Water Environment Research

Embed Size (px)

Citation preview

  • 7/26/2019 Water Environment Research

    1/10

    Importance of Particulate BiodegradableOrganic Compounds in Performance of

    Full-scale Biological PhosphorusRemoval System

    Tolga Tuncal1*, Aysegul Pala2, Orhan Uslu3

    ABSTRACT: In this study, biological treatment performances of two

    parallel treatment lines operating with and without primary sedimentation

    were investigated. The research was carried out in a large-scale enhanced

    biological phosphorus removal (EBPR) process. Influent and effluent oftreatment lines were characterized continuously during the study. In

    addition, anaerobic anoxic and aerobic EBPR activities were investigated

    by batch tests using fresh activated sludge samples. All of the

    environmental and operational conditions of the treatment lines were

    statistically compared. Evaluation of effluent compositions indicated that

    EBPR performances of treatment lines were significantly different. Results

    of the research also indicated that settling characteristics of the activated

    sludge process could be improved significantly with increasing particulate

    biodegradable organic compound (pbCOD) loading rate. Batch test results

    revealed that anaerobic, anoxic, and aerobic biochemical reaction rates of

    activated sludge cultivated on increased pbCOD loading rate were

    significantly higher compared to activated sludge cultivated on soluble

    substrate forms.Water Environ. Res., 81, 886 (2009).

    KEYWORDS: Enhanced biological phosphorus removal, particulatebiodegradable chemical oxygen demand, phosphorous accumulating

    organisms, denitrifying phosphorous accumulating organisms, electron

    acceptor, air entrainment, primary sedimentation, activated sludge, sludge

    volume index.

    doi:10.2175/106143009X407320

    Introduction

    Izmir Bay is one of the great natural bays of the Aegean Sea.

    Total surface area of the bay is 500 km2 and total water volume is

    11.5 billion m3. The bay can be examined in three sections

    outer, middle, and inneraccording to the physical characteristics

    of different water masses. The depth of the water decreases from

    the outer bay to the inner bay and the average water depth in the

    outer bay is 70 m (Kucuksezgin et al., 2005). Scientific

    investigations indicated that eutrophication of the inner bay is a

    serious, year-round problem, and that red tide events are

    becoming more frequent (Kontas et al., 2004; United Nations

    Environment Programme [UNEP], 1993).

    It also was found that inner-bay phosphate concentration was

    higher than values measured in clean waters because of domestic

    wastewaters. The atomic ratio of TNOx to phosphate in the outer

    bay was between 1.8 and 27 and 0.02 and 54 in the middle and

    inner bay. Observed average N/P ratios were lower than optimal

    growth requirement (N/P 5 15/1) in conformity with Redfields

    ratio (N/P516). According to the measured N/P ratios, nitrogen is

    the limiting nutrient in Izmir Bay. Phosphorus, however, also

    becomes a limiting nutrient in the summer period because of

    cyanobacteria activity. Pollution levels in the outer bay were not

    significant, but eutrophication of the inner bay has begun and

    could be spreading (Kucuksezgin et al., 2005).

    To prevent discharge of untreated wastewaters, Izmir Waste-

    water Treatment Plant (WWTP) began operating in early 2000.The plant was designed to treat both domestic and pretreated

    industrial wastewaters of Izmir City. Because nitrogen (N) and

    phosphorus (P) concentrations in the sea were found to be at

    critical levels for potential eutrophication, the plant was designed

    to remove both nutrients using activated sludge process following

    adequate physical treatment including fine screens, aerated grit

    chambers, and circular primary sedimentation primary sedimen-

    tation tanks. The average design capacity of the plant is

    604 800 m3/d. As a result of rapid industrial developments and

    climate changes, total volume of wastewaters collected from the

    city has increased. Capacity enlargement was necessary to protect

    effluent quality even at peak flow rates and to create reserve

    volume for the maintenance works. The Water and SewerageAdministration of the city has decided to increase capacity of the

    plant up to 1 036 800 m3/d in the near future.

    Many scientific investigations focused on enhanced biological

    phosphorus removal (EBPR) were carried out on activated sludge

    cultures cultivated on easily biodegradable carbon forms.

    Phosphorus release and volatile fatty acids (VFAs) uptake

    mechanism occurring in the anaerobic phase cause proliferation

    of phosphorus accumulating organisms (PAOs) within the

    activated sludge (Metcalf and Eddy, 2003; U.S. Environmental

    Protection Agency [U.S. EPA], 1987; Water Pollution Control

    Federation [WPCF], 1983). Most full-scale EBPR processes

    include simultaneous nitrification-denitrification (SND) configu-

    rations. Carbon can be derived from both soluble and particulate

    1*Dokuz Eylul University, The Graduate School of Natural and AppliedSciences, Izmir, Turkey; Izmir Water and Sewerage Administration(IZSU), Cigli Wastewater Treatment Plant, Tuzla Street, Sasal Rd., 8thkm, P.O. Box 7, Cigli/Izmir/Turkey; e-mail: [email protected].

    2 Environmental Engineering Department, Engineering Faculty, DokuzEylul University.

    3 Environmental Engineering Department, Engineering Faculty, Bahce-sehir University, Besiktas, Istanbul, Turkey.

    886 Water Environment Research,Volume 81, Number 9

  • 7/26/2019 Water Environment Research

    2/10

    substrate forms from denitrification reactions that take place in

    SND processes (U.S. EPA, 1993). Hydraulic retention time in

    SND configurations is up to 10 hours, and intracellular carbon

    reserves of PAOs may be depleted at the beginning of the anoxic

    zone. Thus PAOs became exhausted in the later part of remaining

    oxidative zones, resulting in washout. The main objective of this

    study, therefore, was to investigate the effect of increased loading

    rate of particulate biodegradable organic compounds (pbCOD) on

    the EBPR process. To achieve this objective, biological treatment

    performances of two treatment lines operating with and without

    primary sedimentation (primary sedimentation) were investigated

    at the Izmir WWTP for more than one year. In one treatment line

    (A), inflow was fed directly to the anaerobic tank; in the secondline (B), inflow was fed to primary sedimentation before EBPR.

    Field measurements were used to characterize influent and

    effluent of both lines. The EBPR activities also were examined

    by batch tests using fresh activated sludge samples obtained from

    the treatment lines. All data were compared statistically, including

    environmental and operational conditions, effluent, and return

    activated sludge (RAS) characteristics of treatment lines.

    Material and Methods

    Process Configuration. Izmir WWTP pretreats wastewater

    with fine screens, aerated grit chambers, and circular primary

    sedimentation tanks. Following grit removal process, wastewater

    is distributed equally to three treatment lines. On each of these

    independent treatment lines, wastewater optionally can be treated

    in primary sedimentation, which allows for more flexible

    operation. In the period of study, primary sedimentation tanks

    were offline in Line A and online in Line B.

    After physical treatment, wastewater is fed to biological

    treatment facilities composed of anaerobic tanks, aeration basins,

    and circular final clarifiers. The biological process configuration

    of the plant is similar to the five-stage modified BardenphoH

    process. Figure 1 shows simplified process configuration of the

    treatment lines. The volume of one anaerobic tank is 16 800 m3;

    nominal hydraulic retention time in the anaerobic tank is 1.1 hours

    (including return activated sludge [RAS] flow rate: 0.76 Qi at

    average flow (2.3 m3/s).

    The aeration basins were designed for SND with a total volume

    of 100 000 m3

    . Dissolved oxygen concentration in fully aerated

    parts is controlled by an online measurement system. Ratio of the

    aerated and nonaerated zones to total volume is 60% and 40%.

    The nitrate-rich mixed liquor is recycled from the end of the last

    aerated zone to the main anoxic zone in which nitrate is reduced to

    nitrogen gas by metabolizing the influent chemical oxygen

    demand (COD). The ratio of this internal recirculation is 400%

    of the inflow.

    After biological treatment, the activated sludge is settled in the

    circular final clarifiers. Total volume of these tanks (in one

    treatment line) is approximately 38 400 m3. Settled sludge is

    withdrawn from the bottom of the clarifiers to the RAS pumpingstation by gravity. A portion of the activated sludge is then

    transferred to the anaerobic tanks. The RAS pumping rate is

    controlled depending on inflow rate (R/Qi 5 0.76).

    Analytical Methods

    Chemical parameters were determined by colorimetric methods

    using following cuvette test-kits: COD: LCK 114; soluble COD

    (sCOD): LCK 314; VFAs: LCK 365; total phosphorus and PO4-P:

    LCK 114: total nitrogen: LCK 338 (for influent) and LCK 238 (for

    effluent); nitrate: LCK 339 from HACH LANGEH. Hach Lange

    DR 2800 spectrophotometer and Hach Lange LT 200 thermo

    reactor were used for analysis. Suspended solids, mixed-liquor

    suspended solids (MLSS), and mixed-liquor volatile suspended

    solids (MLVSS) were measured according to Standard Methods

    (American Public Health Association, 1998). The pH and

    temperature were measured using well-calibrated manual probes

    from WTWH.

    Readily Biodegradable Soluble Chemical Oxygen

    Demand Concentration

    Carbonaceous constituents in wastewater could be measured by

    COD analyses. Because a portion of COD is not biodegradable, it

    is divided into biodegradable and non-biodegradable fractions.

    The next level of COD categorization is determination of

    dissolved and particulate fractions. Previous studies indicate that

    readily biodegradable soluble COD (rbsCOD) is one of the critical

    Figure 1Simplified process configuration of the treatment lines.

    Tuncal et al.

    September 2009 887

  • 7/26/2019 Water Environment Research

    3/10

    COD components for EBPR. To estimate rbsCOD, it was assumed

    that influent non-biodegradable soluble COD (nbsCOD) is equal

    to effluent soluble COD (activated sludge process, operating with

    hc greater than four days). To measure the concentration of

    rbsCOD, the flocculation/filtration method and ZnSO4and NaOH

    precipitation were used (Metcalf and Eddy, 2003).

    Experimental Procedures for Field Measurements

    Samples were collected periodically and diluted with distilled

    water according to the measurement range of a cell test. The time

    between sampling to analysis was kept as short as possible

    because chemical characteristics of the samples (both wastewater

    and activated sludge) were unstable. Because of the large WWTP

    sampling area, samples for soluble forms, including VFAs, PO4-P,

    sCOD, NO3-N, NH4-N, were filtered immediately at the sampling

    points using both rough filter paper and single-use syringe filter

    with a pore size of 0.45 mm (SartoriusH Minisart RC 25).

    Experimental Procedures for Batch-Scale Investigations

    Barch-scale experiments were conducted using 2-L Woulf

    Bottles (Woulffsche-Flaschen, DURANH, Schott). During batch

    tests, a magnetic stirrer was used for mixing. A diffuser fed N2gas

    through a tube to the reactor in the anaerobic and anoxic batch

    tests. Oxygen demand was provided with an air pump connected

    to a diffuser in aerobic batch tests. All samples were filtered

    immediately using a single-use syringe filter with a pore size of

    0.45 mm (SartoriusH Minisart RC 25).

    Anaerobic Phosphorus Release Batch Test

    Freshly collected effluent and activated sludge samples were

    obtained from the RAS pumping stations of treatment lines (A and

    B). A portion of the effluent, acetate, and activated sludge were

    placed into a 1.5-L reactor. Volumetric mixture ratio between

    effluent, acetate, and activated sludge were determined according

    to the actual food-to-microorganism (F/M) ratio of the treatment

    lines. The reactor was continuously flushed with N2 gas to create

    anaerobic conditions. Samples were taken periodically at 0, 1, 30,

    60, 90, 120, 180, and 210 minutes. Figure 2 shows the

    experimental setup used in anaerobic phosphorus release batch

    tests.

    Excess acetate (100 to 300 mg HAc-C/L or 266 to 800 mg

    sCOD/L) was instantly added to the reactor under anaerobic

    conditions. The PO4-P, sCOD, and MLVSS concentrations were

    measured for 3.5 hours to determine the phosphorus release rate

    (mg P/g VSS?min) and sCOD consumption rate (mg sCOD/g

    VSS?min). At the end of the test, surplus COD always remained in

    the solution to guarantee that COD did not limit phosphorus

    release (Brdjanovic, 1998).

    Anoxic and Aerobic Phosphorus Uptake Batch Tests

    The activated sludge samples collected from the RAS pumping

    stations of treatment lines were exposed first to anaerobic

    conditions in the presence of acetate in order to deplete internal

    poly-P pool and increase the polyhydroxylalkanoate (PHA) level

    of the biomass (Brdjanovic D, 1998). Acetate (2025 mg HAc-C/L) was instantly added to the reactor under anaerobic conditions

    as carbon source. After 3 h- anaerobic period, acetate was fully

    consumed and P was released into solution. At this point, sludge

    was divided into two equal parts to perform anoxic and aerobic

    phosphorus uptake batch tests.

    Anoxic conditions were maintained by the addition of surplus

    amount of nitrate at the beginning of the test (28 mg N/L) and

    continuously flushing the mixed liquor with N2 gas. Figure 3

    illustrates the experimental setup used in the anoxic phosphorus

    uptake batch tests.

    The reactor was mixed for 3.5 hours and samples were taken

    periodically at 0, 1, 30, 60, 90, 120, 180, and 210 minutes. The

    PO4-P, NO3-N, and MLVSS concentrations were measured todetermine the anoxic phosphorus uptake rate (mg P/g VSS?min)

    and nitrate use rate (mg NO3-N/g VSS?min).

    The remaining activated sludge was exposed to aerobic

    conditions for 3.5 hours. Figure 4 shows the experimental setup

    used in the aerobic phosphorus uptake batch. The reactor was

    mixed for 3.5 hours and samples were taken periodically at 0, 1,

    30, 60, 90, 120, 180, and 210 minutes. The PO 4-P and MLVSS

    concentrations were measured to determine the aerobic phospho-

    rus uptake rate as mg P/g VSS?min.

    Statistical Analysis

    Important environmental and operational parameters, design

    aspects, influent - effluent characteristics of the treatment lines,

    Figure 2Experimental setup for anaerobic phosphorus

    release batch tests.

    Figure 3Experimental setup for anoxic phosphorus

    uptake batch tests.

    Tuncal et al.

    888 Water Environment Research,Volume 81, Number 9

  • 7/26/2019 Water Environment Research

    4/10

    nitrate, and dissolved oxygen concentrations of RAS streams were

    compared statistically using independent samples t-test at 0.95

    confidence interval. The SPSSE v13 software, developed by

    SPSSE Inc., was used in the data analysis. Statistical analysis

    results were evaluated using the users manual of the software,

    which explained that the t-test could be used to compare the

    means of two groups. If the significance value for the Levenes

    test is high (typically greater than 0.05), then the results obtained

    assuming equal variances for both groups could be used. If the

    significance value for the Levenes test is low, then the results that

    do not assume equal variances for both groups should be used

    (SPSSE v13 Manuel, 2004).

    Results and DiscussionEnvironmental and Operational Conditions of Treatment

    Lines. Organic and hydraulic loading rates to the biological

    treatment units are one of the important factors that could

    significantly affect the process performance of full-scale EBPR

    systems (Lopez-Vazquez et al., 2008). Table 1 summarizes

    important influent characteristics of the treatment lines. According

    to statistical comparison of inflow rates, the mean difference

    between hydraulic loading rates was 0.005 m3/s and it was not

    significant (two-tailed significance value: r 5 0.782). Hydraulic

    loading rates of the treatment lines were approximately equal in

    this study. Table 2 provides a statistical comparison of the influent

    characteristics of the treatment lines.

    Although hydraulic loading rates were approximately equal,

    different organic loading rates could be expected because some

    part of the particulate organic compounds removed in primary

    sedimentation. In addition, fractions of the COD could be

    influenced by primary sedimentation because of hydrolysis and

    acidogenesis reactions. Previous studies showed that primary

    sedimentation could be used to produce VFAs from settled

    organic materials. In these systems, however, required hydraulic

    retention time (HRT) is greater than 1 day (Barajas et al., 2001;

    Katehis et al., 2003). During this study, HRT in primary

    sedimentation ranged from 1.4 to 2.2 hours, with an average of

    2 hours. As a result, rbsCOD and VFAs concentration in the

    influents of anaerobic tanks were not significantly different

    (rsCOD 5 0.823, rVFAs 5 0.488).

    Table 2 shows that influent total phosphorus concentrations of

    the treatment lines were significantly different (r 5 0.000).

    Average influent total phosphorus concentration of the Line A and

    B was 10.69 62.151 mg/L and 9.51 61.733 mg/L. Because somepart of particulate phosphorus is removed in primary sedimenta-

    tion, the rbsCOD/total phosphorus ratio of the treatment line

    operating with primary sedimentation (Line B) was relatively

    higher. Average influent total nitrogen concentrations were 38

    63.86 and 34 62.86 mg/L for Line A and B, respectively.

    Influent ammonium concentrations were nearly equal (r 5 0.867)

    and within the range of 22 to 28 mg NH4-N/L.

    Sludge age and wastewater temperature can affect EBPR

    performance (Brdjanovic et al., 1997; Converti et al., 1995;

    McClintock et al., 1993; Whang et al., 2006). During the study,

    average wastewater temperature was 17.7 61.42 uC. Sludge ages

    of Line A and B were 9.68 and 10.69 days. Based on these ranges,

    temperature and sludge ages likely did not result in significantdifference in EBPR performances of the treatment lines in the

    period of study. Another important environmental factor is initial

    wastewater pH, which is effective on both anaerobic and aerobic

    PAOs metabolism (Filipe et al., 2001; Jeon et al., 2001; Liu et al.,

    1996; Liu et al., 2007). Initial pH of the treatment lines were

    measured continuously in the period of study. Average influent

    pH of Line A and B were 7.55 60.156 and 7.55 60.157. Table 3

    shows that the effect of primary sedimentation on wastewater pH

    were negligible (r50.947). In this type of EBPR process,

    Figure 4Experimental setup for aerobic phosphorus

    uptake batch tests.

    Table 1Influent characteristics of the treatment lines (COD = chemical oxygen demand; rbsCOD = readily

    biodegradable COD; VFAs = volatile fatty acids; TP = total phosphorus; HAc = acetic acid).

    Parameter Treatment line Data number Mean Standard deviation Standard error mean

    Flow m3/s A 151 2.218 0.174 0.017

    B 151 2.213 0.166 0.016

    COD mg/L A 151 668 132.746 15.128

    B 77 530 155.576 12.661

    rbsCOD mg/L A 77 195.3 39.648 4.518

    B 77 193.8 39.959 4.554

    VFAs mg HAc/L A 77 76.78 1.600 76.78

    B 77 75.19 1.625 75.19

    TP mg/L A 150 10.69 2.151 0.175

    B 77 9.51 1.733 0.197

    rbsCOD/TP A 77 19.26 6.631 0.755

    B 77 21.34 7.525 0.826

    Tuncal et al.

    September 2009 889

  • 7/26/2019 Water Environment Research

    5/10

    optimum MLVSS concentration should be within the range of 2.5

    to 3.0 g/L (Metcalf and Eddy, 2003; U.S. EPA, 1987). MeasuredMLVSS concentrations of the treatment lines also were examined.

    Average MLVSS concentrations were 2.654 60.582 and 2.491

    60.241 g/L for Line A and Line B.

    Previous studies indicated that presence of electron acceptors

    (nitrates and dissolved oxygen) in the anaerobic zone could be

    detrimental to anaerobic EBPR activities (Metcalf and Eddy,

    2003; U.S. EPA, 1987). In the period of study, nitrates and air

    entrainment were monitored continuously in both influent and

    RAS of the treatment lines. Electron acceptor concentrations in

    the RAS and influent were summarized in Table 4. Measurement

    results indicated that RAS of both treatment lines contained

    different concentrations of nitrate, and the mean difference

    between RAS nitrate concentrations was 0.68 mg/L. The RAS

    of Line A contained fewer nitrates, which could be attributed tobetter denitrification efficiency. In addition, influents of both lines

    contained negligible amount of nitrate (,0.2 mg/L). It also was

    reported that 4 to 5 mg rbsCOD are removed by 1 mg/L NO3-N

    (Metcalf and Eddy, 2003; U.S. EPA, 1987). Nitrate concentrations

    in the RAS of the treatment lines were not in a range that would

    have led to serious EBPR deterioration with particular reference to

    influent rbsCOD/TP ratios. Furthermore, the nitrate level of the

    RAS could support the proliferation of denitrifying PAOs

    (DPAOs) (Falkentoft et al., 2002; Ostgaard et al., 1997).

    Measurement results also indicated that there was significant air

    entrainment into influent and RAS of the treatment lines.

    Turbulences in the wastewater distribution chambers caused high

    concentrations of dissolved oxygen in the influent. Dissolved

    oxygen in the RAS originated from the hydraulic design of the

    RAS pumping station. It has been reported that the presence of

    even low levels of oxygen or other oxidizing substances can

    deteriorate anaerobic PAOs activities (Furumai et al., 1999; Kuba

    et al., 1996; Rensik et al., 1997). It could be concluded that such asignificant dissolved-oxygen load into the anaerobic tanks could

    deteriorate the overall EBPR efficiency in the plant.

    Statistical evaluation of the data compiled in this study revealed

    that organic loading rate and air entrainment into the anaerobic

    tanks interfere with EBPR performance. Detailed wastewater

    characterization revealed that differences in organic loading rates

    originated from biodegradable particulate organic compounds as

    well.

    Observed Differences in Overall Performance

    Statistical evaluation of effluent compositions indicated that

    PO4-P concentration of the treatment line operating without

    primary sedimentation (Line A) was lower than the othertreatment line operating with primary sedimentation (Line B).

    Average PO4-P concentration of Line A and Line B were 1.699

    60.505 and 3.822 61.547 mg/L. Mean difference between

    effluent PO4-P concentrations of the two treatment lines were

    also significant (r 5 0.000).

    Good sludge settling characteristics of activated sludge are

    essential for the entire effluent quality and SVI is one of the

    important indicators of sludge settling behavior (Lee et al., 1996;

    Scruggs and Randall, 1998). Results also revealed that SVI values

    of the treatment lines were different from each other (r 5 0.000)

    and mean difference between Line A and Line B was 39 ml/g.

    Average SVI of Lines A and B were 124 634.557 and 163

    618.557 ml/g, respectively. Abundance of filamentous microor-

    ganisms in the process was one of the most serious operational

    problems during the study. Thickness of the foam layer

    (originating from filamentous activity) was significantly lower

    in the treatment line operating without primary sedimentation

    (Line A). Different mass fraction of filamentous organisms in the

    activated sludge could explain this relatively low SVI. It could

    also be concluded that increased organic loading rate (F/M ratio)

    would limit domination of filamentous microorganisms. Prolifer-

    ation of filamentous microorganisms could be attributed to air

    entrainment in the anaerobic tanks. Increased organic loading rate

    would help reduce dissolved oxygen uptake.

    Figure 5 shows both effluent PO4-P concentrations and SVI

    values of the treatment lines. Graphical evaluation of the results

    Table 2Statistical comparison of influent characteristics

    of Lines A and B (COD = chemical oxygen demand;

    rbsCOD = readily biodegradable COD; VFAs = volatile

    fatty acids; TP =total phosphorus; HAc =acetic acid).

    Parameter r* Mean difference

    95% confidence in-

    terval of difference

    Lower Upper

    Flow m3/s 0.782 0.005 20.030 0.044

    COD mg/L 0.000 138.02 97.101 178.94

    rbsCOD mg/L 0.823 1.442 211.232 14.116

    VFAs mgHAc/L 0.488 1.584 22.922 6.09

    TP mg/L 0.000 1.188 0.631 1.746

    rbsCOD/TP 0.066 22.076 24.289 0.135

    * Two-tailed significance value obtained from t-test.

    Table 3Statistical comparison of initial pH, mixed-

    liquor volatile suspended solids (MLVSS) and sludge age

    of treatment lines.

    Parameter r Mean difference

    95% confidence

    interval of difference

    Lower Upper

    Initial pH 0.947 0.001 20.034 0.036

    MLVSS g/L 0.002 0.163 0.062 0.264

    Sludge age days 0.000 21.013 21.549 20.478

    * Two-tailed significance value obtained from t-test.

    Table 4Electron acceptor concentrations in influent

    and return activated sludge (RAS) of treatment lines.

    Line A Line B

    Influent RAS Influent RAS

    Nitrate mg/L minimum ND 0.06 ND 0.20maximum ND 4.13 ND 2.75

    average ND 0.82 ND 1.50

    Dissolved

    oxygen, mg/L

    minimum 1.63 1.76 1.48 1.54

    maximum 2.86 2.83 2.60 2.60

    average 2.18 2.22 1.98 1.82

    ND 5 under measurement range (,0.2 mg/L).

    Tuncal et al.

    890 Water Environment Research,Volume 81, Number 9

  • 7/26/2019 Water Environment Research

    6/10

    clearly show that operation without primary sedimentation

    resulted with lower effluent PO4-P concentrations and better

    sludge-settling characteristic. The Relatively low SVI could also

    be explained with the maintenance of heavier solids in the aeration

    basin in addition to lower filamentous mass fraction within the

    activated sludge.

    The NH4-N and NO3-N concentrations in the effluent of Line A

    varied from 0.38 to 1.80 mg NH4-N/L and 0.06 to 4.10 mg NO3-N

    /L; average NH4-N and NO3-N concentrations were

    0.8060.42 mg NH4-N/L and 0.8260.76 mg NO3-N/L. The

    NH4-N and NO3-N concentrations in the effluent of Line B

    ranged from 0.10 to 2.40 mg NH4-N/L and 0.22 to 2.75 mg NO3-

    N /L; average NH4-N and NO3-N concentrations were 1.10

    60.12 mg NH4-N/L and 1.6 60.52 mg NO3-N /L. Nitrogen

    removal efficiencies of the treatment lines were 85% 62.1 and

    83% 62.1 for Line A and B within the period of study.

    Comparison of EBPR Activities Observed in Batch Tests

    Batch-scale experiments were conducted to compare the

    characteristics of activated sludge samples obtained from the

    two treatment lines and to validate the results of field

    measurements. Because previous studies clearly showed that the

    optimal carbon source for anaerobic microbial metabolism was

    acetate, other carbon sources were not used in these tests

    (Brdjanovic, 1998; Pala and Bolukbas, 2005).

    Representative anaerobic phosphorus release and sCOD (acetate)

    use profiles of the treatment lines are shown in Figures 6 and 7.

    Average phosphorus release rate of Line A and B were 0.27 and

    0.18 mgP/gVSS?min,respectively;averageanaerobic sCODuse rates

    were 1.10 and 0.72 mg sCOD/g VSS?min. It is evident from these

    faster anaerobic biochemical reaction rates that PAOs domination

    withinthe activated sludgesystemcouldalso be supported by pbCOD.

    Representative anoxic phosphorus uptake and denitrification

    profiles of the treatment lines are shown in Figures 8 and 9.

    Anoxic phosphorus uptake batch test results of both treatment

    lines indicated an exponential reaction kinetic with significant

    correlation coefficients (r2

    5 0.98 for Line A; r2

    5 0.99 for LineB). Average anoxic phosphorus uptake rates were 0.10 and

    0.06 mg P/g VSS?min for Line A and B, respectively. Interest-

    ingly, nearly all of the released PO4-P was taken up by the

    activated sludge obtained from Line A during the anoxic

    conditions in less than 2.5 hours.

    Figure 5Difference in effluent PO4-P concentrations and sludge volume index (SVI) of the treatment lines in the

    period of study.

    Figure 6Anaerobic phosphorus release and soluble chemical (sCOD) utilization profile of Line A.

    Tuncal et al.

    September 2009 891

  • 7/26/2019 Water Environment Research

    7/10

    As shown in Figures 8 and 9, denitrification reactions that

    happen simultaneously with anoxic phosphorus uptake indicate a

    linear reaction kinetic with significant correlation coefficients (r2

    5 1.00 for Line A; r2 5 0.99 for Line B). Average denitrification

    rates were 0.04 and 0.03 mgNO3-N/gVSS?min for Line A and B.

    Although DPAOs are able to directly use internally stored PHA as

    electron donor, ordinary denitrifiers have to use external carbon

    sources, which require further cleavage steps. Therefore, this

    faster denitrification rate may be explained by the existence of

    higher DPAOs in the activated sludge that was cultivated on

    higher pbCOD. Furthermore, SND has been found to improve

    when additional carbon is used, particularly in particulate form

    (U.S. EPA, 1993). Results of the batch tests suggest thatproliferation of DPAOs could also be caused by increased pbCOD

    loading rate. Consequently, additional nitrate removal could affect

    the nitrate concentration being recycled to the anaerobic zone,

    which, in turn, will influence the anaerobic phosphorus release

    and overall EBPR efficiency.

    Typical aerobic phosphorus uptake profile of both treatment

    lines is shown in Figure 10. Aerobic phosphorus uptake rate of

    Line A and Line B were 0.10 and 0.06 mg P/g VSS?min,

    respectively. The aerobic phosphorus uptake rate of Line A was

    higher than Line B at approximately 60%, which is exactly the

    same as anoxic uptake. Anoxic and aerobic phosphorus uptake

    rates measured in Line A and Line B were approximately equal.

    This result indicates that significant amount of PAOs that

    dominated both treatment lines could be DPAOs.

    Although batch tests indicated different denitrification rates,

    nitrogen removal performance of the treatment lines were notsignificantly different. This situation could be explained with

    lower nitrogen load (50% 62.54) to the aeration basins compared

    to the assumed nitrogen loading criteria at design level.

    Figure 7Anaerobic phosphorus release and soluble chemical oxygen demand (sCOD) use profile of Line B.

    Figure 8Anoxic phosphorus uptake and nitrate use profile of Line A.

    Tuncal et al.

    892 Water Environment Research,Volume 81, Number 9

  • 7/26/2019 Water Environment Research

    8/10

    The EBPR characteristics of the activated sludge that were

    determined by performing batch tests are summarized in the

    Table 5. The table shows that 4.0 mg sCOD is required for 1 mg

    phosphorus release. Nitrate requirements of the activated sludge

    samples to remove 1 mg phosphorus were different, however. The

    NO3-Nutilized/Puptake ratios of Lines A and B were 0.38 and

    0.49 mg NO3-N/mg P, respectively. This result indicates an

    improved reaction yield of 77%. The higher yield efficiency could

    be explained by domination of DPAOs.

    Results of batch tests also proved that difference in effluent

    PO4-P concentrations were caused by both metabolic phospho-

    rus removal and increased PAOs mass fraction. Because there

    was not sufficient biochemical reaction time for conversion of

    pbCOD in the anaerobic period, these organic materials likely

    were converted to simple substrate forms later in the oxidative

    zones. It also was reported that denitrifiers grown in SND

    processes can use both soluble and particulate carbon forms

    (U.S. EPA, 1993). In addition, a significant amount of nitrate-

    rich activated sludge containing PAOs and enriched with

    polyphosphate is recycled from the aerobic zone to the anoxic

    zone in EBPR processes by internal recirculation. Therefore,

    some of the PAOs could use simple substrate forms that

    remained from endogenous respiration and/or cleavage of

    biodegradable particulate compounds depending on diffusion

    limitations (absence of both dissolved oxygen and nitrates and

    presence of readily biodegradable substrate forms) within the

    floc structure. This mechanism could be executed by the energy

    made available from phosphorus release reactions in the anoxic

    zone. Results of this investigation suggest that PHA storage/

    degradation may continue in the anoxic environment as well.

    This idea is supported by occurrence of anoxic/anaerobic

    conditions within the activated sludge flocks even in aerobic

    zones (Metcalf and Eddy, 2003; U.S. EPA, 1993). By this

    mechanism, DPAOs may survive and proliferate in the activated

    sludge without being exhausted even in the strongly competitive

    (substrate-limited) long oxidative periods.

    Figure 9Anoxic phosphorus uptake and nitrate use profile of Line B.

    Figure 10Aerobic phosphorus uptake profiles of the treatment lines (A and B).

    Tuncal et al.

    September 2009 893

  • 7/26/2019 Water Environment Research

    9/10

    It was reported that settlement of solids in the channel type

    anaerobic tanks could improve performance of full-scale EBPR

    processes (U.S. EPA, 1987). This improvement can be explained

    with fermentation of the settled solids that will further increase the

    anaerobic PAOs activities. Inventory of anaerobic tanks is mixed

    by submersible mixers and, depending on design, horizontal

    velocity of the activated sludge in the channels is approximately

    0.3 m/s in the plant. In this study, all mixers were in operation,

    and influent of both treatment lines contained adequate level of

    readily biodegradable or fermentable substrate forms (rbsCOD

    and VFAs). Nevertheless, settlement of some part of suspended

    particles may also be expected. Therefore, some part of EBPR

    improvement in the treatment line operating without primary

    sedimentation may also be attributed to fermentation of settled

    solids in the anaerobic tank.

    Conclusion

    In this study, biological treatment performance and activated

    sludge characteristics of two treatment lines operating with and

    without primary sedimentation were investigated in a large-scale

    EBPR process. Results indicate that, in addition to soluble

    substrate forms, particulate biodegradable compounds present in

    the wastewater also could affect performance of EBPR processes

    by several mechanisms. Possible pathways of interaction between

    pbCOD and EBPR could be concluded as follows

    (1) Proliferation of DPAOs within the activated sludge culture

    appears important for stability of EBPR performance and the

    entire effluent quality. Increased organic loading rate, even in

    the particulate form, could support survival and proliferation

    of DPAOs in EBPR processes with SND configuration.

    (2) Increased DPAOs mass fraction could increase the denitrifi-

    cation rate. Thus, detrimental effects of nitrates on anaerobic

    EBPR activities would be improved.

    (3) Settlement of particulate solids in the anaerobic tanks also

    may occur when operating without primary sedimentation.

    Fermentation of these settled organic materials would have led

    to increased EBPR performance.

    (4) Insufficient EBPR performance of both treatment lines could

    be attributed to significant air entrainment in anaerobic tanks

    from influent and RAS. In the case of air entrainment to the

    anaerobic zone, the larger organic load would help to improve

    anaerobic conditions and EBPR performance. Significant air

    entrainment into anaerobic tanks also may support prolifera-

    tion of filamentous microorganisms. Larger organic load to

    the anaerobic tank would control the mass fraction of

    filamentous microorganisms and improve overall effluent

    quality.

    Submitted for publication November 21, 2007; revised manu-

    script submitted August 6, 2008; accepted for publication June 2,

    2009.

    ReferencesAMERICAN PUBLIC HEALTH ASSOCIATION (1998) Standard Methods for the

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

    Health Association: Washington, D.C.; American Water Works

    Association: Denver, Colorado; Water Environment Federation:

    Alexandria, Virginia.

    Barajas M. G; Escalas A; Mujeriego R. (2001) Fermentation of Low VFA

    Wastewater in an Activated Primary Tank. Water SA, 28, 8998.

    Brdjanovic, D. (1998) Modeling Biological Phosphorus Removal in

    Activated Sludge Systems. A.A. Balkema Publishers: Leiden, The

    Netherlands.

    Brdjanovic, D.; Van Loosdrecht, M. C. M.; Hooijmans, C. M.; Alaerts, G.

    J. F. R.; Heijnen J. J. (1998) Temperature Effects on Physiology of

    Biological Phosphorus Removal. J. Environ. Eng., 123 (2), 2323

    2328.

    Converti, A.; Rovatti, M.; Del Borghi, M. (1995) Biological Removal of

    Phosphorus from Wastewaters by Alternating Aerobic and Anaerobic

    Conditions. Water Res., 29 (1), 263269.

    Filipe, C. D. M.; Daigger, G. T.; Grady, C. P. L. (2001) Effects of pH on the

    Rates of Aerobic Metabolism of Phosphate-accumulating and Glyco-

    genaccumulating Organisms.Water Environ. Res.,73 (2), 213222.

    Jeon, C. O.; Lee, D. S.; Lee, M. W.; Park, J. M. (2001) Enhanced

    Biological Phosphorus Removal in an Anaerobic-Aerobic Sequencing

    Batch Reactor: Effect of pH. Water Environ. Res., 73 (3), 301306.

    Katehis, D.; Anderson, J.; Carrio, L.; Gopalokrishnan, K.; Stinson, B.;

    Chandran, K. (2003) Viability and Application of Primary Sludge

    Fermentate as a BNR Supplemental Carbon Source. Proceedings of

    the 76th Annual Technical Exhibition and Conference WEFTEC [CD-

    ROM]; Los Angeles, California, Oct 1115; Water EnvironmentFederation: Alexandria, Virginia.

    Kontas, A.; Kucuksezgin, F.; Altay, O.; Uluturhan, E. (2004) Monitoring of

    Eutrophication and Nutrient Limitation in the Izmir Bay (Turkey) before

    and after Wastewater Treatment Plant.Environ. Int.,29, 10571062.

    Kucuksezgin, F.; Kontas, A.; Altay, O.; Uluturhan, E.; Darlmaz, E. (2005)

    Assessment of Marine Pollution in Izmir Bay: Nutrient, Heavy Metal

    and Total Hydrocarbon Concentrations. Environ. Int., 32 (1), 4151.

    Lee, N. M.; Carlsson, H.; Aspegren, H.; Welander, T.; Andersson, B. (1996)

    Stability and Variation in Sludge Properties in Two Parallel Systems for

    Enhanced Biological Phosphorus Removal Operated with and without

    Nitrogen Removal.Water Sci. Technol., 34 (12), 101109.

    Liu, Y.; Chen, Y.; Zhou, O. (2007) Effect of Initial pH Control on

    Enhanced Biological Phosphorus Removal from Wastewater Con-

    taining Acetic and Propionic Acids. Chemosphere, 66, 123129.

    Table 5Summary of observed enhanced biological phosphorus removal activities in batch tests (sCOD = soluble

    chemical oxygen demand).

    Parameter Line A Line B Dimension

    Anaerobic phosphorus release rate 0.27 0.18 mg P/g VSS?min

    sCOD use rate 1.10 0.72 mg sCOD/g VSS?min

    sCOD consumed/Preleased 4.08 4.00 mg sCOD/mg PAnoxic phosphorus uptake rate 0.10 0.06 mg P/gVSS?min

    Denitrification rate 0.04 0.03 mg NO3-N/g VSS?min

    NO3-Nutilized/Puptake 0.37 0.49 mg NO3-N/mg P

    Aerobic phosphorus uptake rate 0.10 0.06 mg P/g VSS?min

    Tuncal et al.

    894 Water Environment Research,Volume 81, Number 9

  • 7/26/2019 Water Environment Research

    10/10

    Liu, W. T.; Mino, T.; Matsuo, T.; Nakamura, K. (1996) Biological

    phosphorus removal process Effect of pH on Anaerobic Substrate

    Metabolism. Water Sci. Technol., 34 (12), 2532.

    Lopez-Vazquez, C. M.; Hooijmans, C. M.; Brdjanovic, Gijzen, H. J; van

    Loosdrecht, M. C. M. (2008) Factors Affecting the Microbial

    Populations at Full-scale Enhanced Biological Phosphorus Removal

    (EBPR) Wastewater Treatment Plants in The Netherlands. Water

    Res., 42 (1011), 23492360.

    McClintock, S. A.; Randall, C. W.; Pattarkine, V. M. (1993) Effects of

    Temperature and Mean Cell Residence Time on Biological Nutrient

    Removal Processes. Water Environ. Res., 65 (5), 110118.

    METCALF AND EDDY (2003) Wastewater Engineering Treatment and Reuse,

    4th ed.; McGrawHill Inc.: New York.

    Pala, A.; Bolukbas, O. (2005) Evaluation of Kinetic Parameters for

    Biological CNP Removal from a Municipal Wastewater through

    Batch Tests. Process Biochem., 40, 629635.

    Schuler, A. J.; Jenkins, D. (2003) Enhanced Biological Phosphorus

    Removal from Wastewater by Biomass with Different Phosphorus

    Contents, Part I: Experimental Results and Comparison with

    Metabolic Models. Water Environ. Res., 75 (6), 485498.

    Scruggs, C. E.; Randall, C. W. (1998) Evaluation of Filamentous

    Microorganism Growth Factors in an Industrial Wastewater

    Activated Sludge System. Water Sci. Technol., 37 (45), 263

    270.

    UNITED NATIONS ENVIRONMENT PROGRAMME (1993) MAP Technical

    Report Series 72. United Nations Environment Programme: Nairobi,

    Kenya.

    U.S. ENVIRONMENTAL PROTECTION AGENCY (1987) Design Manual Phos-

    phorus Removal, EPA/625/187/001; U.S. Environmental Protection

    Agency: Cincinnati, Ohio.

    ENVIRONMENTAL PROTECTION AGENCY (1993) Manual Nitrogen Control,

    EPA/625/93/010; Environmental Protection Agency: Washington,

    D.C.

    Whang, L. M.; Park, J. K. (2006) Competition between Polyphosphate and

    Glycogen-accumulating Organisms in Enhanced Biological Phospho-

    rus Removal Systems: Effect of Temperature and Sludge Age.Water

    Environ. Res., 78 (1), 411.

    WATERPOLLUTIONCONTROL FEDERATION (1983) Nutrient Control, Facilities

    Desig n, Manual of Practice FD7; Water Pollution Control

    Federation: Washington, D.C.

    Tuncal et al.

    September 2009 895