Você está na página 1de 13

DESIGN AND OPTIMISATION OF FINAL CLARIFIER PERFORMANCE WITH CFD MODELLING

D. J. Burt*, BEng, MSc, CEng, MIMechE J. Ganeshalingam+, BSc, MSc, PhD, AMIChemE Presented at the CIWEM / Aqua Enviro joint conference Design and Operation of Activated Sludge Plants 19th April 2005.

ABSTRACT
A Computational Fluid Dynamics (CFD) prediction procedure for computing the internal hydrodynamic behaviour of final clarifiers is presented. Calculations are carried out and presented for comparison with experimental measurements of point velocities in a shallow circular clarifier in operation at a UK waste water treatment works. The calculations compare favourably with the data and the model has subsequently been used for many design studies. A separate study is described where CFD predictions are used to investigate the influence of introducing energy dissipating influents (EDI), varying the stilling well diameter, adding Stamford baffles at the side wall or placing an influent floor baffle (McKinney baffle) below the stilling well. From the study it is possible to determine the sizes and combinations of internal baffling likely to give the best clarification performance across a range of operating conditions. The success of the optimum design is judged by the depth of the settling sludge bed and the magnitude of the effluent suspended solids (ESS) for the state points considered. Key words: Computational Fluid Dynamics (CFD), Energy Dissipation Influent (EDI), Stirred Sludge Volume Index (SSVI), McKinney-baffle, Stamford-baffle. *Senior Engineer, MMI Engineering, Bristol, UK. and Dept of Mechanical Engineering, Queens Building, University of Bristol. + Project Engineer, MMI Engineering, Bristol, UK.

INTRODUCTION
According to the CIWEM handbook(1), Activated sludge plants (ASPs) are responsible for the treatment of about 50% of all sewage treated by biological oxidation in the UK. These plants are able to produce effluents in compliance with the current legislative requirements for suspended solids (SS), biochemical oxygen demand (BOD) and nitrate content (ammoniacal N). However, European legislation is demanding improvements in effluent quality and the requirement to treat larger volumes in our expanding towns and cities, is putting considerable pressure on the existing sewage treatment infrastructure. Consequently, there is much interest in the enhancement and improvement of the performance of the final stages of the activated sludge process. For, it is in the final clarifier that the sludge settles and it is the separation efficiency of the clarifier that largely determines the effluent quality of the ASP. It is well known that the settling efficiency of the clarifier is greatly affected by the hydrodynamic flow paths within the tank. In 1940, Anderson(2) published a paper showing how the internal flows within the tank led to a billowing of suspended solids immediately below the effluent weir. Figure 1 is taken from the work of Anderson and shows a section through a typical circular clarifier of 40m diameter and 3.5m side wall depth. The influent has an axial riser delivering feed into the tank through a central radial diffuser. Flow rates for this tank are in excess of 1000 m3/hr giving inlet velocities of order 0.1 m/s. A conventional stilling well is used to attenuate the turbulent inlet flow; the diameter and depth of this well are significant design parameters. The measurements show the typical density driven current observed in all circular clarifiers. The

dense feed with a concentration of mixed liquor suspended solids (MLSS) typically ranging from 2000 to 6000 mg/l, falls from the influent and generates a radial underflow. In order to balance this momentum, the underflow is matched by a return flow at higher depths in the tank. The resulting re-circulation is one of the reasons why the sediment blanket appears to lift at the side wall below the effluent weir. This early work clearly shows that the flow in a circular clarifier is far from one dimensional but can be considered as close to two dimensional with axis symmetry. The standard technique for designing a secondary clarifier is mass flux theory, sometimes referred to as state point analysis. This method uses a one-dimensional settling model that cannot take account of the complex internal flow patterns flow present in the clarifier. Shortcircuiting of the influent flow, scouring of solids and re-entrainment of solids into the recirculating flow pattern all contribute to the effluent suspended solids (ESS). Excess influent momentum, perhaps from an undersized stilling well, can invoke a complete failure with overflow of the sludge blanket. Even when the clarification surface area is over sized, relative to mass flux theory, the tank may still fail, or perform badly in practice because of these internal flow features. Clearly there is a need to use a new design method for clarifiers that is capable of overcoming the limitations of mass flux theory. In this work, a CFD modelling technique was developed, verified and validated to determine the internal hydrodynamic performance of secondary clarifiers. The model is based on an adaptation of the IAWQ drift flux model part of which is discussed in the Scientific and Technical report No 6 by Ekama et al(3). Results from the model were used to demonstrate several characteristic internal flow features that are thought to be causes of poor clarification performance. A validation study is presented which compares local velocity data with experiment for the Rye Meads clarifier, measured extensively by Richardson et al(4)(5). Also a single design study is presented for a typical UK clarifier where a number of internal modifications have been investigated in order to understand the influence of introducing energy dissipating influents (EDI), varying the stilling well diameter, adding Stamford baffles or placing an influent floor baffle (McKinney) below the stilling well. In the design study, the flow and settling behaviours were calculated for a base case configuration and then for alternative designs. The aim of the work was to determine the internal design likely to give best clarification performance across a range of operating state points. The success of a design was judged by the height of the settled sludge bed and the magnitude of the ESS for all state points.

MASS FLUX THEORY


Assuming sludge settling velocity is a unique function of the solids concentration; flux, the product of settling velocity and solids concentration, can be plotted against solids concentration. This generalized flux curve can be calculated from any of the standard correlations for settled volume index (SVI) or stirred sludge volume index (SSVI), e.g. Pitman(6) and White (7) or Wahlberg and Keinath(8). On this same curve two operating lines, over flow and under flow, are also plotted. If the intersection of these lines is above the flux curve then clarifier failure is predicted. At any state point, the crossing of the underflow and overflow lines must be below the flux curve for safe operation. If the underflow line is below the flux curve, but becomes tangential to it at higher concentrations, then the tank is said to be critically loaded and on the point of failure. There are other definitions of tank failure obtainable from mass flux theory that are not discussed here, see Ekama et al(3). The main difficulty in estimating the limiting situation with mass flux theory is a poor correlation between SSVI and flux, largely due to the flow patterns and physical processes previously discussed. Therefore, a safe design is not usually based on the generalized flux curve but on only 80 % to 90 % of the value at any point. Care must also be taken in use of correlations for

the mass flux curve as there are significant differences in these. Figure 2 shows how correlations can differ greatly when used in a practical selection application.

CFD MODELLING
Most of the CFD modelling work for secondary clarifiers reported in the literature uses a form of the algebraic slip (ASM) or drift flux model of Wallis(9) to represent the two-phase mixture of water and activated sludge. Zhou and McCorquodale(10) describe how the density variations and settling velocity relationships may be modelled. Lakehal and Krebs (11) have extended this model to include variations in fluid mixture rheology. Two recent PhD theses discuses the model in greater detail and include extended rheology functions, see Armbruster(12) and deClercq(13). In this implementation, a modified version of the CFX code was used as the modelling tool with an adaptation of the IAWQ drift flux model(3). The simulations were performed in twodimensional, axis-symmetric co-ordinates with models for sludge mixture density, viscosity, following Bokil and Bewtra(14), or Dahl(15), Dick and Ewing(16), and the double exponential settling function, following Takcs(17). A low Reynolds number k - turbulence model(18) was used in deference to the large variation in mixing time scales present in a clarifier. The model uses a multiphase method where the sludge is able to move independently with respect to the water but only in the direction of a slip vector; where, in this implementation the slip only acts in the direction of gravity.

Boundary Conditions
Even with its limitations, there is still value in using mass flux theory to calibrate the boundary conditions for subsequent CFD analysis. In the process of developing the generic CFD studies reported here it was discovered that, without applying any factors of safety, tanks with sufficient surface area to satisfy mass flux theory always fail before the mass flux limit when modelled with CFD, in other word mass flux theory is the absolute upper bound of performance. Mass flux theory also provides a useful indicator of the likely RAS solids concentration.

Physical Properties for Sludge


The Water Research Council (WRc) standard for clarifier design refers to the application of a method based on 30 minute SSVI settling tests following the Pitman(6) and White (7) correlation. However, in validation studies for this model it was discovered that not all UK sites exhibit a good match to this correlation(19) and a more rigorous assessment of sludge settleability as documented by Ekama et al(3) should be used when obtaining the settling or Vesilind(20) coefficients for use in the CFD model. Coefficients for the rheological constitutive relationships are difficult to obtain as there is no firm agreement on the type of viscosity model that should be applied to activated sludge. Consequently there are no standard test procedures for characterising sludge rheology. In these studies several rheological models were compared and it is thought that models incorporating a yield stress term, see deClercq(13), are the most appropriate.

VALIDATION CASE STUDY


The Rye Meads sewage treatment works (STW) has a population equivalent of 360,000 and a 95 percentile effluent discharge consent of 15:08:03 (ESS:BOD:NH3) mg/l. The works has 3 stages of activated sludge plant with a total of 16 final settlement tanks. The circular final tanks at Rye Meads are 28m diameter with a side wall depth of 2.2m. They are shallow with a floor

inclined at 8, they are equipped with a central feed well, perimeter weir and bridge driven scrapers. The conventional centre well is a 9m diameter, open bottomed, circular baffle arrangement, 1.5m deep (1.3m submerged). The tanks were analysed experimentally by Scriven and Richardson(5) at flow rates representative of the bounds of operation for the tanks. Comparisons between CFD and experiment are presented here for a forward flow of 360 m3/hr, mixed liquors at influent of 2825 mg/l, a RAS ratio of 1.0 and SSVI of 80 ml/g. The experimental work provided information on effluent quality, fluoride tracer break through curves, approximate blanket depth and Acoustic Doppler Velocimetry (ADV) measurements. The ADV data was found to be particularly useful for validating the CFD model results.

Validation against experimental data


The comparisons between CFD and experiment results for point measurements within the tank are presented as a series of 5 profiles. Each profile has a normalised radial position which is shown in Figure 3a. At each radial position, the depth at which the measurement is recorded is normalised to the total depth at that position. For radial velocity the experimental data was extracted as a single component of the velocity vector reading from the ADV probe. At each data point, the average value of the radial velocity component was calculated and normalised to the influent flow velocity (Uin). The normalised radial velocity profile was then plotted against normalised height for all 5 radial locations. Figure 3b compares the ADV data with three alternative CFD solutions each using a different rheological model. Figure 3c compares sludge profiles normalised to the influent concentration (Cin), against a single experimental data point recording the sludge depth taken to be the 850 mg/l contour.

Key Flow Features


Some key flow features were observed in the Rye Meads model common to similar designs. These features are indicated by the vector plot in Figure 3. Density waterfall. At this flow rate and for this size of stilling well, the flow entering at influent does not spread throughout the stilling zone. Rather it spreads locally from influent and then falls into the sludge blanket before forming a radial jet. Radial shear layer. All circular clarifiers show a strong radial shear layer immediately above the settled sludge bed. The shear layer will be constantly depositing and re-suspending solids from the settling sludge bed and it is this mechanism which is largely responsible for determining the effluent solids loading. Shear layer separation. At some point the radial flow breaks away from the bed and turns upwards before re-circulating back near the surface towards the central stilling well. The flow upwards at the side wall and the point of separation both influence the process of solids resuspension and carry over to the effluent weir. Re-entrainment. In this particular case the return flow is not sufficiently separated from the influent stilling well and there is re-entrainment of clarified water back into the mixed liquor zone. This is augmenting the influence of the density current and is likely contributing to higher ESS values. The influence of this surface conveyor phenomenon on the energy balance of the system is discussed by Armbruster(12). Bed conveyor. Solids that enter into the sludge bed are carried back to the RAS return by flowing down the sloping floor. It is not clear whether this return mechanism is influenced by the action of the rotating scraper. However, it is noticeable that some of the flow returning to RAS can be re-entrained by the influent density waterfall resulting in several recycles and increased residence for solids in the bed.

DESIGN STUDY
The design study is for a typical UK clarifier 55 (16.76 m) diameter and 8 (2.438 m) side wall depth. The tank has a 7.5 sloping floor with bridge scrapers and a pumped central hopper RAS extraction. As built, the centre well is only 2m (12%D) diameter and 1.5m deep. Modifications based on a larger stilling well were considered either with a central EDI, similar to that shown in Figure 4a or by mounting a baffle plate, Figure 5c, directly below the enlarged stilling well. This second design, known as a McKinney floor baffle is described in Ekama(3), it is located at a depth close to the side water depth with a gap sized for densimetric Froude number, Frd 0.7, at the highest flow rate. The key design parameter for the McKinney baffle is the slot gap between the bottom of the stilling well and the baffle. This is sized to keep 0.5 < Frd < 0.7 such that re-entrainment into the stilling well is prevented(12),

Frd =

U w gh w

where U is the average velocity through the slot, h is the height of the slot, is the mixture density at influent and w is the density of water, g is the gravitational constant. Alternative designs were analysed at various state points representative of the bounds of operation. Comparisons are presented here for a forward flow of 156.3 m3/hr, with mixed liquors at influent of 3700 mg/l, RAS ratio 0.51 or 1.0 and SSVI of 80 ml/g.
CFD Models As Built EDI McKinney Flow ratio R 0.51 a) 1 0.51 b) 1 0.51 c) 1 ESS Depth below TWL (mg/l) (m) 312 0.245 9 1.450 328 0.255 12 1.925 24 0.825 10 2.170 Status Fail Pass Fail Pass Pass Pass

Table 1: Summary of bed depth and ESS for the model variations. Figure 5 and Table 1 summarise the results for the study. Two state points are considered for three geometries allowing the performance of each design to be compared. The first thing to note from table 1 is that a critically loaded clarifier can be brought back into compliance by simply increasing the RAS rate. Figure 5a shows how the smaller stilling well gives rise to a higher sludge bed but, where the influent flow is now directed through the bed, the ESS is seen to be the lowest of the three configurations. Figure 5c shows how the McKinney baffle acts to separate the stilling zone from the settling zone and this design has the ability to carry the greatest volume flow rate, the results in Table 1 show that, in this case, only the McKinney design is compliant at both of the RAS rates investigated. Comparisons for the EDI design with and without a Stamford baffle showed no difference in the effluent quality, see Figure 6.

DISCUSSION OF INTERNAL DESIGN PARAMETERS


Masss flux theory only gives guidance on the likely surface area required to effect clarification, and, as has been stated, this can be a significant under estimate. There are many designs of circular clarifier using different shapes and depths in use in the UK. Side walls are generally a minimum of 2m in depth but floor angles can vary from flat to very steep in excess of 45.

Although there may be merit in deep designs (this is still a topic of investigation), the modeling techniques applied here have largely been used to obtain maximum performance from flat bottom or shallow angle clarifiers with various forms of RAS removal system. From these studies a number of observations have emerged that have provided guidance on internal baffling which can improve effluent quality and allow the tank performance to approach the mass flux limit.

The Stilling Well


One of the first things to consider in internal design is the diameter and depth of the stilling well. The concept of a flocculating stilling well gained some credibility in the 1990s; however, as can be seen in Figure 2a, a large stilling well is likely to generate significant re-entrainment from the settling region of the tank back into the stilling pond and this can persist even when an EDI is included. If the stilling pond is too small then the resultant down flux of momentum from the influent can be sufficient to disrupt the settled bed leading to higher beds and early failure. This impact of a small stilling well on bed height is shown in Figure 5a. In various studies a stilling well diameter of 20%D has been found to be most effective with depth set to half of water depth at the radius of the stilling well. Any shallow tank which has an influent based on a stilling well only design will suffer from strong density current effects at certain state points.

Improving the Influent


A central EDI, spreads the load out uniformly near the top of the stilling well reducing the density variation in the stilling pond. If it is designed well it can limit the influence of flow reentrainment into the stilling pond and can give good performance across a range of state points. Studies on the influence of vanes suggest that these are largely inconsequential accept when the combination of slot size and vane angle gives rise to excessive swirl and invoke resuspension of the sludge below the stilling pond. The important requirement is to keep the momentum exiting the EDI ports at a level high enough to promote homogenisation but not so high as to produce re-suspension. Some workers have made much of the idea that the stilling pond is contributing to re-flocculation, however, it has been shown in the measurements of deClercq(13) that there is little variation in the Particle Size Distribution (PSD) within the stilling pond. Other CFD studies of Camp number and this authors own concept of a G Scalar history function(21) suggest that levels of G high enough to promote orthokinetic flocculation of activated sludge, as defined by Biggs(22), are only present in the first few tens of seconds following entry into the stilling pond. A McKinney baffle cuts the density current and, if designed correctly, completely separates the stilling and settling zones. However, the McKinney design shows sensitivity to bed depth as it operates most effectively when the baffle sits at the same level as the sludge blanket; in this way the flow exiting the influent forms a level radial jet across the top of the settling bed. For low flow or low SSVI situations a second density waterfall can form at the end of the McKinney baffle which has a degrading effect on tank performance. Ideally, a McKinney baffle would track the bed height in operation or the bed height would be maintained (through RAS control) at the level of the McKinney. When including a McKinney baffle it is usually necessary to increase the depth of the stilling pond to suit the required slot height.

Side Wall Baffles


In other design studies, modifications to the influent were augmented by a variety of side wall baffling options, see Figure 4b. In shallow tanks with low floor angles, no clear benefit from using a Stamford baffle was observed and this is consistent with Figure 6. The CFD predictions indicate that shear layer separation always occurs before the flow reaches the side wall in such

a way that the effluent flow effectively by passes the location of the Stamford. In flat bottomed deep tanks, with low beds, the shear layer tends to persist all the way to the side wall and the Stamford can have a much more significant influence on the effluent quality.

CONCLUSIONS
An integrated CFD and mass flux modelling approach has been developed that may be used to optimise the internal designs of final clarifiers. During the course of the development work a number of key conclusions have arisen both with reference to current clarifier design methodology and to current design practice. Flows in secondary clarifiers are characterised by a strong density current arising from the influent that drives a radial shear layer above the settling sludge bed. Several flow feature exist that must be designed out to optimise the tank performance. The flow is far from one dimensional but it can be considered as close to two dimensional and in a circular clarifier there is axis-symmetry. Tanks with sufficient surface area to satisfy mass flux theory will always fail before the mass flux limit when modelled with CFD, in other words mass flux theory is the absolute upper bound of performance. The stilling well dimensions are important, too large a diameter or too shallow, can serve to enhance the density current momentum through re-entrainment. Two options are currently favoured as a means of breaking the density current. An EDI or a McKinney baffle. A central EDI helps to diffuse the density current by spreading the load uniformly near the top of the stilling well and reducing the density gradients in the stilling pond. A McKinney baffle cuts the density current and, if designed with 0.5 < Frd < 0.7 it distinctly separates the stilling zone from the settling zone introducing the flow as a direct radial jet into the settling zone. The McKinney design tends to favour a limited operating range as it provides maximum benefit only when the baffle sits slightly above the settling sludge bed. In its best operating range it will tend to out perform an EDI. Sidewall baffling has only a limited influence in shallow clarifiers but can be beneficial in flat bottomed clarifiers with deep side walls. It is now possible to check clarifier retrofit and final design options with CFD modelling prior to executing a civil engineering project.

ACKNOWLEDGEMENTS
The author would like to thank Dr Pete Pearce of Thames Water for his help, advice and continued support. Thanks also to other colleagues at Thames Water, United Utilities, Montgomery Watson Harza, Severn Trent and Yorkshire Water who have all supported projects contributing to the overall understanding of internal clarifier flows.

REFERENCES
[1] Chartered Institution of Water and Environmental Management (CIWEM), Activated Sludge Treatment, Handbooks of UK Wastewater practice, London, 1997. [2] Anderson, N.E., Design of Settling Tanks for Activated Sludge, Sewage Works J., 17(1), 50-63, 1945. [3] Ekama, G.A., Barnard, J.L., Gunthert,F.W., Krebs,P., McCorquadale, J.A., Parker, D.S. and Wahlberg, E.J., Secondary Settling Tanks, Theory, Modelling, Design and Operation, International Association of Water Quality, Scientific and Technical Report No 6, 1997. [4] Richardson, D.S., Hydraulic Considerations of Final Settlement Tank Design, Cranfield University School of Water Sciences, M.Sc. Thesis, 1998. [5] Scriven, R. and Richardson, D.S. Rye Meads STW Stage 1 Final Settlement Tanks Process Investigation, Thames Water Technical Report, R19808, August 1998. [6] Pitman, A.R. Settling Properties of Extended Aeration Sludge, J. Wat. Pollut. Control Fed. 52(3), 524-536, 1980. [7] White, M.J.D., Settling of Activated Sludge, Technical Report TR11, Water Research Centre, Stevenage, UK, 1975. [8] Wahlberg, E.J. and Keinath, T.M. Development of settling flux curves using SVI J. Wat. Pollut. Control Fed. 60 (12), pp2095-2100, 1988. [9] Wallis, G.B., One-Dimensional Two Phase Flow, McGraw-Hill, 1st Ed, 1969. [10] Zhou, S. and McCorquodale, J.A., Modelling of Rectangular Settling Tanks, J. Hydr. Eng., ASCE, 118(10), October 1992. [11] Lakehal, D., Krebs, P., Krijgsman, J. and Rodi, W. Computing Shear Flow and Sludge Blanket in Secondary Clarifiers, J. Hydr. Eng., ASCE, 125(3), 1999. [12] Armbruster, M., Untersuchung der mglichen Leistugssteigerung von Nachklrbeken mit Hilfe numerischer Rechungen, PhD thesis, University of Karlesruhe, August 2003. [13] De Clerq, B. Computational Fluid Dynamics of Settling Tanks: Development of Experiments and Rheological, Settling and Scraper Sub Models, PhD Thesis, Dept of Applied Math, Biometrics and Process Control (BIOMATH), University of Ghent, Belgium, 2003. [14] Bokil, S.D. and Bewtra, J.K., Influence of Mechanical Blending on Aerobic Digestion of Waste Activated Sludge, Proc., 6th Int. IAWPRC Conf. on Water Pollution Res., Int. Assoc. on Water Pollution and Control, London, 421-438, 1972. [15] Dahl, C.P., Larsen, T. and Peterson, O., Numerical Modelling and Measurement in a Test Secondary Settling Tank, Water Sci. and Technology., 30(2), 219-228, 1994. [16] Dick, R.I. and Ewing, B., The Rheology of Activated Sludge, J. Water. Pollution Control Fed., 39(4), 543-560, 1967. [17] Takcs, I., Patry, G.G., and Nolasco, D. A Dynamic Model of the Clarification Thickening Process., Water Res, 25(10), 1991

[18] CFX International, CFX-4.4 Solver Manual, Vol 3, AEA Technology, Harwell, 2001. [19] Burt D.J. and Ganeshaligam, J. Validation study for the Witney Clarifier, MMI Engineering Report, MMU035, February 2005. [20] Vesilind, P.A. Theoretical considerations: Design of prototype thickeners from batch settling tests, Water and Sewage Works, 115 (July), 302-307, 1968. [21] Burt, D.J. and Gilbertson, M.A. Flocculation Frameworks and CFD Modelling for Activated Sludge Clarifiers, 5th Particle Technology Forum, University of Sheffield, July 2003. [22] Biggs, C. A. Activated Sludge Flocculation: Investigating the Effect of Shear Rate and Cation Concentration on Flocculation Dynamics, PhD Thesis, Dept of Chem Eng, University of Queensland, Australia, 2000.

Figure 1: Taken from Anderson(1) shows concentration gradients through a circular clarifier and velocity vectors mapped at discrete points.

Rye Meads Final Clarifier


Q 360.00 m3/h Aeration Lane Qin=Q(1+R) 720.00 m3/h Q 360.00 m3/h

RAS = RQ 360.00 m3/h Recycle 360.00 Tank Diameter SST Area (m2) Forward Flow (Q, m3/h) Design Overflow Rate (m/h) SSVI (ml/g) 3 Influent MLSS ( kg/m ) RAS ratio 3 RAS_surplus (m /h) Solids Loading (l/m2h) Estimated RAS Conc. ( kg/m3) 28 615.75 360.00 0.58 100 2.825 1.00 0 220 5.6500 m /h
3

RAS_surplus 3 0.00 m /h Flux theory

160 140 Solids Flux (kg/m2/day) 120 100 80 60 40 20 0 0 1 2 3 4

SSVI = 100 [Pitman and White (1980, 1984)] SSVI = 100 [Wahlberg and Keinath (1998)] Overflow Line Underflow Line Influent Conc.

6
3

10

Sludge Concentration (kg/m )

Figure 2: Mass flux graph comparing Pitman(6) and White (7) with Wahlberg and Keinath(8) for the Rye Meads clarifer at average flow and 100 SSVI.

a)
Density Waterfall 0.35 0.5 0.64 0.78 0.96

Shear Layer Separation Radial Shear Layer Entrainment Bed Conveyor

b)
R / Rmax =0.35

R / Rmax =0.50 0.0 0.1 0.2 0.3 0.4 H / Hmax[-] H / Hmax[-] 0.5 0.6 0.7 0.8 0.9 1.0 -0.2 0.0 0.1 0.2 0.3 0.4

R / Rmax =0.64 0.0 0.1 0.2 0.3 0.4 H / Hmax[-] 0.5 0.6 0.7 0.8 0.9

R / Rmax =0.78 0.0 0.1 0.2 0.3 0.4 H / Hmax[-] 0.5 0.6 0.7 0.8 0.9

R / Rmax =0.96

0.0 0.1 0.2 0.3 0.4


H / Hmax[-]

0.5 0.6 0.7 0.8 0.9 1.0 -0.2

0.5 0.6 0.7 0.8 0.9 1.0 -0.15

0
V / Uin [-]

0.2

0 V / Uin [-]

0.2

-0.05

0.05

0.15

1.0 -0.15

-0.05

0.05

0.15

1.0 -0.05

0 V / Uin [-]

0.05

V / Uin [-] Expt T4 T3 T5 Expt T4

V / Uin [-] T3 T5 Expt T4

Expt T4

T3 T5

Expt T4

T3 T5

T3 T5

c)
0.0 0.1 0.2 0.3 0.4 R / Rmax =0.35

0.0 0.1 0.2 0.3 0.4 H / Hmax[-]

R / Rmax =0.50

0.0 0.1 0.2 0.3 0.4 H / Hmax[-]

R / Rmax =0.64

0.0 0.1 0.2 0.3 0.4 H / Hmax[-] 0.5 0.6 0.7 0.8 0.9 1.0

R / Rmax =0.78

0.0 0.1 0.2 0.3 0.4 H / Hmax[-] 0.5 0.6 0.7 0.8 0.9 1.0

R / Rmax =0.96

H / Hmax[-]

0.5 0.6 0.7 0.8 0.9 1.0 0 1 2 C / Cin [-] T3 T5 T4 Exp 3

0.5 0.6 0.7 0.8 0.9 1.0 0 C / Cin [-]


T3 T5 T4 Exp

0.5 0.6 0.7 0.8 0.9 1.0 0 1 2 3 C / Cin [-]


T3 T5 T4 Exp

1 C / Cin [-] T3 T5 T4 Exp

0.5 C /Cin [-] T3 T5 T4 Exp

Figure 3: CFD analysis for Rye Meads, T3, T4(14) and T5(11) represent alternative rheological models compared with the experimental data of Richardson et al(4)(5). The results show good agreement for radial velocity profiles and bed height.

CROSB Y

VARIATION

PLAIN

McKI NNEY

CANTILEVERED TROUGH

a) Typical flocculating centre well arrangement with central EDI, swirling ports and large, 30% D, stilling well.

b) Various alternatives for baffling the side wall. The Crosby is also sometimes called a Stamford.

Figure 4: a) A central EDI, spreads the load out uniformly near the top of the stilling well reducing the density variation. b) The return current at the side wall can be disrupted by various baffle options.

a) Clarifier as built with a 2m (12%D) diameter stilling well.

b) Revised design with flocculating stilling well at 20%D and enclosing an EDI. The EDI is a variation on Figure 4a.

c) A McKinney baffle design with stilling well at 20%D and gap sized for densimetric Froude number, Frd 1, at the highest flow rate.

Figure 5: A typical UK Clarifier diameter 16.76m. Comparisons between designs are presented here for a forward flow of 156.3 m3/hr, mixed liquors at influent of 3700 mg/l, a RAS ratio of 1.0 and SSVI around 80 ml/g. A logarithmic scale is used to show the solids distribution 1 mg/l to 10,000 mg/l. The lightest shading at 1000 mg/l is approximately at the transition into the settled bed.

With Stamford baffle

Without Stamford baffle

Figure 6: A Stamford baffle was included for the EDI design shown in Figure 5b. The flow patterns near the effluent are redirecated but there is no significant difference in the effluent quality.

Word count: 4794 words

Você também pode gostar