You are on page 1of 17

ASSESSING THE USE OF AUTONOMOUS VALVELESS FILTERS FOR TURBIDITY REMOVAL IN RURAL APPLICATIONS

B.M. Brouckaert1, M.J. Pryor2, P. Pillay2, T.A. Zondi2, A. Amirtharajah3,


1

School of Chemical Engineering, University of Natal, Durban, South Africa Tel: 031 260 1129, Fax: 031 260 118 E-mail brouckaertb@nu.ac.za 2 Process Evaluation Facility, Umgeni Water, Pietermaritzburg, South Africa 3 School of Civil and Environ. Eng., Georgia Institute of Technology, Atlanta, USA 1. BACKGROUND

The choice of a treatment process for potable water production is strongly influenced by the raw water quality, desired final water quantity, physical location and the availability of resources for operation and maintenance. The complexity of the treatment process required to treat raw water to potable standard is not significantly affected by the capacity of the plant. Economies of scale between large municipal water treatment plants and small-scale plants can influence the process selection and therefore the quality of the water supplied to a consumer. When conventional type small-scale plants are operated by central authorities (water boards or municipalities), operating costs are driven up primarily by the traveling time to and from the plant, as well as the wide range of specialised skills that are required to operate the processes from a central base. These circumstances place restrictions on the designers of rural water supply schemes to provide low cost, low maintenance plants without compromising the performance and quality of water produced. Processes that are inherently self-regulating and simple to operate have an obvious advantage in that they will require fewer skilled operator visits to the plant at less frequent intervals. This allows many of the operating and maintenance tasks to be delegated to community level and reduces the reliance of the plant on specialised skills. The autonomous valveless filter (AVF) has the particular advantage that a backwash is initiated automatically by a unique siphoning system once a specific headloss across the media has been reached, without any electronic controls or operator intervention. Umgeni Water initiated a project (funded by the Water Research Commission) to study the performance of an AVF and its potential application to rural water treatment. The objectives of the project were to assess the suitability of the filter for potable water production from surface waters, propose design modifications aimed at improving the reliability and reducing the capital cost of the filter, and propose operating rules by which the unit should be operated to ensure reliability and reduce operator supervision. This paper will focus on critical design aspects identified in this study as well as mudballing effects resulting from an inadequate backwash regime and their impact on long term filter performance.
Paper presented at the Biennial Conference of the Water Institute of Southern Africa (WISA) www.wisa.co.za CD-ROM produced by: Water Research Commission (WRC), www.wrc.org.za 19 23 May 2002, Durban, South Africa ISBN Number: 1-86845-844-X Organised by: Conference Planners

2.

OPERATION OF THE AVF


G

The operation of the AVF is illustrated in Figure 1. Raw water enters the filter, passes through filter media and nozzles into a collector chamber below the filter floor, then rises through the effluent duct and out to service. During the first part of the filter run, water is collected in the backwash storage chamber. As a result of headloss development the water level in the backwash pipe rises until the level initiates a self-actuated primer system which rapidly exhausts air from the backwash pipe and starts a siphoning action that backwashes the filter. Water from the storage chamber siphons through the backwash water ducts, up through the filter nozzles and media and out to waste. At the beginning of the backwash the level in the storage chamber is highest resulting in high initial backwash rates. The rate decreases as the level drops until a siphon breaker allows air into the backwashing pipe, breaks the siphon, and stops backwashing.
S H A O R A
SIPHON BREAKER BACKWASH PIPE OUT TO SERVICE

INLET

B S

K T

TO WASTE

a.

FILTERING Figure 1.

b. BACKWASHING Operation of the Autonomous Valveless Siphoning Filter

The primary advantage of the AVF over other types of filters is the fact that backwash is initiated automatically at a pre-determined terminal headloss without electronic controls, operator intervention or even a backwash pump. Another important advantage is that the rising head system ensures that the filtration rate is maintained constant throughout the filter run without any valve adjustment required. This can result in a substantial reduction in labour requirements and operating costs. However, coagulant addition is usually required to meet drinking water standards. Therefore, in practice, the performance of the AVF will depend on the reliability of the coagulant dosing and, in general, it should not be left to operate without supervision for extended periods of time. The major disadvantage of the AVF is that it relies on water only backwash to clean the media between filter runs. Water only backwash is known to be an inherently weak cleaning process as fluid shear is not very effective for breaking up cohesive deposits (Amirtharajah, 1980) particularly when coagulant is used. Auxiliary backwash involving either air scour or high velocity surface or sub-surface water jets is typically required to prevent the degradation of filter media over the long term. AVFs can be designed with the facility for air scour, however, this requires manual or electronic control which detracts from the inherent simplicity of the filter operation (Stevenson, 1998). One of the key issues, which will determine the suitability of the AVF for use in rural areas, is whether or not the filter media can be adequately cleaned by water only backwash. If the AVF cannot operate indefinitely without intervention to restore the filter bed to an acceptable state, then the feasibility of this process may depend on the nature and frequency of the intervention required to maintain satisfactory performance.

3.

CRITICAL ASPECTS OF FILTER DESIGN, OPERATION AND PERFORMANCE

The AVF is a proprietary item and there is little information on its design, operation and performance in the literature. In many respects it resembles any other downflow rapid filtration process and general design guidelines for such filters apply. However, it also has some unique constraints, particularly in the design of the backwash system, which require special consideration. This section presents general background from the literature on filter design and performance while Section 5 discusses specific design features of the AVF and Section 6 presents experimental results relating to AVF performance. 3.1 Media selection and design

The selection of media for filtration should primarily be based on the degree of purification required, as well as the durability of the media (Kawamura, 1975a). However, the length of the filter run and the ease with which suspended matter is removed from the media during backwashing are equally important factors. The coarser the media used, the longer the filter runs, but also the deeper the bed required to achieve a particular filtrate quality and the higher the backwash rates required. Using finer filter media improves the filtrate quality, but decreases run lengths (Baumann, 1978). Finer media, also appears to be more prone to mudballing than coarser media. In a study of filters at a plant operated by the City Of Kyoto, Japan, Kawamura (1975a) found that 0,5 mm sand filters accumulated 10 times more mud over a four year period than 0,7 mm sand filters, even though the same backwash rate was used in both sets of filters. Using a combination of sand and anthracite (dual media) results in longer run times when compared to fine monomedia sand, while reducing the risk of breakthrough when compared to coarse monomedia filters. However, care must be taken to choose compatible media sizes to prevent excessive intermixing of sand and anthracite. A typical design recommendation is that the d90 (mesh size through which 90 % of the mass of the media would pass) of the anthracite should be ~3 times the d10 size of the sand (Amirtharajah, 1978a). This is based on sand and anthracite densities of 2650 kg/m3 and 1500 kg/m3 respectively, and allows a moderate amount of intermixing at the interface between the sand and anthracite. 3.2 Mudballing

An inadequate backwash regime results in mudballing of the filter media. A mudball is considered to be any structure consisting of media grains glued together or coated with floc, which is not completely broken down by backwashing. Each time the filter is washed the particles of mud remaining trapped in the media become more compact and build up in the filter bed forming greater masses. Mudballs tend to accumulate at or near the surface of the media, but eventually they grow large enough, or attain a specific gravity great enough, to cause them to sink to the bottom of the sand bed during washing (Baylis, 1959; Cleasby, 1990). Within the media they form subsurface solidified inactive regions of the filter bed that are not remedied by normal backwashing. The effective surface area for particle capture is reduced, velocity of flow through the rest of the bed is increased and filter performance eventually deteriorates.

3.3

Backwash Rate and Volume

The efficiency of fluidised bed water only backwash in cleaning a filter bed depends on a number of factors, the most important of which are grain size and density, backwash rate and volume, cohesiveness of the floc, uniformity of wash water distribution and temperature. For water backwash without auxiliary scour, Amirtharajah (1978b) showed that maximum shear occurs at expanded porosities of 0,68 to 0,71. For typical graded sand beds, the optimal porosity is achieved in the top sections of the bed at bed expansions of 40 to 50 %. With dual- and mixed-media filters, greater than normal bed expansion may be necessary to release the entrained floc. Amirtharajah (1978a) recommends 25% expansion of the anthracite bed and 20% expansion of the sand as a guideline. Kawamura (1975b) derived an alternate model of the optimum backwash rate (termed the appropriate backwash rate) based on maximizing the rubbing action between grains. Although, Amirtharajah (1978b) has contested the theoretical basis of this model, it is widely published in filter design texts and is often more convenient to use in calculating appropriate backwash rates than recommendations based on bed expansion. Kawamuras model for the appropriate backwash rate can be reduced to: Vb = d
Vb D
1 3 1 3

for sand ...........................

(1a)

for anthracite ................. Appropriate backwash rate, m/min Average grain size , mm (d60 is usually used as the average grain size) Water viscosity , cp

Vb = 0 ,47 d

(1b)

The backwash volume required to completely remove entrained material from the filter depends on a number of factors including the strength and amount of filter deposits, the type, size and depth of the media, wash rates and use of auxiliary backwash, the distance between the expanded media and the backwash trough and the subsequent filter start-up strategy. Kawamura (2000) recommends termination of filter washing when the backwash turbidity drops below 10 to 15 NTU. Design recommendations for backwash volume vary from 3-4 m3/m 2 (Amirtharajah 1978a) to 6 m3/m 2 (Kawamura, 2001). The recommendation of 3-4 m 3/m 2 is based on a backwash trough height of 0,91m above the media. Larger spacing between troughs and greater heights above the filter media increase the washwater requirements. 3.4 Filter nozzles and backwash water distribution

Proper selection and positioning of filter nozzles is essential to provide even distribution of backwash water. According to Stevenson (1998), the backwash headloss across the filter nozzles needs to be at least one tenth of the media headlosses for the flow to be evenly distributed throughout the media and should ideally be equivalent to the submerged weight of the clogged filter bed to ensure even distribution of flow through the filter floor. The optimum nozzle density depends on the size of nozzle strainer. Stevenson suggests that the optimal configuration would use strainers with a diameter of half the nozzle pitch. The size of the strainer slots should be selected to prevent clogging of the slots by fine media grains.

4. 4.1

MATERIALS AND METHODS Autonomous valveless filter

A 1,2 m diameter AVF was installed at the Umgeni Water Process Evaluation Facility in Durban. This particular model was designed for a 0,6 m bed of 0,7 mm sand and a maximum filtration rate of 7,6 m 3/h. Automatic backwash is initiated at a headloss of approximately 1,5 m across the filter bed. Manual backwash can be initiated at any time by opening a valve which allows water to be drawn directly from the backwash reservoir into the self-priming system. The backwash reservoir capacity is 2,5 m 3 or 2,2 m 3/m2. The backwash rate peaks at 46 m/h and declines linearly with time to 25 m/h over a period of almost 4 minutes. The filter was originally fitted with 105 mm strainers (22 nozzles at 19,5 per m 2). However, when the filter was installed, the original nozzles had been replaced by disc type strainers. These were subsequently found to be prone to clogging and to have reduced the available backwash rates. They were subsequently replaced with KSH type C2 nozzles with 0,2 mm slots and the number of nozzles increased to 39 (34,5 per m 2) by drilling additional holes in the filter floor. 4.2 Laboratory filters

Four PVC laboratory filters (200 mm diameter) were used to simulate filtration and backwashing of the AVF for various media designs. The laboratory filters were equipped with a backwash pump rather than a siphoning system and the AVF backwash was simulated by manually varying the flowrate. Manometer tubes connected to each filter allow the measurement of headloss development. The clear PVC walls of the laboratory filters allowed direct observation of the backwash behaviour of the filter beds. The laboratory filters could be backwashed with air scour and/or combined air and water backwash as well as water only backwash. Backwash with air was used to restore the media to a clean state before the start of each series of experiments simulating the continuous operation of the AVF. Combined air and water wash was also used to recover the accumulated floc not removed by water only backwash at the end of each series of AVF simulations. 4.3 Raw water quality and pretreatment

Both the AVF and the laboratory filters were operated in in-line filtration mode with alum coagulation to produce filtrate turbidities of < 0,5 NTU before filter breakthrough. The raw water used was an impounded surface water fed under gravity from the adjacent Umgeni Water Wiggins Waterworks. The turbidity of the water is usually <10 NTU with seasonal increases to between 15 NTU and 100 NTU after significant rains in the catchment. The pH ranges between 7,5 and 7,7 with an average alkalinity of 44 mg/L as CaCO3.

4.4

Filter media

Filter media properties and design are summarised in Tables 1 and 2. The filter media used was also characterised in terms of clean bed headloss (up flow and downflow) and minimum fluidisation velocity at various temperatures. The media were obtained from local suppliers of silica sand and anthracite in South Africa. Only one size range of anthracite was available and its d90 (2,10 mm) was found to be four times larger than the d10 of the 0,5 mm sand, compared to the recommended design ratio of three for a similar anthracite density (Amirtharajah, 1978a). The anthracite was sieved by hand using a stainless steel screen to remove the fraction >1,7 mm in order to obtain a size ratio of 3,0.
Table 1. Media properties Silica sand d10 d90* Uniformity Coefficient Density 2642 kg/m3 * d10, d90 are the sieve sizes through which 10% and 90 % respectively of the mass of media passes ** density after soaking in water for 24 hrs (Ives, (1990))
*

0,5 mm 0,52 mm 0,91 mm 1,42

0,7 mm 0,72 mm 1,04 mm 1,36

Anthracite (sieved) 0,91 mm 1,60 mm 1,48 1545 kg/m3**

Sand size Sand depth Anthracite size Anthracite depth

Media design for filter operation Laboratory Filters 0,7 mm sand Dual media 0,7 mm 0,5 mm 0,52 m 0,41 m _ 0,9 mm 0,12 m

Table 2.

Full Scale AVF Dual Media 0,5 mm 0,39 m 0,9 mm 0,12 m

4.5

Methodology

Initially the operation of the AVF with and without coagulant was assessed. The use of a blended polymeric coagulant (di-alyl,di-methyl ammonium chloride and poly-aluminium chloride) resulted in rapid mudballing of the filter media with decreased run times within 3 months, whilst operating the filter without coagulant generally resulted in poor filtrate quality (> 1 NTU). All subsequent experiments used alum as the coagulant. In the next phase of the project, the headlosses in the various components of the filter (media, nozzles, pipes and appertenances were characterised as functions of backwash flowrate. A hydraulic model of the AVF backwash was developed and used to predict the effect of various design changes on the backwash (Pryor and Brouckaert, 2001). Four laboratory filters were constructed and used to simulate the operation of the AVF for various media designs. Based on these results, a dual media design was selected for the full-scale AVF. The AVF was operated for a period of 1 months, during which time it was 5 opened on three occasions to assess the state of the filter media. Further laboratory scale trials using dual media and 0,7 mm monomedia sand indicated that some mudballing of the filter media was inevitable in the absence of auxiliary backwash, even if the backwash design was improved. The focus of the investigation then shifted to understanding the mechanisms by which mudballs formed and accumulated in the filter media and how their presence impacted the long-term performance of the filters.

The laboratory filters were operated as consistently as possible for 65 consecutive runs without auxiliary backwash, totaling a cumulative run time of approximately 1 month. Filtration performance was closely monitored and the data analysed to determine whether there were any statistically significant trends in performance, which could be attributed to mudballing of the media. Backwash performance was also analysed in terms of the accumulation of mudballs and the efficiency of wash water usage. 5. ASSESSING THE DESIGN OF THE AVF

The AVFs unique backwashing system can significantly improve the reliability of the filter operation compared to other types of filtration systems, which rely on manual or electronic controls. However, it also imposes a number of constraints on the filter operation. These constraints should be born in mind when considering the AVF for drinking water applications. 5.1 Initiation of backwash

Automatic backwash in the AVF is initiated only when a specific terminal headloss is reached. While this should ensure regular backwash of the filter, it will not necessarily prevent premature turbidity breakthrough at high influent turbidities (Pryor and Brouckaert, 2001). Furthermore, the frequency of backwash depends on the rate of headloss development. Low influent turbidities can result in low rates of headloss development and excessively long run times. Monk and Willis (1987) suggest that irreversible deterioration of filter media can occur during filter runs > 36 h, even for systems with auxiliary backwash. In order to maintain filter performance over a broad range of raw water turbidities and coagulant doses whilst ensuring the filter backwashes regularly, it is desirable to be able to adjust the terminal headloss as required. The position at which the self-actuating primer system connects to the backwash pipe determines the headloss at which the filter backwashes. Multiple ports on the backwash pipe with individual isolation valves could be installed to provide this flexibility. 5.2 Backwash rate and volume

The backwash rates are determined by the hydraulic design of the filter i.e. the pipe and nozzle headlosses and the difference in head between the backwash reservoir and overflow weir in the backwash waste tank. The backwash rate peaks as soon as the evacuation of air from the backwash pipe is complete and declines thereafter as the backwash reservoir empties. The volume of backwash in the AVF is determined by the dimensions of the backwash reservoir and the level at which the filtrate pipe exits the filter (Figure 1). The declining backwash rate may sometimes be an advantage because it facilitates good re-stratification of the media, however, it also means that some portion of the wash water may not effective in cleaning the media. The total backwash volume in the AVF in this study was 2,2 m 3/m 2, which is smaller than the 3-6 m3/m 2 recommended by Amirtharajah (1978a) and Kawamura (2001). The depth / volume of the backwash reservoir and the size of the effluent pipes were identified as design parameters which could affect the backwash performance of the AVF. Figure 2 (a) shows the effect of varying the backwash effluent pipe diameter, backwash reservoir depth and media headloss on the backwash rates and duration. 0,5m of media headloss corresponds to a bed depth of approximately to a 0,6 m of sand or 2,6 m of anthracite (density ~1550 kg/m3).

Whether or not the available volume and backwash rate are adequate or effective for cleaning the media depends on the amount and cohesiveness of the floc to be removed and in part on the size and density of the media. In practice, backwash efficiency can only be determined experimentally.
3.0

70 60

increase backwash effluent pipe to 98 mm diameter

2.5

Velocity (m/h)

Volume (m3)

50 40 30

increase backwash reservoir depth by 0.5 m

2.0 1.5 1.0 0.5 0.0 0

Backwash Headloss
0,3 m 0,5 m

Vmf (DM)

20 10 0 0 1 2 3 4 5

0,7 m 0,9 m

media headloss = 0.4 m

10

20

30

Vmf (0,7)

media headloss = 0.6 m

40

2 x Vmf (DM)
50

Time (min)
Figure 2(a). Effect of varying design parameters on backwash rates

Minimum Fluidisation Velocity (m/h)


Figure 2 (b). Volume available for fluidised bed backwash

5.3

Media selection and design

It is important to note that since the backwash rates are fixed by the physical dimensions of the filter, the range of media sizes which can be used in any given filter is limited. Figure 2 (b) shows the volume of backwash available to fully fluidise the media in the AVF as a function of minimum fluidisation velocity and fluidised bed headloss. It is evident that the 0,7 mm sand for which the AVF was designed is fully fluidised for only part of the backwash. Coarser media may be less prone to mudballing than finer media, however, a deeper filter bed is required to achieve the same filtrate quality as finer media and higher backwash rates are required to achieve fluidisation. This translates into a larger (taller) and more expensive filter design if coarser media is used. Since the media headlosses have little effect on the available backwash rates (Figure 2(a)), the maximum bed height which will not lead to media losses during backwash should be used to improve filtrate quality and delay breakthrough. The maximum bed height can be estimated from the steady state bed expansion at the maximum backwash rate and the minimum expected temperature, however, allowance should be made for the fact the clogged beds overshoot the maximum clean bed height during expansion. 5.4 Backwash nozzles and flow distribution

The filter nozzles in the AVF make a significant contribution to the total backwash headlosses (Pryor & Brouckeart, 2001) and therefore influence the backwash rate. If the nozzles have to be changed, care must be taken to ensure that replacement nozzles have the same headloss characteristics as the original nozzles. The nozzle headloss characteristics were apparently not taken into account when the original filter nozzles were replaced with disc type strainers (Section 4.1) resulting in a reduction in backwash rates. The headlosses through the disc type strainers were nearly 3 times as high as through the original nozzles even before the former became clogged.

The original 105 mm nozzles were no longer manufactured and so the disc type strainers were ultimately replaced with KSH type C2 nozzles which produced approximately the same headloss through the filter floor as the original set up. This restored the backwash rates to their design values. Increasing the nozzle density was intended to improve the backwash flow distribution but it appears that poor flow distribution, especially at the edges of the filter may nonetheless have contributed to tmudballing of the filter bed. This may have been partly due to the internal structure of the filter compartment. There were a number features inside the AVF which appeared to contribute to the formation of dead zones. The most obvious was a box at the back of the filter into which the feed water drained before entering the filter compartment. The top of the box was below the level of the fluidised bed, resulting in a stagnant shelf on which a large pile of media and mud piled up. Other dead zones developed around where the edges of the filter access hatches protruded into the filter bed. The nozzle headlosses were similar to the media headlosses at the beginning of backwash but then decreased to less than half of the media headloss towards the end of backwash. Whether this could have contributed to poor backwash flow distribution is unclear but it is a design aspect which could possibly be improved upon. 6. 6.1 EXPERIMENTAL RESULTS AND DISCUSSION Mudball Formation

During filtration, the media grains in the upper regions of the bed become cemented together with floc to form a low permeability composite plug which subsequently breaks down during backwashing with most the floc being flushed away. However, at sub-optimal backwash rates and in areas of poor flow distribution, the disintegration of the plug may not always be complete. For example, on several occasions during the operation of the 0,7 mm laboratory filter, large chunks (up to 10 cm across) of clogged media were observed sinking into the filter as the top sections of the bed began to fluidise. These chunks were slowly eroded away as the backwash progressed but residue from the original structures was visible through subsequent filter runs. This type of mudballing can usually be avoided by increasing the backwash rates. The maximum backwash rate required appeared to depend on the degree of clogging and possibly also on the characteristics of the floc. Although increasing the backwash rates should ensure the complete disintegration of the plug of floc and media, which develops during filtration, it apparently cannot prevent a coating of floc remaining on some of the media grains. Thiswas particularly evident at the top of the bed where the media grains are small and the shearing forces relatively low. Water only backwashing appears to always leave a layer of floc coated grains on the surface of the bed. Based on the evidence which will be presented below, this surface layer appears to be the source of a mudballing phenomenon which can only be prevented by using auxiliary backwash. The growth and accumulation of mudballs was first observed in the dual media laboratory filter. After as little as two or three consecutive filter runs without air scour, clusters of ~ 0,5 to 1 cm mudballs began to appear in the anthracite layer of the dual media filter. These clusters grew in size from run to run, causing increasing channeling during backwash and disruption of the settled bed surface at the end of backwash. Figure 3 (a) shows a large cluster of mudballs close to the sand- anthracite interface in the dual media filter.

The authors attempted to reduce the mudballing in the laboratory filters by three different means: (1) skimming both the sand and anthracite layers in one of the laboratory filter to remove fines, as suggested by Kawamura (1975a) (2) reducing the coagulant dose and backwashing the filters at 24 hour intervals regardless of whether terminal headloss had been reached (3) increasing the maximum backwash rate to 55 m/h and the volume to 3,2 m 3/m2 (equivalent to adding 1 m to the backwash water reservoir). None of these approaches appeared to have much impact on mudball formation. However, at the same time, the presence of the mudballs did not appear to be having any noticeable impact on filtration performance. It was therefore concluded that some mudballing was inevitable but not necessarily disastrous. The focus of the laboratory scale experiments then shifted to what would happen to the filter media over long-term and how the filter performance would be affected.

Figure 3(a).

Mudball formation in dual media

Figure 3(b)

Mudballs found at the interface in the AVF

Mudball formation in the AVF and the dual media laboratory filter appeared to be confined to the anthracite layer while the 0,5 mm sand layer beneath remained remarkably clean and free of mudballs even after 15 months of operation. Examination of the anthracite from both filters provided clues to the mechanisms involved. A layer of floc coated media at the top surface of the filter bed was clearly visible during backwashing of the dual media laboratory filter. When the bed settled back after backwashing, a thin coating of sludge was evident on the surface of both the full scale and laboratory filter. The top 1 cm o the f bed was also impregnated with sludge. Sludgy areas were distributed throughout the entire anthracite layer. Near the top of the bed, small clusters of anthracite were loosely glued together by patches of relatively soft floc. Deeper in the anthracite, the mudballs became larger, harder and contained increasing amounts of relatively coarse sand. It is thought that this sand may have been carried into the anthracite layer by high velocity channels which developed between the larger mudballs floating at the interface during backwashing. The amount of sand in the anthracite layer of the AVF was greater than in the laboratory filter.

This was possibly the result of more severe mudballing in the former and poor flow distribution resulting from the internal structure of the filter compartment. Mudballs at the interface were entirely coated with sand but their sand content decreased towards their centres where the material was similar to that observed in mudballs near the top of the bed. After 8 months of operation, the interface in the AVF consisted almost entirely of mudballs ranging in size from 4 to 15 cm. Figure 3 (b) shows an example of a mudball removed from the sand/anthracite. The mudball has been sliced open displaying an external coating of sand whilst internally the contents are predominantly anthracite. These observations in dual media filters suggest that mudballing is initiated in the anthracite near the surface of the bed and as the mudballs become larger and heavier they sink deeper into the bed where they accumulate sand. It is likely that these mudballs start off as fairly small structures, comprising individual grains to which floc is attached. After backwashing these partially cleaned grains settle and become compacted together. The floc deposits possibly act as nuclei for the formation of new inter-grain floc bonds which are stronger and more likely to survive the next backwash than new bonds forming on clean media surfaces. Once grains are no longer separated by backwashing the crevices between them tend to fill up with floc resulting in the formation of smooth, rounded structures. Similarly, clusters of small mudballs eventually fuse into larger structures which sink deeper into the media. The increase in the size of mudballs from run to run implies an increase in floc strength with age which allows larger and larger structures to survive backwash. In dual media filters, bulk density effects appear to keep the larger mudballs floating on top of the sand layer, which is why they accumulate a the t interface (Stevenson, 1998). The same mechanisms probably apply to mudball formation in monomedia filters except that the mudballs will eventually sink to the filter floor as they grown in size. Apart from the mudballs, the 0,7 mm sand excavated from the laboratory filter after 1 month of operation was relatively clean below the first 1 cm from the bed surface, compared to the anthracite in the dual media filter. The top surface of the bed was however also coated with a thin layer of sludge. The amount of accumulated mud recovered from the two laboratory filters at the end of one month of operation approximately the same. After 15 months of operation the anthracite layer in the AVF was found to be relatively thin in the middle of the filter while it had piled up against sides of filter. Mudballs in the regions close to the walls and in particular, between the walls and the internal pipework were larger than those in the centre of the filter and in some places large areas of media had become fused together in a cake structure rather than clusters of discrete mudballs. It appears that at this point, increasingly poor fluidisation near the walls was resulting in the deposition of anthracite at the edges of the filter while most of the backwash flow was channeled through the centre of the filter. This phenomenon may have been linked to the ultimate decline in filtration performance discussed in Section 6.3.

6.2

Evaluating the efficiency of wash water usage

The volume of backwash which was effective in cleaning the media was estimated for various raw water and dosing conditions by measuring the backwash turbidity in the laboratory filters as a function of backwash time. The volume of effective backwash was defined to be the volume required for the effluent turbidity to reach 10 NTU. Floc was considered detached from the media once it had traveled up past the surface of the bed. It was assumed to be removed from the filter during backwashing only if it reached the backwash trough in the case of the laboratory filters, or the highest point of the backwash pipe in the case of the AVF before backwashing stopped. If not, it was assumed to settled back onto the filter bed after backwash was terminated. Figures 4 (a) (d) show some of the backwash results obtained for different operating conditions. Backwash results for other filters ( 10 = 0,5 and 0,8 mm monomedia sand filters d and other dual media designs) showed similar results (Pryor & Brouckaert, 2001). Backwash turbidities measured at the laboratory filter backwash pipes are plotted against the cumulative backwash volume at the time the same volume of fluid passed through the top of the expanded bed. This was calculated by subtracting the estimated volume between the bed surface and the sampling point from the cumulative backwash volume at the time the sample was taken. The expanded bed height was assumed to be the average of the maximum and minimum clean bed heights measured over the range of backwash rates utilised. Dispersion effects were not taken into account. In Figure 4(a), backwash was terminated before the turbidity reached 10 NTU. The tail of the curve was extrapolated using the following equation to estimate the additional volume required:

ln (ln (c )) = A V + B
c A, B Effluent turbidty, NTU Fitting parameters

(2)

Based on the limited number of results obtained, the effective backwash volume appeared to depend mainly on the amount and strength of (detachable) floc in the bed and the type of the filter bed. The dual media filters always required approximately 0,5 m 3/m 2 more backwash volume to reach 10 NTU than the monomedia filters. Estimation of the pore volume of the expanded anthracite layer indicated that the travel time from the sandanthracite interface could not have contributed much more than 1 % of this difference, indicating a slower release of floc from the anthracite grains than from the sand.

Turbidity Run A Extrapolated turbidity Run B Vb(d60)

Turbidity Run B Backwash velocity Vmf(d10)

Extrapolated turbidity Run A Vmf(d60) Vb (d10)

Backwash Turbidity Vmf(d60)

Backwash Velocity Vb (d10)

Vb(d60) Vmf (d10)

10000

60

1000

60

Backwash Turbidity (NTU)

50 1000 40

50

Backwash Turbidity (NTU)

Backwash Velocity (m/h)

100

40

100

30

30

20 10 10

10

20

10

1 -0.5

0 0 0.5 1 1.5 2
3 2

2.5

1 -0.5

0 0 0.5 1 1.5 2
3 2

2.5

Cumulative Volume Backwash (m /m )

Cumulative Volume Backwash (m /m )

Figure 4(a). Backwash turbidity leaving filter bed 0,7 mm sand: filter influent 20 120 NTU, 15 30 mg/L alum, Backwash rates: 46 - 25 m/h. 3 2 Backwash volume = 2,2 m /m
Backwash turbidity Vb(d60) Backwash Velocity Vmf (d10) Vmf(d60) Vb (d10)

Figure 4(b): Backwash turbidity leaving filter bed 0,7 mm sand: filter influent 2-7 NTU, 5 mg/L alum, Backwash rates: 46 - 25 m/h. 3 2 Backwash volume = 2,2 m /m
Turbidity Run A Vb(d60) Turbidity Run B Vmf (d10) Backwash Velocity Vb (d10) Vmf (d60)

1000

60

1000

60

50

50

Backwash Turbidity (NTU)

Backwash Turbidity (NTU)

Backwash Rate (m/h)

100

40

100

40

30

30

10

20

10

20

10

10

1 -0.5

0 0 0.5 1 1.5 2
3

2.5
2

3.5

1 -0.5

0 0 0.5 1 1.5 2 2.5 3 3.5

Cumulative Volume Backwash (m /m )

Cumulative Volume Backwash

(m 3/m2 )

Figure 4(c). Backwash turbidity leaving filter bed 0,7 mm sand: filter influent 1 - 3 NTU, 6 mg/L alum, backwash rates: 55 - 25 m/h. 3 2 Backwash volume = 3,2 m /m

Figure 4(d). Backwashing turbidity leaving filter bed Dual media: filter influent 1 - 3 NTU, 6 mg/L alum, backwash rates: 55 - 25 m/h. 3 2 Backwash volume = 3,2 m /m

Depending on the expanded bed height, the backwash volume required to carry floc from the surface of the filter bed to the point of no return in the AVF backwash pipe, was approximately 0,6 to 0,8 m 3/m 2. Based on the results presented above, the maximum effective backwash volumes were estimated to be 2,5 m 3/m 2 and 3,1 m 3/m 2 respectively for monomedia sand and dual media filters operating at low influent turbidities, and 3,1 m 3/m2 and 3,6 m 3/m 2 respectively for high influent turbidities. While these values are relatively low compared to typical design specifications for conventional filters, it appears that additional wash water would be wasted unless the efficiency of floc detachment is improved.

Backwash Rate (m/h)

Backwash Velocity (m/h)

Kawamuras recommendations (1975b) for optimum backwash rates (Eqn 1) are based on a theoretical analysis of the maximum cleaning action on the average grain size (d60). However, mudballing appears to be initiated in the top layer of the bed where the finest grain sizes accumulate. Therefore it is possible that backwash efficiency may not be significantly impacted if the backwash velocity towards the end of backwashing drops below the optimum backwash rate or even the minimum fluidisation velocity for the average grain size provided that conditions remain close to optimal for the finest grain sizes. Figures 4 (a) (d) also show the minimum fluidisation velocities (Vmf) and appropriate backwash velocities for the d10 and d60 sizes of both the 0,7 mm sand and the anthracite. There were no discernible trends in volume of effective backwash with respect to the optimum and minimum fluidisation velocities for different grain sizes and more experimentation is needed in this area. Ideally, the average backwash rate should be the optimum rate (Vb(d60)) and the minimum rate should not be less than ( b(d10)). However, V since efficiency of backwash is probably not very sensitive to the backwash rate in the region of the optimum (Amirtharajah, 1978b), there may not be much advantage in using such high rates. The backwash rate should however, not drop below the minimum fluidisation rate of the bed. 6.3. Long term filtration performance

If backwash is ineffective, filtration performance is ultimately affected. In order to understand whether the appearance of mudballs in the media necessarily implied a decline in filter performance, the 0,7 mm sand and dual media laboratory filters were closely monitored while being operated as consistently as possible over a total run time of approximately one month. The pilot plants were backwashed with 3,2 m3/m2 instead of 2,2 m 3/m2 and 55 to 25 m/h instead of 46 to 25 m/h for this series of the experiments. This was equivalent to increasing the height of the full-scale AVF by 1 m. During these experiments, the influent and effluent turbidity, raw and finished pH, filtration rate, filter run time to terminal headloss and filtrate volume produced per run, coagulant dose, temperature and fixed bed (down flow) headloss profile at the end of backwashing were monitored. Significant accumulation of mudballs was observed and the setteld bed heights of both the dual and monomedia filters were found to increase (~2 cm in the dual media filter over the period of a month). It was however not immediately obvious whether the filter performances were adversely affected. Figure 5 shows the influent and filtrate turbidities and volume of filtrate produced per run (including the volume in the backwash tank) plotted as function of cumulative volume filtered for the dual media laboratory filter. (The laboratory filters were thoroughly cleaned using combined air and water wash as well as fluidised bed washing before the start of the experiment). Filtrate turbidities produced by the 0,7 mm sand filter were similar to the dual media filter, but the filter run volumes were slight lower on average (~ 140 m 3/m 2 compared to ~160 m 3/m 2 for dual media).

Coagulated Water NTU

8.0 6.0 4.0 2.0 0.0 1.2 1.0 0.8 0.6 0.4 0.2 0.0 200.0 180.0 160.0 140.0 120.0 100.0 0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000

500

1000

1500

2000

2500

3000

3500

4000

4500

5000

Volume filtered per run

Filtrate NTU

500

1000

1500

2000

2500

3000

3500

4000

4500

5000

Cumulative Volume Filtered

(m3/m2 )

Figure 5. Extended pilot-scale filtration using dual media (filtration rate = 6 7 m/h, alum dose = 5,5 6,5 mg/L, filtrate pH = 7,2 8,5)

Since long term trends could have been obscured by short term fluctuations in filtration rate, coagulant dose and raw water characteristics, the data was analysed using stepwise multiple regression (a model selection procedure which determines which variables are significantly correlated with certain specified performance parameters). The analysis was carried out using the statistical software package SPSS 9.0 for Windows. Variations in filtrate turbidity were analysed in terms of changes in cumulative volume filtered, volume filtered since the last backwash, coagulant dose, influent turbidity and filtration rate. The data was divided into four regions based on the filter ripening curve so that the relationship between filtrate turbidity and volume filtered could be approximated as linear for each region. Variations in filter run volume (volume of filtrate produced per filter run) were analysed in terms of changes in cumulative volume filtered, dose, average filtration rate and water viscosity. The primary aim of the statistical analysis was to determine whether there was any discernable decline in filter performance (decrease in filter run time or in increase in filtrate turbidity) with changes in bed structure over time (represented by cumulative volume filtered) once the contribution of other factors such as influent turbidity had been accounted for. However, the model selection results suggested that either there was no change in performance due to mudball accumulation over the monitoring period, or that in fact the filter performance may have improved over time. It was therefore concluded that the AVF could probably operate satisfactorily for an extended period of time, even with sub-optimal backwash.

This was confirmed by results obtained at full scale. The AVF was operated for a period of 15 months during which time the raw water turbidity remained < 10 NTU. For most of this period, filtrate turbidities < 1 NTU and adequate run times (> 15 hours) were consistently achieved, provided that the coagulant dosing system was operating properly and coagulant doses were adjusted to accommodate changes in raw water. However, after 14 months of operation the filter run times dropped to less than 10 hours in spite of decreases in coagulant dose. Inspection of the filter media suggested that increasingly poor backwash flow distribution was depositing anthracite from the centre of the filter in relatively stagnant and mudballed regions at the edges. The decrease in run times could have been related to a decrease in effective filter area and/or loss of anthracite from the active regions of the filter. 7. CONCLUSIONS AND RECOMMENDATIONS Coagulation is required to achieve filtrate quality standards in the AVF. The operation of the filter will therefore only be as reliable as the coagulant dosing system and local operators have to be trained to adjust coagulant doses when necessary. Water only backwashing in the AVF is not very efficient and mudballing of the media is inevitable, however, the filter may still perform adequately for up to a year or more before any intervention is required. Improving the backwash design in the AVF may extend the life of the filter bed. Backwash volumes of at least 3,1 m 3/m 2 and 3,6 m 3/m 2 respectively for monomedia sand and dual media beds are recommended for the AVF. Particular attention should be paid to ensuring that the available backwash rates are compatible with the media selected and that backwash water distribution is as even as possible. The AVF should be designed with the option of manually controlled air scour. Air scour could be carried out by a skilled operator during a regular (e.g. quarterly) inspection of the plant in order to reverse filter mudballing. Provision should be made in the mechanical design of the AVF to be able to select an appropriate terminal headloss to prevent excessively long run times at low raw water turbidities, and turbidity breakthrough before backwashing at high raw water turbidities. Local operators should be trained to initiate manual backwash if the AVF does not backwash automatically at least once a day or if turbidity breakthrough occurs, however, skilled personnel would be required to determine when the terminal headloss setting should be changed. The suitability and cost-effectiveness of using AVF's for rural water supply may depend on the frequency of skilled operator intervention required. ACKNOWLEDGEMENTS

8.

The authors would like to thank Umgeni Water for the use of the Process Evaluation Facility and allowing the presentation of this paper, as well as the Water Research Commission of South Africa for their financial support.

REFERENCES 1. AMIRTHARAJAH A. (1978a) Design of granular media filtration units, Chapter 28 in Water Treatment Plant Design, Sanks RL (ed.), Butterworth-Heinemann, Stoneham, MA. AMIRTHARAJAH A. (1978b) Optimum backwashing of Sand Filters, J. Envir. Eng. Div., ASCE, 104(10), 917 932. AMIRTHARAJAH A, (1980) Closure of Optimum backwashing of sand filters. J. Envir. Eng. Div. ASCE, 106(4), 859. BAUMANN ER, (1978) Granular-Media Deep-Bed Filtration, Chapter 12 in Water Treatment Plant Design, Sanks RL (ed.), Butterworth-Heinemann, Stoneham, MA. BAYLIS JR (1959) Review of filter bed design and methods of washing, J. AWWA, 51(11), 1433-1454. CLEASBY JL, (1990) Filtration, Chapter 8 in Water Quality and Treatment: A Handbook of Community Water Supplies, 4th ed., AWWA, pp 455 560 IVES KJ, (1990) Testing of Filter Media, J. Water SRT-Aqua, 39(3), pp 144-151 KAWAMURA S, (1975a) Design and operation of high-rate filters Part 1., J. AWWA, 67(10), 535-544. KAWAMURA S, (1975b) Design and operation of high-rate filters Part 2., J. AWWA, 67(11), 653-662. KAWAMURA S, (2000) Integrated Design and Operation of Water Treatment Facilities, 2nd ed., John Wiley and Sons, Inc., New York. KAWAMURA S, (2001) Design and Operation of Filters with Granular Media, International Conference on Advances in Rapid Granular Filtration in Water Treatment, CIWEM, 4-6 April, London, pp 149-160 MONK RDG and WILLIS JF, (1987) Designing Water Treatment Facilities, J. AWWA, 79 (2), pp 45 57 PRYOR M AND BROUCKAERT BM, (2001) Research on optimisation of an autonomous valveless filter for the cost-effective production of potable water for rural areas. (K5/919), Draft Report to the South African Water Research Commission, June 2001. STEVENSON DG, (1998) Water treatment unit processes, Imperial College, London.

2. 3. 4. 5. 6. 7. 8. 9. 10. 11.

12. 13.

14.

You might also like