You are on page 1of 10

Nitrification and Denitrification in the Wastewater Treatment System

pp. 145-158 in Traditional Technology for Environmental Conservation and Sustainable Development
in the Asian-Pacific Region

Proceedings of the UNESCO - University of Tsukuba International Seminar on Traditional


Technology for Environmental Conservation and Sustainable Development in the Asian-Pacific
Region, held in Tsukuba Science City, Japan, 11-14 December, 1995.

Editors: Kozo Ishizuka, D. Sc. , Shigeru Hisajima, D. Sc. , Darryl R.J. Macer, Ph.D.

Copyright 1996 Masters Program in Environmental Sciences, University of Tsukuba. Commercial rights are reserved, but this book may be reproduced for limited educational
purposes. Published by the Master's Program in Environmental Science and Master's Program in Biosystem Studies, University of Tsukuba, 1996.

P. Y. Yang and Zhi-Qin Zhang


Dept. of Biosystems Engineering, University of Hawaii at Manoa, Honolulu, Hawaii 96822, USA

Abstract

Various forms of nitrogen that contribute to the deterioration of the water resources and environment was presented. Review
on the nitrification and denitrification process for the removal of nitrogen through various existing treatment alternatives
was conducted. Research studies at the University of Hawaii on the application of nitrification and denitrification processes
for nitrogen removal from dilute pig wastewater and synthetic wastewater (diluted and concentrated nitrogen concentration)
were presented. Separated and combined nitrification and denitrification processes from these studies were presented and
discussed. Application of various treatment technologies, either traditional or innovative, depends on the location, social,
and economic considerations. Also, it may vary for small and large wastewater treatment plants for the application of
nitrification and denitrification.

Key Words: Source of nitrogen; separate process; combined process; activated sludge;

fixed film, rotating bio-contact; fluidized bed; immobilized cell.

Introduction

Carbon and nitrogen are the major pollution sources that contribute to environmental quality problems. All of the pollution
sources; i.e., municipal, industrial, and agricultural, must be managed in order to reduce the carbon and nitrogen
concentration within a certain level to improve the quality of the environment. Problems which are associated with carbon
and nitrogen are: 1) imbalance of natural ecological systems and increase of eutrophication; 2) depletion of dissolved
oxygen in surface waters which kills fish and create septic conditions; 3) odor problems; 4) contaminants that complicate
water treatment, such as ammonia used for water supplies that require an increase of chlorine dosage to achieve a free
chlorine residual in the process of disinfection; and 5) increase risks to human health, such as NO3-N concentration in the
groundwater for potable use.

Nitrate is regarded as an undesirable substance in public water. Although it occurs naturally in water, elevated levels of
nitrate in groundwater usually result from human activities, such as over use of chemical fertilizers in agriculture and
improper disposal of human and animal wastes (Nugent, et al., 1988). High nitrate concentration in drinking water may
cause serious problems in humans and animals (Sawyer and McCarty, 1978; Darimont, 1988; Nugent, et al., 1988). In order
to protect against this effort, the United States Environmental Protection Agency (USEPA) has established the maximum
contamination level of nitrate in drinking water at 10 mg N03-N/L, which corresponds to the maximum allowance
recommended by the World Health Organization (WHO) and maximum acceptable limit in Canada (Sayre, 1988).

The general treatment alternatives available for the treatment of wastewater can be divided into two major categories: 1)
physical/chemical treatment systems and 2) biological treatment systems. Physical treatments include screening,
sedimentation, filtration and flotation. Chemical treatments include disinfection, adsorption, and precipitation. The major
biological processes used for wastewater treatment can be separated into five major groups: aerobic process, anoxic process,
anaerobic process, combined aerobic-anoxic-anaerobic processes, and pond processes. The principal application of the
processes are for 1) the removal of the carbonaceous organic matter in wastewater; 2) nitrification; 3) denitrification; 4)
phosphorus removal; and 5) waste stabilization. The biological processes are considered the most effective and economic
process in the field of wastewater treatment (Metcalf and Eddy, 1991).

Nitrification and Denitrification Process

Nitrification

Nitrification is the biological process by which ammonia is first converted to nitrite and then to nitrate. Nitrification can be
achieved in any aerobic-biological process at low organic loadings and where suitable environmental conditions are
provided. Nitrifying bacteria are slower growing than the heterotrophic bacteria, which comprises the greater proportion of
the biomass in both fixed film and suspended growth systems. The key requirement for nitrification to occur, therefore, is
that the process should be so controlled that the net rate of accumulation of biomass, and hence, the net rate of withdrawal of
biomass from the system, is less than the growth rate of the nitrifying bacteria (Barnes and Bliss, 1983). The processes
currently used in the treatment of wastewater for nitrification are presented as follows.

1. Trickling filters

The extent of nitrification in trickling filters depended on a variety of factors; including temperature, dissolved oxygen, pH,
presence of inhibitors, filter depth and media type, loading rate, and wastewater BOD (Parker and Richards, 1986). Low-rate
trickling filters allowed the development of a high-nitrifying population. For rock media filters, organic loading should not
exceed 0.16 kg BOD5/m3/day (USEPA, 1975). Higher loading rates (0.36 kg BOD5/m3/day) were allowable in plastic
media trickling filters because of the higher surface area of the plastic media (Stenquist, et al., 1974). If two filters were
used, heterotrophic growth occurs in the first filter and nitrification in the second filter (Bitton, 1994). Boller and Gujer
(1986) conducted a pilot plant study of tertiary trickling filters, recommending a media surface loading rate of 0.4 g NH3-
N/m2/day for complete nitrification (effluent NH3-N < 2.0 mg/L) at a water temperature of 10 C.

2. Rotating biological contractor

RBC biofilm has an initial adsorption of microorganisms to the disk surface to form 1-4 mm thick biofilm that is responsible
for BOD removal in rotating biological contractors. The rotating disks provided a large surface area for the attached
biomass. The first stages of an RBC mostly removed organic materials, whereas subsequent stages removed NH3-N as a
result of nitrification, when the BOD5 was low enough. Ammonia oxidizers could not effectively compete with the faster-
growing heterotrophs that oxidize organic matter. Nitrification occurs only when the BOD was reduced to approximately 14
mg/L, and increases with rotation speed (Weng and Molof, 1974). RBC performance was negatively affected by low
dissolved oxygen in the first stages and by low pH in the later stages where nitrification occurred (Hitdlebaugh and Miller,
1981).

An innovative operational process using recirculation in RBC was developed and used to improve nitrification (Klees and
Silverstein, 1992). They found that it could improve nitrification at all hydraulic loading rates; the positive effect of
recirculation on nitrification was due to the dilution of influent organic carbon. Degradable organic carbon inhibits
nitrification at concentrations greater than 15-20 mg/L BOD5; extremely low concentrations of influent BOD5 (less than 10
mg/L) did not improve nitrification.

Figueroa and Silverstein (1992) studied the effect of particulate organic matter on biofilm nitrification in a pilot RBC. For a
range of 12 and 82 mg/L total BOD to the pilot RBC, particulate BOD was found to inhibit nitrification to the same degree
as soluble BOD. Total influent organic matter was found to be a better predictor of nitrification than soluble organic matter
concentration. The inhibition of nitrification by particulate BOD suggested that clarified influent should be used for
nitrifying the biofilm process.

Dunn, et al. (1984), investigated the nitrification process with a biofilm fluidized reactor. The bed was designed cone-shaped
and filled with 0.2-0.3 mm diameter quartz sand. With three years of operation under a loading rate of 5.2 l/hr, recycle rate
of 84 l/hr and NH4-N concentration of 45 mg/L, the maximum nitrification rate was observed as 134 mg N/g MLSS/hr for
Nitrosomonas, and 120 mg N/MLSS/hr for Nitrobacter. Optimum pH value was found to be around 7.8; optimum
temperature was found between 30-40 C.

Tijhuis, et al. (1992), studied the development of nitrifying biofilms, as well as the short- and long-term influence on the
nitrification capacity in a Biofilm-Airlift-Suspension-Reactor. They found that the specific nitrification capacity during start-
up was constant, 1 g N/g TOC/d, which was high compared to the activated sludge process. The influence of the temperature
on the nitrification rate was much less than could be expected from the pure culture experiments, and the maximum
nitrification rate during the experiment was 6 kg N/m3/d, which was also relatively high compared to the activated sludge
process.

3. Fixed bed reactor

Jimenez, et al. (1987), conducted a pilot scale research by using a fixed bed reactor for nitrification of the effluent from an
extended aeration sewage treatment plant. The non-settled sewage influent of COD of 373 mg/L, NH4-N of 45 mg/L, SS of
297 mg/L and pH 8.1 were used in this study. With an HRT of 4 to 6 hours and recycle ratio of 3.5, removal efficiencies of
70% of COD, 67% of SS and 95% of NH4-N were obtained.

Less than 10 mg/L effluent BOD and ammonia removal efficiency of 65-85% were achieved in a four stage aerobic
submerged biofilm reactor containing ball-rings (Gonzalez-Martinez and Duqu-Luciano, 1992). Nitrification was achieved
in all 4 reactors operated at the loading rate of 6-21 g/m2/d and the HRTs of 130-290 minutes.

4. Conventional activated sludge processes at low loadings

Weismann (1994) studied the nitrification in a conventional activated sludge system and found that it was relatively low for
carbon removal and nitrification of sewage because carbon removal and nitrification occurred in the same reactor with an
activated sludge system. This resulted in a population mixture of mainly heterotrophs and few autotrophs. In this kind of
treatment system, it was not possible to enrich the autotrophic bacteria because the slower growing autotrophs were removed
with the surplus sludge. It was necessary to separate the autotrophic from the heterotrophic biomass in order to increase the
specific nitrification rate.

Suwa, et al. (1989), conducted a research on simultaneous organic carbon removal-nitrification by an activated sludge
process with cross-flow filtration. Because of the recycle of sludges with cross-flow filtration, this process made the sludge
retention time very long; simultaneous carbon removal-nitrification was achieved quite well under the loading rate of about
0.10 g BOD/g VSS/d. The efficiency of dissolved organic carbon removal was more than 95%, and nitrification was
sufficient (NH3-N was not detected in the effluent).

5. Two-stage activated sludge systems with separate carbonaceous oxidation and nitrification systems

The nitrification process requires a slow-growing nitrifying bacteria with sludge that has been aged for a long time and high
dissolved oxygen concentration. In addition, they were susceptible to inhibition by a wide range of compounds at
concentrations so low as not to affect the heterotrophic bacteria. For these reasons, it would seem sensible to separate the
processes of carbonaceous removal and nitrogen removal into separate reactors (Horan, 1989).

Rimer and Woodward (1972) investigated separate carbonaceous oxidation and nitrification systems and found that this
process minimized sludge washout with the nitrification stage, and the process could be operated successfully at a shorter
detention time, lower MLSS and solid retention time.

Denitrification
Denitrification is the biological process by which nitrate is converted to nitrogen and other gaseous end products. The
requirements for the denitrification process are: a) nitrogen present in the form of nitrates; b) an organic carbon source, and
c) an anaerobic environment. The processes currently used for biological denitrification are presented as follows.

1. Continuous flow stirred reactor (CFSR)

Moore and Schroeder (1971) investigated effects of nitrate feed rate and cell residence time on complete mixed continuous-
flow stirred-reactor (CFSR) operated at the steady state. They concluded that denitrification processes could be operated at
near maximum unit removal rates and still obtained acceptable nitrogen conversion (less than 2 mg of NO3-N/L in the
denitrified effluent). Stensel, et al. (1973), also indicated that cell retention time in the reactor would depend on the organic
carbon requirement and nitrate removal efficiency. An SRT of at least 4 days was recommended for design at 20 C and 30 C.
An SRT of at least 8 days was recommended for design at 10 C.

2. Activated sludge systems

Lesouef, et al. (1992), demonstrated a test on a two zone in activated sludge systems and showed to be capable of removing
75% of the total N from about 30 mg TN/L in the feed to

< 10 mg TN/L in the effluent. The multiple anoxic zones with a step feed process had recently been modeled and appeared
to be the most cost effective denitrification option because it made the fullest use of the carbon that was present in the feed
as the carbon source for step feed denitrification.

3. Fixed film reactor

Nurizzo and Mezzanatte (1992) studied denitrification results from a fixed-film anoxic sand reactor used for the treatment of
drinking water. The anoxic reactor was operated downflow at 20 m3/h with a nitrate loading of 0.4-1.5 kg NO3-N/m3/d.
They found that nitrate removal rates using sugar or glucose syrup as organic carbon sources were usually greater than 95%.

4. Fluidized bed biofilm reactor

Denitrification was performed in a fluidized-bed biofilm reactor using activated carbon particles (1.7 mm diam) as the
carrier and molasses as the carbon source (Coelhoso, et al., 1992). The experimental results illustrated that the average cell
residence was in the range 6.7-15.4 hours, which means biofilm was developed in a fast way; nitrogen removal was 5.3-8.6
kg NO3-N/m3/day. Experimental profiles of nitrate and nitrite were modeled by two consecutive zero-order reactions
coupled with substrate diffusion into the biofilm.

Fluidized Bed Biofilm Reactor (FBBR) was also studied by Jesis and Owen (1977). It was found that the use of small,
fluidized media enabled the FBBR to retain high biomass concentrations and, thereby, operated at significantly reduced
hydraulic retention times. It was also reported that when the volatile solid concentrations were between 30,000 to 40,000
mg/L (when the pilot-scale denitrification was employed), 99% of influent nitrates could be removed at an empty bed with
hydraulic retention times as low as 6 minutes.

Carbon and Nitrogen Removal Process

Currently, the processes used for carbon and nitrogen removal can be divided into two major groups: separated stage and
single stage processes. For multiple stages of carbon and nitrogen removal, there is a disadvantage for denitrification which
occurs either in the addition of external carbon or the recycle part of the effluent of nitrifying bacteria. Carbon and nitrogen
removal occurring in a single unit is a possibility to overcome these disadvantages. Multiple and single processes for the
removal of carbon and nitrogen are presented as follows.

Separated Stage Process

The carbon removal and nitrification stages were performed in an aerobic activated sludge tank, while the denitrification
was completed by either an anaerobic stirred tank reactor or an anaerobic plug-flow reactor (Murphy and Sutton, 1975). To
optimize the removal of nitrogen by the denitrification process, an external source of carbon and energy was normally
required. A two-stage nitrification and denitrification process in an activated sludge system with recycle of mixed liquid
from the nitrification stage to the denitrification stage was studied (ten Have, et al., 1994). It was concluded that full
denitrification was impossible with the pig manure supernatant as the only carbon source under the circumstances tested. It
was suggested that an extra carbon source, molasses, must be used.

Cecen, et al. (1992), conducted a subsequent nitrification-denitrification of high strength nitrogenous wastes in two upflow
submerged filters. The nitrified effluent of the nitrification filter was combined with diluted molasses and fed into the
submerged denitrification filter. It was found that the conversion of ammonium nitrogen to oxidized nitrogen was about 98%
in nitrification, but it strongly depended on the amount of dissolved oxygen which must be maintained at 4-5 mg/L. This
finding was consistent with a number of researchers (Gînenc and Harremoâ, 1985; Okey and Albertson, 1989). Molasses
was proved to be a good alternative for methanol in denitrification. The maximum denitrification rates can only be achieved
at an influent COD/NOx-N ratio of approximately 5.

Carbon and nitrogen removal by submerged biofilm columns attached to fibrous carriers with effluent recycle in subsequent
anaerobic and aerobic processes using domestic sewage was studied (Wang, et al., 1992). This process consisted of an
anoxic column and two aerobic columns in series. Four ropes of fibrous carriers were packed in each column, and four
diffusers were installed in the bottom of each aerobic column. The hydraulic retention times in the anoxic column and two
aerobic columns were 7h, 12h and 5h, respectively. In the aerated columns, sufficient air flow rate was supplied to maintain
DO at 3 mg/L. This process was shown to be capable of removing 80% of the total nitrogen and 90% and 98% of COD and
BOD5, respectively.

A Biological Nutrient Removal (BNR) process was studied by deBarbadillo, et al. (1995). A secondary treatment at the
Little Patuxent Water Reclamation Plant (LPWRP) consisted of a two-stage aeration system. The first stage was a
conventional activated-sludge system for BOD removal, followed by a separate-stage nitrification system. To implement the
BNR process, the existing aeration basins were divided into three functional zones: anaerobic, anoxic, and oxic (aerobic).
The anaerobic and oxic zones performed anaerobic selection and biological oxidation of BOD and ammonia, respectively.
The anoxic zone, with a recycle stream of nitrified mixed liquor from the oxic zone, provided biological denitrification
between the anaerobic and oxic zones. Data illustrated that the denitrification rate was lower for this process since the
internal recycle contained significant DO; thereby, inhibiting the denitrification reaction. Therefore, a certain anaerobic zone
was constructed after the last oxic zone to remove oxygen. Bench-scale tests showed that a detention time of 10 minutes
would yield a 1.5 mg/L drop in the mixed liquor DO concentration. A third anoxic zone was also provided to increase the
available anoxic detention time. The results also indicated that effluent phosphorus and nitrogen concentrations were
achieved to the levels of l mg/L TP, 6 mg/L NO3-N, and 8 mg/L TN.

A new moving bed biofilm reactor was developed (Rusten, et al., 1995). The biofilm grew on the small, free-floating plastic
elements with a large surface area and a density slightly less than 1.0 g/cm3. In the pilot test, one option was nitrification in
a recirculated system with predenitrification, and the other was the nitrification of preprecipitated wastewater in a once-
through system with post-denitrification. Predenitrification, which used untreated wastewater as a carbon source with less
than 100 mg SCOD/L and 25 mg TN/L, was proven to be carbon limited, and only 50% to 70% of total N removal was
obtained at a recirculation ratio of approximately 2.0 and a total empty bed hydraulic residence time of approximately 6
hours in this biofilm reactor. Post-denitrification, which used acetate as an external carbon source, could easily achieve 80%
to 90% of total N removal efficiency at total empty bed hydraulic residence time of less than 3 hours and organic loading
rate of less than 4.0 g/m2/d of SCOD. This system was operated at a temperature between 7 to 18 C.

Single Stage Process

Experiments on an Attached-Growth Circulating Reactor (AGCR) were conducted to investigate its efficiencies on organic
carbon and nitrogen removal (Karnchanawong and Polprasert, 1990). The optimal COD loading rate was found to be 5
g/m2/d corresponding to the TN loading rate of 0.54 g/m2/d. At this loading rate, the removal rates of COD and TN of 4.8
and 0.43 g/m2/d, respectively (or 96% COD removal and 79% TN removal efficiencies), could be achieved. The overall
AGCR performance was limited by the nitrification efficiency at the high TN loading rates of 0.54 g/m2/d.

A combined anaerobic-aerobic system with internal recirculation of effluent in a single fluidized bed reactor had
demonstrated simultaneous removal of organic carbon and nitrogen. With the loading rate of organic carbon < 1.2 kg/m3/d
and nitrogen < 0.2 kg/m3/d and HRT of 24 hours, the levels of purification could reach COD removals of > 80% and the
effluent concentration of (BOD5)S < 10 mg/L, NOx-N < 5 mg/L, NH3-N < 1 mg/L (Fdez-Polanco, et al., 1994).

A study was conducted on whether intermittent aeration of an aerated lagoon could provide effective organic carbon and
nitrogen removal for domestic wastewater. The influent concentrations of TCOD and TKN were 137.1 mg/L and 14.29
mg/L, respectively. The result demonstrated that intermittent aeration of an aerated lagoon did not affect removal efficiency
of organic substances. BOD removal ranged from 69% to 86% at a detention time of 2 to 6 days. The BOD level of effluent
from the pilot plant was in the range of 9.23-19.66 mg/L. Total nitrogen removal of 45% could be obtained from this
experiment (Koottatep and Araki, 1993).

Watanabe, et al. (1992), investigated the Simultaneous Nitrification and Denitrification (SND) in RBC. The experimental
result showed that SND was strongly influenced by C/N ratio of wastewater, hydraulic loading and oxygen partial-pressure
(Po) in the phase. The maximum nitrogen removal efficiency was achieved at C/N ratio of 6.0, HRT of 5.5 hours and a Po of
0.10 atm. Nitrification mainly occurred in a biofilm rotating through the air phase, while denitrification mainly occurred in a
biofilm rotating through the water phase. The removal efficiency, due to the SND, increased with a decrease in the disk
rotating speed up to the optimum speed of around 3 rpm.

The sequencing batch reactor (Irvine, et al., 1979) and the intermittent operation of the extended aeration process (Goronszy,
1979) had exhibited their ability of accommodating nitrification, denitrification, biological oxidation, sedimentation and
flocculation all in a single unit for domestic sewage.

Kondo, et al. (1992), developed simultaneous removal of BOD and nitrogen with an anoxic/oxic porous biomass support
system. In the full-scale field test, the system was used for the treatment of sewage and operated with a schedule of 1.5
hours aeration and 0.5 hours agitation. Less than 20 mg/L of the effluent BOD (93% BOD removal) and 15 mg/L of total
nitrogen could be achieved at the organic loading rate of 0.8 kg-BOD/m3/day. At the low temperature of 13 C, the
nitrification rate was slow and the removal efficiency of nitrogen fell to 65%. Total nitrogen removal efficiency could reach
as much as 75% with an annual average temperature of 13 C-31 C.

Nitrification and Denitrification Studies at the University of Hawaii

Dilute Pig Wastewater

A field 2.5 m3 Intermittently Aerated Bio-Carousel Reactor (IABR), as shown in Figure 1, was investigated for the
treatment of dilute swine wastewater (Yang and Koba, 1988). Two alternate modes of operation; namely, no sludge recycling
and with sludge recycling, were tested. For better removal of COD and TKN, process operation with sludge recycling for
increase of SRT and decrease of HRT was recommended. Combined biological oxidation, nitrification and denitrification,
and settling all in a single unit could be achieved using the IABR process. At the conditions of HRT = 2.7 days, SRT > 150
days, TCOD loading rate of 1.2 g/L/day and TKN loading rate of 0.17 g/L/day, the TCOD and TKN removal efficiencies
could achieve 87% and 95%, respectively. The starting up of this process requires 2 weeks.

Figure 1: Schematic of intermittently aerated bio-carousel reactor.

Figure 2: Combined bio-fixed film and aquatic plant treatment systems (serial ponds).

A combined bio-fixed film and aquatic plant (CBFFAP) system (as shown in Figure 2) was investigated for treatment of
dilute swine wastewater and anaerobically digested effluent under a tropical climate and land-limited conditions (Yang and
Chen, 1994). It was found that the CBFFAP process could remove more than 90% of COD and 95% of TKN under 0.1 g
COD/L/day. Because it is simple in design and operation and low in maintenance, the process can be very beneficial for
application in tropical areas with moderately limited land. Also, the almost zero energy input for the operation of the system
has very important implications for energy conservation. Based on a 1000-pig operation, the land requirement is about 0.077
ha. Compared to the facultative pond and aquatic treatment systems, it is considered as moderately land-saving. More
importantly, this system actively combines various types of biological treatment processes, including anaerobic/aerobic bio-
fixed film, facultative/aerobic p onds, nitrification/denitrification, and aquatic plant processes. Consequently, the reduction
of organic solid/liquid, organic nitrogen, NH4-N and NO3-N can occur simultaneously when all of the anaerobic, anoxic,
and aerobic conditions are maintained in one system.
A prototype of a 300-pig waste treatment system, as shown in Figure 3, including a solid/liquid separation (12 m3), 2
anaerobic reactors (10 m3 each), 1 aeration unit (20 m3) and 1 sedimentation tank (20 m3) were investigated (Yang, et al.,
1994). The concept of prefabrication of each treatment component was used. The present two stages of the anaerobic
reactors are able to properly treat the high solid content (10% of TS) of the swine wastes. The aeration unit provides a
TCOD and NH3-N reduction of 94% and 91%, respectively. The process performance of the aeration and rock filter units
are shown in Table 1. Apparently, the nitrification and denitrification processes occur in the aeration and rock filter units,
respectively.

Synthetic Wastewater

A denitrification process was examined by using the Entrapped-Mixed-Microbial Cells Immobilization (EMMCI) process
(Nitisoravut and Yang, 1992; Yang, et al., 1995). The experimental set up is shown in Figure 4. At an HRT of 1.8 hours, the
maximum nitrate-N loading rate of 591.6 g/m3.hr (based on carrier volume) was achieved with an 88.3% denitrification
efficiency. Under a high range of NaHCO3 concentrations from 10 to 20 g/L, the denitrification efficiency above 96% could
be obtained. The process was found to be an economical alternative for treatment of brine wastes and exhibits a great
potential to be coupled with an ion exchange process for complete nitrate removal from groundwater contaminated with
nitrate.

An Entrapped-Mixed-Microbial-Cell (EMMC) process (as shown in Figure 5) was studied to investigate its simultaneous
removal of carbon and nitrogen in a single reactor which was operated on an alternate schedule of aerobic, anoxic and
anaerobic condition (Zhang, 1995). Two different sizes of carriers with dimensions of 20 x 20 x 20 mm3 (large carriers) and
10 x 10 x 10 mm3 (medium carriers) were employed in this study. The reactors were fed with simulated domestic
wastewater with concentrations of 250 mg/L of chemical oxygen demand (COD) and 26 mg/L of NH3-N. The optimum
operating conditions for simultaneous carbon and nitrogen removal were achieved at the hydraulic retention time (HRT) of 9
hours at 1 hour of aeration and 2 hours of no aeration in the medium carrier system. At the organic and nitrogen loading of
0.67 kg COD/m3 of void volume/day and 0.069 kg/NH3-N/m3 of void volume/day, the removal efficiencies of soluble
chemical oxygen demand (SCOD) and total nitrogen (Ntotal-N) of 96% and 84%, respectively, were achieved in the
medium carrier system. The EMMC process also demonstrated that the secondary sedimentation tank could be eliminated
because of the low concentration of total suspended solid, TSS (< 30 mg/l), in the effluent.

Since the dissolved oxygen plays an important role in the nitrification and denitrification processes, it is necessary to
provide the optimum demands of DO for the development of nitrification and denitrification. The experiment was conducted
to investigate if the change of no aeration time could affect the carbon removal and to optimize the operating conditions by
controlling different aerobic/anaerobic periods for developing the optimum efficiency of nitrification and denitrification.

For the large carriers (as shown in Table 2), compared to the results at the condition of 1 hour of aeration and 1 hour of no
aeration at HRT = 9 hours, the results indicated that SCOD, SBOD5 removal efficiencies and concentrations in the effluent
are not affected by the increase of no aeration time from 1 hour to 2 hours. The nitrification efficiency is decreased from
98% to 78%, which respects to the increase of NH3-N concentration from 0.46 mg/L to 5.55 mg/L in the effluent. The
NO3-N concentration in the effluent is decreased from 5.7 mg/L to 2.61 mg/L. The nitrogen removal efficiency is decreased
from 76% to 67%. The TSS concentration in the effluent is not affected and maintained less than 30 mg/L. The experimental
results also demonstrate that the increase of no aeration time can decrease the nitrification efficiency. The overall total
nitrogen removal is reduced due to the decrease of nitrification efficiency, although the NO3-N concentration is decreased in
the effluent.

Figure 3: Swine waste treatment components.

Figure 4: Experimental set-up for denitrification using entrapped-mixed-microbial cells immobilisation (EMMCI) reactor.

Table 1: Process Performance of Aeration and Rock Filter Systems

Table 2: Process Performance of Large and Medium Carriers with One Hour of Aeration and One Hour of No Aeration at
HRT = 9 hours
Table 3: Process Performance of Large and Medium Carriers with One Hour of Aeration and Two Hours of No Aeration at
HRT = 9 hours

Figure 5: Schematic diagram of the EMMC process.

For the medium carriers, compared to the results at the condition of 1 hour of aeration and 1 hour of no aeration at HRT = 9
hours, the SCOD and SBOD removal efficiencies and concentrations in the effluent are also not affected by an increase of
no aeration time from 1 hour to 2 hours. Moreover, the nitrification efficiency was still maintained at 99%, which respects to
the 0.2 mg/L of NH3-N concentration in the effluent although the no aeration time increased from 1 hour to 2 hours. As the
N03-N concentration is decreased from 7.04 mg/L to 3.77 mg/L, the overall nitrogen removal efficiency is increased from
72% to 84% in the effluent. The TSS concentration in the effluent is also less than 30 mg/L.

The DO concentration, which was monitored by a DO probe in the reactor, shows that it ranges from 0-4.5 mg/L by
increasing the no aeration time from 1 hour to 2 hours. This indicates that the anaerobic zone is formed in the reactor at this
condition. The decrease of DO concentration results in the decrease of NO3-N concentrations in the effluent for both large
and medium carriers.

In general, the nitrification and denitrification processes can be applied for the removal of nitrogen contained in the
wastewater. The concentration of nitrogen in the wastewater can be from low to high contents, which require an appropriate
treatment process for the removal. The application of a certain biological treatment method or process will vary, depending
on the location, land availability, social and economic concerns. Thus, both traditional and innovative technologies for the
nitrification and denitrification processes for nitrogen removal from the wastewater need to be evaluated and applied
properly. Thus, the protection of our environment and conservation of our resources can be warranted.

Acknowledgement

The funding for this work was supported, in part, by the United States Department of Agriculture under CSRS Special Grant
no. 84-CSRS-2-2387 and No. 91-34135-6150, managed by the Pacific Basin Advisory Group (PBAG); the Research
Foundation of the American Water Works Association to the University of North Carolina at Charlotte with a subcontract to
the Department of Biosystems Engineering, University of Hawaii at Manoa; and the Mini and Vision Grants of the Hawaii
Institute of Tropical Agriculture and Human Resources, University of Hawaii.

References

Barnes, D. and Bliss, P.J. 1983. Biological control of nitrogen in wastewater treatment, 1st edition, E. & F.N. Spoon, Ltd.,
New York.
Bitton, G. 1994. Wastewater Microbiology, Wiley-liss, New York.
Boller, M. and Gujer, W. 1986. Nitrification in tertiary trickling filter nitrification towers, J. Wat. Poll. Contr. Fed. , 58:60.
Cecen, F. and Gînenc, E. 1992. Nitrification-denitrification of high-strength nitrogen wastewater in two up-flow submerged
filters, Wat. Sci. Tech. 26(9-11):2225-2228.
Coelhoso, L., Boaventura, R. and Rodrigues, A. 1992. Biofilm reactors: an experimental and modeling study of wastewater
denitrification in fluidized-bed reactors of activated carbon particles, Biotech. Bioengr., 40:625.
Darimont, T., Schwabe, R., Sonneborn, R. & Schulze, G. 1988. Nitrate in drinking water in the West Germany wine-growing
areas of Baden and WÅrttemberg, Analysis and Chemistry of Water Pollutions, Gordon and Breach Science Publisher, Inc.,
109-110.
DeBarbadillo, C., Kharkar, S.M. and Jaworski, L.P. 1995. Optimal nutrient, Water and Environ. & Tech., 7(2):40-44.
Dunn, I.J., Tanaka, H., Uzman, S. and Denac, M. 1984. Biofilm fluidized bed reactors and their application to wastewater
nitrification, Biotech. Bioengr., 23:1683-1696.
Fdez-Polanco, F.J.R. and Garcia, P.A. 1994. Behavior of an anaerobic/aerobic pilot scale fluidized bed for the simultaneous
removal of carbon and nitrogen, Wat. Sci. Tech., 29(10-11):339-346.
Figueroa, L.A. and Silverstein, J. 1992. The effect of particulate organic matter on biofilm nitrification, Wat. Environ. Res.,
64:728-733.
Gînenc, J.E. and Harremoâ, P. 1985. Nitrification in rotating disk System-I-Criteria for transition from oxygen to ammonia
rate limitation, Wat. Res., 19:1119-1127.
GonzÝlez-Martinez, S. and Duqu-Luciano, J. 1992. Aerobic submerged biofilm reactors for wastewater treatment, Wat.
Res., 26, 825.
Goronszy, M.C. 1979. Intermittent operation of the extended aeration process for small systems, J. Wat. Poll. Contr. Fed.,
51:274-287.
Hitdlebaugh, J.A. and Miller, R.D. 1981. Operational problems with rotating biological contactors, J. Wat. Poll. Contr. Fed.,
53:1283-1293.
Horan, N.J. 1989. Biological Wastewater Treatment Systems, John Wiley & Sons, Rochester, New York.
Irvine, R.L., Miller, G. and Bhamrah, A.S. 1979. Sequencing batch treatment of wastewater in rural areas, J. Wat. Poll.
Contr. Fed., 51:244-254.
Jesis, J.S. and Owen, R.W. 1977. Biological fluid-bed treatment for BOD and nitrogen removal, J. Wat. Poll. Contr. Fed.,
49:816.
Jimenez, B., et al. 1987. Design considerations for nitrification and denitrification process using two fixed bed reactors in
series, Wat. Sci. Tech., 19:139-149.
Karnchanawong, S. and Polprasert, C. 1990. Organic carbon and nitrogen removal in attached-growth circulating reactor,
Wat. Sci. Tech., 22(3-4):179-186.
Klees, R. and Silverstein, J. 1992. Improved biological nitrification using recirculation in rotating biological contactors, Wat.
Sci. Tech., 26(3-4):545-553.
Kondo, M., Hozo, S. and Inamori, Y. 1992. Simultaneous removal of BOD and nitrogen with anoxic/oxic porous biomass
support systems, Wat. Sci. Tech., 26:2003-2006.
Koottatep, C.L. and Araki, H. 1993. Intermittent aeration for nitrogen removal in small aerated lagoon, Wat. Sci. Tech.,
28(10):335-341.
Lesouef, A., Payraudeau, M., Rogalla, F. and Kleiber, B. 1992. Optimizing nitrogen removal reactor configurations by on-
site calibration of the IAWPRC activated sludge model, Wat. Sci. Tech., 25(6):105-123.
Moore, S.F. and Schroeder, E.D. 1971. The effect of nitrate feed rate on denitrification, Wat. Res., 5:445-452.
Murphy, K.L. and Sutton, P.M. 1975. Pilot scale studies on biological denitrification, Prog. Wat. Tech., 7(2):317-328.
Nitisoravut, S. 1991. Denitrification with entrapped-microbial cell immobilization technology. Master's Thesis, University
of Hawaii at Manoa, 103 pp.
Nitisoravut, S. and Yang, P.Y. 1992. Denitrification of nitrate-rich water using entrapped-mixed-microbial cells
immobilization technique, Wat. Sci. Tech. G.B., 26:923-931.
Nugent, M., Kamrin, M., Wolfson, L. and D'ltri, F.M. 1988. Nitrate _ a drinking water concern. A Publication of the
Community Assistance Program in Environmental Toxicology (Capet), Center for Environmental Toxicology and the
Institute of Water Research, Michigan State University.
Nurizzo, C. and Mezzanatte, V. 1992. Groundwater biodenitrification on sand fixed film reactor using sugars as organic
source, Wat. Sci. Tech., 26:827.
Okey, R.W. and Albertson, O.E. 1989. Diffusion's role in regulating rate and masking temperature effects in fixed-film
nitrification, J. Wat. Poll. Contr. Fed., 61:550-559.
Parker, D.S. and Richards, T. 1986. Nitrification in trickling filters, J. Wat. Poll. Contr. Fed., 58:896-902.
Rimer, A.E. and Woodward, R.L. 1972. Two-stage activated sludge pilot-plant operations at Fitchburg, Massachusetts, J.
Wat. Poll. Contr. Fed., 44:101-116.
Rusten, B., Hem, L.J. and Odegaard, H. 1995. Nitrogen removal from dilute wastewater in cold climate using moving-bed
biofilm reactors, Wat. Environ. Res., 67(1):65-74.
Sawyer, C.N. and McCarty, P.L. 1978. Chemistry for Environmental Engineering, 3rd edition, McGraw-Hill, New York.
Sayre, I.M. 1988. International standards for drinking water, Journal AWWA, 80, 53-60.
Stenquist, R.J., Parder, D.S. and Dosh, T.J. 1974. Carbon oxidation-nitrification in synthetic media trickling filters, J. Wat.
Poll. Contr. Fed. , 46:2327-2339.
Stensel, H.D., Loehr, R.C. and Lawrence, A.W. 1973. Biological kinetics of suspended-growth denitrification, J. Wat. Poll.
Contr. Fed., 45:249-261.
Suwa, Y., Yamagishi, T., Urushigawa, Y. and Hirai, M. 1989. Simultaneous organic carbon removal-nitrification by an
activated sludge process with cross-flow filtration, J. of Fermentation and Bioengineering, 67(2):119-125.
ten Have, P.J.W., Willers, H.C. and Derikx, P.J.L. 1994. Nitrification and denitrification in an activated-sludge system for
supernatant from settled sow manure with molasses as an extra carbon source, Bioresource Tech., 47:135-141.
Tijhuris, L., van Loosdrecht, M.C.M. and Heijnen, J.J. 1992. Nitrification with biofilms on small suspended particles in
airlift reactors, Wat. Sci. Tech., 26:2207-2211.
U.S. EPA. 1975. Process Design Manual for Nitrogen Control. Office of Technology Transfer, Washington, D.C.
Wang, B., Li, G., Yang, Q. and Liu, R. 1992. Nitrogen removal by a submerged biofilm process with fibrous carriers, Wat.
Sci. Tech., 26:2039-2042.
Watanabe, Y., Matsuda, S. and Ishiguro, M. 1992. Simultaneous nitrification and denitrification in micro-aerobic biofilms,
Wat. Sci. Tech., 26(3-4):511-522.
Weng, C.N. and Molof, A.H. 1974. Nitrification in the biological fixed film RBC, J. Wat. Poll. Contr. Fed., 46:1674-1685.
Weismann, U. 1994. Biological nitrogen removal from wastewater. In: Advances in Biochemical Engineering/Biotechnology,
A. Fiechter (ed.), 51:113-154.
Yang, P.Y. and Koba, B.H. 1988. Field IABR process for the treatment of dilute swine wastewater, Trans. of ASAE,
31:202-207.
Yang, P.Y., et al. 1994. A prototype animal waste treatment system in Hawaii, presented at the ASAE Summer Meeting,
Paper No. 94-4062, Kansas City, Missouri.
Yang, P.Y. and Chen, H. 1994. A land limited and energy saving treatment system for dilute swine wastewater, Bioresource
Tech., 49:129-137.
Yang, P.Y., et al. 1995. Nitrate removal using a mixed-culture entrapped microbial cell immobilization process under high
salt conditions, Wat. Res., 29(6):1525-1532.
Zhang, Z.Q. 1995. Entrapped-mixed-microbial-cell process for removal of carbon and nitrogen in one single reactor. M.S.
Thesis, University of Hawaii, Honolulu, Hawaii, USA, 151 pp.

Back to contents list


To the next chapter
Back to books published by Eubios Ethics Institute
Back to Eubios Ethics Institute Home Page

You might also like