T. Lim, E. B. Shuy and S. K. Tan. (2009). Controlled field studies on soil aquifer treatment in
a constructed coastal sandfill. Water Science and Technology, 60(5), 1283-1293
Mallin M.A., Johnson V.L., Ensign S.H. (2009). Comparative impacts of stormwater runoff on water
quality of an urban, a suburban, and a rural stream. Environmental Monitoring and Assessment,
159 (1-4), 475-491.
Marla C. M., Lee-Hyung K. (2010). Long-Term Monitoring of Infiltration Trench for Nonpoint
Source Pollution Control Water, Air and Soil Pollution, 212 (1-4), 13-26
Mays D.C., Asce P.E.M. (2010).Contrasting Clogging in Granular Media Filters, Soils, and Dead-End
Membranes, Journal of Environmental Engineering, 136 ( 5), 475-480
Nivala J, Knowles P, Dotro G, Garca J, Wallace S. (2012). Clogging in subsurface-flow treatment
wetlands: measurement, modeling and management. Water Research, 46 (6), 1625-1640.
Oe. T., Hiroyuki K., Hiroyuki H., Katsumi O.(1996). Performance of membrane filtration system used
for water treatment, Desalination, 106 (1-3), 107-113
Pedretti D., Barahona-Palomo M., Bolster D., Fernndez-Garcia D., Sanchez-Vila X., Tartakovsky D.
M. (2012). Probabilistic analysis of maintenance and operation of artificial recharge
ponds.Advances in Water Resources;36, 23-35
Prez-Paricio A. (2001). Integrated modelling of clogging processes in artificial groundwater
recharge. Ph.D.Thesis, Technical University of Catalonia (UPC), Barcelona.
Raimbault G., Nadji D., Gautier C.(1999). Storwater infiltration and porous materia clogging. 8th
ICUSD, Sydney, 1016-1024
12

Ravi V., Williams J.R.(1998). Estimation of infiltration rate in vadose zone: Application of selected
mathematical models. Report EPA/600/R-97/128a, Vol.I.
Rinck-Pfeiffer S., Ragusa S., Sztajnbok P., Vandevelde T. (2000). Interrelationships between
biological, chemical, and physical processes as an analog to clogging in aquifer storage and
recovery (ASR) wells, Water Research, 34 (7), 2110-2118
Rupak A., Sarvanamuthu V., Jaya K., Ravi N. (2010).Urban stormwater quality and treatment. Korean
Journal of Chemical Engineering, 27 (5), 1343-1359
Seki K., Miyazaki T. (2001). A mathematical model for biological clogging of uniform porous media.
Water Resources Research, 37 (12), 2995-3000
Skolasiska K. (2006). Clogging microstructures in the vadose zone - laboratory and field studies.
Hydrogeology Journal. 14, 1005-1017.
Taylor S.W., Jaffe P.R.(1990). Biofilm growth and the related changes in the physical properties of a
porous medium.1.Experimental Investigation. Water Resources Research. 26(9), 2153-2159.
Taylor S.W., Milly P.C.D., Jaffe P.R.(1990). Biofilm growth and the related changes in the physical
properties of a porous medium.2. Permeability. Water Resources Research. 26(9), 2161-2169.
Tesaik I.(1980). Separcia suspendovanch astic pri prave vody. VEDA, vydavatelstvo Slovenskej
akadmie vied, Bratislava.
Thorolfsson S.T.(1998). A new direction in the urban runoff and pollution management in the city of
Bergen, Norway, Water Science and Technology. 38 (10), 123-130
Vandevivere P., Baveye P., Sanchez de Lozada D., Deleo P.(1995). Microbial clogging of saturated
soils and aquifer materials:evaluation of mathematical models.; Water Resources Research.
31(9), 2173-2180.
Vigneswaran S., Suazo R.B. (1987). A detailed investigation of physical and biological clogging
during artificial recharge. Water, Air and Soil Pollution. 35,119-140.
Williams J.R., Ouyang Y., Chen J.S., Ravi V.(1998). Estimation of infiltration rate in vadose zone:
Application of selected mathematical models. Report EPA/600/R-97/128b Vol. II..
Ying M.; Guangyao G.; Zailin H.; Shaoyuan F.; Dongyuan S. (2010). Modeling water infiltration in a
large layered soil column with a modified Green-Ampt model and HYDRUS. Computers and
electronics in agriculture, 71, 40-47
Zhuanxi L. , Tao W., Meirong G., Jialiang T.,Bo Z. (2012) Stormwater runoff pollution in a rural
township in the hilly area of the central Sichuan Basin, China. Journal of Mountain Science, 9
(1), 16-26

13

9th International Conference on Urban Drainage Modelling


Belgrade 2012

Modelling sediment bed aggradation in storm


water/combined sewers
Alberto Campisano1, Carlo Modica2
1
2

Dept. Civil and Env. Eng., University of Catania, V.le A. Doria 6, 95125 Catania, Italy, acampisa@dica.unict.it
Dept. Civil and Env. Eng., University of Catania, V.le A. Doria 6, 95125 Catania, Italy, cmodica@dica.unict.it

ABSTRACT
In this paper a numerical investigation to model bed aggradation due to the
entrance of large amounts of sediments in sewers is presented. A semi-coupled
modelling approach for uniform sediments based on 1D-shallow water De Saint
Venant-Exner equations was adopted to describe the temporal evolution of the bed
in the deposition and transport phenomena associated to aggradation process. Three
well-established bed-load transport formulas were used to evaluate the sediment
discharge over the aggrading bed. Simulations enabled the comparison of results
obtained with the selected formulas against experimental measurements from
literature. Results have shown the adopted model to be able to successfully predict
the evolution of sediment bed profiles during the aggradation experiments
depending on the used sediment transport formula.

KEYWORDS
Sewer system modelling, Sediment aggradation, Bed-load transport

INTRODUCTION

Storm water and combined sewer systems serving peri-urban catchments are often plagued by
problems associated to the entrance and deposition of large amounts of sediments. Deposition
phenomena in sewer pipes can assume troublesome sizes in basins where sewer inlets are not
adequately protected by specific sediment control measures. For instance, Ashley et al. (2004) report
about severe sewer restrictions due to introduction of solids from upstream artificial and natural
streams in Buenos Aires, Argentina and in catchments of So Paulo, Brazil.
Kolsky (1998) reports about rapid blockages in the storm water sewer system of Indore, India due to
the absence of installations to intercept street debris and to the widespread presence of wrong
connections in the collecting system.
Also, the level of sewer protection from the entrance of solids significantly affects the size of
sediments which are encountered in pipes. Open drains (typical of developing countries) have shown
to allow easier ingress in sewers of the coarser (sediment type A according to Crabtree (1989)

taxonomy) granular solids (Kolsky and Butler, 2000). However, large percentages of gravel in sewer
deposits are also found in sewers of developed countries (such as UK and France) where specific inlet
structures are installed (Crabtree et al., 1991; Bachoc, 1992).
Under previous conditions, sewer operation often shows the existing mismatch between catchment
sediment input rate and in-pipe sediment transport capacity. During rain events, in fact, incoming
material entering the sewer can result much more than sediment transported downstream by sewer
flows, thus determining deposition and bed aggradation phenomena in pipes.
Depending on geometrical and hydraulic characteristics of the pipes, aggradation phenomena can be
responsible for the increase in the overall hydraulic roughness of the pipe and, more seriously, for flow
capacity restrictions or total blockage of the sewer, having also a detrimental impact on flow quality
(Ashley et al. 2004).
Aggradation processes have been studied by laboratory experiments in flumes in the last decades. Soni
et al. (1980) firstly investigated the effects of deposition in flume streams due to sediment
overloading. Besides, the influence of the bed sediment sizes on the aggrading process was explored
by Yen et al. (1992).
At the end of the nineties, a comprehensive experimental campaign on the analysis of sediment bed
aggradation was performed by Seal et al. (1997) at the Saint Anthony Falls Laboratory (SAFL),
University of Minnesota. Experiments aimed at measuring bed profile evolution and selective
transport of sediment mixtures and were performed in a rectangular flume considering various values
of the width and longitudinal bottom slope. Sediments used in the experiments are representative of
the coarser portion of deposits in storm water/combined sewers.
Modelling approaches to the analysis of the aggradation process have mainly developed with regard to
fluvial engineering (Ferguson and Church, 2009). Literature shows the application of both models for
uniform sediments (Miglio et al., 2009) and for sediment mixtures, these last models enabling the
description of the relative mobility of grain sizes in the bed mixture (Wu et al., 2004). Also, complex
depth-averaged 2D models have also recently been used in order to describe in detail the behaviour of
rivers in terms of hydraulics and morphological evolution (Wu, 2004).
Different approaches have also been adopted in the literature as far as the numerical solution of the
equations governing flow and sediment is concerned. Specifically, uncoupled, semi-coupled and
coupled models have been developed to describe the mutual interactions between water flow and
sediment transport. Uncoupled models do not take into account the interactions between water flow
and sediment bed within each simulation time step (Ferreira and Leal, 1998). Despite the benefits due
to the ease of these models, numerical limits of their application to aggradation processes have been
highlighted, especially when super-critical flow conditions are considered (Cao et al. 2002). On the
other hand, coupled models allow for the simultaneous solution of water flow and sediment bed
equations. However, full coupling can result difficult to implement and it is accomplished if governing
equations are written in a not conservative form, leading to numerical errors in correspondence to high
gradients in water flow variables (Ferreira and Leal, 1998). Finally, with semi-coupled models (Creaco
et al., 2010) the initial solution obtained by a decoupled approach is iteratively refined in order to take
account of the mutual interactions between water flow and sediment transport within each time step.
The use of complex models to describe aggradation processes can arouse minor interest in sewer
systems since flows are basically one-dimensional and the limited range of sediment grain sizes
permits the use of models for uniform sediments. Moreover, the description of the evolution of bed
deposit profiles is often sufficient to evaluate malfunctioning due to pipe restrictions or blockages,
regardless of the analysis of specific phenomena associated to the relative mobility of grain sizes.
2

During the last decades, various models have been developed to successfully simulate sediment
deposition and transport in sewer systems (Ashley et al. 2004). However, most models in the literature
are based on numerical schemes which do not enable an accurate description of in-pipe phenomena
such as steep mobile bed fronts and mobile hydraulic jumps commonly occurring during rapidly
varied flow conditions in sewers.
Various sediment transport formulas adapted or specifically developed for sewer systems have been
extensively implemented in these models (Ackers et al., 1996; De Sutter et al., 2003) in order to
simulate transport and deposition in pipes during normal flows and during unsteady flow conditions
determined by storm water events. However, such models have not been tested specifically for the
simulation of aggradation processes in sewer systems.
To handle this question, an appropriate semi-coupled numerical model for uniform sediments was
specifically developed. The model is based on the shock-capturing solution of the 1D De Saint
Venant-Exner equations and results of application to aggradation processes are presented in this paper.
Three bed-load transport formulas commonly used for the simulation of sewer system sediment
transport were implemented within the model and tested against experimental measurements from
SAFL runs.

2
2.1

MODEL DESCRIPTION
Mathematical equations

The adopted numerical model for uniform sediments is based on the use of the fully-dynamic 1D De
Saint Venant-Exner equations, and it has been already validated in other cases concerning the
transport of sediments in sewer under unsteady flow conditions (Campisano and Modica, 2003;
Campisano et al., 2007). Equations are written in the following conservative vector form:

U FU

DU
t
x

(1)

with the vectors U, F and D given by:

A
U Q ;
As

Fh

;
FU V Q

Qs
1 p

Fh

DU
g A i J

qls
1 p

(2)

where x [m] and t [s] are the spatial and temporal independent variables respectively, A [m2] is the
wetted area corresponding to water level h [m], Q [m3/s] and V [m/s] are the water flow discharge and
flow velocity respectively, Fh [N] is the hydrostatic force over the cross section, Fw [N/m] is the force
due to channel width variations, [kg/m3] is the water density, g [m/s2] is the gravity acceleration, i [-]
and J [-] are the bed slope and the energy friction slope, respectively. Moreover p [-] is the sediment
porosity, As [m2] is the sediment cross section area, Qs [m3/s] is the sediment discharge and qls [m2/s] is
the sediment input for unit channel length.

The energy friction slope is evaluated by the Strickler equation:

Q2
k eq2 A2 R 4 / 3

(3)

where R [m] is the hydraulic radius and keq [m1/3/s] is the Strickler composite roughness coefficient
evaluated by the Einstein formula (Einstein and Banks, 1950).
Vector equation (1) enables description of uniform sediment transport by evaluating the vector U of
the unknown variables as a function of x and t. The resolution of equation (1) with vectors given by
equations (2) entails using a sediment transport formula for the estimation of the sediment discharge
Qs as a function of the flow and sediment properties. To this end, the following three transport
relationships were selected for the simulation of the SAFL experiments:

qs

8 cr

32

s 1 g d 50

(Meyer Peter-Mller, 1948)

(4)

where qs [m2/s] is the sediment volumetric discharge per unit width; s [-] is the relative density of the
sediment; d50 [m] is the sediment mean diameter; b s g d 50 is the Shields
dimensionless shear stress in which b [N/m2] is the bottom shear stress; s [kg/m3] is the sediment
density; cr = 0.047 is the dimensionless critical shear stress. The range of application provided by the
author is 0.4 mm < d50 < 30 mm.

qs

s 1 g d 503

10.4

1.5


1 cr

2.5

(Suszka, 1991)

(5)

with cr assumed equal to 0.045. The range of application is 3 mm < d50 < 44 mm.
n

1n

d V B R
qs q Ggr 50 * se
R V A

(Ackers, 1984)

(6)

where the effective bed width Bse [m]= min0.2 0.33 hs D 0.01Bs ,0.5Bs ; hs [m] is the height of
the sediment bed; D is the pipe size; R is the hydraulic radius; V* [m/s] is the flow shear velocity. The
range of application is 0.04 mm < d50 < 4.94 mm (Ghani, 1993).
The non-dimensional transport parameter:

Fgr

G gr H
1
Agr

is to be calculated as function of the mobility number Fgr:

Fgr

V*

V 1n

g s 1 d
12 R

32 log10
d

50

1n

with:

n 1.0 0.56 log10 Dgr ;


4

Agr 0.14

m 1.34

0.23
Dgr

9.66
;
Dgr

log10 H 2.86log10 Dgr log10 Dgr 3.53


2

where Dgr [-] is the dimensionless grain size to be evaluated as:

g s 1
Dgr d 50
2

13

with [m2/s] being the kinematic viscosity of the water.


For course sediment (Dgr>60) it is assumed: n = 0, Agr= 0.17, m = 1.5, H = 0.025.
Formulas (4-6) were originally derived with specific regard to river sediment transport and their
applicability in storm water sewer systems for the evaluation of the sediment discharge is still a
discussed question. For instance, De Sutter et al. (2003) indicated that the direct transfer to sewers of
the Ackers formula (Ackers, 1984) may result as inappropriate since it can provide a significant
over/under estimation of sediment discharge for finer/coarser sediment respectively. However,
selected formulas have been extensively used in literature over the past decades to model sediment
transport in sewers in many processes, also under rapidly varied flow conditions (Mark et al., 1995;
Ashley and Verbanck, 1996; Lin and Le Guennec, 1996).
The application of previous sediment transport formulas requires the bottom shear stress b due to the
water flow over the sediment bed to be evaluated. This can be accomplished using the following
relationship:

k eq
b g R J
ks

3/ 2

(7)

where the sediment roughness coefficient ks [m1/3/s] is evaluated as a function of the mixture grain size
d90 according to Meyer-Peter and Mller:

ks

26

(8)

16

d 90

In agreement with experimental protocols used at SAFL (see next section), no specific formulas were
taken into account for separately modelling the suspended-load transport. In addition, to maintain the
model as simple as possible, no loading law equations for the lag between sediment transport capacity
and sediment discharge have been considered (hypothesis of immediate adaptation), the total sediment
transport being modelled by bed-load formulas (Armanini, 1999).
2.2

Numerical scheme

The finite difference scheme of MacCormack is used for the numerical integration of equation (1)
(MacCormack, 1971). This scheme is a well-known shock-capturing one, i.e. it is able to describe
discontinuities such as mobile hydraulic jumps and steep mobile bed fronts and then particularly
suitable for the analysis of aggradation processes. The scheme is two-step predictor-corrector and has
5

a second order accuracy in both space and time. A third step based on the Total Variation Diminishing
(TVD) theory (Garcia-Navarro et al., 1992) is applied to the hydraulic variables in order to reduce the
numerical oscillations due to high gradients in water levels and sediment bed heights.
The adopted conservative-law form of equation (1) does not provide the full coupling of flow and
sediment equations. Then, the presented model belongs to the category of semi-coupled schemes.
However, the two-step nature of the MacCormack scheme makes it possible to enforce the coupling
between hydraulic and sediment variables, enabling the solution of the De Saint Venant-Exner
equations at the same time step. This condition provides nearly identical performances in comparison
to coupled models (Ferreira and Leal, 1998).
Boundary conditions for the variables A, Q and As are assigned to the upstream and downstream limits
of the integration domain. In particular, for sub-critical flows, one hydraulic condition and one
condition for the sediment are prescribed at the upstream end; a further hydraulic condition is
prescribed at the downstream end. Instead, for super-critical flows, two hydraulic conditions are
assigned upstream; moreover, the condition for the sediment is assigned at the downstream end.
Cao et al. (2002) have shown that the use of semi-coupled models can entail conditions of
mathematically ill posed modelling in super-critical flows when a sediment input from the upstream
boundary is considered. Such a condition was reported by various authors to lead the simulation to
fail, as aggradation process starts at the upstream boundary.
Then, a simple solution to model the sediment input was considered in order to overcome encountered
numerical problems. Specifically, the sediment input into the flume was simulated as lateral sediment
inflow qls for unit of flume length. The feeding was assumed to occur in the flume over four
consecutive x, enabling both a more realistic description of the experimental feeding conditions and
the reduction of the numerical oscillations at the feed point.
A limitation of the numerical scheme concerns the simulation of the hydraulic jump. Such a jump is
allowed by the scheme to develop over two successive sections of the spatial domain, and can lead to
very steep slope of the deposit front.

3.1

SAFL EXPERIMENTS AND COMPARISON WITH NUMERICAL


RESULTS
Simulation framework

The experimental work performed at the Saint Anthony Falls Laboratory, University of Minnesota
consisted of six experiments (runs) concerning the processes of aggradation and selective transport of
sediment mixtures (Seal et al., 1997).
Main results presented in this paper concern the application of the numerical model to run 3, for which
more detailed measurements are available. In particular, the numerical results obtained with the three
selected sediment transport formulas were compared against experimental data to evaluate the
behaviour of the model in reproducing time-varying mobile bed profiles.
The experiment was performed in a 55 m long laboratory flume with a 0.305 m wide rectangular cross
section with a PVC membrane covered bottom and sidewalls. The flume slope was set to 0.2%.

-5

Entering flow rate (Q = 0.049 m3/s) and sediment feeding (Qs = 1.193 10 m3/s) were maintained
constant during the entire experiment. Also the downstream water level hv was kept constantly equal to
0.5 m.
The material used in the experiments was made up of a sediment mixture (gravel and sand) with a
density s = 2650 kg/m3. In particular, the used sediments had median grain size d 50 4.63 mm and

standard deviation 5.57 mm. The characteristics of the sediments adopted in the experiments are
similar to those featured by the coarser part of sewer deposits. During the initial part of the
aggradation process, the authors observed that the sandy portion of the mixture was transported as
suspended load, ending up depositing in the downstream part of the flume without taking a role in the
aggradation process. Accordingly, particles with grain size lower than 2 mm were originally excluded
by the experimental results (Seal et al., 1997) in order to take account of the bed-load transport
exclusively. Then, the modified grain size distribution curve was finally characterized by d 50 12.8
mm, d 90 36 mm and 3.62 mm.
Measurements performed during the experiment concerned the sediment bed elevations at several
flume sections (bed profiles) and at different time instants. Obtained measurements are reported with
dots in Figure 1. Full documentation can be found in Seal et al. (1997) with a complete listing of the
data.
According to selected sediment transport formulas, the characteristic grain sizes d50 was chosen to
represent the sediment material in the simulations. Such a d50 value falls within the application ranges
of Meyer Peter-Mller and Suszka formulas and results a bit larger than Ackerss formula
experimental range. Moreover, the sediment roughness coefficient (equal to 45 m1/3/s) was calculated
as a function of d90 using equation (8). The roughness coefficient for the flume PVC bottom and
sidewalls was set at 110 m1/3/s.
All the simulations were performed using a spatial step x = 0.25 m. The spatial domain scheme was
made of a total of 221 sections. The sediment input was assumed to occur as lateral inflow qls at
sections 21-24 (progressive 0, 0.25, 0.50, 0.75). The time step t was derived in correspondence to a
Courant stability number equal to 0.8.
As for the boundary conditions, according to the spatial numerical scheme, the experimental flow Q(t)
and sediment discharge Qs(t)=0 were prescribed at the flume upstream end (section 0, progressive -5
m) in case of sub-critical flows; the experimental water level hv(t) was fixed in correspondence to the
flume downstream end (section 221, progressive 50 m). Instead, in the case of super-critical flows, the
experimental flow Q(t) and the critical water level hcr(Q) were prescribed upstream; the sediment
condition As(t)=0 was assigned at the downstream end.
The initial condition of simulations was that of a bare bottom with the water flow featuring backwater
effects due to the presence of the fixed weir close to the downstream end.
Model simulations were also carried out for the other five SAFL runs characterised by different flume
slope and width. Results of such simulations are synthetically reported in the paper in terms of root
mean square errors (RMSE) between experimental and numerical bed and water surface elevations.
3.2

Results of the simulations and discussion

The graphs of Figure 1 relate to run 3 and show the comparison between simulated (solid lines) and
experimental (dots) elevations for the deposit bed (at 6, 8, 14, 26, 32, 46 and 64 hours after the
beginning of the experiment) and for the water surface at the end of the experiment (64 hours).

bed and water surface elevations (m)

1,2

a)

Meyer-Peter and Mller (1948)

1,0
0,8

water surface at 64h

0,6
0,4
0,2

bed at 6h

8h

14h

26h

32h

46h

64h

0,0
0

10

15

20

25

30

35

40

45

50

x (m)

bed and water surface elevations (m)

1,2

b)

1,0

Suszka (1991)

0,8
water surface at 64h

0,6
0,4
0,2

bed at 6h

8h

14h

26h

32h

46h

64h

0,0
0

10

15

20

25

30

35

40

45

50

x (m)

bed and water surface elevations (m)

1,2

c)

Ackers (1984)

1,0
0,8
water surface at 64h

0,6
0,4
0,2

8h

bed at 6h

14h

26h 32h

46h

64h

0,0
0

10

15

20

25

30

35

40

45

50

x (m)

Figure 1. Model simulation of SAFL run 3. Comparison of sediment bed and water surface profiles
obtained using a) Meyer-Peter and Mller, b) Suszka, c) Ackers sediment transport formulas.

Plotted results enable a quick qualitative evaluation of the differences obtained with the various
sediment transport formulas. In particular, looking at bed profiles at successive times, differences are
encountered for aggradation slope. Clearly, Meyer-Peter and Mller formula shows to underestimate
the bed aggradation slope with reduced bed sediment elevations in the upstream segment of the flume.
Consequently, in the same segment, water surface elevations result underestimated too. On the
contrary, Ackers formula provides increased bed deposit and water surface profile slopes with
overestimation of upstream deposition. Intermediate results were obtained with the formula by Suszka.
As expected, the model provides downstream front of the bed steeper than the experimental one,
regardless the used sediment transport relationship. Instead, models that are not shock-capturing can
provide bed fronts flatter than the experimental one (Wu, 2004). Furthermore, the introduction of an
adaptation length parameter can allow an improved description of the bed front shape (Wu et al.,
2004). However, the introduction of such a parameter requires specific calibration and determines
increased complexity of the model.
Also, simulation results presented in Figure 1 show that the model reproduces the hydraulic jump
correctly. In particular, Suszka formula appears to provide estimation of the hydraulic jump position
very close to that observed in the experiment (about 37 m downstream of the feeding point).

cum. sediment volume through section 5 m (m3)

An evaluation of the sediment transport rate obtained with the three formulas is reported in Figure 2.
The figure shows cumulated volumes of sediment passed through progressive section 5 m downstream
the feeding point. Basically, according to the constant sediment input and water inflow into the flume,
the figure shows quasi-linear trends of sediment volumes as time increases. In agreement with De
Sutter et al. (2003), Ackers formula is observed to slightly underestimate the transport of sediment.
15

Meyer Peter and Mller

12

Suszka

Ackers
Experiment
s

0
0

10

20

30

40

50

60

70

80

90

100

time (h)

Figure 2. Cumulated volumes of sediment passed through progressive section 5 m from the feeding
point. Comparison between experimental data and numerical results.
In Table 1 calculated RMSE values of bed elevations (average RMSE at different times) and water
surface elevations (at the end of the experiment) are summarised for the three used formulas and for
the six SAFL runs.

Results in the table show relatively low values of RMSE for all the simulations confirming the adopted
model to be able to successfully reproduce the experimental evolution of sediment bed profiles and
water surface elevation. In particular, small differences among the tested bed-load transport formulas
were obtained for both downstream sub-critical and super-critical flows.

Table 1. Values of RMSE for bed elevations (BE) and water elevations (WE).

Formula

Run 1
BE WE

Run 2
BE WE

Run 3
BE WE

Run 4
BE WE

Run 5
BE WE

Run 6
BE WE

Meyer-Peter and
Mller (1948)

0.042

0.037

0.033

0.037

0.061

0.053

0.015

0.013

0.015

0.016

0.021

Suszka (1991)

0.020

0.026

0.028

0.040

0.029

0.036

0.016

0.014

0.015

0.012

0.012

Ackers (1984)

0.012

0.022

0.034

0.040

0.058

0.065

0.012

0.022

0.160

0.018

0.024

CONCLUSIONS

A numerical investigation to model bed aggradation in stormwater/combined sewers resulting from the
entrance of large amounts of sediments in pipes is presented in the paper. A shock-capturing model
solving the 1D-De Saint Venant-Exner equations was developed and applied to aggradation
experiments carried out at Saint Anthony Falls Laboratory, University of Minnesota.
Three sediment transport formulas commonly used for the simulation of sediment transport in sewers
were implemented in the model. Numerical results were compared against experimental results in
terms of temporal evolution of the deposit shape, front advancing and transported volumes of
sediment.
The adopted numerical scheme based on a conservative-law form of equations allowed to properly
simulate the presence of mobile hydraulic jumps over the deposit and to enforce the coupling between
water flow and sediment transport relationships. A simple solution to model the sediment input as
lateral sediment inflow was adopted to overcome numerical problems which were reported in the
literature to lead to erroneous sediment transport rate evaluation under super-critical flow conditions.
Despite the simplicity of the approach chosen, results have shown the adopted model to be able to
describe successfully the evolution of sediment bed profiles during the aggradation experiments
depending on the used sediment transport relationship. Root mean square errors of simulated and
experimental results were evaluated for both sediment bed and water surface elevations using the three
formulas.
According to the size of the sediment mixture used for SAFL experiments, the validity of the obtained
results remains limited to the coarser portion of sediments forming the deposits of storm
water/combined sewers.

10

REFERENCES

Ackers, P. (1984). Sediment transport in sewers and the design implications. Proc. of the Int. Conf.
on Planning, Construction, Maintenance and Operation of Sewerage Systems. Reading,
England, BHRA (Cranfield), pp. 215-230.
Ackers, J.C., Butler, D., and May, R.W.P. (1996). Design of sewers to control sediment
problems. Rep. No. 141, Construction Industry Research and Information Association
(CIRIA), London.
Armanini, A. (1999). Principi di Idraulica fluviale. Editoriale BIOS, Cosenza.
Ashley, R.M., Bertrand-Krajewski, J.L., Hvitved-Jacobsen, T., and Verbank, M. (2004). Solids in
Sewers. Scientific and Technical Report No. 14, IWA Publishing, 360 pp.
Ashley, R.M. and Verbanck, M.A. (1996). Mechanics of sewer sediment erosion and transport,
Journal of Hydraulic Research, 34 (6), 753769.
Bachoc, A. (1992). Location and general characteristics of sediment deposits into man-entry
combined sewers, Wat. Sci. Tech., 25(8), 47-55.
Campisano, A., Creaco, E., and Modica, C. (2007). Dimensionless Approach for the Design of
Flushing Gates in Sewer Channels Journal of Hydraulic Engineering, 133(8), 964-972.
Campisano, A., and Modica, C. (2003). Flow velocities and shear stresses during flushing operations
in sewer collectors Water Science and Technology, 47(4), 123-128.
Cao, Z., Day, R., and Egashira S. (2002). Coupled and Decoupled Numerical Modelling of Flow and
Morphological Evolution in Alluvial Rivers. Journal of Hydraulic Engineering, ASCE, 128(3),
306-321.
Crabtree, R.W. (1989). Sediment in sewers, Journal IWEM, 3(6), 569-578.
Crabtree, R.W., Ashley, R.M., and Saul, A.J. (1991). Review of research into sediments in sewers
and ancillary structures, Foundation for Water Research, Report No.FR 0205.
Creaco, E., Campisano, A., Khe, A., Modica, C., and Russo, G. (2010). Head Reconstruction Method
to Balance Flux and Source Terms in Shallow Water Equations Journal of Engineering
Mechanics, ASCE, 136(4), 517-523.
De Sutter, R., Rushforth, P., Tait, S., Huygens, M., Verhoeven, R., and Saul, A. (2003). Validation of
Existing Bed Load Transport Formulas Using In-Sewer Sediment, Journal of Hydraulic
Engineering, ASCE, 129(4), 325-333.
Einstein, H.A., and Banks, R.B. (1950). Fluid resistance of composite roughness. Transactions of
the American Geophysical Union, 31, 603-610.
Ferguson, R.I., and Church, M. (2009). A critical perspective on 1-D modelling of river processes:
Gravel load and aggradation in lower Fraser River, Water Resources Research, 45, W11424, 115.
Ferreira, R.M.L., and Leal, J.G.A.B. (1998). 1D Mathematical modelling of the instantaneous DamBreak Flood Wave Over Mobile Bed: Application of TVD and Flux-Splitting Schemes.
Proceedings of Union Europeenne, CADAM meeting, Munchen, Germany, 215227.
Garcia-Navarro, P., Alcrudo, F., and Saviron, J.M. (1992). 1-D Open-Channel Flow Simulation
Using TVD-MacCormack Scheme. Journal of Hydraulic Engineering, ASCE, 118(10), 13591372.
Ghani A.A. (1993). Sediment Transport in Sewers. PhD Thesis, University of Newcastle Upon Tyne,
November 1993.

11

Kolsky, P. (1998). Storm drainage: an engineering guide to the low-cost evaluation of system
performance IT Publisher, London.
Kolsky, P. and Butler, D. (2000). Solids size distribution and transport capacity in an Indian drain,
Urban Water, 2(2000), 357-362.
Lin, H. and Le Guennec, B. (1996). Sediment transport modelling in combined sewer, Water
Science and Technology, 33(9), 6167.
MacCormack, R.W. (1971). Numerical Solution of the Interaction of a Shock Wave with a Laminar
Boundary Layer. Proc.of the 2nd International Conference on Numerical Method in Fluid
Dynamics, Berlin, Germany, 151-163.
Mark, O., Appelgren, C. and Larsen, T. (1995). Principles and approaches for numerical modeling in
sewers, J. Environ. Polym. Degrad., 31(7), 107-115.
Meyer-Peter, E., and Mller, R. (1948). Formulas for Bed-Load Transport. Proc. of IAHR, Second
Meeting, Stockholm.
Miglio, A., Gaudio, R., and Calomino, F. (2009). Mobile-bed aggradation and degradation in a
narrow flume: Laboratory experiments and numerical simulations. Journal of Hydroenvironment Research, 3(2009), 9-19.
Seal, R., Paola, C., Parker, G., Southard, J.B., and Wilcock, P.R. (1997). Experiments on
Downstream Fining of Gravel: I. Narrow Channel Runs. Journal of Hydraulic Engineering,
ASCE, 123(10), 874-884.
Soni, J.P., Garde, R.J., and Ranga Raju, K.G. (1980). Aggradation in streams due to overloading.
Journal of Hydraulic Division, ASCE, 106(HY1), 117-132.
Suszka, L. (1991). Modification of transport rate formula for steep channels. Lecture Notes in Earth
Sciences, Vol. 37, 59-70, Springer-Verlag, Berlin, Germany.
Yen, C., Chang, S., and Lee, H.Y. (1992). Aggradation-Degradation Process in Alluvial Channels.
Journal of Hydraulic Engineering, ASCE, 118(12), 1651-1669.
Wu, W. (2004). Depth-Averaged Two-Dimensional Numerical Modeling of Unsteady Flow and
Nonuniform Sediment Transport in Open Channels, Journal of Hydraulic Engineering, ASCE,
130(10), 1013-1024.
Wu, W., Vieira, D.A., and Wang, S.S.Y. (2004). One-dimensional Numerical Model for Nonuniform
Sediment Transport under Unsteady Flows in channel Networks, Journal of Hydraulic
Engineering, ASCE, 130(9), 914-923.

12

You might also like