Fluid Heat Food Process Digging Tools

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

Abstract: An anaerobic continuously stirred tank reactor (CSTR) integrated with an external tubular nanofiltration membrane module was

operated for 466 days. Studies were conducted to compare the efficiencies in flux and operational permeability recovery of two different
chemical cleaning processes. Chemically enhanced backwashes (CEBs) were applied for approximately 176 days of the study and their
influence on chemical cleaning performance and membrane fouling rate following chemical cleanings were also investigated. The anaerobic
nanofiltration membrane bioreactor used in the study was fed a high-strength, potato starch–based synthetic wastewater (average chemical
oxygen demand of 83.4 g=L and average total suspended solids of 5.8 g=L). Important findings were that a static, one-phase cleaning process
using caustic cleaning chemicals outperformed a dynamic, three-phase cleaning process that also incorporated acidic cleaning chemicals; the
acidic phase of the three-phase cleaning had a detrimental effect on the overall cleaning procedure’s performance; CEBs enhanced the per-
formance of cleaning processes; and, finally, caustic cleaning solutions provided benefits to both membrane cleaning effectiveness and rate of
fouling in the short term but may lead to an increased rate of membrane fouling in the long term. DOI: 10.1061/(ASCE)EE.1943-
7870.0001524. © 2019 American Society of Civil Engineers.
Author keywords: Anaerobic membrane bioreactor; Chemical cleaning; Chemically enhanced backwash; Membrane fouling;
Nanofiltration membrane.

Introduction anaerobic systems such as high concentration of biomass retention


(as the effectiveness of biomass retention becomes independent of
With an ever-growing population, sustainable water and waste- sludge settleability due to membranes’ near-complete solid–liquid
water treatment technologies will become essential in solving water separation capabilities), compact design, and, most importantly, the
scarcity problems prone to develop in upcoming years. Anaerobic ability to produce a superior quality effluent for possible recycle
membrane bioreactors (AnMBRs) offer a promising means of or high-quality discharge to meet stringent wastewater disposal
sustainable and value-added wastewater treatment, coalescing the regulations (Skouteris et al. 2012).
benefits of anaerobic digestion for the treatment of municipal and What continues to remain the most significant factor limiting the
industrial wastewaters with the almost 100% solid–liquid separa- cost effectiveness of membranes in wastewater treatment applica-
tion potential of membrane integrated wastewater treatment tech- tions (both municipal and industrial) is membrane fouling (Ramos
nologies. Anaerobic digestion for wastewater treatment provides et al. 2014). Membrane fouling may be defined as the combination
certain benefits over more commonly used conventional aerobic of processes leading to flux deterioration and transmembrane pres-
technologies, such as high-rate suspended-growth systems, very sure (TMP) increase due to surface or internal pore blockage of the
high removal efficiencies of biochemical oxygen demand (BOD) membrane (Lousada-Ferreira et al. 2014). Surface or internal pore
and chemical oxygen demand (COD), resource recovery (value- blockage is generated through the contact and deposition of sus-
added products such as methane in its biogas and the possible pended particles, colloids, and solutes present in the mixed liquor
recovery of select biochemicals such as volatile fatty acids in its on the membrane surface or in its pores, henceforth referred to as
mixed-liquor and effluent), lower sludge yield, and low carbon foulants (Iorhemen et al. 2016). Membrane fouling increases TMP
footprint (Kinnunen et al. 2015; Meng et al. 2009; Ozgun et al. and decreases flux, while the control of foulant deposition increases
2013; Pilli et al. 2016; Skouteris et al. 2012). Through the integra- manpower and energy requirements, demands costly chemical
tion of membrane separation technologies within anaerobic di- cleaning (that further requires waste handling), and may also
gestion, further benefits are provided over non-membrane-based result in increased deterioration of the membrane material, affecting
its lifespan (Ramos et al. 2014). Foulant mitigation for a membrane
1
Graduate Student, Dept. of Civil Engineering, Univ. of New module is a two-pronged approach: first, the membrane-fouling
Brunswick, 17 Dineen Dr., Fredericton, NB, Canada E3B 5A3 (corre- rate should be minimized, and second, technologies to restore the
sponding author). Email: h557d@unb.ca flux of a fouled membrane must be implemented (Liao et al. 2006).
2
Professor, Dept. of Civil Engineering, Univ. of New Brunswick, 17 Physical cleaning processes (e.g., gas scouring, backwashing,
Dineen Dr., Fredericton, NB, Canada E3B 5A3. Email: singhk@unb.ca relaxation) have only proven to be effective at removing some pore-
3
Process Engineer, ADI Systems, Inc., 370 Wilsey Rd., Fredericton,
clogging and cake-layer membrane foulants (Liao et al. 2006).
NB, Canada E3B 6E9. Email: gustavo.zanatta@evoqua.com
Note. This manuscript was submitted on July 18, 2018; approved on Chemical cleanings, using low- to moderate-strength corrosive or
October 26, 2018; published online on March 6, 2019. Discussion period caustic solutions, are highly effective at removing both the gel layer
open until August 6, 2019; separate discussions must be submitted and cake layer developed through solids deposition on the mem-
for individual papers. This paper is part of the Journal of Environmental brane surface and pore-blocking foulants, making them typically
Engineering, © ASCE, ISSN 0733-9372. more efficient at recovering membrane flux. These chemical

© ASCE 04019018-1 J. Environ. Eng.

J. Environ. Eng., 2019, 145(5): 04019018


cleanings can be performed either as a preventive maintenance op- are, to the authors’ best knowledge, no studies investigating the
eration to maintain permeate flux [through incorporating chemical impact of daily CEBs on the effectiveness of chemical cleaning
cleanings into regular physical cleaning processes, for example processes.
chemically enhanced backwashes (CEBs)], or as a recovery oper-
ation to recover membrane flux from a severely fouled membrane
module (Ramos et al. 2014). Chemically enhanced backwashes Materials and Methods
combine conventional physical backwashing with chemical clean-
ing in one process, which could be of high importance for the
Reactor Configuration
design and operation of membrane bioreactor (MBR) systems
for wastewater treatment. Recent studies using CEBs have typically A 1,000 L stainless steel anaerobic CSTR was equipped with an
been performed at long intervals (weekly or monthly cleaning proc- external nanofiltration tubular membrane module and operated in
esses) (Zhou et al. 2014). There have been few comparative studies cross-flow mode. A simplified process flow diagram (PFD) of the
on different cleaning methods and conditions, and scant informa- configuration may be viewed in Fig. 1. The anaerobic CSTR was
tion exists on CEBs, in general, in the literature (Ramos et al. 2014; equipped with a heat wrap, which maintained the reactor temper-
Zhou et al. 2014). Snowdon et al. (2018) observed that one 60 s ature at an average value of 36.2°C (0.5°C) throughout the dura-
CEB every 24 h resulted in increased permeate flux, operational tion of the study. The operational temperature of approximately
permeability, and net flux, demonstrating CEBs’ effectiveness as 36°C was selected since it is the upper limit of mesophilic anaerobic
a preventive maintenance operation. digestion (25°C–35°C) and maximized biogas generation, which
This research project compared the efficiency of two different was desired (Metcalf & Eddy, Inc. 1991).
chemical cleaning processes applied to one membrane module Reactor-mixed liquor was continuously recirculated from the
integrated within a continuously operated pilot-scale AnMBR sys- CSTR to the membrane module and back into the CSTR by a pro-
tem. The two chemical cleaning processes were applied to the gressive cavity pump (Nemo, New Brunswick, Canada). This pump
membrane module only once it was severely fouled, acting as a maintained a mixed-liquor flow rate of 6,000 L=h, which resulted
recovery operation. As a secondary objective, the project investi- in the targeted membrane surface cross-flow velocity of 2.5 m=s
gated the impact of daily CEBs on the effectiveness of each and a complete recirculation of reactor contents every 10 min.
chemical cleaning process. The novel pilot-scale AnMBR system A cross-flow velocity of 2.5 m=s was selected since it is deemed
consisted of a 1,000 L anaerobic continuously stirred tank reactor to be the minimum operational cross-flow velocity that must be
(CSTR) integrated with an external nanofiltration tubular mem- maintained for the membrane module based on manufacturer rec-
brane module. The reactor and membrane configuration provided ommendations. It was desired to operate at the lower spectrum of
benefits over conventional cleaning processes in that the membrane the operational cross-flow velocity to minimize energy demands.
module did not need to be removed for chemical cleanings (since it The membrane module was cylindrical in shape, composed of poly-
could be isolated from the CSTR through a recirculation loop). This vinylidene fluoride (PVDF), with an outer diameter of 50 mm and
resulted in minimal downtime and maintenance work required for a total length of 1.436 m. The module contained 13 tubular mem-
chemical cleaning processes. With few direct comparative studies brane channels, each 8 mm in diameter, which resulted in a total
on different chemical cleanings applied to the same membrane membrane area of 0.42 m2 . The membrane surfaces had a mean
module, and little information in the literature on CEBs, there pore size of 30 nm.

Fig. 1. Simplified PFD of experimental setup.

© ASCE 04019018-2 J. Environ. Eng.

J. Environ. Eng., 2019, 145(5): 04019018


Table 1. Chemical cleaning conditions
Cleaning Recirculation Soak time Number of backwashes and
process Cycle time (min) (min) duration (number and min) Chemical used
1 1 30 N/A 2 and 2 min 1% sodium hydroxide
2 N/A 20 N/A
3 30 N/A 2 and 2 min
4 30 N/A 2 and 2 min 1% citric acid
5 N/A 20 N/A
6 30 N/A 2 and 2 min
7 30 N/A 2 and 2 min 1% sodium hydroxide and sodium hypochlorite
8 N/A 20 N/A
9 30 N/A 2 and 2 min
2 1 30 N/A N/A 1% sodium hydroxide and sodium hypochlorite
2 N/A 60 N/A
3 30 N/A N/A
3 1 30 N/A N/A 1% sodium hydroxide and sodium hypochlorite
2 N/A 60 N/A
3 30 N/A N/A
4 1 30 N/A N/A 1% sodium hydroxide and sodium hypochlorite
2 N/A 60 N/A
3 30 N/A N/A
5 1 30 N/A N/A 1% sodium hydroxide and sodium hypochlorite
2 N/A 60 N/A
3 30 N/A N/A
6 1 30 N/A 2 and 2 min 1% sodium hydroxide
2 N/A 20 N/A
3 30 N/A 2 and 2 min
4 30 N/A 2 and 2 min 1% citric acid
5 N/A 20 N/A
6 30 N/A 2 and 2 min
7 30 N/A 2 and 2 min 1% sodium hydroxide and sodium hypochlorite
8 N/A 20 N/A
9 30 N/A 2 and 2 min

chemical cleaning process. Finally, the chemicals used outline the module were also measured. These values allowed the MLF
exact concentration of chemical cleaning solutions applied to the and MLOP to be calculated. The mixed-liquor recoveries were
membrane module during each cleaning condition. calculated by dividing the post–chemical cleaning MLF or post–
Clean water flux was measured by isolating the CSTR recircu- chemical cleaning MLOP values by the initial MLF or initial
lation line from the CSTR itself, allowing a 50-L tank to be filled MLOP values measured on Day 2 of operation, and then multiply-
with clean water (maintained at around 30°C) and filtered through ing by 100. Equations used to calculate clean water and mixed-
the membrane module. Each clean water test recirculated the clean liquor recoveries are displayed in Eqs. (2) and (3), respectively:
water at five flow rates in increasing order (790, 1,580, 2,360,
3,150, and 3,940 L=h), and the resultant membrane fluxes and J; OPPC
TMPs were measured. Five different increasing flow rates were se- CW J;OP ð%Þ ¼ × 100 ð2Þ
J; OPo
lected to provide incremental increases in flux and TMP, reaching
the maximum flow rate of the recirculation loop at approximately
3,940 L=h. Operational permeability is defined as membrane flux
divided by instantaneous operational TMP, in other words pressure- J; OPPC
MLJ;OP ð%Þ ¼ × 100 ð3Þ
averaged flux, and may be defined by Eq. (1) (Hamden de Andrade J; OPo
et al. 2013):
Ji In Eq. (2), CW J;OP is the clean water flux or operational per-
OPi ¼ ð1Þ
ΔPi meability recovery (%), J; OPPC is the post–membrane cleaning
clean water membrane flux or clean water operational permeability
In Eq. (1), OPi is the instantaneous operational permeability (L=m2 =h, L=m2 =h=bar), and J, OPo is the initial clean water mem-
(L=m2 =h=bar), J i is the instantaneous membrane flux (L=m2 =h), brane flux or clean water operational permeability (L=m2 =h,
and ΔPi is the instantaneous transmembrane pressure (bar). Opera- L=m2 =h=bar). In Eq. (3), MLJ;OP is the mixed-liquor flux or opera-
tional permeability was considered so that TMP could be factored tional permeability recovery (%), J; OPPC is the post–membrane
into the flux recovery analysis. Clean water recoveries were calcu- cleaning mixed-liquor membrane flux or mixed-liquor operational
lated by dividing the post–chemical cleaning clean water flux or permeability (L=m2 =h, L=m2 =h=bar), and J; OPo is the initial
post–chemical cleaning clean water OP by the initial clean water mixed-liquor membrane flux or mixed-liquor operational per-
flux or initial OP values measured on Day 1 of operation, and then meability (L=m2 =h, L=m2 =h=bar).
multiplying by 100. Once the membrane module was placed back Membrane resistance coefficients were calculated using the
inline for mixed-liquor recirculation, and after 24 h of recirculation resistance-in-series model [Eq. (4)]. This equation was adopted and
at a flow rate of 6,000 L=h, the flux and TMP of the membrane developed from other studies (Ruigomez et al. 2017):

© ASCE 04019018-4 J. Environ. Eng.

J. Environ. Eng., 2019, 145(5): 04019018


cleanings can be performed either as a preventive maintenance op- are, to the authors’ best knowledge, no studies investigating the
eration to maintain permeate flux [through incorporating chemical impact of daily CEBs on the effectiveness of chemical cleaning
cleanings into regular physical cleaning processes, for example processes.
chemically enhanced backwashes (CEBs)], or as a recovery oper-
ation to recover membrane flux from a severely fouled membrane
module (Ramos et al. 2014). Chemically enhanced backwashes Materials and Methods
combine conventional physical backwashing with chemical clean-
ing in one process, which could be of high importance for the
Reactor Configuration
design and operation of membrane bioreactor (MBR) systems
for wastewater treatment. Recent studies using CEBs have typically A 1,000 L stainless steel anaerobic CSTR was equipped with an
been performed at long intervals (weekly or monthly cleaning proc- external nanofiltration tubular membrane module and operated in
esses) (Zhou et al. 2014). There have been few comparative studies cross-flow mode. A simplified process flow diagram (PFD) of the
on different cleaning methods and conditions, and scant informa- configuration may be viewed in Fig. 1. The anaerobic CSTR was
tion exists on CEBs, in general, in the literature (Ramos et al. 2014; equipped with a heat wrap, which maintained the reactor temper-
Zhou et al. 2014). Snowdon et al. (2018) observed that one 60 s ature at an average value of 36.2°C (0.5°C) throughout the dura-
CEB every 24 h resulted in increased permeate flux, operational tion of the study. The operational temperature of approximately
permeability, and net flux, demonstrating CEBs’ effectiveness as 36°C was selected since it is the upper limit of mesophilic anaerobic
a preventive maintenance operation. digestion (25°C–35°C) and maximized biogas generation, which
This research project compared the efficiency of two different was desired (Metcalf & Eddy, Inc. 1991).
chemical cleaning processes applied to one membrane module Reactor-mixed liquor was continuously recirculated from the
integrated within a continuously operated pilot-scale AnMBR sys- CSTR to the membrane module and back into the CSTR by a pro-
tem. The two chemical cleaning processes were applied to the gressive cavity pump (Nemo, New Brunswick, Canada). This pump
membrane module only once it was severely fouled, acting as a maintained a mixed-liquor flow rate of 6,000 L=h, which resulted
recovery operation. As a secondary objective, the project investi- in the targeted membrane surface cross-flow velocity of 2.5 m=s
gated the impact of daily CEBs on the effectiveness of each and a complete recirculation of reactor contents every 10 min.
chemical cleaning process. The novel pilot-scale AnMBR system A cross-flow velocity of 2.5 m=s was selected since it is deemed
consisted of a 1,000 L anaerobic continuously stirred tank reactor to be the minimum operational cross-flow velocity that must be
(CSTR) integrated with an external nanofiltration tubular mem- maintained for the membrane module based on manufacturer rec-
brane module. The reactor and membrane configuration provided ommendations. It was desired to operate at the lower spectrum of
benefits over conventional cleaning processes in that the membrane the operational cross-flow velocity to minimize energy demands.
module did not need to be removed for chemical cleanings (since it The membrane module was cylindrical in shape, composed of poly-
could be isolated from the CSTR through a recirculation loop). This vinylidene fluoride (PVDF), with an outer diameter of 50 mm and
resulted in minimal downtime and maintenance work required for a total length of 1.436 m. The module contained 13 tubular mem-
chemical cleaning processes. With few direct comparative studies brane channels, each 8 mm in diameter, which resulted in a total
on different chemical cleanings applied to the same membrane membrane area of 0.42 m2 . The membrane surfaces had a mean
module, and little information in the literature on CEBs, there pore size of 30 nm.

Fig. 1. Simplified PFD of experimental setup.

© ASCE 04019018-2 J. Environ. Eng.

J. Environ. Eng., 2019, 145(5): 04019018


Wat. Sci. Tech. Vol. 28. No.2. pp. 165-176. 1993. 0273-1223/93 $6'00 0'00
Printed in Great Britain. All rights reserved. Copyright © 1993 IAWQ

ANAEROBIC-AEROBIC TREATMENT OF
POTATO-STARCH WASTEWATER

U. Abeling and C. F. Seyfried

Institut for Siedlungswasserwirtscha/t und Abfalltechnik (Water Supply and Treatment of


Solid Wastes). University of Hannover. Welfengarten I. D-3000 Hannover I. Germany

ABSTRACT

Production of potato starch produces wastewaters witb high concentrations of COD, nitrogen and
phosphorus. The best known and most economical solution for tbe nearly full elimination of tbese substances
is tbe two-stage anaerobic-aerobic treatment. The anaerobic pretreatment must only go so far as to maintain
enougb COD for biological nitrogen and phosphorus elimination in tbe aerobic stage. To optimize tbis
process, several tests have been carried out on nitrification/denitrification by means of tbe intermediate
product nitrite. The carbon consumption amounts to only 60% in comparison witb denitrification via nitrate.
The essential parameter for regulating tbe process is tbe concentration of free ammonia in tbe reactor.
Concentrations of I to 5mg NH31l inhibit tbe nitratation but not tbe nitritation. A separation of botb partial
steps is possible. The content of ammonia was controlled by means of continuous NH4 and pH measuring.

The process is suitable for many wastewaters witb low CODrrKN ratios and higb nitrogen concentrations. If
tbere are undesired nitrite peaks in an aerobic treatment plant caused by bigh pH values and temperatures,
specific nitrogen elimination by means of nitrite is a reliable treatment system.

KEYWORDS

Potato-starch wastewater; anaerobic-aerobic treatment; high-strength ammonium wastewater;


nitrification/denitrification via nitrite; nitritation; nitratation; industrial wastewater; nitrogen removal.

INTRODUCTION

In Europe, about 4.7 Mio. t starch and starch decay products are produced each year. The main raw materials
used are com (60% of the production) and potatoes and wheat (nearly 20% each). The largest product group
is the true starches with nearly 30%, followed by glucose (23%), isoglucose (13%) and modified starch
(12%). Most starch is employed in the food industry (for example sweets, drinks, fruit manufacturing - total
57%), the paper and paperboard industry (25%), as well as the chemical and pharmaceutical industry (10%).
Whereas a short time ago most of the attention was directed to production processes and commercializing,
nowadays wastewater avoidance and wastewater treatment are becoming more and more important. The cost
share of advanced wastewater treatment in the total production costs amounts to nearly 15%.

The starch wastewaters have high nitrogen and phosphorus concentrations. In Germany both nutrients must
be largely eliminated, and this will soon apply to the whole of the European Community as far as the
harmonization of environmental protection is concerned.

For purification of industrial wastewaters, especially for starch factories, more and more two-stage
anaerobic-aerobic processes have been used during the last few years. This is useful - for energetic reasons
- when highly concentrated wastewaters are to be treated, as a substantial part of the organic load can be

1 65
Potato-starch wastewater 167

t:----I:;;.;;;,;::;;;;�)_---__1 transport and


'---.--./ washing water

starch/potato - protein -coagulation )..-----l potato juice


juice seperation

S��t-----1 process water

Figure I. Flow diagram of potatO staJeh production

TABLE 1. Potato-Starch Production - Wastewater Characteristics and Specific volumes

washing water process water potato juice


(after protein-coagulation)

pH-Value 6 - 7 6 - 7 5 - 6
COD [mg/l] 2,000 - 4,000 4,000 - 8,000 50,000 - 60,000 (25,000 - 30,000)
TKN [mg/l] 150 - 30 150 - 600 3,600 - 4,400 (1,800 - 2,200)
tot. P [mg/l] 20 - 40 20 - 40 800 - 1,000 (400 - 500)

m3/t potatoes 0. 3 - 1.0 0.6 - 1.3 0.8 - 1.0

The washing water resulting from potato washing is recycled. The solid matter (earth, plant parts etc.) is
eliminated in a sedimentation tank, and any lost part of the water is substituted by fresh water continuously.
The longer the campaign lasts, COD and TKN concentration increases, because the potato quality decreases
for the reason of storage. The process water (starch washing water) contains the contamination component
fme starch, which cannot be gained. The main component of the high strength potato juice is protein. It goes
into a protein extraction, where about 50% of the protein is coagulated, separated, dried and packed for sale.
A still highly contaminated wastewater with 25-30g CODn remains.

Modern manufacturers use circular movement of the water during washing and starch extraction to reduce
the specific wastewater quantities to keep the values preferably in the low regions of the range in Table 1.
The application of an effective technique for starch extraction also reduces the concentrations. Anyway, in
the field of potato juice the technical limits are reached by reducing the concentration by 50% by means of
protein extraction. The wastewater percentage is also layed down by the water content of the potatoes and
changes only slightly. Table 2 shows the specific wastewater quantities and concentrations of a potato starch
factory which was opened in 1987 in Germany (Liichow). The treatment capacity amounts to 300 000
t1campaign (end August until mid March).
Bioresource Technology 98 (2007) 3298–3308

trification, (2) the distinction between the anoxic heterotrophic yield and the aerobic heterotrophic yield, respectively equal to 0.53 and 0.6
and (3) the first-order hydrolysis of the slowly biodegradable fraction. The calibration and the validation of the model was performed
using experimental data from three experiments with two piggery wastewaters. A set of kinetic and stoichiometric parameters emerged
from these tests. Except the kinetic of hydrolysis of the slowly biodegradable organic matter varying from 6 to 25 g COD (g COD day)1,
all other parameters were similar for all experiments. The dissolved oxygen concentration was identified as the main variable influencing
the nitrite accumulation during nitrification. In the calibrated model, the oxygen half-saturation coefficient of the ammonium oxidisers
(0.3 g O2 m3) was lower than for the nitrite oxidisers (1.1 g O2 m3), leading to nitrite accumulation when the dissolved oxygen concen-
tration was low. Simulations with the proposed model could be very useful for improved design and management of biological treatment
of piggery wastewaters, particularly in case of partial nitrification to nitrite directly followed by denitrification.
 2006 Elsevier Ltd. All rights reserved.

Keywords: Piggery wastewater; Modelling; Nitrification; Denitrification; Carbon oxidation; AMA


F. Béline et al. / Bioresource Technology 98 (2007) 3298–3308 3299

Nomenclature

SNH NHþ 4 þ NH3 nitrogen (g N m )


3
SS readily biodegradable organic matter (g O2 m3)
SND soluble biodegradable organic nitrogen XS slowly biodegradable organic matter (g O2 m3)
(g N m3) SI soluble inert organic matter (g O2 m3)
XND particulate biodegradable organic nitrogen XI particulate inert organic matter (g O2 m3)
(g N m3) XBH active heterotrophic biomass (g O2 m3)
XNI particulate inert organic nitrogen (g N m3) XBAI active NHþ 3
4 oxidisers biomass (g O2 m )
SNOI NO 3
2 nitrogen (g N m ) XBAA 
active NO2 oxidisers biomass (g O2 m ) 3

SNOA NO3 nitrogen (g N m3) XP particulate inert organic matter from biomass
SO oxygen (g O2 m3) decay (g O2 m3)

wastewater, a model allows to quickly screen many poten- chemical oxygen demand (COD) hydrolysis could signifi-
tial design and management alternatives. cantly affect the denitrification.
For the evaluation and the optimization of the treat- The objective of this work was to develop and to cali-
ment of urban wastewaters, various models have been brate a mathematical model for the simulation of the mod-
developed and edited by the IWA task group (Henze ified SBR, treating piggery wastewaters. First, an analysis
et al., 1987), such as the activated sludge models (ASM). of the processes involved in nitrogen removal was carried
These models are largely used for domestic sewage but out based on the experimental results. Then, a mathemati-
the modelling is overlooked in the treatment of livestock cal model was built and calibrated using the experimental
wastes and industrial effluents (Orhon, 1998). The available data.
activated sludge models were specifically designed and cal-
ibrated for urban wastewater treatment, and therefore, the
extension of such models to more concentrated influents
such as piggery wastewaters, appears particularly interest-
ing but requires specific experiments and research.
In contrast to the domestic sewage, a very limited
amount of data on the kinetic and stoichiometric processes
and on the influent fractionation are available for piggery
wastewaters (Andreottola et al., 1997). The ASM models
describe the processes of nitrification and denitrification
as one-step processes. However, nitrogen removal for con-
centrated wastewaters, as the piggery wastewaters, might
result in the inhibition of nitrification by free ammonia
and nitrous acid (Anthonisen et al., 1976), and nitrification
might be limited by the dissolved oxygen (DO) concentra-
tion (Wiesmann, 1994). Therefore, the simplified assump-
tion, used in the ASM models, is no longer applicable
for the description of nitrogen removal processes in this
case. In fact, the free ammonia inhibition and/or the
nitrous acid inhibition and/or the dissolved oxygen limita-
tion frequently lead to nitrite accumulation. High temper-
atures can also lead to nitrite accumulation (Hellinga
et al., 1998). Consequently, more sophisticated model
assumptions are required for an adequate description of
the system efficiency. In this case, each process must
be described as a two-step process including nitrite as
intermediate (Andreottola et al., 1997). Moreover, some
authors studied and compared the heterotrophic yields
under aerobic and anoxic conditions (Gujer et al., 1999;
Orhon et al., 1996; Boursier et al., 2004). The observed
lower anoxic heterotrophic yield leads to a lower COD to
NO 3  N ratio for the denitrification. That is more partic-
ularly important for the wastewaters with high N concen-
tration where the kinetics and the stoichiometry of the
3300 F. Béline et al. / Bioresource Technology 98 (2007) 3298–3308

known amount of piggery wastewater was added in the the reactor reached steady-state conditions (after more
reactor through a buffer tank where the quantity added than three hydraulic residence times), the characterisation
was weighed and recorded. The reactor was equipped with of the effluent was done (COD, soluble COD and total
a foam breaker working during the aerobic phases and a nitrogen) and was followed by a study of the evolution
fine bubbles diffuser fed with a compressor (2–5 m3 h1). of NHþ  
4 , NO2 and NO3 in the reactor, during one
For the mixing, a centrifugal pump (30 m3 h1) circulated cycle.
the activated sludge, during the anoxic phases from the Immediately after run 1, a second study of the evolution
 
bottom to the top of the reactor. Water was continuously of NHþ 4 , NO2 and NO3 , during one cycle, was performed
added in the reactor (65 l per day) to compensate for the (run 2). At the beginning of this cycle, 1 l of ammonium
evaporation caused by the high temperature and the aera- carbonate solution (13.5 kg N m3) was added in the reac-
tion. At the end of each cycle, an amount of the activated tor during the feeding phase, together with the piggery
sludge equal to the piggery wastewater added at the begin- wastewater (PW1). For this experiment, the other experi-
ning of the cycle was withdrawn from the reactor. The bio- mental conditions were similar to run 1.
logical reactor was equipped with on-line measurements of For run 3, as for run 1, the reactor was initially filled
DO (WTW Oxi197 oxygen meter), temperature (Ponselle with 50 l of activated sludge coming from the same farm
PONAPF-TTRANS), oxidation–reduction potential (Pon- and 50 l of tap water. About 600 l of raw piggery waste-
selle PONAPF-EHTRANS) and pH (Ponselle PONAPF- water was collected in an experimental farm near Rennes
PHTRANS) allowing measurements at 10 min. intervals. (France) and was mechanically sieved (1 mm) using a
The system operated without sludge recirculation nor rotary screen. The liquid fraction called piggery wastewater
decantation in the reactor, leading to a solids retention 2 (PW2) was stored at 4 C and was used for run 3. For this
time equal to the hydraulic retention time. For each exper- run, the duration of a cycle was equal to 12 h, consisting of
iment, each whole-cycle of the modified SBR was scheduled 7 h of anoxic conditions and 5 h of aerobic conditions. At
as follows: the beginning of each cycle, 4.4 l of PW2 was added in the
reactor, leading to a working volume of 104.4 l and a HRT
• a feeding phase, equal to 11.9 days. At the end of each cycle, the same
• an anoxic phase, amount of activated sludge was withdrawn. A characterisa-
• an aerobic phase, tion of the effluent (COD, soluble COD and total nitrogen)
 
• a withdrawal phase. and a study of the evolution of NHþ 4 , NO2 and NO3 dur-
ing a cycle, were performed when more than 3 hydraulic
residence times had passed.
2.2. Experimental The main characteristics of each piggery wastewater
used in this study are presented in Table 2.

Runs PW HRT (days) P1 (h) P2 (h) P3 (h) P4 (h) Observations


1 1 11.0 0.1 3.9 19.9 0.1 Effluent characterisation and study of a cycle
2 1 11.0 0.1 3.9 19.9 0.1 Study of a cycle (ammonium carbonate was added)
3 2 11.9 0.1 6.9 4.9 0.1 Effluent characterisation and study of a cycle
PW: piggery wastewater, P1: feeding phase, P2: anoxic phase, P3: aerobic phase, P4: withdrawal phase.
Total solids (TS), volatile solids (VS), COD and total
nitrogen (TN) were analysed according to standard
methods (APHA, 1992). The soluble COD (CODs) was
analysed after centrifugation at 4000 rpm during 15 min.
Volatile fatty acids (VFA) were analysed using HPLC
(Peu et al., 2004). NHþ
4 was analysed by steam distillation
(Bremner and Keeney, 1965) directly on the effluent
(NHþ 4TOT ) and on the liquid fraction after centrifugation
 
at 4000 rpm during 15 min (NHþ 4SOL ). NO2 and NO3 were
analysed by ion chromatography on a Dionex DX120
system equipped with a Dionex AS9HC column. Eluant
was a 10 mM Na2CO3 solution, flowing at 1 ml min1.
Respirometry, widely recognised as the most convenient
tool to determine the biodegradable wastewater fractions
(Sollfrank and Gujer, 1991), was used for the COD frac-
tionation during this study. The oxygen uptake rates
(OUR) were measured for each wastewater and allowed
to determine the readily biodegradable COD (SS) and the
slowly biodegradable COD (XS) as described by Orhon
and Okutman (2003); Boursier et al. (2005). The inert
COD fraction was calculated as the difference between
total COD and biodegradable COD (SS + XS).

2.4. Modelling

For the modelling, the mathematical equations were


implemented in the Scilab software using the Scicos
interface. This software is available as a toolbox called
AS-Model1.0 (see Web references). The integration routine
was chosen automatically by the software. A run time of
more than three times the HRT was used to establish the
pseudosteady-state conditions. For the calibration of the
model, the parameters were estimated manually by com- As shown in Fig. 2, the NO 3 concentration decreased
paring the simulated output with the experimental data quickly from 210 to 0 mg N l1 during the anoxic phase.
1
until the best fit was achieved. Concurrently, an accumulation of NO 2 up to 115 mg N l
occured. During the aerobic phase, an increase of NO 3
3. Results and discussion from 0 to 220 mg N l1 was obtained and an accumulation
1
of NO 2 up to 110 mg N l was also observed. The same
3.1. Experimental results evolution of nitrogen, particularly the accumulation of
NO 2 during the nitrification and the denitrification, was
The evolution of the nitrogen species during a cycle was observed for all runs and was previously described by
used for a better characterisation of processes involved in several authors (Andreottola et al., 1997; Kim et al.,
the nitrogen transformations. As an example, the evolution 2004; Huang and Hao, 1996).
of the nitrogen forms during run 1 is presented in Fig. 2. Concerning the denitrification process, various environ-
At the beginning of the cycle, the addition of piggery mental factors were found to be responsible for nitrite
wastewater in the reactor led to an increase of NHþ4TOT from
accumulation: the type and quantity of organic substrate,
163 to 465 mg N l1 and an increase of NHþ 4SOL from 0 to oxygen, pH, nitrate availability, and temperature (Wilderer
170 mg N l1. After the addition of piggery wastewater, et al., 1987; Oh and Silverstein, 1999).
both NHþ þ
4TOT and NH4SOL concentrations were quite stable
During the nitrification process, a partial inhibition of
during the anoxic phase. During the aerobic phase, both the NO 2 oxidisers by the free ammonia or by the nitrous
concentrations decreased simultaneously, probably due acid (Andreottola et al., 1997; Abeling and Seyfried,
to the nitrification process. At the beginning of the cycle, 1992) could induce the observed accumulation of NO 2.
the quantity of NHþ þ
4 found as soluble (18 gNH4SOL  N)
During our experiments, and more generally during the
in comparison with the added quantity as soluble treatment of piggery wastewaters, the pH (higher than
(30 gNHþ 4SOL  N) showed that a part was probably
7.5–8) led to nitrous acid concentrations lower than the
3302 F. Béline et al. / Bioresource Technology 98 (2007) 3298–3308

threshold of inhibition. On the contrary, the free ammonia Table 4


concentrations observed during the treatment of piggery Nitrogen mass balance during run 1 and 3
wastewaters, were higher than the threshold of inhibition Runs Influent (g N) Effluent (g N) N2 (g N) Mass balance (%)
(0.1–10 mg N l1) found in the literature (Anthonisen 1 46.0 18.0 25.0 93.5
et al., 1976). Temperature, used in the SHARON process, 3 18.9 4.5 14.5 100.5
could also influence nitrite accumulation (Hellinga et al.,
1998). Indeed, above 20 C, the NO 2 oxidisers growth rate was equal to 100.5%. According to these nitrogen mass bal-
is faster than the NHþ 4 oxidisers growth rate. This differ- ances, the nitrogen not taken into account was close to the
ence could partially explain the accumulation of NO 2 experimental errors (67%). These results indicate clearly
experimentally observed. DO could also limit significantly that the nitrification during the aerobic phase followed by
1
the NO 2 oxidisers below 0.5–1.0 mg O2 l (Wiesmann, the denitrification during the anoxic phase were the domi-

½NHþ   10pH able fraction was replaced by a simplified first-order


½NH3  ¼  4  ð1Þ
6344 hydrolysis process (Boursier et al., 2004),
exp þ 10pH
273 þ T • the distinction between the anoxic heterotrophic yield
and the aerobic heterotrophic yield (Gujer et al., 1999;
In addition to the study of the evolution of nitrogen in the Orhon et al., 1996; Boursier et al., 2004),
þN 2
reactor, a nitrogen mass balance (defined as N Effluent  • the limitation by NHþ
N Influent 4 of the growth of heterotrophic
100Þ was calculated for run 1 and 3 (Table 4). The N2, pro- and autotrophic biomass.
duced by denitrification during each cycle, was estimated as
the difference between the NO 
2 plus NO3 at the beginning The model distinguishes 15 components and 11 biol-
and at the end of the anoxic phase. For run 1, the feeding ogical processes for the oxidation of organic matter, the
provided 46 g of nitrogen in the reactor while 18 g were nitrification and the denitrification (Appendix A.1). The stoi-
withdrawn at the end of the cycle. Considering the calcu- chiometry of these processes is shown in Appendix A.2.
lated N2 produced by denitrification, the nitrogen mass For the heterotrophic biomass, the kinetic and stoichi-
balance was equal to 93.5%. For run 3, the mass balance ometric values are identical to the ASM1 (Henze et al.,
2000), except the aerobic and anoxic heterotrophic yield
Table 3 and the maximum specific hydrolysis rate (Boursier et al.,
Comparison between the average NO2 oxidation rate (r1), NH3 and 2004). Concerning the autotrophic biomass, the kinetic
DO concentrations during the first part of the aerobic phase (0.2
and stoichiometric values suggested by Hao et al.
days 6 Time 6 0.6 days)
(2002a,b) were used. All values are reported in Appendix
Runs r1 (mg N l1 h1) NH3 (mg N l1) DO (mg O2 l1)
A.3. Several kinetic parameters depend on the temperature.
1 7.5 35.6 0.14 The basic Arrhenius kinetic equation for the dependency
2 16.8 53.4 0.28
on temperature, implemented in the model, is as follows:
3304 F. Béline et al. / Bioresource Technology 98 (2007) 3298–3308

Fig. 5. Experimental and simulated nitrogen concentrations during the cycles of run 1, 2 and 3.

anoxic heterotrophic yield of 0.53 was used instead of aero- forms of nitrogen (NO 
2 and NO3 ) was not correctly simu-
bic heterotrophic yield of 0.6, leading to a lower need of lated during the anoxic phase of run 3 (Fig. 5). For a better
COD for denitrification. Even if this difference seems low, modelling result, a distinction between the heterotrophic
it is not negligible and particularly important with nitrogen biomass reducing NO 
3 to NO2 and the heterotrophic

concentrated effluent and sequential batch reactors. More- biomass reducing NO2 to N2 would probably be necessary
over, the difference of oxygen half-saturation constant for (Huang and Hao, 1996). The influence of organic matter
ammonium and nitrite oxidisers previously observed (Wies- limitation on nitrite accumulation during denitrification (Oh
mann, 1994) was confirmed during this work. and Silverstein, 1999) could also be considered in order to
However, some model limitations were observed during improve the simulations. Finally, during this work, the
this work and would require more research. During this developed model was calibrated and validated using only
study, the simulations were performed using a constant two piggery wastewaters with different compositions. Addi-
temperature while the temperature varied in the experimen- tional works using others piggery wastewaters and longer
tal reactor. The variation of temperature, mainly during run HRT would allow to improve the model.
F. Béline et al. / Bioresource Technology 98 (2007) 3298–3308 3305

4. Conclusions waters. It could be used also for other concentrated waste-


waters after calibration. Taking into account nitrite as
In this study, a mathematical model was developed and intermediate of the nitrification and the denitrificiation
was calibrated for the simulation of the modified SBR, could be an interesting tool allowing the determination of

treating piggery wastewaters. This model based on conditions for partial nitrification of NHþ4 to NO2 directly
ASM1, allowed to simulate nitrite as intermediate of nitri- followed by denitrification. For this, more research is
fication and also of denitrification. The results obtained needed to include and model all effects that can potentially
during this study identified the DO concentration, rather influence NO 2 accumulation.
than NH3, as the main factor influencing the NO 2 oxida-
tion rate, and leading to accumulation of NO 2 during 5. Web references
nitrification. The experimental data coming from three
runs carried out with two different piggery wastewaters ASModel1.0: http:/scilabsoft.inria.fr/contribution/display-
were used for the calibration and the validation of the Category.php?category=SCICOS or http://beline. homeli-
model. Finally, a set of model parameters was determined, nux.net/fabrice/rubrique.php3?id_rubrique=15
and allowed to describe the behaviour of the nitrogen
during the biological treatment of piggery wastewaters. Acknowledgements
However, the results seem to indicate a variation between
piggery wastewaters of the maximum specific hydro- This study was financially supported by the French
lysis rate, which ranged between 6 and 25 g COD national program ‘‘Porcherie Verte’’. We thank professor
(g COD day)1. P. Lessard (University Laval, Québec) and Dr. A. Tré-
This model could be used for improving the design and mier (Cemagref) for the help and the comments on this
operation of the biological treatment of piggery waste- paper.

Appendix A

A.1. Process kinetics for organic matter oxidation, nitrification and denitrification

j Process rate Definition


   
SS SO S NH Aerobic growth of heterotrophs
q1 lH X BH
SS þ K S S O þ K OH S NH þ K NH
     
SS K OH S NH S NOI S NOI
q2 lH gG X BH Aerobic growth of heterotrophs
SS þ K S S O þ K OH S NH þ K NH S NOI þ K NOI S NOI þ S NOA
with SNOI
     
SS K OH S NH S NOA S NOA
q3 lH gG X BH Aerobic growth of heterotrophs
SS þ K S S O þ K OH S NH þ K NH S NOA þ K NOA S NOI þ S NOA
with SNOA
   
S NH SO S NH
q4 lAI X BAI Aerobic growth of NHþ
4
S NH þ K NAI S O þ K OAI S NH þ K NH
oxidisers
   
S NOI SO S NH
q5 lAA X BAA Aerobic growth of NO
2
S NOI þ K NAA S O þ K OAA S NH þ K NH
oxidisers

q6 bH Æ XBH Decay of heterotrophs


q7 bAI Æ XBAI Decay of NHþ 4 oxidisers
q8 bAA Æ XBAA Decay of NO 2 oxidisers
q9 KA Æ SND Æ XBH Ammonification of soluble
      organic nitrogen
SO K OH S NOI S NOI
q10 K H þ gH Hydrolysis of entrapped
S O þ K OH S O þ K OH S NOI þ K NOI S NOI þ S NOA
   i organics
K OH S NOA S NOA
þ gH XS
S O þ K OH S NOA þ K NOA S NOI þ S NOA
 
X ND
q11 q10 Hydrolysis of entrapped
Xs
organic nitrogen
3306 F. Béline et al. / Bioresource Technology 98 (2007) 3298–3308

A.2. Process stoichiometry for organic matter oxidation, nitrification and denitrification

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
SI SS XI XS XBHI XBAI XBAA XP SO SNOI SNOA SNH SND XND XNI
1 ð1Y H Þ
q1 YH 1 YH  iBN

1 1Y HD
q2 Y HD 1  1:71Y HD
 iBN

1 1Y HD 1Y HD
q3 Y HD 1  1:14Y HD
 1:14Y HD
iBN

q4 1  3:43Y
Y AI
AI 1
Y AI iBN  Y1AI

q5 1  1:14Y
Y AA
AA
 Y 1AA 1
Y AA iBN
q6 1  fP 1 fP iBN  fP fP Æ iBN
Æ iBN
q7 1  fP 1 fP iBN  fP fP Æ iBN
Æ iBN
q8 1  fP 1 fP iBN  fP fP Æ iBN
Æ iBN
q9 1 1
q10 1 1
q11 1 1

A.3. Default values at 20 C for the model and parameters used for simulating run 1, 2 and 3 (-: default value was used)

Symbol Definition Default Run 1 Run 2 Run 3 Ref.


value
YH Aerobic yield heterotrophic 0.6 – – – Boursier et al. (2004)
biomass (g COD g COD1)
YHD Anoxic yield heterotrophic 0.53 – – – Boursier et al. (2004)
biomass (g COD g COD1)
lH Maximum specific growth rate for 6 – – – Henze et al. (2000)
heterotrophic biomass (day1)
bH Decay coefficient for heterotrophic 0.62 – – – Henze et al. (2000)
biomass (day1)
KS SS half-saturation coefficient for 20 – – – Henze et al. (2000)
heterotrophic biomass (g COD m3)
KOH SO half-saturation coefficient for XBH 0.2 0.05 0.05 0.05 Henze et al. (2000)
(g COD m3)
KNH SNH half-saturation coefficient 0.05 – – – Henze et al. (2000)
for biomass (g N m3)
KNOI SNOI half-saturation coefficient for XBH 0.5 – – – Henze et al. (2000)
under anoxic conditions (g N m3)
KNOA SNOA half-saturation coefficient for 0.5 – – – Henze et al. (2000)
under anoxic conditions (g N m3)
gG Correction factor for lH under anoxic 0.8 – – – Henze et al. (2000)
condition (dimensionless)
gH Correction factor for hydrolysis under 0.4 – – – Henze et al. (2000)
anoxic condition (dimensionless)
KH Maximum specific hydrolysis rate 3 25 25 6 Boursier et al. (2004)
(g COD (g COD day)1)
KA Ammonification rate 0.08 – – – Henze et al. (2000)
(m3 COD (g day)1)
(continued on next page)
3308 F. Béline et al. / Bioresource Technology 98 (2007) 3298–3308

treatment using sequencing batch reactors. Water Res. 38, 3340– Peu, P., Béline, F., Martinez, J., 2004. Volatile fatty acids analysis from
3348. pig slurry using high performance liquid chromatography. Int. J.
Lee, S.M., Jung, J.Y., Chung, Y.C., 2000. Measurement of ammonia Environ. An. Ch. 84 (13), 1017–1022.
inhibition of microbial activity in biological wastewater treatment Sollfrank, U., Gujer, W., 1991. Characterisation of domestic wastewater
process using dehydrogenase assay. Biotechnol. Lett. 22, 991–994. for mathematical modelling of the activated sludge process. Water Sci.
Nielsen, P.H., 1996. Adsorption of ammonium to activated sludge. Water Technol. 23, 1057–1066.
Res. 30 (3), 762–764. Villaverde, S., Fdz-Polanco, F., Garcia, P.A., 2000. Nitrifying biofilm
Oh, J., Silverstein, J., 1999. Acetate limitation and nitrite accumulation acclimatation to free ammonia in submerged biofilters: start-up
during denitrification. J. Environ. Eng. ASCE 125 (3), 234–242. influence. Water Res. 34, 602–610.
Orhon, D., 1998. Evaluation of industrial biological treatment design on Wiesmann, U., 1994. Biological nitrogen removal from wastewater.
the basis of process modelling. Water Sci. Technol. 38 (4–5), 1–8. Adv. Biochem. Eng. Biot. 51, 113–154.
Orhon, D., Okutman, D., 2003. Respirometric assessment of residual organic Wilderer, P.A., Jones, W.L., Dau, U., 1987. Competition in denitrification
matter for domestic sewage. Enzyme Microb. Tech. 32 (5), 560–566. systems affecting reduction rate and accumulation of nitrite. Water
Orhon, D., Sozen, S., Artan, N., 1996. The effect of heterotrophic yield on Res. 21, 239–245.
the assessment of the correction factor for anoxic growth. Water Sci.
Technol. 34 (5–6), 67–74.
166 U. ABELING and C. F. SEYFRIED

removed anaerobically, hereby producing methane gas. Only the pollution finally remaining has to be
eliminated aerobically by using certain aeration energy.

Minimizin� the Eneuy Consumption

The combined anaerobic-aerobic technology is the most economical solution for as long as only the removal
of carbon compounds is concerned. With regard to the elimination of nitrogen and phosphorus, however, this
process is problematic, as no nitrogen and phosphorus removal worth mentioning takes place during the
elimination of the greater part of the carbon compounds in the anaerobic pretreatment. The BODIN ratio
decreased from 9:1 to nearly 1:1, the BOD:P ratio from 50:1 to 5:1. It is no problem to nitrify the
ammonium in the following aerobic treatment stage, but denitrification and biological phosphorus removal
are often incomplete, possibly due to the low content of organic carbon compounds. Phosphorus may be

eliminated by precipitation with lime or metal ions, whereas the chemicaVphysical methods for ammonium
elimination cause many problems with the resulting products. A total nitrification, however, as well as a
denitrification as far as possible and an effluent concentration of 1 or 2mg PIl, is prescribed in Germany for
most industrial wastewaters (alcohol and food production (pectin, potato, starch and sugar processing
industry, slaughter-houses etc.), pulp and paper industry, animal waste).

In order to achieve a total denitrification sometimes external carbon sources (methanol, acetic acid) are used
(very expensive). In practice, anaerobic pretreatment is often evaded by means of a bypass, but this causes a
lower gas production and also sometimes problems concerning the sludge composition (bulking) in the
aerobic stage.

Our investigations aimed to realize a process which enables total denitrification in spite of far-reaching
anaerobic pre-elimination. In contrast to the process which has been used so far, nitrification/denitrification
was carried out with assistance of the intermediate product nitrate. The carbon consumption amounts to only
60% in comparison with denitrification via nitrate. The essential parameter for regulating the process is the
concentration of free ammonia in the reactor (controlled by means of continuous NH4 and pH measuring).
Concentrations of 1 to 5mg NHyt inhibit the nitratation but not the nitritation.

The experiments were carried out in half-technical pilot plants for a period of three years, using the
wastewater of a potato starch factory. At the same time detailed examinations were made in laboratories
with regard to the kinetics and to the inhibition of nitrification/denitrification by NH3 and HN02. The new
process is available for all wastewaters with high ammonium concentrations and pH > 7.5, which often
cause problems concerning the inhibition of nitrogen removal dependent on high nitrite or ammonium
concentrations.

TREATMENT OF POTATO-STARCH WASTEWATER

Wastewater Production

During potato starch production, there are three partial wastewater streams with different concentrations of
COD, TKN and P. Figure 1 shows a flow diagram of potato starch production. The concentrations and
specific wastewater quantities are listed in Table 1.
168 U. ABELING and C. F. SEYFRIED

TABLE 2. Specific volumes and Concentrations - Potato-Starch Factoe' Llichow

m3/t potatoes m3/h concentration [mg/l]


COD TKN tot. P .

washing-water 0.43 25 2,000 150 20


processwater 0.64 38 4,000 150 20
potato juice 0.86 50 30,000 1,700 450

I 1.93 113

Treatment Process

The biological treatment of process and washing waters poses no considerable problems. The treatment is
for example carried out in wastewater ponds or activated sludge plants (inside the factory or together with
domestic sewages). Because of the nematodes contained in the washing water (parasites of the potatoes),
agricultural utilization of sewage sludge which had not been subject to anaerobic treatment is only possible
on areas which are not used for potato cultivation.

The high contents of organic carbons and the nutrients nitrogen and phosphorus in the potato juice usually
require multi-stage treatment processes. The application by rain gun of the potato which used to happen in
the past, is nowadays possible within certain limits for reasons of ground water protection. The largest
wastewater quantities are produced in the winter i.e. in the vegetationless period, and cannot be absorbed by
the plants. To reach the necessary input values, an aerobic biological treatment stage occurs. The following
processes are suitable for pretreatment:

- Membrane fIltration
- Vaporization
- Anaerobic treatment

Upgrading of the potato juice is very energy-consuming and is economical only if the final waste disposal
product is commercial in the long term. Direct utilization of the vaporization concentrate as cattle food fails
for example because of the high salt contents. The precipitation of phosphorus and ammonium (e.g.
magnesium-ammonium-phosphate) is only possible if first all organic bound nitrogen and phosphorus are
transferred to the mineral form. In these respects, the anaerobic treatment gains more and more importance.
On the one hand, the mineralization of the nutrients takes place in the anaerobic process, on the other hand
the anaerobic technique gives the possibility to reduce the COD concentration by extraction of biogas as far
as to allow an economical aerobic post-treatment. Examples of such combined anerobic-aerobic processes
are described in Nanninga and Gottschal (1986); Seyfried and Austermann-Haun (1990) and Austermann­
Haun (1990).

For the wastewater treatment of the potato starch factory built in 1987 in Llichow, an anaerobic plant has
been constructed (fixed bed circulating reactors, system KWU). The pretreated wastewater is being treated
in the new-built municipal treatment plant with biological nitrogen and phosphorus elimination to reach
recipient qUality. Figure 2 shows the process scheme of the anaerobic plant.

The washing water is led directly to the aerobic stage, while the coagulated potato juice and the process
water are subject to anaerobic pretreatment. Until now, the planned degrading percentage (nCOD 70-80%)
=

could not yet be reached, considering the space load of 20-30kg COD/m3.d. The efficiency is between 50%
and 60%. The solid matter separation (coagulated protein) after acidification causes problems. By
application of the second lamella separator, a certain improvement was achieved. Furthermore, the plant has
been extended by a buffer tank (l000m3). For the purposes of capacity extension, in 1992 the construction of
an additional anaerobic reactor (V 1800m3, fixed fIlm with plastic media) was started. This reactor was
=
170 U. ABELING and C. F. SEYFRIED

TABLE 3. Operatinl: Results (Ayeral:inl: values 1988-1990.

INFLUENT AEROBIC STAGE EFFLUENT


anaerobic stage wash-water and
influent effluent domestic wastew.
mg/l mg/l mg/l mg/l

COD 16,000 - 18,000 5,000 - 10,000 1,400 80 - 140


BODS 10,700 - 12,000 2,000 - 5,000 700 5 - 15
TKN 900 - 1,200 800 - 1,100 100 0 - 2
N0 2-N <
N03-N 5 - 20
��ot 180 -
1,000 -
240
2,000
170 -
800 -
230
1,500
30
500
0.5 - 10

Q m3/d 2,100 2,100 2,100 4,200

The four aerated tanks are loaded simultaneously via a sludge contact tank. The aeration of the tanks occurs
according to an interval process (3 h nitrification, 6 h denitrification), so that an intermittent denitrification
takes place.

In spite of the high loads from the anaerobic plant, until now a complete nitrification as well as large
reduction of the carbon compounds could be achieved. For the denitrification and the biological phosphorus
elimination (the sludge contact serves as anaerobic basin) the bad elimination rate of the anaerobic stage has
a positive effect, especially by the high content of organic acids.

As the insufficient anaerobic elimination rate must be improved for economical reasons (high energy
consumption and increased surplus sludge production in the aerobic stage), optimization of the
denitrification potential was indispensable. Thus, the process of nitrogen elimination by means of nitrate was
developed.

NITROGEN REMOVAL VIA NITRITE

Nitrification happens in two partial steps, i.e. nitritation (oxidation from ammonium to nitrite by
Nitrosomonas a.o.) and nitratation (oxidation from nitrite to nitrate by Nictobacter a.o.). Both organism
groups are aerobic and need CO2 as carbon source. They are characterized by low growth rates, high
sensibility to pH values and temperature deviations as well as toxic substances. Furthermore, there are
inhibitions because of the substrates ammonium and nitrite. In this respect, the substrate inhibition with
Haldane kinetics can be clearly described, where the non-dissociate forms, i.e. ammonia and nitric acid,
serves as substrates.

In practice, the product inhibition of Nitrosomonas is very important. It seems that inhibition of
Nitrosomonas by nitric acid can be described as non-competitive inhibition. Especially upon start-up of a
plant or appearance of load peaks, there are higher nitrite concentrations. A certain percentage comes out as
nitric acid, depending on temperature and pH value (Figure 4). Nitric acid then causes inhibition of the
Nitrosomonas, so that concentration of ammonium, respectively ammonia, increases. On the one hand,
ammonia can release a substrate inhibition of Nitrosomonas, on the other hand, it can start a non-competitve
inhibiton of Nitrobacter. In bad conditions, there may be a complete breakdown of nitrification.
3300 F. Béline et al. / Bioresource Technology 98 (2007) 3298–3308

known amount of piggery wastewater was added in the the reactor reached steady-state conditions (after more
reactor through a buffer tank where the quantity added than three hydraulic residence times), the characterisation
was weighed and recorded. The reactor was equipped with of the effluent was done (COD, soluble COD and total
a foam breaker working during the aerobic phases and a nitrogen) and was followed by a study of the evolution
fine bubbles diffuser fed with a compressor (2–5 m3 h1). of NHþ  
4 , NO2 and NO3 in the reactor, during one
For the mixing, a centrifugal pump (30 m3 h1) circulated cycle.
the activated sludge, during the anoxic phases from the Immediately after run 1, a second study of the evolution
 
bottom to the top of the reactor. Water was continuously of NHþ 4 , NO2 and NO3 , during one cycle, was performed
added in the reactor (65 l per day) to compensate for the (run 2). At the beginning of this cycle, 1 l of ammonium
evaporation caused by the high temperature and the aera- carbonate solution (13.5 kg N m3) was added in the reac-
tion. At the end of each cycle, an amount of the activated tor during the feeding phase, together with the piggery
sludge equal to the piggery wastewater added at the begin- wastewater (PW1). For this experiment, the other experi-
ning of the cycle was withdrawn from the reactor. The bio- mental conditions were similar to run 1.
logical reactor was equipped with on-line measurements of For run 3, as for run 1, the reactor was initially filled
DO (WTW Oxi197 oxygen meter), temperature (Ponselle with 50 l of activated sludge coming from the same farm
PONAPF-TTRANS), oxidation–reduction potential (Pon- and 50 l of tap water. About 600 l of raw piggery waste-
selle PONAPF-EHTRANS) and pH (Ponselle PONAPF- water was collected in an experimental farm near Rennes
PHTRANS) allowing measurements at 10 min. intervals. (France) and was mechanically sieved (1 mm) using a
The system operated without sludge recirculation nor rotary screen. The liquid fraction called piggery wastewater
decantation in the reactor, leading to a solids retention 2 (PW2) was stored at 4 C and was used for run 3. For this
time equal to the hydraulic retention time. For each exper- run, the duration of a cycle was equal to 12 h, consisting of
iment, each whole-cycle of the modified SBR was scheduled 7 h of anoxic conditions and 5 h of aerobic conditions. At
as follows: the beginning of each cycle, 4.4 l of PW2 was added in the
reactor, leading to a working volume of 104.4 l and a HRT
• a feeding phase, equal to 11.9 days. At the end of each cycle, the same
• an anoxic phase, amount of activated sludge was withdrawn. A characterisa-
• an aerobic phase, tion of the effluent (COD, soluble COD and total nitrogen)
 
• a withdrawal phase. and a study of the evolution of NHþ 4 , NO2 and NO3 dur-
ing a cycle, were performed when more than 3 hydraulic
residence times had passed.
2.2. Experimental The main characteristics of each piggery wastewater
used in this study are presented in Table 2.
During this study, three experiments were carried out
using the pilot-scale modified SBR (Table 1).
Table 2
For run 1, the reactor was filled with 50 l of activated Characteristics of the piggery wastewaters
sludge coming from a farm equipped with a modified
Variable PW1 PW2
SBR at the beginning of the experiment. The reactor
3
was supplemented with 50 l of tap water. At the same TS, kg m 54.0 (0.4) 31.1 (–)
VS, kg m3 42.5 (1.0) 20.1 (–)
time, about 600 l of raw piggery wastewaters were col-
COD, kg O2 m3 50.6 (0.3) 39.4 (1.0)
lected and were manually sieved (5 mm mesh). The liquid CODs, kg O2 m3 21.8 (0.1) 14.8 (0.4)
fraction called piggery wastewater 1 (PW1) was stored VFA, kg O2 m3 10.7 (0.2) 6.2 (–)
3
at 4 C and was used for this experiment. For this run, NHþ 4  N (or SNH), kg N m 3.0 (0.1) 3.4 (0.1)

the duration of a whole-cycle was equal to 24 h consisting NO2  N (or SNOI), kg N m3 0 0
NO 3  N (or SNOA), kg N m
3
0 0
of 4 h of anoxic conditions and 20 h of aerobic conditions.
TN–N, kg N m3 4.6 (0.1) 4.3 (0.05)
At the beginning of each cycle, 10 l of PW1 were added SS, kg O2 m3 5.2 (0.5) 6.0 (1.0)
in the reactor, leading to a working volume of 110 l and XS, kg O2 m3 14.1 (3.6) 23.9 (1.4)
a HRT equal to 11 days. At the end of each cycle the Inert COD, kg O2 m3 31.3 (3.4) 9.5 (1.4)
same amount of activated sludge was withdrawn. When Standard errors shown in parenthesis, n = 5.

Table 1
Summary of the conditions during the three experiments carried out with the pilot-scale modified SBR
Runs PW HRT (days) P1 (h) P2 (h) P3 (h) P4 (h) Observations
1 1 11.0 0.1 3.9 19.9 0.1 Effluent characterisation and study of a cycle
2 1 11.0 0.1 3.9 19.9 0.1 Study of a cycle (ammonium carbonate was added)
3 2 11.9 0.1 6.9 4.9 0.1 Effluent characterisation and study of a cycle
PW: piggery wastewater, P1: feeding phase, P2: anoxic phase, P3: aerobic phase, P4: withdrawal phase.
F. Béline et al. / Bioresource Technology 98 (2007) 3298–3308 3301

2.3. Analytical methods

Total solids (TS), volatile solids (VS), COD and total


nitrogen (TN) were analysed according to standard
methods (APHA, 1992). The soluble COD (CODs) was
analysed after centrifugation at 4000 rpm during 15 min.
Volatile fatty acids (VFA) were analysed using HPLC
(Peu et al., 2004). NHþ
4 was analysed by steam distillation
(Bremner and Keeney, 1965) directly on the effluent
(NHþ 4TOT ) and on the liquid fraction after centrifugation
 
at 4000 rpm during 15 min (NHþ 4SOL ). NO2 and NO3 were
analysed by ion chromatography on a Dionex DX120
system equipped with a Dionex AS9HC column. Eluant
was a 10 mM Na2CO3 solution, flowing at 1 ml min1.
Respirometry, widely recognised as the most convenient
tool to determine the biodegradable wastewater fractions Fig. 2. Nitrogen species concentrations during the run 1 cycle (F: feeding,
(Sollfrank and Gujer, 1991), was used for the COD frac- W: withdrawal).
tionation during this study. The oxygen uptake rates
(OUR) were measured for each wastewater and allowed
to determine the readily biodegradable COD (SS) and the adsorbed to the activated sludge flocs, as previously
slowly biodegradable COD (XS) as described by Orhon observed by Nielsen (1996). However, NHþ 4 was available
and Okutman (2003); Boursier et al. (2005). The inert almost entirely for nitrification and the adsorption did not
COD fraction was calculated as the difference between significantly limit the nitrification process. Therefore, the
total COD and biodegradable COD (SS + XS). process of adsorption of NHþ 4 on the flocs was neglected.
In the remainder of this paper, the nitrifiable ammonium
2.4. Modelling (NHþ þ þ
4NIT ), rather than NH4SOL or NH4TOT , was considered.
þ
NH4TOT found at the end of the nitrification was considered
For the modelling, the mathematical equations were as the NHþ þ
4 not available for the nitrification (NH4NO–AV ).
implemented in the Scilab software using the Scicos NH4NIT was calculated as the difference between NHþ
þ
4TOT
interface. This software is available as a toolbox called and NHþ 4NO–AV (Nielsen, 1996). For all runs, the NH4NIT
þ

AS-Model1.0 (see Web references). The integration routine corresponded to the amount added at the beginning of
was chosen automatically by the software. A run time of each cycle. The NHþ 4NO–AV was probably organic nitrogen
more than three times the HRT was used to establish the considered as NHþ 4 because of an analytical error (steam
pseudosteady-state conditions. For the calibration of the distillation).
model, the parameters were estimated manually by com- As shown in Fig. 2, the NO 3 concentration decreased
paring the simulated output with the experimental data quickly from 210 to 0 mg N l1 during the anoxic phase.
1
until the best fit was achieved. Concurrently, an accumulation of NO 2 up to 115 mg N l
occured. During the aerobic phase, an increase of NO 3
3. Results and discussion from 0 to 220 mg N l1 was obtained and an accumulation
1
of NO 2 up to 110 mg N l was also observed. The same
3.1. Experimental results evolution of nitrogen, particularly the accumulation of
NO 2 during the nitrification and the denitrification, was
The evolution of the nitrogen species during a cycle was observed for all runs and was previously described by
used for a better characterisation of processes involved in several authors (Andreottola et al., 1997; Kim et al.,
the nitrogen transformations. As an example, the evolution 2004; Huang and Hao, 1996).
of the nitrogen forms during run 1 is presented in Fig. 2. Concerning the denitrification process, various environ-
At the beginning of the cycle, the addition of piggery mental factors were found to be responsible for nitrite
wastewater in the reactor led to an increase of NHþ4TOT from
accumulation: the type and quantity of organic substrate,
163 to 465 mg N l1 and an increase of NHþ 4SOL from 0 to oxygen, pH, nitrate availability, and temperature (Wilderer
170 mg N l1. After the addition of piggery wastewater, et al., 1987; Oh and Silverstein, 1999).
both NHþ þ
4TOT and NH4SOL concentrations were quite stable
During the nitrification process, a partial inhibition of
during the anoxic phase. During the aerobic phase, both the NO 2 oxidisers by the free ammonia or by the nitrous
concentrations decreased simultaneously, probably due acid (Andreottola et al., 1997; Abeling and Seyfried,
to the nitrification process. At the beginning of the cycle, 1992) could induce the observed accumulation of NO 2.
the quantity of NHþ þ
4 found as soluble (18 gNH4SOL  N)
During our experiments, and more generally during the
in comparison with the added quantity as soluble treatment of piggery wastewaters, the pH (higher than
(30 gNHþ 4SOL  N) showed that a part was probably
7.5–8) led to nitrous acid concentrations lower than the
3302 F. Béline et al. / Bioresource Technology 98 (2007) 3298–3308

threshold of inhibition. On the contrary, the free ammonia Table 4


concentrations observed during the treatment of piggery Nitrogen mass balance during run 1 and 3
wastewaters, were higher than the threshold of inhibition Runs Influent (g N) Effluent (g N) N2 (g N) Mass balance (%)
(0.1–10 mg N l1) found in the literature (Anthonisen 1 46.0 18.0 25.0 93.5
et al., 1976). Temperature, used in the SHARON process, 3 18.9 4.5 14.5 100.5
could also influence nitrite accumulation (Hellinga et al.,
1998). Indeed, above 20 C, the NO 2 oxidisers growth rate was equal to 100.5%. According to these nitrogen mass bal-
is faster than the NHþ 4 oxidisers growth rate. This differ- ances, the nitrogen not taken into account was close to the
ence could partially explain the accumulation of NO 2 experimental errors (67%). These results indicate clearly
experimentally observed. DO could also limit significantly that the nitrification during the aerobic phase followed by
1
the NO 2 oxidisers below 0.5–1.0 mg O2 l (Wiesmann, the denitrification during the anoxic phase were the domi-
1994; Bernet et al., 2001). nant processes leading to nitrogen removal, as previously
During the experiments, the average NH3 concentra- observed (Choi and Eum, 2002). The contribution of other
tions, during the first part of the aerobic phase (0.2 days 6 processes, such as ammonia volatilisation or the simulta-
Time 6 0.6 days), were estimated for runs 1 and 2, using neous nitrification–denitrification during the aerobic phase,
Eq. (1) (Lee et al., 2000). The results are shown in the Table were low and could probably be neglected.
3. At the same time, the average NO 2 oxidation rate and
the average DO concentrations were calculated for the 3.2. Model development
same period and are also presented in Table 3. During both
runs, the pH and temperature were similar but the calcu- According to previous works concerning the modelling
lated NH3 concentration was higher for run 2 due to the of biological wastewater treatment processes (Henze
ammonium carbonate addition. Although this concentra- et al., 2000; Boursier et al., 2004) and characterisation of
tion was higher for run 2 than for run 1, the NO 2 oxidation piggery wastewaters (Boursier et al., 2005) and taking into
rate, experimentally observed with the same activated account the results described in the previous section, a
sludge and for the same temperature, was also higher for model was developed. The main structural differences with
run 2 than for run 1. Such a result seems to indicate that the well known activated sludge model ASM1, edited by
the NO 2 oxidisers were not inhibited by the NH3. In fact, the IWA task group, are:
previous studies demonstrated that the oxidation of NO 2
could occur even with high concentrations of NH3, partic- • nitrite as intermediate of nitrification and also of denitri-
ularly when the bacteria are acclimated (Villaverde et al., fication, in order to simulate the experimentally observed
2000), like during the piggery wastewaters treatment. On accumulation of nitrite. For denitrification, only one het-
the contrary, the higher concentration of oxygen (DO), erotrophic biomass was considered and was responsible
observed during run 2, could partially explain the increase for both steps. For nitrification, two kinds of autotrophic
of the oxidation rate. biomass were considered,
• the surface-limited hydrolysis of the slowly biodegrad-
½NHþ   10pH able fraction was replaced by a simplified first-order
½NH3  ¼  4  ð1Þ
6344 hydrolysis process (Boursier et al., 2004),
exp þ 10pH
273 þ T • the distinction between the anoxic heterotrophic yield
and the aerobic heterotrophic yield (Gujer et al., 1999;
In addition to the study of the evolution of nitrogen in the Orhon et al., 1996; Boursier et al., 2004),
þN 2
reactor, a nitrogen mass balance (defined as N Effluent  • the limitation by NHþ
N Influent 4 of the growth of heterotrophic
100Þ was calculated for run 1 and 3 (Table 4). The N2, pro- and autotrophic biomass.
duced by denitrification during each cycle, was estimated as
the difference between the NO 
2 plus NO3 at the beginning The model distinguishes 15 components and 11 biol-
and at the end of the anoxic phase. For run 1, the feeding ogical processes for the oxidation of organic matter, the
provided 46 g of nitrogen in the reactor while 18 g were nitrification and the denitrification (Appendix A.1). The stoi-
withdrawn at the end of the cycle. Considering the calcu- chiometry of these processes is shown in Appendix A.2.
lated N2 produced by denitrification, the nitrogen mass For the heterotrophic biomass, the kinetic and stoichi-
balance was equal to 93.5%. For run 3, the mass balance ometric values are identical to the ASM1 (Henze et al.,
2000), except the aerobic and anoxic heterotrophic yield
Table 3 and the maximum specific hydrolysis rate (Boursier et al.,
Comparison between the average NO2 oxidation rate (r1), NH3 and 2004). Concerning the autotrophic biomass, the kinetic
DO concentrations during the first part of the aerobic phase (0.2
and stoichiometric values suggested by Hao et al.
days 6 Time 6 0.6 days)
(2002a,b) were used. All values are reported in Appendix
Runs r1 (mg N l1 h1) NH3 (mg N l1) DO (mg O2 l1)
A.3. Several kinetic parameters depend on the temperature.
1 7.5 35.6 0.14 The basic Arrhenius kinetic equation for the dependency
2 16.8 53.4 0.28
on temperature, implemented in the model, is as follows:
F. Béline et al. / Bioresource Technology 98 (2007) 3298–3308 3307

Appendix (continued)
Symbol Definition Default Run 1 Run 2 Run 3 Ref.
value
fp Fraction of biomass leading to inert 0.08 – – – Henze et al. (2000)
particulate products (dimensionless)
iBN Mass of nitrogen per mass of 0.086 – – – Henze et al. (2000)
COD (g N g COD1)
YAI Yield for NHþ 4 oxidisers autotrophic 0.15 – – – Hao et al. (2002b);
biomass (g COD g N1) Hao et al. (2002a)
lAI Maximum specific growth rate for NHþ 4 0.8 – – – Hao et al. (2002b);
oxidisers autotrophic biomass (day1) Hao et al. (2002a)
bAI Decay coefficient for XBAI autotrophic 0.05 – – – Hao et al. (2002b);
biomass (day1) Hao et al. (2002a)
KOAI SO half-saturation coefficient for XBAI 0.6 0.3 0.3 0.3 Hao et al. (2002b);
biomass (g COD m3) Hao et al. (2002a)
KNAI SNH half-saturation coefficient for XBAI 2.4 5 5 5 Hao et al. (2002b);
biomass (g N m3) Hao et al. (2002a)
YAA Yield for NO 2 oxidisers autotrophic 0.041 – – – Hao et al. (2002b);
biomass (g COD g N1) Hao et al. (2002a)
lAA Maximum specific growth rate for NO 2 0.79 – – – Hao et al. (2002b);
oxidisers autotrophic biomass (day1) Hao et al. (2002a)
bAA Decay coefficient for XBAA autotrophic 0.033 – – – Hao et al. (2002b);
biomass (day1) Hao et al. (2002a)
KOAA SO half-saturation coefficient for XBAA 2.2 1.1 1.1 1.1 Hao et al. (2002b);
biomass (g COD m3) Hao et al. (2002a)
KNAA SNOI half-saturation coefficient for XBAA 5 2.5 2.5 2.5 Hao et al. (2002b);
biomass (g N m3) Hao et al. (2002a)

References Bremner, J.M., Keeney, D.R., 1965. Steam distillation methods for
determination of ammonium, nitrate and nitrite. Anal. Chim. Acta
Abeling, U., Seyfried, C.F., 1992. Anaerobic–aerobic treatment of high- 32, 485–495.
strength ammonium waste-water nitrogen removal via nitrite. Water Choi, E., Eum, Y., 2002. Strategy for nitrogen removal from piggery
Sci. Technol. 26 (5–6), 1007–1015. waste. Water Sci. Technol. 46 (6–7), 347–354.
Andreottola, G., Bortone, G., Tilche, A., 1997. Experimental validation of Coelho, M.A.Z., Russo, C., Araujo, O.Q.F., 2000. Optimization of a
a simulation and design model for nitrogen removal in sequencing sequencing batch reactor for biological nitrogen removal. Water Res.
batch reactors. Water Sci. Technol. 35 (1), 113–120. 34, 2809–2817.
Anthonisen, A.C., Loehr, R.C., Prakasam, T.B.S., Srinath, E.G., 1976. Gujer, W., Henze, M., Mino, T., van Loosdrecht, M.C.M., 1999.
Inhibition of nitrification by ammonia and nitrous acid. J. Water Activated sludge model N3. Water Sci. Technol. 39 (1), 183–193.
Pollut. Con. F. 48 (5), 835–852. Hao, X., Heijnen, J.J., Van Loosdrecht, M.C.M., 2002a. Model-based
APHA, 1992. Standard methods for the examination of water and evaluation of temperature and inflow variations on a partial nitrifica-
wastewater. 18th ed., American Public Health Association, Washing- tion-anammox biofilm process. Water Res. 36, 4839–4889.
ton DC, USA. Hao, X., Heijnen, J.J., van Loosdrecht, M.C.M., 2002b. Sensitivity
Béline, F., Daumer, M.L., Guiziou, F., 2004. Biological aerobic treatment analysis of a biofilm model describing a one-stage completely
of pig slurry in france: nutrients removal efficiency and separation autotrophic nitrogen removal (canon) process. Biotechnol. Bioeng.
performances. Trans. ASAE 47 (3), 857–864. 77 (3), 266–277.
Bernet, N., Dangcong, P., Delgenes, J.P., Moletta, R., 2001. Nitrification Hellinga, C., Schellen, A.A.J.C., Mulder, J.W., van Loosdrecht, M.C.M.,
at low oxygen concentration in biofilm reactor. J. Environ. Eng. ASCE Heijnen, J.J., 1998. The SHARON process: an innovative method
127 (3), 266–271. for nitrogen removal from ammonium-rich wastewater. Water Sci.
Boursier, H., 2003. Etude et modélisation des processus biologiques au Technol. 37 (9), 135–142.
cours du traitement aérobie du lisier de porcs en vue d’une optimisa- Henze, M., Grady, C.J., Gujer, W., Marais, G.V., Matsuo, T. 1987.
tion du procédé (modelling of biological processes during aerobic Activated sludge model N1. IWAPRC Scientific and Technical
treatment of pig slurry aiming at process optimisation). PhD thesis, Report N1, London.
Institut National des Sciences Appliquées de Toulouse, France. Henze, M., Gujer, W., Mino, T., van Loosdrecht, M.C.M., 2000.
Boursier, H., Béline, F., Paul, E., 2004. Activated sludge model no. 1 Activated sludge models ASM1, ASM2, AsM2d and ASM3. Tech.
calibration for piggery wastewater using respirometry. Water Sci. Rep. 9, IWA Scientific and Technical Report, London.
Technol. 49 (5–6), 389–396. Huang, J., Hao, O.J., 1996. Alternating aerobic-anoxic process for nitrogen
Boursier, H., Béline, F., Paul, E., 2005. Piggery wastewater characterisation removal: dynamic modeling. Water Environ. Res. 68 (1), 94–104.
for biological nitrogen removal process design. Bioresource Technol. Kim, J.H., Chen, M., Kishida, N., Sudo, R., 2004. Integrateds real-
96, 351–358. time control strategy for nitrogen removal in swine wastewater
3308 F. Béline et al. / Bioresource Technology 98 (2007) 3298–3308

treatment using sequencing batch reactors. Water Res. 38, 3340– Peu, P., Béline, F., Martinez, J., 2004. Volatile fatty acids analysis from
3348. pig slurry using high performance liquid chromatography. Int. J.
Lee, S.M., Jung, J.Y., Chung, Y.C., 2000. Measurement of ammonia Environ. An. Ch. 84 (13), 1017–1022.
inhibition of microbial activity in biological wastewater treatment Sollfrank, U., Gujer, W., 1991. Characterisation of domestic wastewater
process using dehydrogenase assay. Biotechnol. Lett. 22, 991–994. for mathematical modelling of the activated sludge process. Water Sci.
Nielsen, P.H., 1996. Adsorption of ammonium to activated sludge. Water Technol. 23, 1057–1066.
Res. 30 (3), 762–764. Villaverde, S., Fdz-Polanco, F., Garcia, P.A., 2000. Nitrifying biofilm
Oh, J., Silverstein, J., 1999. Acetate limitation and nitrite accumulation acclimatation to free ammonia in submerged biofilters: start-up
during denitrification. J. Environ. Eng. ASCE 125 (3), 234–242. influence. Water Res. 34, 602–610.
Orhon, D., 1998. Evaluation of industrial biological treatment design on Wiesmann, U., 1994. Biological nitrogen removal from wastewater.
the basis of process modelling. Water Sci. Technol. 38 (4–5), 1–8. Adv. Biochem. Eng. Biot. 51, 113–154.
Orhon, D., Okutman, D., 2003. Respirometric assessment of residual organic Wilderer, P.A., Jones, W.L., Dau, U., 1987. Competition in denitrification
matter for domestic sewage. Enzyme Microb. Tech. 32 (5), 560–566. systems affecting reduction rate and accumulation of nitrite. Water
Orhon, D., Sozen, S., Artan, N., 1996. The effect of heterotrophic yield on Res. 21, 239–245.
the assessment of the correction factor for anoxic growth. Water Sci.
Technol. 34 (5–6), 67–74.

You might also like