Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Contaminant-Removal from Wastewater in

a Secondary Sedimentation Tank


Fluid Flow in Process Units

Otso Seppälä
Alina Artman

Introduction
Secondary sedimentation tanks (SST) are mainstream technology in wastewater treatment. Typical
configuration is to have first primary sedimentation tank, then aeration tank and then SST. Figure 1
shows the configuration concerning modelling (without primary settler). Overflow of SST goes to
receiving water body or for tertiary treatment. Underflow or the activated sludge returns to process so
that biological reactor would have higher concentration of biomass to purify wastewater. The process is
called activated sludge, because the sludge is sort of “trained” to settle. All the sludge that does not
settle, discharges. There are of course SSTs in many applications other than CASP systems, like in mine
industry to settle hydroxide sludge or in potable water treatment prior to sand filtration and other
process units.

Even though much effort has been focused on modelling the biological part, SSTs do not have well-
established models. For example, International Water Association’s (IWA) activated sludge model (ASM)
1/2/3 are now a practice in industry, but they are mainly focused on the biological part (Ploz et al.,
2012). However, SSTs were one of the first units to be modelled first in 1D already in the 80’s and as
computing capacity increased 2D and even 3D CFD models have been implemented in 21 st century
(Samstag et al., 2016). The problem when comparing the models is that they do not account all the
phenomena and it is still unclear which of the phenomena govern sedimentation.

Many SSTs are sensitive to flow patterns, particle size distribution and are coupled with the biological
part. A complete CFD modelling would involve many physical, chemical and biological phenomena like,
turbulence, flocculation, sludge rheology, hydrodynamics, temperature and biokinetics. Main practical
problem with SST modeling nowadays is not lack of computation capacity or model accuracy, but the
measurement quality the models obtain their data. Even high-end CFD models have some empirical
functions, that would need specialized data for calibration. (Ploz et al., 2012)

CFD is superior to other models when complex geometries are involved. For example, Guo et al. (2017)
used COMSOL to simulate sedimentation enhanced with lamellas and an adjustable baffle in a small
tank. They used finite element method, 2D geometry and Euler-Euler approach to describe fluid flow
and particle motion. Turbulence model was basic Reynolds Average Navier (RANS, or k-). Once
simulation was done, they tracked 9000 particles using a model that accounts particle’s dynamic
behavior. The scientists were happy with the results after model validation and recommend CFD for
design of SSTs.

Ramin et al. (2014) developed a new settling velocity model for SST. The idea was to account hindered,
transient and compression settling velocities separately and develop an easy method to gather required
data for model calibration on sludge rheology. Finally, a 2D axisymmetric model was validated with a
full-scale settling tank. The model performed very well except in case of compression velocity that
affected strongly sludge blanket height. However, the model produced considerably better results in
case of wet-weather flow than previously used models. The research group strongly recommends use of
the model and measurement methods related to it.

Goal of the Simulation


This quick modelling work follows tutorial of COMSOL. First, the model is built and computed to obtain
steady state. Then inflow is increased temporarily and particle fraction in dispersion is decreased to
study wet weather flow. Finally, we suggest some modifications to geometry to mitigate effect of wet
weather flow in dispersion phase and model the modification to see if it has any effect.

Figure 1 Modelling approach on CASP (Patziger et al., 2012)


Model Setup
This study was conducted using an axisymmetric 2D model. The geometry and the dimensions of the
clarifier model are illustrated in figure X. The inlet is located in the middle of the tank whereas the
sludge outlet is in the bottom of the tank and the peripheral outlet extends across the outer edges. The
inlet flow consists of liquid mixed with circular, uniformly sized solid particles. The separation process of
this multiphase flow relies on gravity to separate the solid flocs to the bottom of the clarifier where they
are drained through the sludge outlet. Comparably, the peripheral outlet on the top is used to remove
excess liquid.

The multiphase flow model uses momentum equation (1-2) and mass conservation of each phase (3) to
describe the motion of the fluid and particles.

𝜌𝒖𝑡 + 𝜌(𝒖 ∙ ∇)𝒖 = −∇p − ∇ ∙ (𝜌𝑐𝑑 (1 − 𝑐𝑑 )𝒖𝑠𝑙𝑖𝑝 𝒖𝑠𝑙𝑖𝑝 ) + ∇ ∙ τ𝐺𝑚 + 𝜌𝒈 (1)


(𝜌𝑐 − 𝜌𝑑 )[∇ ∙ (ϕ𝑑 (1 − 𝑐𝑑 )𝒖𝑠𝑙𝑖𝑝 − 𝐷𝑚𝑑 ∇ϕ𝑑 )] + 𝜌𝑐 (∇ ∙ 𝐮) = 0 (2)
𝜕 (3)
(ϕ 𝜌 ) + ∇ ∙ (ϕ𝑑 𝜌𝑑 𝒖𝑑 ) = 0
𝜕𝑡 𝑑 𝑑
Here u, 𝜌, p, c, g, 𝜙, and τGm denotes mixture velocity, density, pressure, mass fraction, gravity vector,
volume fraction, and sum of viscous and turbulent stress respectively. The sub-index slip refers to the
relative velocity between the phases while sub-indices d and c refer to dispersed and continuous phase
respectively. Furthermore, the model uses Hadamard-Rybczynski drag law for the solid particles and it is

(𝜌 − 𝜌𝑑 )𝑑𝑑2 (4)
𝒖𝑠𝑙𝑖𝑝 = − ∇p,
18𝜌𝜂
applied to the slip velocity equation (4)

where 𝑑𝑑 and 𝜂 denote for particle diameter and fluid viscosity respectively. The turbulence was

𝜂𝑇 (5)
𝐷𝑚𝑑 = −
𝜌𝜎𝑇
described using the k-ε model to obtain the dispersion coefficient 𝐷𝑚𝑑 (5) in equation (2).

The sub-index T denotes turbulent parameters and 𝜎 is Schmidt number.

The inlet and sludge outlet velocities were fixed as well as the inlet solid phase concentration. The
peripheral outlet was set to constant pressure to avoid numerical instabilities. The symmetry boundary
condition was utilized for the axial symmetry line and the water surface while all the solid walls used a
logarithmic wall function. All the walls are set to no slip and water surface to slip. The water surface is
assumed to be horizontal because the inlet flow is low. Initially, velocity and solid phase volume fraction
is zero (the clarifier is filled with treated water).

Mesh is extra fine and around walls and on the area of high shear rate close to inlet it is extremely fine.
Total simulation time is 12 h, and initial time step is 0.01 s for the first second, then for the next 9 s it is 1
s, then 100 s for the next 10 steps and then the rest of the time it is 1800 s.

The first simulation is done with above configuration.

The second simulation shows effect of a baffle on the fluid dynamics and solid removal. Baffle are
normal practice to reduce turbulence and help solids to settle. However, they are not very easy to build
because clarifiers have normally a rotating scraper bridge. If they are designed incorrectly or there are
fundamental problems with the whole clarifier geometry they may not help at all and just increase
capital costs.

For this simulation, the mesh needed to be set only finer because there were problems to run it.

The third simulation shows effect of temporarily increasing flow that could be caused by a precipitation
in municipal WWTP for example. However, solids concentration stays roughly the same, because
municipal systems have large aeration basin before the clarifiers and sludge recirculation rate is about
200 %. Longer rain events could decrease solids concentration at the inlet of clarifier, but only
temporarily. Solids concentration of the clarifier inflow is controlled by removing sludge from clarifier
underflow rather than recirculating it back to aeration tank. In case of long rain, the sludge removal
could be just decreased or stopped totally to ensure stable solid concentration.

This simulation uses the solution of the first model as initial conditions and same geometry. Velocity at
the inlet increases to 2.5 m/s for two hours and then goes back. Velocity in the bottom outlet is also
doubled to 0.5 m/s for 2 hours and then it returns back to initial value.
Results and Discussion
Normal Geometry and Flow
Figure 2 shows a slice of the clarifier after 12 hours when it is in steady state. Velocity streamlines shows
the velocity field of the mixture and colour concentration of the solids in dispersion. As expected, high
velocity at the inlet mixes the dispersion phase so that the concentration is uniform and average. The
heaviest particles settle very fast and form a point concentration close to inlet. As the model has
turbulence, the concentration gradient is smooth where velocity is large. This can be seen in the first
cycle in the velocity field because it is almost uniformly yellow. The other cycle, however, has high
concentration in the bottom and low on the surface. This is means the design is good for this kind of
flow and dispersion phase concentration. The bottom has largest concentration, which is also a good
thing. In general, turbulence has negative effect on particle settling and clarifier designers try to mitigate
its effects. Clarifier modelling without adding turbulence in the model would be waste of time when
assessing solid removal efficiency.

According to scale in legend, the initial guess about range of the concentration of dispersion phase was
good and the model works as it should. If the whole slice would be abnormally red or blue the model
would not find a solution. Maximum packing density is set to 0.62 %, so the model seems to work well.

Figure 2 Streamlines of the mixture velocity and the solid phase concentration in kg/m3 after 12 hours
Figure 3 shows the swept volume of the clarifier with streamlines for both continuous (water) and
dispersed (solids) phase. There are now two distinctive recirculation zones. The first is created by the
fast flow of the inlet and locates next to the inlet. The second is created by falling particles and when
water flows to the peripheral outlet. Solids follow the lower edge of this zone but most of them fall to
the bottom edge and slowly move to the central outlet. Some solids (light and fastest) still follow the
flow to the peripheral outlet. Some baffle could possible block these or the clarifier should be even
larger.

Figure 3 Cut through volume of the clarifier with streamlines for the dispersed (blue) and continuous (white) velocity fields.

Figure 4 shows mass flows of dispersed phase at inlet, central outlet and peripheral outlet. 12 h seems
to be good simulation time since the conditions are getting stable in the end of the simulation. Solids
reduction in this clarifier is:

𝑚𝑖𝑛𝑙𝑒𝑡 − 𝑚𝑏𝑜𝑡𝑡𝑜𝑚 𝑜𝑢𝑡𝑙𝑒𝑡 0.52 − 0.42


= = 82 %
𝑚𝑖𝑛𝑙𝑒𝑡 0.52
This is good reduction for a secondary clarifier.

Figure 4 Mass flow of dispersed phase during the simulation. Blue is at the inlet. Yellow at the central outlet and violet at
peripheral outlet.

Geometry with a Baffle


Figure 5 shows how baffle changes flow pattern and dispersion phase concentrations. Seems that the
baffle works extremely well, but actually the best quality water is just behind the baffle, not on the edge
outlet. Now there are 3 distinctive recirculation zones on the surface and one on the bottom for solids.
The largest should be positioned a bit closer to center so that the water on the edge would be cleaner.
Figure 5 2D with baffle

Figure 6 has streamlines for both continuous and dispersed phase and volume fraction of dispersed
phase. Seems that the baffle conducts effectively solids to the bottom, but the sludge recirculation zone
in the bottom can be a threat as stated before.

Based on figure 7, solids removal efficiency is a whopping 96 %. However, denitrifying and other
biological reactions producing gases can deteriorate clarifiers ability to separate solids. The blue line for
peripheral mass flow rate shows that the flow stabilizes rapidly. This means the design allows lower
startup time and is not so prone to external disturbances, like increase in flow rate.

𝑚𝑖𝑛𝑙𝑒𝑡 − 𝑚𝑏𝑜𝑡𝑡𝑜𝑚 𝑜𝑢𝑡𝑙𝑒𝑡 0.52 − 0.49


= = 96 %
𝑚𝑖𝑛𝑙𝑒𝑡 0.52
Figure 6 3D swept out model for streamlines for both phases and volume fraction concentration.

Figure 7 Mass flow rates in outlets and inlets of dispersed phase.

Increase in Flow
The simulation was run with normal parameters and normal geometry for 10 hours and then inlet
velocity was doubled for last 2 hours. Figure 8 shows streamlines of continuous phase velocity field and
concentration of dispersed phase at t = 12 h. The flow in clarifier is dominated by one big recirculation
zone and turbulence has smoothed the dispersed phase all over the clarifier. Seems that some of the
heaviest particles are trapped next to inlet and do not sink to the bottom.
Figure 8 Streamlines of the mixture velocity and the solid phase concentration in kg/m3 after 12 hours with increased inlet
velocity

Figure 9 has streamlines for continuous and dispersed phase. Seems that solids just flow on the surface
and most of them exit the clarifier from the peripheral outlet, not the bottom. This is not good.

Figure 10 shows mass flow rates of the outlets. As the Inlet velocity increases peripheral mass flow rate
almost equals the bottom flow rate. Some accumulated solids also start to flow out of the clarifier due
to increasing flow. This can be seen from the peak of the mass flow rates just after the increase in
velocity. After peak mass flow rate there is a decline to a level where the inlet equals bottom and
peripheral outlet mass flow rates.
Figure 9 3D swept out model for streamlines for both phases and volume fraction concentration when velocity has been
increased

Figure 10 Mass flow rates in outlets and inlets of dispersed phase.


Model Limitations
A real circular clarifier has typically a rotating scraper to help sludge flowing to the bottom outlet and
not accumulating in the clarifier. The incline of the bottom could also be increased, but the larger the
clarifier the more expensive it is because the clarifier would get very deep in the bottom. Very large
(diameter over 50 meter clarifiers have a suction scraper to enhance sludge removal even more. All
these rotating devices create flow patterns, even though they move at a very low velocity and only in
the bottom. The model in this study does not take this kind of devices in account.

The model does not account any flows originating from external forces, like wind or temperature
change. If weather warms up quickly, sun shines brightly, retention time is long in the clarifier and
original temperature is 14 (which is normally is), water on the surface can heat up significantly before it
exits the clarifier. However, there are nowhere near such large flows caused by heating as in lakes.
Nonetheless, as the water heats on the surface, it gets lighter and heavier water will more easily
dispatch it resulting unwelcome flow patterns.

The model does not have any biology or chemistry, which are always present in biological wastewater
treatment. Some gas may release from biological reactions or electrochemical forces may form larger
flocs or they may break them. The effect of these forces is small in well- designed clarifiers but if
sludge’s retention time in the bottom grows excessively large they may deplete oxygen and release
gases from anaerobic reactions.

The model is just 2D axisymmetric, which does not represent the real situation, especially if there is
some rotating scraper on the bottom.
Conclusions
All the models were well built and produced reasonable results. Only model with increase in flow should
have been started from the equilibrium after 12 h of simulation. We did not find the way to do it in
Comsol even though it for sure has it.

Normal mode produced good reduction for a clarifier and showed what kind of flow patterns form in the
clarifier. Simulation with baffle produced extremely well results but the design is more expensive and
sludge recirculation zone may cause sludge to produce gases.

The model has limitations. It neglects biology and chemistry and effect of external forces. It is also just
2D and does not account any scraper on the bottom. However, it is sufficient for preliminary design.
References
Plósz, B., Nopens, I., Rieger, L., Griborio, A., Clercq, J. D., Vanrolleghem, P. A., ... Ekama, G. A. (2012). A
critical review of clarifier modelling: State-of-the-art and engineering practices. Paper presented at 3rd
IWA/WEF Wastewater Treatment Modelling Seminar, Mont-Sainte-Anne QC, Canada. Available:
http://orbit.dtu.dk/ws/files/9851890/prod21341690900325.WWTmod2012_SST_critical_review_PLosz_
et_al.pdf

Hong Guo, Seo Jin Ki, Seungjae Oh, Young Mo Kim, Semyung Wang, Joon Ha Kim, 2017, Numerical
simulation of separation process for enhancing fine particle removal in tertiary sedimentation tank
mounting adjustable baffle, Chemical Engineering Science, Volume 158, Pages 21-29, ISSN 0009-2509,
https://doi.org/10.1016/j.ces.2016.09.022

Elham Ramin, Dorottya S. Wágner, Lars Yde, Philip J. Binning, Michael R. Rasmussen, Peter Steen
Mikkelsen, Benedek Gy. Plósz, 2014, A new settling velocity model to describe secondary sedimentation,
Water Research, Volume 66, Pages 447-458, ISSN 0043-1354,
https://doi.org/10.1016/j.watres.2014.08.034.

M. Patziger, H. Kainz, M. Hunze, J. Józsa, 2012, Influence of secondary settling tank performance on
suspended solids mass balance in activated sludge systems, Water Research, Volume 46, Issue 7, Pages
2415-2424, ISSN 0043-1354, https://doi.org/10.1016/j.watres.2012.02.007.

You might also like