WER 2008 Levine Cloth Media

Embed Size (px)

Citation preview

  • 8/3/2019 WER 2008 Levine Cloth Media

    1/13

    Pathogen and Indicator OrganismReduction Through Secondary

    Effluent Filtration: Implications forReclaimed Water Production

    Audrey D. Levine1*, Valerie J. Harwood2, Samuel R. Farrah3, Troy M. Scott4, Joan B. Rose5

    ABSTRACT: The reduction of pathogens and indicator organisms

    through secondary effluent filtration was investigated at six full-scale

    treatment facilities, ranging in capacity from 0.04 to 1 m3 /s (1 to 25 mgd).

    Grab samples were assayed for pathogens (cultivable enteric viruses,Giardia, a nd Cryptosporidium) and indicator organisms (coliforms,

    enterococci, Clostridium perfringens, and coliphages) quarterly under peak

    flow conditions from each facility over the course of 1 calendar year. Log10removals resulting from filtration averaged 0.3 to 0.8 log10 for cultivable

    enteric viruses, 0.4 to 1.5 log10 for protozoan parasites, 0.01 to 3.7 log10 for

    indicator bacteria, and 0.3 to 1.1 log10 for coliphages. In addition to filter

    design (cloth, monomedium shallow- or deep-bed, or dual-media filters),

    differences in reduction of pathogens and indicators could be attributed to

    the combined effects of hydraulic loading rates, chemical addition practices,

    backwashing and postbackwashing operating strategies, and the effective-

    ness of upstream biological treatment and sedimentation. Water Environ.

    Res., 80, 596 (2008).

    KEYWORDS: reclaimed water, pathogens, wastewater filtration,

    prechlorination.

    doi:10.2175/106143008X266742

    Introduction

    With increasing use of reclaimed water for public access

    applications, such as landscape irrigation, urban water uses,

    wetlands and surface water augmentation, and recirculating cooling

    water for buildings and other industrial applications, the effective-

    ness of treatment practices for control of pathogens has become

    more important. In many facilities, filtration is used between

    secondary biological treatment and disinfection (either chemical or

    photochemical) to reduce the concentration of suspended solids and

    particle-associated microorganisms that may interfere with disin-

    fection effectiveness. Because it is now recognized that some

    pathogens, such as Cryptosporidium, are resistant to disinfection,

    secondary effluent filtration represents a critical step to control

    pathogens that are not removed by biological treatment. The

    effectiveness of pathogen reduction through filtration is influenced

    by filter characteristics and operating practices, microbial properties

    (size, surface properties, and degree of association with other

    microorganisms or particles), and water quality variables. A wide

    variety of filters has been adopted for water reclamation applica-

    tions, including conventional deep-bed monomedia or dual-media

    filters, continuously backwashed upflow filters, disk (cloth) filters,

    fuzzy filters, and membranes (Bourgeous et al., 2003; England

    et al., 1994). To meet requirements for process reliability and

    maintenance of flow, secondary effluent filtration typically consists

    of a minimum of two filters operated in parallel, in either a declining

    rate or constant rate mode. Typically, effluents from individual

    filters are combined before disinfection. Direct monitoring of the

    microbial characteristics of filter effluents is not widely practiced, asthe point of compliance for most reclaimed water systems is after

    disinfection.

    The purpose of this paper is to evaluate the role of filtration for

    removal of pathogens and indicator organisms during production

    of reclaimed water and the degree to which traditional models of

    surface water filtration can be applied to reclaimed water filtration

    scenarios. The data presented in this paper are derived from testing

    conducted at six full-scale water reclamation facilities that use either

    granular media filtration (anthracite and/or sand) or cloth filters

    (Rose et al., 2004). Relationships between filter characteristics

    (depth, hydraulic loading, and backwashing practices); water qual-

    ity; and removal or persistence of viruses, protozoan pathogens,

    conventional indicator bacteria, and alternative microbial indicatorsthrough full-scale secondary effluent filtration are evaluated.

    Background

    Individually and collectively, treatment processes provide

    a combination of physical, chemical, and biological barriers to

    reduce pathogen levels in reclaimed water. To verify treatment

    effectiveness, disinfected effluents are routinely assayed for

    bacterial indicators and sampled intermittently for viruses and/or

    protozoan pathogens (U.S. EPA, 2004; York and Walker-Coleman,

    2000). Coliform bacteria (either the total coliform group or

    thermotolerant [fecal] coliforms) are used almost universally to

    benchmark the microbiological quality of reclaimed water. In the

    United States, monitoring requirements (microbial analyte[s] and

    1 Office of Research and Development, U.S. Environmental ProtectionAgency, Washington D.C.

    2 Department of Biology, University of South Florida, Tampa, Florida.

    3 Department of Microbiology & Cell Science, University of Florida,Gainesville, Florida.

    4 BCS Laboratories Inc., Miami, Florida.

    5 Department of Crop and Soil Sciences, Michigan State University, EastLansing, Michigan.

    * Office of Research and Development, U.S. Environmental ProtectionAgency, 1200 Pennsylvania Avenue NW 8101R, Washington D.C. 20460;e-mail: [email protected].

    596 Water Environment Research, Volume 80, Number 7

  • 8/3/2019 WER 2008 Levine Cloth Media

    2/13

    frequency) and specific numerical limits are set by state and local

    regulatory agencies and, typically, are based on daily monitoring of

    disinfected effluents.Filtration Removal Mechanisms. Removal of microorgan-

    isms through granular media filtration is governed by particle-media

    interactions that result from diffusion, interception, and sedimen-

    tation (Camp, 1964; Darby et al., 1991; Elimelech, 1992; Ghosh

    et al., 1975; Glasgow and Wheatley, 1999; Habibian and OMelia,1975; Hall and Fitzpatrick, 2000; Iwasaki, 1937; Jegatheesan and

    Vigneswaren, 2000; Kavanaugh et al., 1978; OMelia and Ali, 1978;

    Stevenson, 1997; Tchobanoglous and Eliassen, 1970; Trussell and

    Chang, 1999; Yao et al., 1971). The deposition of microorganisms

    and other particles in filters is dependent on effective transport to the

    filter media coupled with retention within the filter pores or at-

    tachment to the media surface.

    The single collector efficiency concept has been widely used to

    evaluate the overall efficiency of filtration for removal of particles

    by defining a nondimensional single collector efficiency, g, which

    reflects the ratio of particles striking the filter media to the total

    particle flux by incorporating the combined effects of diffusion

    (gD), interception (gI), and gravity sedimentation (gG) on particle-media interactions, as defined in eq 1 (Camp, 1964; Elimelech,

    1992; Habibian and OMelia, 1975; Harvey and Garabedian, 1991;

    Iwasaki, 1937; Logan et al., 2001; Redman et al., 2001; Yao et al.,

    1971).

    g gD gI gG 0:9kT

    ldpdv

    !2=3

    1:5dp

    d

    2qp qgd

    2

    p

    18lv1

    Where

    g 5 overall collision frequency coefficient;

    gD 5 collision frequency coefficient resulting from diffusion;gI 5 collision frequency coefficient resulting from interception,

    gD;

    gG 5 collision frequency coefficient resulting from gravitational

    sedimentation;

    k5 Boltzman constant (1.38 3 10223 J/K);

    T 5 absolute temperature (K);

    l 5 dynamic viscosity of water (N-s/m2);

    dp 5 effective particle (microorganism) diameter (m);

    d5 effective diameter of media grain (m);

    v 5 water velocity 5 flowrate/surface area (m/s);

    q 5 density of water (kg/m3); and

    qp 5 particle (microorganism) density (kg/m3).

    In this model, transport by diffusion is inversely proportional to

    the media grain size (dp) and filtration velocity (v). Transport by

    interception is controlled by the ratio of the microorganism size (dp)

    to the grain size (d) and is independent of filtration velocity, while

    transport by sedimentation is proportional to the ratio between the

    particle settling velocity and the filtration velocity (v) or hydraulic

    loading rate. The model does not directly incorporate water quality

    variables, such as pH, ionic strength, or organic content.

    A comparison of the theoretical single collector removal effi-

    ciency for control of protozoa, bacteria, and viruses through sand

    and anthracite media typically used for production of reclaimed

    water suggests that individual bacterial cells are likely to have the

    lowest removal efficiency compared with freely suspended viruses,

    protozoa, or aggregates (flocs) of microorganisms and other particu-

    lates (Figure 1). Essentially, removal of freely suspended viruses

    and other nanoscale particles is controlled by diffusion, while re-

    moval of protozoa is influenced through the combined effects of

    differential sedimentation and interception. Removal of freely sus-

    pended bacterial cells is influenced by the cumulative effects of

    diffusion, differential sedimentation, and interception. The effective

    grain size of the media affects the collector efficiency for removal of

    viruses and bacteria, whereas removal of protozoa and/or microbialaggregates is more affected by hydraulic loading rates than by

    media characteristics. It should also be noted that removal of bac-

    teria through filtration is likely to be less sensitive to the influence of

    variable hydraulic loading rates than removal of viruses or protozoa

    (see eq 1).

    These calculations suggest that the use of bacterial indicators may

    provide a conservative assessment of filter performance for removal

    of microorganisms, if bacteria are present as discrete particles (not

    associated with other particulate matter), and disinfectant chemicals

    are not added upstream. From a process design perspective, this

    model suggests that smaller grain size media would affect removal

    of freely suspended viruses and other nanoparticles, whereas lower

    hydraulic loading rates would be more effective at improvingremoval efficiency for protozoan pathogens. In addition, the use of

    upstream coagulation to aggregate freely suspended viral and bac-

    terial particles should be designed in conjunction with optimization

    of hydraulic loading rates. This model does not address the devel-

    opment of headloss through filter runs, the interplay of media size

    and hydraulic loading rates on headloss development, or potential

    effects of coagulant chemicals and water quality on headloss and

    filter run length.

    While the collector efficiency concept describes the likelihood

    of particles colliding with filter media, the net removal depends

    on retention within the filter. Factors that are likely to influence

    attachment include the net surface charge on the filter media and

    microbial surfaces; media properties (type, size, and depth);hydraulic loading rates; upstream chemical use (oxidants and/or

    coagulants); water quality variables; flow control; and backwashing

    and post-backwashing practices. The characteristics of secondary

    effluent that might affect filter and microbial surface characteristics

    include pH, ionic strength, temperature, particle characteristics, and

    characteristics of dissolved organics (concentration, molecular size,

    and charge density). High ionic strength and residual organics act

    to decrease the electric double layer surrounding microorganisms,

    particles, and filter media, thereby increasing the potential for

    particle attachment (Brown and Abramson, 2006; Elimelech and

    OMelia, 1990; Hsu and Huang, 2002; Hsu et al., 2001). Ionic

    strength varies with total dissolved solids levels in reclaimed water,

    which can range from 200 to 2500 mg/L, depending on the char-

    acteristics of the service area source water quality, water usage

    patterns, stormwater effects (infiltration and inflow), and sources of

    dissolved solids from industrial discharges and point-of-use water

    treatment discharges. The presence of residual organics may affect

    adsorption to filter media or may serve to improve transport through

    filtration (Foppen et al., 2006).

    Typically, secondary effluent contains active biomass that

    consists of microorganisms that are either freely dispersed or

    entrapped in microbial agglomerates (flocs) that can colonize the

    filter media. During a filter run, the buildup of accumulated material

    within the filter coupled with the development of biofilms can serve

    to augment filter efficiency; however, a side effect of particle

    accumulation and biofilm formation is an increase in specific

    velocities within pores, resulting in pore blockage, channeling

    Levine et al.

    July 2008 597

  • 8/3/2019 WER 2008 Levine Cloth Media

    3/13

    effects, higher effective hydraulic loading rates, and particle remo-bilization (Ahmad and Amirtharajah, 1998; Hall and Fitzpatrick,

    2000; Hozalski and Bouwer, 1998). Some particles/pathogens may

    be released during filtration, depending on the duration of the filter

    run, hydraulic inconsistencies, and effectiveness of upstream treat-

    ment (Ahmad and Amirtharajah, 1998; Darby et al., 1991; Glasgow

    and Wheatley, 1999; Jegatheesan and Vigneswaran, 2000; Jiminex

    et al., 2000).

    Backwashing and postbackwashing practices vary with filter

    designs and plant operating practices. Backwashing operations are

    designed to dislodge and scour accumulated particles from filter

    media, by fluidizing the media and applying a combination of high-

    velocity water with or without air scour (Ahmad and Amirtharajah,

    1998; Hall and Fitzpatrick, 2000; Hozalski and Bouwer, 1998).

    Operating variables, such as backwashing frequency and duration,

    degree of fluidization, backwash source water (chlorinated or non-

    chlorinated), degree to which air scour is used and monitored, and

    approach for returning a filter to service following backwashing, can

    affect the efficiency of subsequent filtration cycles. In general, back-

    washing is not designed to remove biofilms (Wang et al., 1995).

    Particle/pathogen release may occur following backwashing, re-

    sulting from either remnant particles (Amburgey et al., 2004) or

    hydrodynamic inconsistencies in the filter media (Glasgow and

    Wheatley, 1999).

    Modeling of filtration effectiveness for reclaimed water applica-

    tions is complicated by the heterogeneity of the particles in

    secondary effluent coupled with the lack of tools to quantify the

    effects of particle accumulation and filter porosity on interstitial

    fluid velocities (Redman et al., 2001). One approach that has beenwidely used is to assume that pathogen deposition through filtration

    can be modeled as pseudo-first-order removal, as shown in eq 2,

    where Cis the number concentration of particles or microorganisms

    entering a unit volume of the filter, k is a filter coefficient, and L

    is depth (Brown and Abramson, 2006; Camp, 1964; Harvey and

    Garabedian, 1991; Iwasaki, 1937; Reddi, 1997; Trussell and Chang,

    1999). The filter coefficient, k, reflects the composite effects of filter

    media characteristics, hydrodynamic factors, and water quality

    variables.

    dC

    dL kC 2

    Where

    C5 number concentration of microorganisms (#/100 mL or

    #/100 L),

    L5 depth from surface of filter (m), and

    k 5 filter coefficient (m21).

    Under steady-state conditions, the concentration of microorganisms

    in the filter effluent can be estimated as follows:

    C Coekz 3

    Where

    Co 5 concentration of particles or microorganisms applied to the

    surface of the filter (#/100 mL or #/100 L), and

    z 5 filter depth (m).

    Figure 1Predicted removal efficiency of viruses, bacteria, and protozoa based on hydraulic loading rates (1.0 to

    3.4 L/s) and media size (0.3, 0.5, and 1 mm) typical of secondary effluent filtration. Calculations based on eq 1.

    Assumed sizes were viruses 2 lm.

    Levine et al.

    598 Water Environment Research, Volume 80, Number 7

  • 8/3/2019 WER 2008 Levine Cloth Media

    4/13

    This model suggests that, for a given filter coefficient, k, filter

    efficiency should increase with depth. However, in practice, filter

    depth and filtration run length are limited by headloss and other

    practical constraints. Another way of increasing filter performance

    is to increase the effective filter coefficient, k, through upstream

    chemical addition of coagulants, flocculants, and/or oxidants.Review of Pathogen Removal Through Surface Water

    Filtration in the Context of Water Reclamation. Pathogen re-moval through filtration has been studied extensively in the produc-

    tion of drinking water from surface water sources (Amburgey et al.,

    2004, 2005; Bustamante et al., 2001; Edzwald et al., 2000; Emelko

    et al., 2003; Harrington et al., 2003; Hsu and Huang, 2002; Logan

    et al., 2001; Nasser et al., 1995; Nieminski and Ongerth, 1995;

    Patania et al., 1995; Swertfeger et al., 1999; Xagoraraki et al., 2004).

    In potable water systems, it hasbeen observed that removal ofGiardia

    cysts and Cryptosporidium oocysts is influenced by the degree of

    filter maturation; use of coagulant chemicals (Koivunen et al., 2003;

    Mosher and Hendricks, 1986; Patania et al., 1995); ionic strength;

    pH; zeta potential (Bustamante et al., 2001; Hsu and Huang, 2002);

    and filter grain size (Logan et al., 2001; Stevenson, 1997). In general,

    the surface charge of cysts and oocysts is negative under neutral andbasic pH conditions, with the zeta potential for Cryptosporidium

    approximately double that ofGiardia at neutral pH (Hsu and Huang,

    2002; Ongerth and Pecoraro, 1996), suggesting that removal of

    Cryptosporidium through filtration maynot parallel Giardia removal,

    particularly under low-ionic-strength and/or alkaline pH conditions.

    While reclaimedwaterfiltration is similar to surface water filtration, in

    principle, limited data are available on the relative roles of pH, ionic

    strength, and organic matter in reclaimed water applications. Because

    indicator organisms (coliform bacteria and coliphages) are used in

    conjunction with indirect measures of microbial concentrations

    (suspended solids and turbidity) to assess the performance of re-

    claimed water filtration, it is important to identify the appropriate

    parameters that should be monitored to facilitate assessment andprediction of pathogen removal through filtration.Comparison of Regulatory Approaches for Surface Water

    Filtration and Reclaimed Water Production. Because of con-

    cerns about the protozoan pathogens Giardia and Cryptosporidium

    in water supplies, the U.S. Environmental Protection Agency

    (Washington, D.C.) (U.S. EPA) has implemented rules that mandate

    that surface-water-treatment facilities include robust filtration sys-

    tems that meet stringent turbidity requirements and achieve at least

    a 2-log reduction of protozoan pathogens (40 CFR 141.721 [U.S.

    EPA, 2006]). For example, under the Long-Term 2 Enhanced Sur-

    face Water Treatment Rule enacted in 2006, at least 95% of the

    combined filter effluent turbidity measurements must be less than

    or equal to 0.15 NTU (40 CFR 141.727) (U.S. EPA, 2003). While

    turbidity levels do not correspond to pathogen concentrations, the

    rule is based on conclusive evidence from a wide array of surface

    water filtration studies, which demonstrated that turbidity levels

    below 0.2 NTU are likely to correspond to low or nondetectable

    levels of Giardia and Cryptosporidium (Edzwald et al., 2000;

    Emelko et al., 2005; Xagoraraki et al., 2004). In contrast, turbidity

    (or suspended solids) requirements for water reclamation facilities

    are based on monitoring of the disinfected effluent, with typical

    turbidity limits ranging from 2 to 5 NTU or suspended solids limits

    of 5 mg/L, depending on permit requirements (U.S. EPA, 2004;

    York and Walker-Coleman, 2000). Thus, surface-water-filtration

    practices cannot be directly extrapolated to reclaimed water produc-

    tion, as a result of the order of magnitude differences in turbidity

    limits and other confounding variables.

    Role of Filtration in Reclaimed Water Production. In

    reclaimed water production, filtration is intended to provide for

    removal of microorganisms and associated turbidity, as a pretreat-

    ment to disinfection. Disinfection systems are designed to inactivate

    pathogens that are not removed through upstream processes. His-

    torically, wastewater treatment facilities have relied on the use of

    chlorine as a disinfectant. However, because of increasing concerns

    about the presence of disinfection byproducts in treated wastewaterand the need to dechlorinate effluents that are discharged to

    receiving waters, there has been a shift towards the use of UV and

    other alternative disinfectants. In addition, some pathogens are re-

    sistant to chlorine. As disinfection systems become more stream-

    lined, the role of filtration becomes more important for pathogen

    control. In addition, pathogen removal through filtration is even

    more critical for control of pathogens that are resistant to chemical

    or photochemical disinfection.

    Methodology

    The reduction of pathogens through filtration was assessed

    through sampling of six wastewater treatment facilities under peak-flow conditions. A summary of the characteristics of the treatment

    facilities is given in Table 1. The filters tested in this project

    included granular media (deep-bed and shallow-bed) and cloth

    filters. Prechlorination was applied in one of the facilities (S11);

    cationic polyelectrolyte was applied upstream of filtration in another

    (A48); and a third facility used alum in the secondary clarifier

    (A24S48), while the remaining facilities did not use any upstream

    chemicals during the testing program. Backwashing practices varied

    among the facilities, but four of the six facilities used chlorinated

    effluent on a routine basis, as detailed in Table 1. A minimum of

    four sampling events were conducted at each facility, over a 1-year

    period, to assess pathogen and indicator concentrations associated

    with untreated wastewater and effluents from biological treatment(postsedimentation), filtration, and disinfection. For each sample

    event, filter operating data were compiled, including hydraulic

    loading rates, backwashing history, and operational anomalies.

    Routine water quality monitoring data, such as the 5-day test for

    carbonaceous biochemical oxygen demand (C-BOD5), total sus-

    pended solids (TSS), turbidity, ammonia, and phosphorus were

    obtained from each facility; however, all facilities did not monitor all

    parameters on the same schedule as the microbial sampling program.

    A side-by-side comparison of the effects of prechlorination on

    removal of indicator organisms was conducted over a 10-hour filter

    run at the facility that prechlorinates (S11). In addition, intensive

    testing of a single filter run, to evaluate the consistency of removal

    of bacterial indicators, was conducted at three of the facilities

    (S11, A48, and A24S48). The effect of backwashing on protozoan

    pathogen concentrations (Giardia and Cryptosporidium) in filter

    effluents immediately before and after putting a filter back in service

    was evaluated at two facilities (A48 and A24S48).

    Nine independent measurements of microbiological water quality

    were used to assess filtration performance. A summary of the charac-

    teristics of the indicator organisms (coliform bacteria, enterococci,

    Clostridium perfringens, and coliphages) and pathogens (cultivable

    enteric viruses, Giardia, and Cryptosporidium) tested in the influent

    and effluent from each filter is shown in Table 2. The frequency

    with which each of the tested microorganisms was detected in each

    filter effluent is also reported in Table 2.

    All microbial assays were carried out as previously reported

    (Harwood et al., 2005; Rose et al., 2004). One-liter grab samples

    Levine et al.

    July 2008 599

  • 8/3/2019 WER 2008 Levine Cloth Media

    5/13

    were collected for analysis of the bacterial and viral indicators,

    while the parasite and virus samples were collected and concen-

    trated by pumping up to 100 L of water through appropriate filters.

    All chlorinated samples were dechlorinated by adding Na2S2O3to a final concentration of 0.001% immediately after collection.

    Cartridge filters (1MDS, Cuno Inc., Meriden, Connecticut) were used

    for collection and concentration of viruses, while Pall-Gelman

    Envirochek HV capsule filters (Pall Gelman Laboratory, Ann Arbor,

    Michigan) were used for collection and concentration of parasites.

    Bacteria were quantified using membrane filtration (47 mm;

    0.45-lm pore size). When necessary, samples were diluted with

    phosphate-buffered saline, and all were filtered in triplicate. Total

    coliforms and fecal coliforms (FC) were enumerated according to

    Standard Methods (APHA et al., 1998)total coliforms on mEndo-

    LES agar at 378C and fecal coliforms on mFC agar at 44.58C.

    Enterococcus spp. were enumerated using method 1600 (U.S. EPA,

    1996) on mEI agar at 418C. Clostridium perfringens was enumer-

    ated on m-CP plates (Acumedia Manufacturers Inc., Baltimore,

    Maryland) using anaerobic gas packs (BBL GasPak, Beckton

    Dickinson, Cockeysville, Maryland) at 458C for 24 hours, followed

    by exposure to ammonium hydroxide fumes. Colonies that turned

    red or dark pink were enumerated as Clostridium perfringens

    (C. perfringens) (Bisson and Cabelli, 1979).

    Coliphages were analyzed by the agar overlay method of Adams

    (1959) using two E. coli host strains in separate assaysE. coli HS

    (pFamp) R (ATCC #700891) and E. coli C3000 (ATCC #15597)

    and were enumerated as plaque-forming units (PFU)/100 mL.

    Viruses were concentrated by filtration followed by organic floccu-

    lation (U.S. EPA, 1996). Cultivable enteric viruses were detected by

    the observation of cytopathic effects on recently passed (,4 days)

    Buffalo Green Monkey, Rhabdosarcoma, and MA-104 cells.

    Giardia and Cryptosporidium were concentrated by filtration with

    Table 1Comparison of preliminary treatment, primary treatment, secondary treatment and filtration at wastewater

    reclamation facilities included in this study.*

    Filter label Cloth S11 A48 SU40 A30S10 A24S48

    Average capacity,

    m3 /s (mgd)

    0.04 to 0.11

    (0.9 to 2.6)

    0.61 to 0.71

    (13.9 to 16.2)

    0.42 to 0.45

    (9.6 to 10.3)

    0.06 to 0.09

    (1.3 to 2.1)

    0.5 to 1.1

    (11 to 25)

    0.06 to 0.1

    (1.3 to 2.4)

    Preliminary/primary

    treatment

    Grit removal Grit removal Grit removal

    Primaryclarifier

    Grit removal

    Equalizationbasin

    Grit removal Grit removal

    Biological treatment

    Process type Activated

    sludge

    Activated

    sludge

    Activated

    sludge

    Biological

    nutrient

    removal

    Activated

    sludge

    Biological

    nutrient

    removal

    Mean cell residence

    time (days) 6 to 8 3.5 to 6 1.6 to 2.7 8.7 to 13.3 3 to 5 8 to 16

    Filtration

    Filter type Cloth Monomedium

    automatic

    backwashing

    filter

    Monomedium Continuously

    backwashed

    upflow filter

    Dual media Dual media

    Filter media(depth)

    Fabric(3.8 mm)

    Sand (0.3 m) Anthracite(1.2 m)

    Sand (1.2 m) Anthracite (0.8 m)Sand (0.25 m)

    Anthracite(0.6 m)

    Sand (1.2 m)

    Media grain

    size (mm)

    Nominal Pore

    size: 0.01

    0.55 to 0.65 1.4 0.9 Anthracite

    Sand: 0.65

    Anthracite: 2

    Sand: 1.8

    Uniformity

    coefficient

    1.5 Anthracite: 1.3

    Sand: 1.35

    Chemical use None Prechlorinate Cationic

    polyelectrolyte

    None None Alum added to

    secondary

    clarifier

    Hydraulic loading

    rate, L/s (gpm/ft2)

    1.0 to 3.9

    (1.5 to 5.8)

    0.4 to 0.7

    (0.6 to 1)

    0.7 to 1.2

    (1.1 to 1.8)

    0.7 to 1.2

    (1.2 to 2)

    1.5 to 3.9

    (2.2 to 5.8)

    0.9 to 1.0

    (1.3 to 1.5)

    (L/d) 545 857 1300 Sand: 384 Anthracite: 300

    Sand: 667

    Volume per

    backwash,

    L/filter (gal/filter)

    4542 (1200) 11 355 (3000) 90 840 (24 000) 8% of flow 946 250

    (250 000)

    227 100

    (60 000)

    Backwashing

    frequency (hours)

    72 to 84 Automatic

    (daily)

    48 Continuous 48 to 168 48 to 168

    Backwash

    water source

    Filtered

    effluent

    Chlorinated

    effluent

    (6 mg/L

    as Cl2)

    Chlorinated

    filter effluent

    (30 mg/L

    as Cl2)

    Final effluent

    (unchlorinated);

    shock

    chlorinated

    periodically

    Chlorinated effluent

    (3 to 6 mg/L);

    shock chlorinated

    33/year

    Chlorinated

    effluent

    (5 mg/L

    as Cl2)

    * Based on operating data during sampling.

    Levine et al.

    600 Water Environment Research, Volume 80, Number 7

  • 8/3/2019 WER 2008 Levine Cloth Media

    6/13

    Pall-Gelman Envirochek HV filters and processed by immunomag-

    netic separation (Dynal Inc., Burlington, Ontario, Canada) and

    immunofluorescent antibody detection (method 1623, U.S. EPA,

    2005). Concentrated samples were enumerated by fluorescent

    microscopic counts (Giardia and Cryptosporidium) (U.S. EPA,

    1995, 1996) and by a viability assay (Cryptosporidium only) (Slifko

    et al., 1999).

    Data were sorted by filter type and microorganism concentration.

    Microbial data were log10-transformed before data analysis. In cases

    where values were below detection limits, the detection limit was

    used for statistical analyses. The performance of the individual

    filters was evaluated based on calculating microbial persistence

    through filtration (C/Co) and log10 reduction (log10 Co 2 log10 C).

    The validity of using parametric statistics to evaluate

    Table 2Comparison of characteristics of indicator organisms and pathogens tested in this study.

    Indicator/pathogen

    Example species

    or description Cell wall Shape Size (lm)

    Frequency of detection in filter effluents (%)

    Cloth S11 A48 SU40 A30S10 A24S48

    Bacterial indicators used in this study

    Total coliform Escherichia, Klebsiella,

    Citrobacter,Enterobacter

    Gram-negative,

    non-spore-forming

    Rod 0.5 to 2.0 80 100 100 100 100 100

    Fecal coliform Escherichia, some

    Klebsiella spp.,

    and other

    thermotolerant

    coliforms

    Gram-negative,

    non-spore-

    forming

    Rod 0.5 to 2.0 60 0 50 100 100 75

    Enterococci Enterococcus

    faecalis,

    Enterococcus

    faecium

    Gram-positive,

    non-spore-

    forming

    Cocci 0.5 to 1.0 80 75 50 100 100 100

    Clostridium

    perfringens

    Highly heat-resistant,

    opportunistic

    pathogen,

    producesenterotoxin

    Gram-positive,

    spore-forming

    Rod 0.6 to 1.3 3

    2.4 to 19.0

    40 100 75 75 70 100

    Coliphages Viruses that infect

    E. coli and other

    coliform bacteria

    0.025 to 0.20

    Coliphages

    that infect

    E. coli 700891

    (Famp)

    Male-specific (F1)

    coliphages, can

    only replicate when

    bacterial host cell

    is in logarithmic

    growth phase

    at .308C

    No cell wall,

    coat protein

    protects

    nucleic acid

    Icosahedral

    capsid

    0.025 40 100 100 100 14 60

    Coliphages

    that infect

    E. coli 15597

    Male-specific (F1)

    and somatic

    coliphages thatinfect E.coli

    15597 (ATTC*)

    No cell wall,

    coat protein

    protectsnucleic acid

    Icosahedral

    capsid

    0.025 40 100 100 100 14 40

    Cultivable enteric

    viruses

    Genus within the

    family Picornaviridae

    includes poliovirus,

    coxsackievirus,

    echovirus, hepatitis

    A virus;

    single-strand (ss)

    RNA genome

    No cell

    wall, Non-

    enveloped

    protein coat

    Icosahedral

    capsid

    0.025 to 0.030 60 20 80 25 67 60

    Protozoan

    parasites

    Complex life cycle.

    Zoonotic

    (animal to human)

    transmissionGiardia

    intestinalis

    Flagellated protozoan,

    Phylum Mastigophora

    Ovoid cyst ;8.5 3 10.0 100 100 83 100 50 100

    Cryptosporidium

    parvum

    Coccidian protozoan,

    Phylum Apicomplexa

    Ovoid oocysts 4.0 to 6.0 100 80 67 50 67 67

    * ATTC 5 American Type Culture Collection.

    Levine et al.

    July 2008 601

  • 8/3/2019 WER 2008 Levine Cloth Media

    7/13

    log10-transformed microbial data was tested using the Kolmogorov-

    Smirnov test for normality, with an alpha value of 0.05, using

    Graphpad Prism statistical software (Graphpad 2000, GraphPad

    Software Inc., San Diego, California). Statistically significant

    differences among filters were identified using either analysis ofvariance (ANOVA) or the Kruskal Wallis test (nonparametric).

    Post-tests included Tukey, Newman-Kuels, and Bonferroni for

    normally distributed data or the Dunns test for nonparametric

    analyses. Pairwise comparisons of filter performance were con-

    ducted using either t-tests (normally distributed data) or the Mann-

    Whitney test (nonparametric).

    Comparison of Filtration Performance

    To facilitate comparison of the filters, they were grouped into two

    categories. Group I includes filters that either use chemicals directly

    upstream of filtration (S11 [chlorine] and A48 [polymer]) or are

    influenced by surface filtration mechanisms (CLOTH), while group

    II includes deep-bed filters that contain sand (SU40, A30S10, andA24S48 [alum used in secondary clarifier]). Differences among the

    filters were evaluated based on influent concentrations and removal

    of cultivable enteric viruses, protozoan parasites (Giardia and

    Cryptosporidium), and bacterial indicators. Detailed analyses were

    conducted on group II filters to assess the effects of hydraulic

    loading rate and sand depth on reduction of pathogens.

    A summary of the water quality associated with sampling of each

    facility is given in Table 3. As shown, C-BOD5 values were

    typically below 5 mg/L for most of the facilities, except A48. TheTSS levels were typically below 4 mg/L. Nitrogen (ammonia,

    nitrite, and nitrate) and phosphorus levels varied with the extent of

    nutrient removal.

    The overall performance of the individual treatment facilities for

    reduction of pathogens and indicators was assessed by comparing

    the average log10 reduction for removal of each microbial parameter

    through each filter, as shown in Table 4. A different pattern was

    observed for removal of pathogens compared with the indicator

    organisms. The highest log10 removal was associated with the filter

    that has the deepest layer of sand, hydraulic loading rates ranging

    from 30 to 36 m3/m2 min (1.3 to 1.5 gpm/ft2), and uses coagulationduring secondary sedimentation (A24S48). Filtration without

    prechlorination achieved average cultivable enteric virus removalsof 0.3 to 0.8 log10 and parasite removals of 0.4 to 1.5 log10. In

    contrast to surface-water-filtration systems designed to provide at

    least a 2-log reduction of protozoan pathogens (40 CFR 141.721),

    the log reductions associated with the filters tested in this project

    Table 3Summary of water quality data for each filter effluent. Filter descriptions are given in Table 1.

    Parameter (units)

    Cloth S11 A48 A30S10 SU40 A24S48

    Average Range Average Range Average Range Average Range Average Range Average Range

    C-BOD5 (mg/L) 2.4 1.2 to 3.2 2.8 2.0 to 4.0 7.5 5.9 to 9.1 2.7 2.3 to 3.4 2.0 ,2 to 2 2.0 ,2 to 2

    TSS (mg/L) 0.8 0.4 to 1.1 1 ,1 to 2 1.2 0.2 to 2.7 0.7 0.3 to 1.2 2.7 0.8 to 4.4 1.1 ,0.5 to 1.2

    Ammonia-nitrogen(mg/L) 10.9 8.3 to 13.5 8.5 4.6 to 15.3 15.8 11 to 21.9 6.4 2.4 to 10.5 0.2 0.09 to 0.2 0.5 0.1 to 1.2

    Nitrite-nitrogen 1

    nitrate-nitrogen

    (mg/L as nitrogen) 0.4 0.2 to 0.5 3.8 1.3 to 6.9 4.6 3.0 to 6.0 1.5 1.2 to 1.7 5.2 3.9 to 7.3 1.0 0.2 to 2.4

    Orthophophate (mg/L

    as phosphorus) 5.3 4.3 to 6.3 0.7 0.56 to 0.9 0.9 0.4 to 1.3 0.2 0.1 to 0.3

    Table 4Summary of log10 reductions of indicators and pathogens through each filtration system.

    Microorganism

    Cloth S11 A48 A30S10 SU40 A24S48

    Ave ra ge Ran ge Averag e Ran ge Ave rag e Ran ge Av era ge Ra nge Ave rag e Ran ge Av era ge Ra nge

    Indicators

    Total coliform 3.7 2.3 to 6.2 3.8 0.4 to 5.7 1.7 0.9 to 2.2 0.2 21.1 to 1 1.8 0.2 to 4.5 3.1 2.8 to 3.3

    Fecal coliform 3.0 1.9 to 4.3 4.2 0.3 to 5.6 2.5 2.3 to 2.9 0.1 21.3 to 1.1 0.9 20.4 to 2.5 2.6 1.7 to 3.4

    Enterococci 2.4 0.8 to 4.5 3.1 1 to 4.8 1.4 0 to 2.7 0.01 20.5 to 0.4 0.8 0.1 to 1.6 1.3 1 to 2.5

    Clostridium

    perfringens 1.0 0 to 2.8 2.0 1.4 to 2.8 0.4 20.3 to 0.8 0.7 20.3 to 2.1 0.3 0 to 0.6 0.8 0.4 to 1.3

    Coliphage 1 0.6 20.1 to 1.8 1 0.6 to 1.4 1.3 0.1 to 5.5 1.3 0 to 2.9 0.6 0 to 1.1 0.3 0 to 1

    Coliphage 2 0.4 20.1 to 1 0.3 20.1 to 0.9 0.4 20.4 to 1.3 0.8 20.1 to 2 0.3 21 to 1.5 0.2 0 to 0.6

    Pathogens

    Cultivable

    enteric viruses 0.5 0.3 to 0.8 1.2 0.8 to 1.8 0.8 0.3 to 1.4 0.3 0 to 0.9 0.6 0.2 to 1.3 0.7 20.3 to 1.3

    Giardia (total) 0.5 20.4 to 1.3 0.7 20.4 to 1.3 1.3 0.4 to 3.2 0.8 0 to 1.7 0.7 20.1 to 1.4 1.5 0.3 to 2.6

    Cryptosporidium

    (total) 0.4 0.1 to 0.6 1.0 0.7 to 1.4 0.4 22 to 2.3 0.1 20.3 to 0.6 0.6 0 to 1.3 1.3 0.5 to 2.2

    Cryptosporidium

    (infectious) 0.5 0.3 to 0.7 1.0 0.7 to 1.2 0.8 0.5 to 1.1 0.3 0.3 to 0.5 0.8 0.3 to 1.7 1.1 0.6 to 1.3

    Levine et al.

    602 Water Environment Research, Volume 80, Number 7

  • 8/3/2019 WER 2008 Levine Cloth Media

    8/13

    were typically below 2-log, and, in some cases, higher concen-

    trations were observed in filter effluents than influents, resulting in

    negative log reductions.

    In general, log10 reduction of bacteria (coliforms, enterococci,

    and Clostridium) was 2- to 9-fold greater than log10 reduction of

    pathogens, suggesting that, in contrast to Figure 1, the use of

    bacterial indicators as a monitoring tool may overpredict pathogen

    reduction. The highest overall removal of the indicator organisms

    was associated with S11, the prechlorinated filter (cumulative log10reduction of indicators of 13.7); this filter ranked second (after

    A28S48) for cumulative pathogen reduction. The cloth filter, which

    performs as a depth filter, had the second highest log10 reduction

    of the tested indicators (cumulative log10 reduction of indicators

    of 10.6) and the lowest reduction of protozoan pathogens. Among

    the other filters, the two filters that used coagulants either during

    secondary sedimentation (A28S48) or upstream of filtration (A48)

    had greater log10 reduction of indicators and pathogens than the

    filters with no chemical pretreatment (A30S10 and SU40). In

    general, the weakest removal of indicators and pathogens was

    associated with the filter with the shallowest bed of sand and higher

    hydraulic loading rates (1.5 to 3.9 L/s [2.2 to 5.8 gpm/ft2]), which

    did not apply upstream chemicals (A30S10).Comparison of Pathogen Removal Through Deep-Bed Filters

    that Contain Sand. Further analysis of data collected fromfacilities that operated deep-bed filters that contain sand (SU40,

    A30S10, and A24S48) was conducted to evaluate the filter

    collection efficiency, g, and filter coefficient, k, for cultivable

    enteric viruses, Giardia, and Cryptosporidium (Figure 2). Based on

    theoretical values of collection efficiency (Figure 1), one would

    expectg values for protozoan pathogens to be higher than observed

    values for viruses, as a result of differences in particle size and

    dominant removal mechanisms. In addition, variations in g values

    among the filter types may reflect differences in apparent particle

    size, as a result of association with larger particles. The collection

    efficiency, g, was significantly higher for Giardia and Cryptospo-

    ridium associated with the dual-media filter A30S10 compared with

    the other filters (p , 0.05), perhaps because of the lower hydraulic

    loading rates. It is important to note that, even though the collection

    efficiency was high for A30S10, the overall log10 reduction was

    lower than that of the other filters (see Table 4). In contrast, the filter

    coefficients, k, were more variable and were not significantly

    different among the filters.

    Relationships between filter coefficients, k, and hydraulic loading

    rates for cultivable enteric viruses, Giardia, and Cryptosporidium

    are shown in Figure 3 (for hydraulic loading rates below 1.4 L/s [2

    gpm/ft2]). Even though the range of hydraulic loading rates that

    were tested in this project was controlled by operating parameters at

    the individual facilities (as opposed to systematic variation through

    experimental design), filter coefficients were inversely correlated to

    the hydraulic loading rate (p , 0.05) for the tested pathogens (i.e.,

    increased removal was associated with lower hydraulic loading

    Figure 2Comparison of calculated filter collection efficiency, g, and filter coefficients, k, for removal of cultivable

    enteric viruses (VIR), Giardia (GIA), and Cryptosporidium (CRY) through three deep-bed granular media filters. Filter

    descriptions are given in Table 1.

    Figure 3Relationship between filter coefficients, k, and

    hydraulic loading rates for cultivable enteric viruses,

    Giardia, and Cryptosporidium, for pooled data from

    A24S10, SU40, and A24S48, with hydraulic loading rates

    below 0.0015 m/s (2 gpm/ft2). Correlation coefficients, R2,

    were 0.8 for enteroviruses and 0.6 for Giardia and

    Cryptosporidium. The slopes of the regression lines were

    similar for all tested pathogens (2300 s/m).

    Levine et al.

    July 2008 603

  • 8/3/2019 WER 2008 Levine Cloth Media

    9/13

    rates). Conversely, there was no significant correlation between

    hydraulic loading rates and removal of the bacterial indicators

    (coliforms, enterococci, and C. perfringens), as predicted by eq 1

    and Figure 1. In addition, microbial sampling for this project was

    conducted under daily peak-flow conditions, which correspond to

    higher hydraulic loading rates. Additional sampling would be

    needed to account for the effects of daily and seasonal variations in

    flowrates (and loading rates) on removal of pathogens and

    indicators.

    The data were also evaluated to determine if there were anyrelationships between the depth of filtration media and log10reduction of indicators and pathogens from deep-bed filters operated

    without upstream chemical addition. For Giardia and Cryptospo-

    ridium, weak but significant correlations were observed. Increased

    depth was associated with improved log10 reduction, which ranged

    from approximately 0.005 log10 removal for Giardia (total) to

    approximately 0.01 log10 removal for Cryptosporidium (total) per

    incremental centimeter of sand. While these correlations were

    statistically significant, these results suggest that it would require an

    additional 100 cm (3.3 ft) of sand to provide filtration capacity for

    an additional 0.5 log10 reduction of Giardia and 1 log10 reduction

    of Cryptosporidium. The additional headloss associated with this

    increased depth makes this concept impracticable for control ofprotozoan pathogens, except in the context of groundwater recharge

    and soil-aquifer treatment. Sand depth did not have a significant

    effect on removal of cultivable enteric viruses or indicators (p .

    0.05). No other significant correlations were observed between

    depth and microbial removal (total depth or anthracite depth).Comparisons Between Indicator Organisms and Pathogens

    in Filter Effluents. Coliphages have been widely used as

    indicators for viral pathogens because of their similarity in size.

    To evaluate the validity of this relationship for the filtration data

    collected in this project, ANOVA tests were conducted to compare

    log10 reductions and C/Co for coliphages and cultivable enteric

    viruses for all filters tested in this project. No significant differences

    were observed for the individual organisms among the filters;

    therefore, the data were pooled for further comparisons. A

    comparison of log10 reductions of coliphages and cultivable enteric

    viruses is given in Figure 4 in a boxplot format. In general, higher

    log10 reductions were associated with coliphages than cultivable

    enteric viruses. In addition, coliphages and cultivable enteric viruses

    did not display consistent patterns of detect versus nondetect (see

    Table 2). The highest frequency of nondetects for cultivable enteric

    viruses (80%) was associated with S11, while the highest frequency

    of nondetects for coliphages (86%) was associated with A30S10.

    These discrepancies may be the result of the characteristics of the

    wastewater associated with each facility coupled with removalmechanisms upstream of filtration, resulting from differences in the

    biological treatment system.

    To test the efficacy of using coliphage concentrations to predict

    cultivable enteric virus concentrations, filter effluent concentrations

    of cultivable enteric viruses are shown as a function of coliphage

    concentrations in Figure 5. The data are weakly correlated (R2 5

    0.4); however, in approximately 44% of the cases where cultivable

    enteric viruses were detected, coliphages were not detected (blue

    squares to the left of coliphage detection limit line). Thus, for this

    study, the absence of coliphages did not correspond to the absence

    of viruses. Perhaps this relationship could be improved if larger

    volume assays for phage were used (1 L versus 10 mL). Improve-

    ments in coliphage detection limits might provide an opportunity forrevisiting this issue. However, because the sources of enteric viruses

    in wastewater differ from the sources of coliphages, the use of

    coliphages as a viral indicator should be verified for individual

    facilities.Effect of Prech lorin ation on Removal of In dicator

    Organisms. To evaluate the effect of prechlorination, two filters

    at the facility that prechlorinates (S11) were isolated and operated in

    parallel. One filter received secondary effluent amended with

    approximately 2 mg/L chlorine, and the second filter received

    nonchlorinated secondary effluent. Hydraulic loading rates were

    similar for the two filters. Grab samples were collected at 2-hour

    intervals and evaluated for selected indicators. A comparison of

    C/Co for fecal coliforms, enterococci, coliphages, and C. perfrin-

    gens with and without prechlorination is shown in Figure 6.

    Figure 4Comparison of log10 reduction of coliphages and cultivable enteric viruses using pooled data from six full-scale filters.

    Levine et al.

    604 Water Environment Research, Volume 80, Number 7

  • 8/3/2019 WER 2008 Levine Cloth Media

    10/13

    Prechlorination resulted in significantly greater removal of fecal

    coliforms and enterococci (p , 0.05); however, prechlorination did

    not have a significant effect on removal ofC. perfringens, probably

    because it is less susceptible to chlorine than the other bacterial

    indicators (Sobsey, 1989). Minimal removal of coliphages occurredwith or without prechlorination. These results suggest that these

    more resistant organisms may be better indicators for assessing

    filtration efficacy.Effect of Backwashing Practices on Removal of Protozoan

    Pathogens. To evaluate the effects of backwashing practices on

    removal of protozoan pathogens, samples were collected before and

    after backwashing of two filtersA48 and A24S48 (Figure 7). In

    each case, chlorinated filter effluent was used as a source of

    backwash water, and the filters were backwashed and returned to

    service using normal protocols. For both filters (A48 and A24S48),the concentrations of Giardia and Cryptosporidium increased

    following backwashing. There was almost a 1 log10 increase

    associated with the dual media deep-bed filter (A24S48). While the

    cost of pathogen testing precluded additional testing of this issue

    in the context of this project, similar findings have been reported

    for filters used to process drinking water, where the presence of

    remnant particles or the lack of filter ripening can result in

    pathogen breakthrough following backwashing (Ahmad and

    Amirtharajah, 1998; Amburgey et al., 2004, 2005; Darby et al.,

    1991; Emelko et al., 2003; Hall and Fitzpatrick, 2000; Hozalski and

    Bouwer, 1998; Jegatheesan and Vigneswaran, 2000; Logan et al.,

    2001). Optimization of backwashing procedures by using extended

    terminal subfluidization backwash and/or more gradual postback-washing startup procedures, such as delayed start or slow start, have

    been suggested for improving the consistency of microbial removal

    through filtration (Amburgey et al., 2004).Water Quality Variables. This study was based on evaluation

    of full-scale wastewater reclamation facilities that produce re-

    claimed water that is used for nonpotable public access reuse. In

    general, all of the facilities produce a high-quality effluent with

    BOD5 and TSS values below 10 mg/L (see Table 3). Two of the

    facilities are designed for biological nutrient removal (SU40 and

    A24S48). Water quality data from each facility were evaluated to

    determine if there were any relationships between reduction of

    pathogens and indicators through filtration and secondary effluent

    characteristics. Removal of indicator bacteria, coliphages, cultivable

    enteric viruses, and Giardia by filtration were not significantly

    Figure 5Comparison of cultivable enteric virus concentrations and coliphage concentrations in pooled data from filter

    effluents tested in this project. n represents detected levels of cultivable enteric viruses, s represents samples where

    enteroviruses are not detected (plotted as detection limit). The line is an exponential fit to the data. The correlation

    coefficient, R2, was 0.4 in both cases (p< 0.05). Detection limits for coliphages ranged from 5 to 10 PFU/100 mL. MPN 5

    most probable number.

    Figure 6Comparison of persistence of fecal coliforms,

    enterococci, coliphages, and C. perfringensthrough S11

    filters operated with and without prechlorination (sam-

    pling conducted every 120 minutes over a 10-hour filter

    run). The first box in each section represents the

    prechlorinated filter (labeled as prechlorine) and the

    second box (unshaded) represents the filter that received

    secondary effluent without upstream chlorination (non-

    prechlorinated).

    Levine et al.

    July 2008 605

  • 8/3/2019 WER 2008 Levine Cloth Media

    11/13

    correlated to any of the monitored water quality parameters (p .

    0.05). However, effluent concentrations of Cryptosporidium were

    weakly correlated to effluent BOD5, TSS, ammonia, and phosphate

    (i.e., increased removal was associated with higher water quality)

    (linear correlation coefficients, R2, were 0.2 [p 5 0.02] for BOD5,

    0.6 [p5 0.0002] for TSS, and 0.3 [p5 0.03] for NH3-N and PO4-P).

    Summary and RecommendationsGranular media filtration of secondary effluent is widely used for

    production of reclaimed water. Key process variables include the

    filter media characteristics (i.e., depth and particle size), use of

    upstream chemicals (oxidants and/or coagulants), hydraulic loading

    rates, and backwash practices. In this study, data from six different

    full-scale wastewater reclamation facilities were compared to

    evaluate filter performance. Because of the nature of the grab

    sampling conducted in this project and differences in operating

    parameters of the individual filters, it was not possible to sample

    each filter under similar operating conditions, thus introducing

    confounding variables that may affect data interpretation. For

    example, the facility that uses prechlorination (S11) also had the

    lowest hydraulic loading rates and smallest media, making itdifficult to separate the effects of prechlorination and hydraulic

    loading rate under routine plant operations.

    Filtration of secondary effluent differs from current practices that

    have been adopted for filtration of surface water. Key differences

    include the degree to which upstream chemical amendments are

    used, turbidity limits, and filter backwashing and postbackwashing

    protocols. While the higher ionic strength and organic content of

    secondary effluents may serve to enhance particle removal through

    filtration, the log10 reduction of pathogens observed through this

    study was significantly lower than values reported for surface water

    filtration. Based on the data from this study and others, some degree

    of pathogen reduction was observed through all of the filters that

    were tested; however, the results suggest that optimization offiltration operations (including pretreatment, loading rates, and

    backwashing practices) can improve pathogen reduction through

    filtration and thereby improve the overall microbial safety of re-

    claimed water. In addition, because of differences in source charac-

    teristics and removal mechanisms through biological treatment and

    filtration, removal patterns associated with traditional bacterial

    indicators (coliforms) and coliphages do not correspond to reduc-

    tion of protozoan pathogens and cultivable enteric viruses. Adop-

    tion of alternative monitoring approaches and extrapolation of design

    and operation strategies from surface-water filtration may result in

    more robust filtration systems for the production of reclaimed water.

    ConclusionsBased on the data derived from this project, filter loading rates,

    use of upstream chemical addition, backwashing practices, and

    water quality of the secondary effluent can influence the efficiency

    of granular media filtration for removal of pathogens from sec-

    ondary effluent. The major conclusions are as follows:

    (1) The use of traditional models of single collector efficiency and

    filter coefficients can be applied to provide a qualitative

    comparison of pathogen removal from granular media filtration

    of secondary effluent; however, supplemental water quality data

    and operations information are needed to optimize filter

    performance.

    (2) Higher rates of removal through granular media filtration were

    observed for conventional bacterial indicators compared with

    pathogens, suggesting that the use of bacterial indicators is notan effective predictor of filtration performance for pathogen

    reduction.

    (3) The suite of microbiological indicators tested in this project

    did not correlate to concentrations, removal (log reduction),

    or presence/absence of cultivable enteric viruses, Giardia, or

    Cryptosporidium in the filtered secondary effluents tested in

    this project. Coliphage removal or concentrations did not

    correspond to removal or concentrations of cultivable enteric

    viruses.

    (4) For deep-bed filters that contained sand and were not

    prechlorinated, removal of pathogens was inversely propor-

    tional to hydraulic loading rates and directly proportional to the

    depth of sand.

    (5) The use of prechlorination resulted in improved removal of

    microorganisms that are susceptible to chlorine inactivation

    (coliforms and enterococci) but had minimal effect on chlorine-

    resistant bacteria, such as the endospore-forming Clostridium

    perfringens.

    (6) The return of filters to service directly following backwashing

    may result in higher effluent concentrations of Giardia and

    Cryptosporidium until filter ripening occurs. The extent of the

    increase varies with filter characteristics (media, depth, and

    upstream chemical use). Continuous backwash filters may not

    experience these performance fluctuations.

    (7) Water quality variables that were correlated to removal of

    Cryptosporidium through granular media or cloth filtration

    include effluent concentrations of BOD5, TSS, ammonia

    Figure 7Comparison of (a) concentrations of total

    Cryptosporidiumand total Giardiain filter effluents before

    and after backwashing from two full-scale filtersA48

    and A24S48, and (b) ratio of effluent to influent concen-

    tration (C/Co) of total Cryptosporidium and total Giardiain

    filter effluents before and after backwashing from two full-

    scale filters.

    Levine et al.

    606 Water Environment Research, Volume 80, Number 7

  • 8/3/2019 WER 2008 Levine Cloth Media

    12/13

    nitrogen, and phosphate. No statistically significant water

    quality relationships were associated with removal of the other

    microorganisms tested in this project (cultivable enteric viruses,

    Giardia, and indicator organisms).

    (8) For the filters tested in this project, removal of conventional

    indicator organisms through filtration tended to exhibit different

    patterns than those observed for viral and protozoan pathogens,

    suggesting that alternative monitoring practices may be neededto assess and manage microbial health risks associated with

    reclaimed water applications.

    Credits

    This study was funded by the Water Environment Research

    Foundation (Alexandria, Virginia) as project 00-PUM2T. Personnel

    from wastewater reclamation facilities in Phoenix, Arizona; Santa

    Barbara, California; and Eustis, St. Petersburg, and Hillsborough

    County, Florida assisted with sample collection and data compila-

    tion. Vasanta Chivulaka, Molly McLaughlin, Stefica Depovic, and

    Angela Gennaccaro conducted microbiological testing for the

    project.Submitted for publication November 12, 2006; revised manuscript

    submitted September 9, 2007; accepted for publication December

    11, 2007.

    The deadline to submit Discussions of this paper is October 15,

    2008.

    References

    Adams, M. H. (1959) Bacteriophages ; Interscience Publishers: New

    York.

    Ahmad, R.; Amirtharajah, A. (1998) Detachment of Particles During Filter

    Backwashing. J. Am. Water Works Assoc., 90 (12), 7485.

    Amburgey, J. E.; Amirtharajah, A.; Brouckaert, B. M.; Spivey, N. C. (2004)

    Effect of Washwater Chemistry and Delayed Start on Filter Ripening.

    J. Am. Water Works Assoc., 96 (1), 97110.

    Amburgey, J. E.; Amirtharajah, A.; York, M. T.; Brouckaert, B. M.; Spivey,

    N. C.; Arrowwood, M. J. (2005) Comparison of Conventional and

    Biological Filter Performance for Cryptosporidium and Microsphere

    Removal. J. Am. Water Works Assoc., 97 (12), 7791.

    American Public Health Association; American Water Works Association;

    Water Environment Federation (1998) Standard Methods for the

    Examination of Water and Wastewater, 20th ed.; American Public

    Health Association: Washington, D.C.

    Bisson, J. W.; Cabelli, V. J. (1979) Membrane Filter Enumeration for

    Clostridium Perfringens. Appl. Environ. Microbiol., 37 (1), 5566.

    Bourgeous, K. N.; Riess, J.; Tchobanoglous, G.; Darby, J. L. (2003)

    Performance Evaluation of a Cloth Media Filter for Wastewater

    Reclamation. Water Environ. Res., 75 (6), 532538.Brown, D. G.; Abramson, A. (2006) Collision Efficiency Distribution of

    a Bacterial Suspension Flowing Through Porous Media and Implica-

    tions for Field Scale Transport. Water Res., 40, 15911598.

    Bustamante, H. A.; Shanker, S. R.; Pashley, R. M.; Karaman, M. E. (2001)

    Interaction Between Cryptosporidium Oocysts and Water Treatment

    Coagulants. Water Res., 35 (13), 31793189.

    Camp, T. R. (1964) Theory of Water Filtration. ASCE J. Sanit. Eng. Div.,

    90 (SA4), 130.

    Darby, J. L.; Lawler, D. F.; Wilshusen, T. P. (1991) Depth Filtration of

    Wastewater: Particle Size and Ripening. J. Water Pollut. Control Fed.,

    63 (2), 228238.

    Edzwald, J. K.; Tobiason, J. E.; Parento, L. M.; Kelley, M. B.; Kaminski,

    G. S.; Dunn, H. J.; Galant, P. B. (2000) Giardia and Cryptosporidium

    Removals by Clarification and Filtration Under Challenge Conditions.

    J. Am. Water Works Assoc., 92 (12), 7084.

    Elimelech, M. (1992) Predicting Collision Efficiencies of Colloidal Particles

    in Porous Media. Water Res., 26 (1), 18.

    Elimelech, M.; OMelia, C. R. (1990) Kinetics of Deposition of Colloidal

    Particles in Porous Media. Environ. Sci. Technol., 24, 15281536.

    Emelko, M. B.; Huck, P. M.; Coffey, B. M. (2005) A Review of

    Cryptosporidium Removal by Granular Media Filtration. J. Am. Water

    Works Assoc., 97 (12), 101115.

    Emelko, M. B.; Huck, P. M.; Douglas, I. P. (2003) Cryptosporidium and

    Microsphere Removal During Late-in-Cycle Filtration. J. Am. WaterWorks Assoc., 95 (5), 173182.

    England, S. K.; Darby, J. L.; Tchobanoglous, G. (1994) Continuous-

    Backwash Upflow Filtration for Primary Effluent. Water Environ. Res.,

    60 (2), 145152.

    Foppen, J. W.; Okletey, S.; Schijven J. F. (2006) Effect of Goethite Coating

    and Humic Acid on the Transport of Bacteriophage PRD1 in Columns

    of Saturated Sand. J. Contam. Hydrol., 85 (34), 287301.

    Ghosh, M.; Jordan, T. A.; Porter, R. L. (1975) Physicochemical Approach to

    Water and Wastewater Filtration. ASCE J. Sanit. Eng. Div., 101 (EE1),

    7186.

    Glasgow, G. D. E.; Wheatley, A. D. (1999) Observations of Rapid Flow

    Rate Disturbances in Drinking Water Filters and Their Effect on Solids

    Removal. J. Water SRT-Aqua, 48, 257262.

    Habibian, M. T.; OMelia, C. R. (1975) Particles, Polymers, andPerformance in Filtration. ASCE J. Environ. Eng. Div., 101 (EE4),

    567583.

    Hall, D.; Fitzpatrick, C. S. B. (2000) Suspension Concentration Profiles

    During Rapid Gravity Filter Backwashing. Chem. Eng. J., 80, 197201.

    Harrington, G. W.; Xagoraraki, I.; Assavasilavasukul, P.; Standridge, J. H.

    (2003) Effect of Filtration Conditions on Removal of Emerging

    Pathogens. J. Am. Water Works Assoc., 95 (12), 95104.

    Harvey, R. W.; Garabedian, S. P. (1991) Use of Colloid Filtration Theory in

    Modeling Movement of Bacteria Through a Contaminated Sandy

    Aquifer. Environ. Sci. Technol., 25 (1), 178185.

    Harwood, V. J.; Levine, A. D.; Scott, T. M.; Chivukula, V.; Lukasik, J.;

    Farrah, S. R.; Rose, J. B. (2005) Validity of the Indicator Organism

    Paradigm for Pathogen Reduction in Reclaimed Water and Public

    Health Protection. Appl. Environ. Microbiol., 71 (6), 31633170.

    Hozalski, R. M.; Bouwer, E. J. (1998) Deposition and Retention of Bacteria

    in Backwashed Filters. J. Am. Water Works Assoc., 90 (1), 7185.

    Hsu, B.-M.; Huang, C. (2002) Influence of Ionic Strength and pH on

    Hydrophobicity and Zeta Potential of Giardia and Cryptosporidium.

    Colloids Surf. A, 201, 210206.

    Hsu, B.-M.; Huang, C.; Pan, J. R. (2001) Filtration Behaviors of Giardia

    and Cryptosporidium Ionic Strength and pH Effects. Water Res.,

    35 (16), 37773782.

    Iwasaki, T. (1937) Some Notes on Sand Filtration. J. Am. Water Works

    Assoc., 29 (10), 15911602.

    Jegatheesan, V.; Vigneswaran, S. (2000) Transient Stage Deposition of

    Submicron Particles in Deep Bed Filtration Under Unfavorable

    Conditions. Water Res., 34 (7), 21192131.

    Jiminex, B.; Chavex, A.; Leyva, A.; Tchobanoglous, G. (2000) Sand and

    Synthetic Medium Filtration of Advanced Primary Treatment Effluentfrom Mexico City. Water Res., 14 (2), 473480.

    Kavanaugh, M. C.; Toregas, G.; Chung, M.; Pearson, E. A. (1978)

    Particulates and Trace Pollutant Removal by Depth Filtration. Prog.

    Water Technol., 10 (5/6), 197213.

    Koivunen, J.; Siitnonen, A.; Heinonen-Tanski, H. (2003) Elimination of

    Enteric Bacteria in Biological-Chemical Wastewater and Tertiary

    Filtration Units. Water Res., 37, 690698.

    Logan, A. J.; Stevik, T. K.; Siegrist, R. L.; Ronn, R. M. (2001) Transport

    and Fate of Cryptosporidium parvum Oocysts in Intermittent Sand

    Filters. Water Res., 35 (18), 43594369.

    Mosher, R. R.; Hendricks, D. W. (1986) Rapid Rate Filtration of Low

    Turbidity Water Using Field-Scale Pilot Filters. J. Am. Water Works

    Assoc., 78 (12), 4251.

    Nasser, A.; Weinberg, D.; Dinoor, N.; Fattal, B.; Adin, A. (1995) Removal

    of Hepatitis A Virus (HAV), Poliovirus, and MS2 Coliphage by

    Levine et al.

    July 2008 607

  • 8/3/2019 WER 2008 Levine Cloth Media

    13/13

    Coagulation and High Rate Filtration. Water Sci. Technol., 31 (56),

    6368.

    Nieminski, E. C.; Ongerth, J. E. (1995) Removing Giardia and

    Cryptosporidium by Conventional Treatment and Direct Filtration.

    J. Am. Water Works Assoc., 87 (9), 96106.

    OMelia, C. R.; Ali, W. (1978) The Role of Retained Particles in Deep Bed

    Filtration. Prog. Water Technol., 10 (5/6), 167182.

    Ongerth, J. E.; Pecoraro, J. P. (1996) Electrophoretic Mobility of

    Cryptosporidium Oocysts and Giardia Cysts. J. Environ. Eng., 122 (3),228231.

    Patania, N. L.; Jacangelo, J. G.; Cummings, L.; Wilczak, A.; Riley, K.;

    Oppenheimer, J. (1995) Optimization of Filtration for Cyst Removal;

    AWWA Research Foundation: Denver, Colorado.

    Reddi, L. N. (1997) Particle Transport in Soils: Review of Significant

    Processes in Infrastructure Systems. J. Infrastruct. Syst., 3 (2), 78

    86.

    Redman, J. A.; Grant, S. B.; Olson, T. M.; Estes, M. K. (2001) Pathogen

    Filtration, Heterogeneity and the Potable Reuse of Wastewater.

    Environ. Sci. Technol., 35 (9), 17981805.

    Rose, J. B.; Farrah, S. R.; Harwood, V. J.; Levine, A. D.; Lukasik, J.;

    Menendez, P.; Scott, T. M. (2004) Reduction of Pathogens, Indicator

    Bacteria, and Alternative Indicators by Wastewater Treatment and

    Reclamation Processes, 00-PUM-2T; Water Environment Federation:Alexandria, Virginia.

    Slifko, T. R.; Huffman, D. E.; Rose, J. B. (1999) A Most-Probable-Number

    Assay for Enumeration of Infectious Cryptosporidium Parvum Oocysts.

    Appl. Environ. Microbiol., 65 (9), 39363941.

    Sobsey, M. D. (1989) Inactivation of Health-Related Microorganisms in

    Water by Disinfection Processes. Water Sci. Technol., 21, 179195.

    Stevenson, D. G. (1997) Flow and Filtration Through Granular MediaThe

    Effect of Grain and Particle Size Dispersion. Water Res., 31 (2),

    310322.

    Swertfeger, J.; Metz, D. H.; DeMarco, J.; Braghetta, A.; Jacangelo, J. A.

    (1999) Effect of Filter Media on Cyst and Oocyst Removal. J. Am.

    Water Works Assoc., 91 (9), 90100.

    Tchobanoglous, G.; Eliassen, R. (1970) Filtration of Treated Sewage

    Effluent. ASCE J. Sanit. Eng. Div., 96 (SA2), 243265.

    Trussell, R. R.; Chang, M. (1999) Review of Flow Through Porous Media

    as Applied to Head Loss Through Filters. J. Environ. Eng., 125 (11),

    9981006.

    U.S. Environmental Protection Agency (2004) Guidelines for Water Reuse,

    EPA-625/R-04-108; U.S. Environmental Protection Agency: Wash-

    ington, D.C.

    U.S. Environmental Protection Agency (1995) ICR Protozoan Method for

    Detecting Giardia Cysts and Cryptosporidium Oocysts in Water by

    a Fluorescent Antibody Procedure, EPA-814/B-95-003; U.S. Environ-

    mental Protection Agency: Washington, D.C.

    U.S. Environmental Protection Agency (2003) Long Term 2 Enhanced

    Surface Water Treatment Rule Toolbox Guidance Manual, EPA-815/

    D-03-009; U.S. Environmental Protection Agency: Washington, D.C.

    U.S. Environmental Protection Agency (2005) Method 1623: Cryptospo-

    ridium and Giardia in Water by Filtration/IMS/FA, EPA-815/R-05-

    002; U.S. Environmental Protection Agency: Washington, D.C.

    U.S. Environmental Protection Agency (1996) Microbial Laboratory

    Manual. ICR: Information Collection Rule. U.S. Environmental

    Protection Agency: Cincinnati, Ohio.

    U.S. Environmental Protection Agency (2006) Ultraviolet Disinfection

    Guidance Manual for the Final Long-Term 2 Enhanced Surface Water

    Treatment Rule, EPA-815/R-06-007; U.S. Environmental ProtectionAgency: Washington, D.C.

    Wang, J. Z.; Summers, R. S.; Miltner, R. J. (1995) Biofiltration

    Performance: Part 1, Relationship to Biomass. J. Am. Water Works

    Assoc., 87 (12), 5563.

    Xagoraraki, I.; Harrington, G. W.; Assavasilavasukul, P.; Standridge, J. H.

    (2004) Removal of Emerging Waterborne Pathogens and Pathogen

    Indicators by Pilot Scale Conventional Treatment. J. Am. Water Works

    Assoc., 96 (5), 102113.

    Yao, K.; Habibian, M. T.; OMelia, C. R. (1971) Water and Waste Water

    Filtration: Concepts and Applications. Environ. Sci. Technol., 5 (11),

    11051112.

    York, D. W.; Walker-Coleman, L. (2000) Pathogen Standards for Reclaimed

    Water. Water Environ. Technol., 12 (1), 5861.

    Levine et al.

    608 Water Environment Research, Volume 80, Number 7