Você está na página 1de 20

PII: S0043-1354(98)00371-6

Wat. Res. Vol. 33, No. 7, pp. 15591578, 1999 # 1999 Elsevier Science Ltd. All rights reserved Printed in Great Britain 0043-1354/99/$ - see front matter

REVIEW PAPER THE USE OF THE ANAEROBIC BAFFLED REACTOR (ABR) FOR WASTEWATER TREATMENT: A REVIEW
M M WILLIAM P. BARBER* and DAVID C. STUCKEY**

Department of Chemical Engineering and Chemical Technology, Imperial College of Science, Technology and Medicine, Prince Consort Road, London SW7 2BY, U.K. (First received May 1998; accepted in revised form August 1998) AbstractA review concerning the development, applicability and possible future application of the anaerobic baed reactor for wastewater treatment is presented. The reactor design has been developed since the early 1980s and has several advantages over well established systems such as the upow anaerobic sludge blanket and the anaerobic lter. These include: better resilience to hydraulic and organic shock loadings, longer biomass retention times, lower sludge yields, and the ability to partially separate between the various phases of anaerobic catabolism. The latter causes a shift in bacterial populations allowing increased protection against toxic materials and higher resistance to changes in environmental parameters such as pH and temperature. The physical structure of the anaerobic baed reactor enables important modications to be made such as the insertion of an aerobic polishing stage, resulting in a reactor which is capable of treating dicult wastewaters which currently require several units, ultimately signicantly reducing capital costs. # 1999 Elsevier Science Ltd. All rights reserved Key wordsanaerobic baed reactor, anaerobic digestion, reactor development, performance, solids retention, molids odelling, full-scale.

INTRODUCTION

The successful application of anaerobic technology to the treatment of industrial wastewaters is critically dependent on the development, and use, of high rate anaerobic bioreactors. These reactors achieve a high reaction rate per unit reactor volume (in terms of kg COD/m3 d) by retaining the biomass (Solids Retention Time, SRT) in the reactor independently of the incoming wastewater (Hydraulic Residence Time, HRT), in contrast to Continually Stirred Tank Reactors (CSTRs), thus reducing reactor volume and ultimately allowing the application of high volumetric loading rates, e.g. 1040 kg COD/m3 d (Iza et al., 1991). High rate anaerobic biological reactors may be classied into three broad groups depending on the mechanism used to achieve biomass detention, and these are xed lm, suspended growth, and hybrid. There are currently 900 full-scale installations in the world today (Habets, 1996), and they are distributed as follows: Upow Anaerobic Sludge Blanket (UASBsuspended growth) 67% (Lettinga et al., 1980); CSTR 12%; Anaerobic Filter (AFxed lm) 7% (Young and McCarty, 1969); other 14%. The highest load*Author to whom all correspondence should be addressed. [Tel. +44-171-594-5591; Fax: +44-171-594-5604, Email: d.stuckey@ic.ac.uk].

ing rates achieved during anaerobic treatment to date are attributed to the ``Anaerobic Attached Film Expanded Bed'' (AAFEB) reactor (120 kg COD/m3 d, Switzenbaum and Jewell (1980)), but its inherent complexity and high operating costs limit its practical use on a wide scale. Around the same time as Lettinga developed the UASB, McCarty and co-workers at Stanford noticed that most of the biomass present within an anaerobic Rotating Biological Contactor (RBC, Tait and Freidman (1980)) was actually suspended, and when they removed the rotating discs they developed the Anaerobic Baed Reactor (ABR, McCarty (1981)). However, baed reactor units had previously been used to generate a methane rich biogas as an energy source (Chynoweth et al., 1980). Although not commonly found on a large scale, the ABR has several advantages over other well established systems, and these are summarised in Table 1. Probably the most signicant advantage of the ABR is its ability to separate acidogenesis and methanogenesis longitudinally down the reactor, allowing the reactor to behave as a two-phase system without the associated control problems and high costs (Weiland and Rozzi, 1991). Two-phase operation can increase acidogenic and methanogenic activity by a factor of up to four as acidogenic bacteria accumulate within the rst stage

1559

1560

William P. Barber and David C. Stuckey

Table 1. Advantages associated with the anaerobic baed reactor Advantage Construction 1 Simple design 2 No moving parts 3 No mechanical mixing 4 Inexpensive to construct 5 High void volume 6 Reduced clogging 7 Reduced sludge bed expansion 8 Low capital and operating costs Biomass 1 No requirement for biomass with unusual settling properties 2 Low sludge generation 3 High solids retention times 4 Retention of biomass without xed media or a solid-settling chamber 5 No special gas or sludge separation required Operation 1 Low HRT 2 Intermittent operation possible 3 Extremely stable to hydraulic shock loads 4 Protection from toxic materials in inuent 5 Long operation times without sludge wasting 6 High stability to organic shocks

(Cohen et al., 1980, 1982), and dierent bacterial groups can develop under more favourable conditions. The advantages of two-phase operation have been extensively documented (Pohland and Ghosh, 1971; Ghosh et al., 1975; Cohen et al., 1980, 1982). These benets have catalysed the development of other staged reactor congurations such as the ``Multiplate Anaerobic Reactor'' (ElMamouni et al., 1992; Guiot et al., 1995), ``Upow Staged Sludge Bed (USSB)'' (van Lier et al., 1994, 1996) and the ``Staged Anaerobic Filter'' (Alves et all., 1997), all of which have showed considerable potential for wastewater treatment. Disadvantages of the baed reactor design at pilot/full-scale include the requirement to build shallow reactors to maintain acceptable liquid and gas upow velocities, and problems with maintaining an even distribution of the inuent (Tilche and Vieira, 1991). However, despite its many potential advantages over other high rate anaerobic reactor designs, and the ever-increasing number of publications, there has never been any attempt to collate all this information in a review. Hence, the objective of this paper is to review the currently available literature on the ABR, focusing on reactor development, hydrodynamics, performance, biomass characteristics and retention, modelling, full-scale operation and a comparison with other well established alternatives. Finally, based on the review, a closing section will discuss future prospects for the ABR.
REACTOR DEVELOPMENT

The ABR is a reactor design which uses a series of baes to force a wastewater containing organic pollutants to ow under and over (or through) the baes as it passes from the inlet to the outlet (McCarty and Bachmann, 1992). Bacteria within

the reactor gently rise and settle due to ow characteristics and gas production, but move down the reactor at a slow rate. The original design is shown in Fig. 1(C), although Fig. 1(A) is more commonly recognised. However, in order to improve reactor performance several modications have been made (Fig. 1(B and DJ)). The main driving force behind reactor design has been to enhance the solids retention capacity, but other modications have been made in order to treat dicult wastewaters (e.g. with a high solids content, Boopathy and Sievers (1991)), or simply to reduce capital costs (Orozco (1997), Fig. 1(F)). A summary of the main alterations is shown in Table 2. In 1981, Fannin et al. (1981) added vertical bafes to a plug-ow reactor treating high solids sea kelp slurry (Fig. 1(C)) in order to enhance the reactor's ability to maintain high populations of slowly growing methanogens, which were being replaced by the inuent solids. With a constant loading rate of 1.6 kg COD/m3 d methane levels increased from 30 to over 55% with a methane yield of 0.34 m3/kg VSS after the baes were added. In a later study, Bachmann et al. (1983) compared the performance of two baed reactors before and after narrowing the downow chambers and slanting the bae edges (Fig. 1(A) and Table 2). Although methane production rates and reactor eciency were improved in the modied design, a decrease in the methane content of the biogas was also noted. Despite the alterations, the performance of both reactors was inferior to an anaerobic lter and rotating biological disc operated under the same conditions. COD removal eciencies were 82, 92 and 90% for the modied bae, anaerobic lter, and rotating biological disc reactors respectively. The next major change occurred with the development of the rst of several hybrid designs (Tilche and Yang, 1987, Fig. 1(E)). The motivation behind the alterations was based on enhancing solids retention for high strength wastewater treatment. The reactor was signicantly larger than those used previously, and incorporated a solids settling chamber after the nal compartment. Solids washed out from the baed reactor were collected in the settling chamber and subsequently recycled to the rst compartment. Packing was also positioned at the liquid surface of each compartment with randomly packed Pall rings in the rst two chambers, and a deeper, structured, modular corrugated block which had a high voidage in the third chamber. Bioocs, which became buoyant due to a reduction in density caused by high gas production, were retained in the rst chamber due to the packing. Higher loading rates were possible with this structure due to minimal solids washout during elevated gas mixing. Each gas chamber was separated permitting the measurement of gas composition and production from each compartment. Although benecial in this regard, the separation of the gas can

The anaerobic baed reactor: a review

1561

Fig. 1. Variations of the baed reactor. (A) Single gas headspace, (B) individual gas headspace, (C) vertical, (D) horizontal, (E) hybrid with settling zone, (F) open top, (G) enlarged rst compartment, (HJ) various packing arrangements: (H) up-comers, (I) down-comers, (J) entire reactor. Key: W = Wastewater, B = Biogas, E = Euent, S = Solids, (shaded areas represent random packing).

also enhance reactor stability by shielding syntrophic bacteria from the elevated levels of hydrogen which are found in the front compartments of the baed reactor. In order to treat swine wastewater containing a high content of small particulate material, Boopathy and Sievers (1991) further modied the baed reactor. The main problems associated with the treatment of swine wastewater in a baed reactor were the inability to produce a oating sludge layer which would enhance solids retention, and, the high velocities associated with the baes caused signicant washout of solid material. Therefore, the baed reactor was modied to reduce upow liquid

velocities and to accept whole waste. The rst compartment of a two-chamber unit was doubled in size to 10 l and this was followed by a second compartment of 5 l (Fig. 1(G)). Performance characteristics and solids retention capabilities were compared with a three-chamber unit with equal volume chambers. The additional chamber in the three-compartment unit, together with physical modications, provided a longer solids retention time and superior performance than the reactor with only two compartments. This was in contrast to earlier ndings (Sievers, 1988), when no dierence was found in treatment eciency compared with compartment number in unmodied reactors.

1562

William P. Barber and David C. Stuckey


Table 2. Development of the ABR

Fig. 1(C) 1(A)

Modication addition of vertical baes to a plugow reactor (i) downow chambers narrowed (ii) slanted edges on baes (40458)

Purpose enhances solids retention to allow better substrate accessibility to methanogens (i) encourages cell retention in upow chambers (ii) routes ow towards centre of compartment encouraging mixing (i) enhances solids retention (ii) prevents washout of solids (iii) ease and control of gas measurement, provides enhanced reactor stability better treatability of high solids wastewater

Ref. Fannin et al., 1981 Bachmann et al., 1983

1(E)

(i) settling chamber (ii) packing positioned at top of each chamber (iii) separated gas chambers enlargement of rst chamber

Tilche and Yang, 1987

1(G)

Boopathy and Sievers, 1991

The larger compartment in the two-compartment reactor acted as a natural lter and provided superior solids retention for the small particles. The reactor collected twice the amount of solid material (20.9 g/l) than the reactor with three chambers. This was further substantiated in the solids washout data, which was lower in the two-compartment reactor despite showing lower treatment eciency. Further analysis showed that despite losing more solids, the three-compartment reactor was more ecient at converting the trapped solids to methane.

REACTOR HYDRODYNAMICS

Flow patterns The hydrodynamics and degree of mixing that occur within a reactor of this design strongly inuence the extent of contact between substrate and bacteria, thus controlling mass transfer and potential reactor performance. In 1992, Grobicki and Stuckey conducted a series of residence time distribution studies by tracking the fate of an inert tracer (Li+) in the euent of a number of baed reactors (48 chambers), both with and without biomass, at various HRTs, and incorporated the data into ``Dispersion'' and ``Tanks In Series'' models previously described by Levenspiel (1972). The models provided a useful method to calculate the degree of mixing and the amount of unused volume (known as ``dead space'') within the reactor. They found low levels of dead space (<8% hydraulic dead space in an empty reactor) in comparison with other anaerobic reactor designs, e.g. 5093% in an anaerobic lter (Young and Young, 1988), and >80% in a CSTR (Stuckey, 1983). Dead space increased to 18% on the addition of 8 g VSS/l, however, no direct correlation between hydraulic dead space and HRT could be drawn. At low HRT, the presence of biomass had no signicant eect on hydraulic dead space, which was found to be a function of owrate and number of baes. This contrasted with biological dead space, which was found to be a function of biomass con-

centration, gas production, and owrate, and which increased with increasing owrates. At high loading rates caused by low HRT, gas production as well as increased owrates kept sludge beds partly uidised. Therefore, the contradictory eects of hydraulic and biological dead space prevented a correlation being derived between HRT and overall dead space. Biological dead space was established as the major contributor to overall dead space at high HRT, but its eect decreased at lower HRT since gas production disrupted channelling within the biomass bed. Severe channelling, caused by large hydraulic shocks, was found to be benecial since most of the biomass was not entrained in the ow, and this resulted in low washout and a fast recovery in performance (Grobicki and Stuckey, 1992; Nachaiyasit and Stuckey, 1997c). Nevertheless, investigations of the hydrodynamics to date have not taken into account various other factors which are probably important, and these include biogas mixing eects, viscosity changes due to extracellular polymer production, and biomass particle size. In addition, no work has been done of the rate at which solid particles/biomass move down the reactor. Eect of euent recycle Recycling the euent stream tends to reduce removal eciency because the reactor approaches a completely mixed system, and therefore the mass transfer driving force for substrate removal decreases despite a small increase in the loading rate. The eect of loading rate and increasing recycle ratios on performance is shown in Table 3. Chynoweth et al. (1980) observed a positive eect caused by recycling twenty percent of the euent, when the methane yield increased by over 30%. The addition of a recycle stream was also found to alleviate the problems of low pH caused by high levels of volatile acids at the front of the reactor, and discourage gelatinous bacterial growth at the reactor inlet for the treatment of a complex protein carbohydrate wastewater (Bachmann et al., 1983). Another benet of recycle is the dilution of toxicants and reduction of substrate inhibition in the

The anaerobic baed reactor: a review


Table 3. Reactor performance vs increasing recycle ratio Recycle ratio 0 0 0.1 0.25 0.5 0.5 1 2.2 2 3 5 6 9.6 11.7 13.8 Reactor volume (l) 13 10 10 10 13 10 10 13 10 13 13 13 13 13 13 Inuent COD (g/l) 8 4 4 4 8 4 4 8 4 8 8 8 8 8 8 Organic loading rate (kg/m3 d) 2.70 4.80 4.80 4.81 2.86 4.87 4.94 3.85 5.18 3.42 5.76 6.83 11.01 16.92 17.62 COD removal (%) 93b 99 98 97 88c 97 97 81c 96 91 77 75 68 55 60 Ref.

1563

Bachmann et al., 1985a Nachaiyasit and Stuckey, 1995b Nachaiyasit and Stuckey, 1995b Nachaiyasit and Stuckey, 1995b Bachmann et al., 1985a Nachaiyasit and Stuckey, 1995b Nachaiyasit and Stuckey, 1995b Bachmann et al., 1985a Nachaiyasit and Stuckey, 1995b Bachmann et al., 1985a Bachmann et al., 1985a Bachmann et al., 1985a Bachmann et al., 1985a Bachmann et al., 1985a Bachmann et al., 1985a

a Recycle ratios calculated from data supplied based on R = 0 for retention time of 71 h, organic loading rates converted from hydraulic loading rates supplied.bLoading rates calculated from recycle ratio data.cNutrient limited conditions.

inuent. (Bachmann et al., 1985; Grobicki and Stuckey, 1991). From theoretical considerations, recycle should have a negative eect on reactor hydrodynamics by causing increased mixing (which encourages solids loss, and disrupts microstructures of bacteria living in symbiotic relationships (Henze and Harremoe s, 1983)) and enhancing the amount of dead space (Grobicki and Stuckey, 1992; Nachaiyasit, 1995). In her thesis, Nachaiyasit (1995) showed that dead space doubled to approximately 40% when the recycle ratio was increased from zero to 2. The author also reported a sudden loss of solids when the recycle ratio was doubled. Increasing recycle has also been linked to an increase in the sludge volume index using anaerobic lters (Matsushige et al., 1990). Mixing caused by recycle has also been found to cause a return to single phase digestion, therefore the benets arising from the separation of acidogenic and methanogenic phases are partially lost. Bachmann et al. (1985) noticed that methanogenic activity was more uniformly distributed over the whole reactor after recycle was used. The consequences of this observation are scavenging bacteria (such as Methanosaeta) will end up at the front of the reactor where harsh conditions of high substrate concentration, high hydrogen partial pressure and low pH will make them relatively inactive, and poorly scavenging acid producing bacteria pushed towards the rear of the reactor will be starved since less substrate will be available. Nachaiyasit (1995)

discovered a fall in both gas production and methane composition down the reactor when the recycle ratio was increased. The overall benets of recycle are unclear, and ultimately its use will depend on the type of waste being treated. If pH problems are severe, the inuent has high levels of toxic material, or high loading rates are preferred then recycle will be benecial. However, as can be seen in Table 4, the disadvantages of recycle show that it should be used with caution, and only when absolutely necessary.

REACTOR PERFORMANCE

Start-up The overall objective of start-up is the development of the most appropriate microbial culture for the waste stream in question. Once the biomass has been established, either as a granular particle or a oc, reactor operation is quite stable. The important factors governing the start-up of anaerobic reactors have been summarised in the literature (Stronach et al., 1986; Weiland and Rozzi, 1991; Hickey et al., 1991), and will not be discussed here. A collection of data obtained during reactor startup is shown in Table 5. Initial loading rates should be low so that slow growing micro-organisms are not overloaded, and both gas and liquid upow velocities should be low so that occulent and granular growth is encouraged. The recommended initial loading rate is ap-

Table 4. Advantages and disadvantages of euent recycle Advantages Disadvantages

1 Front pH increased 1 Overall eciency reduced 2 Reduction of inuent toxicity and substrate 2 Increased solids loss inhibition 3 Higher loading rates possible 3 Increased hydraulic dead space 4 Better substrate/biomass contact 4 Disruption of bacterial communities and bioocs 5 Encourages one-phase digestion

1564

William P. Barber and David C. Stuckey


Table 5. Start-up data for the ABR

LRa initial 1 2 0.4 4.33 1.2 0.97 2.2 13.04 4.35 1.2 NG 1.2

Timeb initial LR

LR increased (ramp increase) (ramp increase) 0.53 10.26 2.4 NG 2.6 2.4 (ramp increase) 2.4

Time increased LR

LR nal 4 20 0.8 12.25 4.8 12.25 3.5 4.8 20 4.8

Start-up timec (d) 57 >60 >60 62 77 78 90 >95 >100 128

Initial VSS (g/l) NGd NG NG 4.01 8.77 4.01 NG 18 18 18 NG 18

Ref. Boopathy and Sievers, 1991 Bachmann et al., 1983 Yang and Moengangongo, 1987 Boopathy and Tilche, 1991 Grobicki, 1989 Boopathy and Tilche, 1992 Boopathy et al., 1988 Nachaiyasit, 1995 Nachaiyasit, 1995 Nachaiyasit, 1995 Fox and Venkatasubbiah, 1996 Barber and Stuckey, 1997

NG 40 7 NG 90 failed failed NG 53

NG 22 10 NG 135 NG 24

a LR = loading rate in kg COD/m3 d.bThe amount of time spent at each loading rate (d).cStart-up time quoted is the time required for reactor to reach steady state.dNG = data not given.

proximately 1.2 kg COD/m3 d (Henze and Harremoe s, 1983), however, successful start-up of a pilot scale ABR has been achieved at signicantly higher primary loading rates (Table 5, Boopathy and Tilche (1991)). Although Nachaiyasit (1995) originally noted adequate performance with an initial loading rate of 13 kg COD/m3 d, an accumulation of intermediate products caused reactor souring and eventual failure after two weeks of operation. A possible way to prevent failure by overloading was employed in 1980 by Chynoweth and coworkers. In order to stimulate the growth of methanogenic archea, pulses of methane precursors (acetate and/or an acetate/formate mixture) were added directly before raising loading rates, and these were eective in minimising the shock caused by a sudden increase in organic loading. Alternative methods to prevent failure include the adjustment of pH in the rst compartment (Grobicki, 1989). A recent study (Barber and Stuckey, 1997) has shown that maintaining an initially long detention time (80 h) which is reduced in a stepwise fashion during which time substrate concentration is kept constant, provides greater reactor stability and superior performance than a reactor started-up with a constant and low detention time coupled to a stepwise increase in substrate concentration. These ndings were linked to better solids accumulation, promotion of methanogenic populations, and faster recovery to hydraulic shocks in the reactor started at the longer retention time. Treatment applications This section reviews the performance of the baed reactor while treating a variety of wastewaters, in particular, low and high strength, low temperature, high inuent solids and sulphate containing waste. Tables 6 and 7 and Fig. 2 summarise the available literature. Low strength treatment. Various authors have treated low strength wastewaters eectively in the ABR, as shown in Table 8. Dilute wastewaters inherently provide a low mass transfer driving force between biomass and substrate, and subsequently biomass activities will be greatly reduced according

to Monod kinetics. As a result, treatment of low strength wastewaters has been found to encourage the dominance of scavenging bacteria such as Methanosaeta in the ABR (Polprasert et al., 1992). Hassouna and Stuckey (1998), have shown that no substantial change occurred in the population of acid producing bacteria down the length of a reactor treating dilute milk waste, indicating the lack of signicant population selection at low COD concentrations. It appears that biomass retention is enhanced signicantly due to lower gas production rates suggesting that low hydraulic retention times (6 4 2 h) are feasible during low strength treatment. Orozco (1988) noted decreasing overall gas production with increasing HRTs, and this implied possible biomass starvation in later compartments at longer retention times. Another important consequence of low retention times when treating dilute wastewaters is an increase in hydraulic turbulence, which can lower apparent Ks values (Kato et al., 1997) thus enhancing treatment eciency. Witthauer and Stuckey (1982) observed irregular COD removal in baed reactors run at low loading rates and long retention times when treating dilute synthetic greywater. These problems were associated with low sludge blankets (inoculum contained less than 3 g VSS/l) caused after long periods of biomass settling. Channels were formed within the low blankets and this resulted in low gas productivity in most of the sludge blanket except for around the channels. Hence, biogas mixing was greatly reduced and this resulted in minimal biomass/substrate mass transport. In contrast, anaerobic lters, operated under the same conditions, outperformed the baed reactors, even after their suspended biomass was ushed out in a hydraulic shock experiment. The authors recommended that when treating dilute wastewater, baed reactors should be started-up with higher biomass concentrations (than used in their study) in order to obtain a suciently high sludge blanket (and better gas mixing) in as short a time as possible.

Table 6. Performance data on baed reactor systemsa HRT (h) 360 288336 1200 4884 4.871 60 Temperature (8C) Ref. Chynoweth et al., 1980 Fannin et al., 1981, 1982

Substrate 600036,000

Volume (l) Chambers Biomass (g VSS/l) Inlet COD (mg COD/l) Loading rate (kg/m3/d) COD removal (%)

Undiluted sea kelp Diluted sea kelp

5.3

30 4.01 4.01

The anaerobic baed reactor: a review

08.5 4.11 and 7.21 18 2025 18 18 18 18

Carbohydrateprotein Synthetic greywater Carbohydrateprotein Diluted swine wastewater Molasses wastewater Sucrose Whisky distillery wastewater Carbohydrateprotein Carbohydrateprotein Molasses wastewater Molasses wastewater Swine manure Municipal wastewater Slaughterhouse wastewater Carbohydrateprotein Molasses wastewater Molasses wastewater Carbohydrateprotein Pharmaceutical wastewater Phenolic Glucose Carbohydrateprotein Domestic sewage/industrial waste Carbohydrateprotein Carbohydrateprotein Carbohydrateprotein 10 20 1.24.8 20 1.672.5 220 1.24.8 0.85 1.24.8 4.89.6 4.818 4075 >70 98, 93 3668 8394 7299 98 70 7583, 9397, 96 9098 5298 612 360 20 2080 138850 H138 360 4.815 2.526 180 24144 H140 20, 80 24 H24 12 2080 10.3 20, 20, 20 20 120

9.8 10 10 10 6.3 8 6.3 20 150 75 6.3 10 7.810.4 150 150 15 350 5.16 10 150 150 10 10 7982 6384 5593 75 98 8593 90 99 95 4988 70 6269 90 7590 67,20089,600 80,000 71007600 480 8000 <5000 500010,000 344500 51,600 4000 4000 115,771990,000 115,771990,000 58,500 264906 450550 4000 115,771990,000 115,771990,000 4000 20,000 22003192 100010,000 10004000 315c 4000 4000 4000

0.42.4 1.6 5.66.4 1.6 220 0.10.4 2.536 1.8 5.5 0.72 2.23.5 1.24.8b 1.24.8 4.328 20 4 2.17 0.94.73

6 10 394,000 10 10 10

5 4 4 4 5 6 5 3 11 5 8 48 3 3 23 3 4 48 3 3 8 5 5 5 8 8 8 8 8

35 35 35 35 35 2533 35 30 37 1316 30 35 35 37 37 35 1828 2530 35 37 37 35 35 21 35 35 15 15, 25, 35 35 35

Bachmann et al., 1983 Witthauer and Stuckey, 1982 Bachmann et al., 1985 Yang and Moengangongo, 1987 Yang et al., 1988 Orozco, 1988 Boopathy et al., 1988 Grobicki and Stuckey, 1989 Grobicki and Stuckey, 1991 Boopathy and Tilche, 1991 Boopathy and Tilche, 1992 Boopathy and Sievers, 1991 Garuti et al., 1992 Polprasert et al., 1992 Grobicki and Stuckey, 1992 Xing and Tilche, 1992 Xing et al., 1991 Nachaiyasit and Stuckey, 1995 Fox and Venkatasubbiah, 1996 Holt et al., 1997 Bae et al., 1997 Barber and Stuckey, 1998 Orozco, 1997 Nachaiyasit and Stuckey, 1997a Nachaiyasit and Stuckey, 1997b Nachaiyasit and Stuckey, 1997c

Contains calculated results, either from graphs or from supplied data.bAlso with shock loading of 96 kg/m3 d.cBOD5 value.

1565

1566

William P. Barber and David C. Stuckey


Table 7. Potential methane yields from baed reactors

Wastewater Swine manure Swine manure Carbohydrate/protein Carbohydrate/protein Sea kelp Molasses Phenol Slaughterhouse

OLR (kg/m3 d) 48 1.8 4.8 4.8 2.4 20 1.672.5 1.824.73

Methane yield (m3/kg VSS/d) 0.761.28 0.27 0.11 0.22 0.35 1.25 0.260.34 0.130.18

Ref. Boopathy and Sievers, 1991 Yang and Moengangongo, 1987 Nachaiyasit and Stuckey, 1995 Grobicki, 1989 Chynoweth et al., 1980 Boopathy and Tilche, 1991 Holt et al., 1997 Polprasert et al., 1992

High strength treatment. Whereas low retention times are possible and even necessary for dilute wastewaters, the opposite applies when treated concentrated waste. This is mainly due to the high gas mixing caused by improved mass transfer between the biomass and substrate. This will result in high biomass wastage, and has led to modications in the reactor design in order to enhance solids retention (see Section 2). A brief summary of the literature available on high strength treatment is shown in Table 9. When Boopathy and Tilche (1991) changed the inuent to a 150 l hybrid reactor from 115 g COD/l molasses alcohol stillage with a loading rate of 12.25 kg/m3 d to raw alcohol molasses (990 g COD/l, OLR = 28 kg COD/m3 d) they noticed an increase in overall gas production of over 65% within 3 weeks, a drop in COD removal of 20%, a fall in the methane composition of the biogas by 20% for one week which then recovered (implying initial overloading of

methanogens), and an approximate increase in volatile suspended solids of 50% within 3 weeks. Higher levels of gas production increased sludge bed expansion, but the improved settling ability of the biomass may have reduced the eects of solids loss caused by the gas (Boopathy and Tilche, 1991). This observation was partially conrmed in an earlier study (Boopathy et al., 1988) where no increase in solids loss or decrease in performance were noted when loading rates were increased from 2.6 to 3.5 kg COD/m3 d. However, minimal solids were lost to the euent at equally low loading rates in the work by Boopathy and Tilche (1991), but levels increased to 17 g VSS/l at higher loading rates. (The reactor contained approximately 1.25 g VSS/l of reactor in a 150 l volume.) According to kinetic considerations, high substrate concentrations will encourage both fast growing bacteria, and organisms with high Ks values, and methane production will be derived mainly from acetate decarboxylation by

Fig. 2. Performance eciency against various loading rates.

The anaerobic baed reactor: a review


Table 8. Selected low strength performance data Wastewater HRT (h) COD (mg/l) inuent euent Greywater Greywater Greywatera Sucroseb Sucroseb Sucroseb Slaughterhouse Slaughterhouse Slaughterhouse
a

1567

COD removal (%) OLR (kg/m3/d) Gas produced (v/v/d)

Ref.

84 48 84 6.8 8 11 26.4 7.2 2.5

438 492 445 473 473 441 730 550 510

109 143 72 74 66 33 80 110 130

75 71 84 74 86 93 89 80 75

0.13 0.25 0.13 1.67 1.42 0.96 0.67 1.82 4.73

0.025 0.05 0.025 0.49 0.43 0.31 0.72 0.33 0.43

Witthauer and Stuckey, 1982 Witthauer and Stuckey, 1982 Witthauer and Stuckey, 1982 Orozco, 1988 Orozco, 1988 Orozco, 1988 Polprasert et al., 1992 Polprasert et al., 1992 Polprasert et al., 1992

Temperature at 258C.bTemperatures lower than 168C. All other work shown in table performed in a mesophilic temperature range

Methanosarcina sp. and hydrogen scavenging methanogens (such as Methanobrevibacter and Methanobacterium). Subsequently Methanosarcina sp. was observed as the dominant bacterial species in bioocs formed during high strength treatment (Boopathy and Tilche, 1991). (See Section 5.2.) Low temperature treatment. At low/ambient temperatures van Lier et al. (1996), found signicant advantages with respect to reactor performance for staged reactors when compared with completely mixed systems. From Table 6 it can be seen that the vast majority of work done so far on the baed reactor has been conducted in the mesophilic temperature range. However, the baed reactor has been run as low as 138C (Orozco, 1988), although the most extensive study at low temperatures in the baed reactor was carried out by Nachaiyasit and Stuckey (1997a, Table 10). Generally, biochemical reactions double in relative activity for every 108C increase in temperature in accordance with the van `t Ho rule over a restricted temperature range. In spite of this, Nachaiyasit (1995), found no signicant reduction in overall COD removal eciency when the temperature of an ABR was dropped from 35 to 258C, with steady state reached after only two weeks. However, lower catabolic rates caused by elevated Ks values (according to Arrhenius kinetics) at the front of the reactor caused a shift in acid production towards the rear, although overall removal was unaected. An increase in VFA production caused a simultaneous reduction in pH and an initial increase in gas phase hydrogen that quickly returned to below background levels. The deeper penetration of the VFAs down the reactor should

potentially improve the growth yields of the methanogens in the latter compartments. The results showed that slower growing organisms exhibited a greater sensitivity to a fall in temperature compared to bacteria with faster growth kinetics, and this is in accordance with literature ndings (Cayless et al., 1989; Kotsyurbenko et al., 1993; Borja et al., 1994; Speece, 1996). Similar high treatment eciencies at ambient temperature have also been noted for a medium strength phenolic wastewater (Holt et al., 1997). Nachaiyasit and Stuckey (1997a) further reduced the temperature to 158C, and a fall in overall eciency of 20% was noted after one month. Changes in performance down the reactor occurred over a long period of time in contrast to CSTRs. This is advantageous since the slow response would inherently provide more protection to shocks than in other reactor systems. However, despite the fact that the reactors were kept for long periods of time at reduced temperatures (12 weeks) their performance did not improve despite the increased intermediate acid concentrations, which according to Monod kinetics should encourage more biomass growth to compensate for the increased substrate levels. This may be due to the fact that Ks increases substantially as temperature falls, (Lawrence and McCarty, 1969) leaving low levels of VFAs that cannot be degraded. This study also found that the fraction of VFAs in the euent in terms of COD had reduced signicantly. VFAs contributed to approximately a third of the COD at 158C, and two thirds at 258C, indicating that the production of refractory material (termed as Soluble Microbial Products (SMPs),

Table 9. Selected high strength treatment data Wastewater Inuent COD (g/l) HRT (h) Reactor volume (l) Temperature (8C) OLR (kg/m3 d) COD removal (%) Biogas production (v/v/d) Ref. Raw molasses 990 850 150 37 28 50 >5 Boopathy and Tilche, 1991 Molasses alcohol stillage 115.8 138636 150 37 4.320 7088 >2.3 Boopathy and Tilche, 1992 Swine waste 58.5 360 15 35 4 6269 2.93.2 Boopathy and Sievers, 1991 Whisky distillery 51 360 6.3 30 2.23.46 >90 1.23.6 Boopathy et al., 1988

1568

William P. Barber and David C. Stuckey


Table 10. Low temperature treatment

Temperature (8C) 35 25 15 1316


a

Inlet concentration (mg COD/l) 4000 4000 4000 500

Reactor volume (l) 10 10 10 75

COD removed (%) 96 9397 7583 8492

Biogas (v/v/d) 2.78 2.36 1.74 0.310.50a

Relative reaction rateb 1 0.676 0.391 0.391

Ref. Nachaiyasit and Stuckey, 1997a Nachaiyasit and Stuckey, 1997a Nachaiyasit and Stuckey, 1997a Orozco, 1988

Calculated from theoretical gas production based on COD removal.bReaction rate relative to that at 358C calculated from typical Q10 values for anaerobic processes (Sawyer et al., 1994).

Rittmann et al. (1987)) increased substantially at lower temperatures. In conclusion, the work found that a combination of decreased catabolic rates, increased Ks, and higher levels of refractory material caused inferior performance at 158C, but that a drop in temperature from 35 to 258C had negligible eects on overall reactor performance despite predictions from the van `t Ho rule. This has been observed before in biolm/oc based reactors where mass transfer limited biomass activity (Hickey et al., 1987). However, Nachaiyasit's work did not consider the eects of nutrient (especially iron) bioavailability, which may be reduced at lower temperatures (Speece, 1996), nor did it investigate the signicance of temperature on ionisation equilibria which inevitably controls the potential toxicity of materials, some of which may be tolerated at higher temperatures (Sawyer et al., 1994). High solids treatment. In early work, Chynoweth's group in Illinois (1980, 1981) used baed reactors to generate methane from sea kelp as an alternative energy source. Although the COD of the kelp was not quoted, the feed contained 15% total solids, which were ground and chopped. Practical problems associated with feeding solids were overcome by applying the substrate by syringe. During a particular run, signicant solids build-up was observed in the rst compartment after 2 weeks of operation. The solids build-up reduced microorganism contact with the substrate therefore minimising hydrolysis and subsequent bioconversion. Performance was signicantly improved after manually agitating the reactor for a short time period. Solid material was also found to physically displace biomass within the reactor indicating that modications to the ABR would be required for high solids treatment. In 1991, Boopathy and Sievers modied the baed reactor (see Section 2) to treat high strength swine waste (see Table 6) containing 51.7 g/l total solids. When a loading rate of 4 kg COD/m3 d with a retention time of 15 d was applied, removal rates for COD (70 and 80%), and total solids (60 and 74%) were achieved for two- and three-compartment reactors respectively. Solids retention times were experimentally determined to be over 20 d in both reactors. The study found that the majority of the protein fraction of the solids was retained within the reactor, compared with a lower retention

of cellulose/hemicellulose, and a virtual loss of all lipid material, although the authors oered no explanation to the cause. Previous work in the same laboratory had shown protein to be dicult to degrade but a great potential source of methane, hence its detention proved to be signicant in reactor performance. Sulphate treatment. Fox and Venkatasubbiah (1996), investigated the eects of sulphate reduction in the ABR by treating a sulphate containing pharmaceutical wastewater up to a nal strength of 20 g COD/l with a COD:SO4 ratio of 8:1. At steady state, 50% COD removal and 95% sulphate reduction was possible with a detention time of 1 day. Reactor proles showed that sulphate was almost completely reduced to sulphide within the rst chamber, and a concomitant increase in sulphide levels down the reactor indicated that sulphate redirected electron equivalents to hydrogen sulphide in preference to methane. After altering the COD:SO4 ratio by adding glucose, isopropanol and sulphate, the authors noted a fall in potential sulphate reduction from >95% at COD:SO4=150:1 to <50% at COD:SO4=24:1. Increasing sulphate concentrations with glucose and isopropanol present showed inhibition of sulphate reduction caused by elevated sulphide concentrations. Increasing the inlet concentration from 2 to 8 g COD/l (COD:SO4 at 8:1) over 100 d caused an increase in the total euent sulphide to toxic levels (200 or 80 mg/l unionised H2S assuming pH 7, pH data not supplied), with COD removal dropping to below 20%. VFA levels as high as 4500 mg/ l were observed during inhibition and these contributed to a maximum of 35% of the reactor euent COD. The major contributor to the euent VFA was acetate indicating inhibition of acetoclastic methanogenesis and a distinct lack of acetate cleaving sulphidogenesis. A recycle stream (recycle ratio 10:1) was employed to overcome sulphide inhibition of both sulphate reducing bacteria and methanogenic archea. The euent was oxidised in a trickling thin lm reactor in the presence of an enriched culture of Thiobacillus sp., which converted the sulphide to elemental sulphur. After employing recycle, total euent sulphide levels decreased to below 75 mg/l (or 30 mg/l unionised H2S after pH correction) with an increase in COD removal to 50%.

The anaerobic baed reactor: a review


Table 11. Bacterial observations in the ABR No. 1 2 3 Observations Methanosarcina predominant at front of reactor with Methanosaeta found towards rear active methanogenic fraction within biomass highest at front of reactor and lowest in last chamber bacteria resembling Propionibacterium, Syntrophobacter and Methanobrevibacter found in close proximity within granules Methanosaeta and colonies of Syntrophomonas also observed large numbers of Methanobacterium at front of ABR along with Methanosarcina covered granules; subsequent chambers consisted of Methanosaeta coated ocs virtually all biomass activity (>85%) occurred in the bottom third of each compartment where biomass was concentrated; highest activity (92%) found in bottom of rst chamber mainly Methanosaeta observed with some cocci; no Methanosarcina observed irregular granules with gas vents covered by single rod shaped bacteria; no predominant species observed bacteria resembling Methanobrevibacter, Methanococcus, and Desulfovibrio found wide variety of bacteria observed at front of reactor Technique SEM, TEM, LLM ATA TEM Ref.

1569

Boopathy and Tilche, 1991, 1992; Tilche and Yang, 1987; Garuti et al., 1992; Yang et al., 1988 Bachmann et al., 1985; Orozco, 1988 Grobicki, 1989

EP

Tilche and Yang, 1987

ATPA

Xing et al., 1991

6 7 8 9

SEM SEM ATPA, SEM, EP SEM, TEM

Polprasert et al., 1992 Holt et al., 1997 Boopathy and Tilche, 1992 Boopathy and Tilche, 1991; Barber and Stuckey, 1997 microscopy,

Abbreviations: ATA = anaerobic toxicity assay, ATPA = ATP analysis, EP = (phase contrast) epiuorescence LLM = light level microscopy, SEM = scanning electron microscopy, TEM = transmission electron microscopy. BIOMASS CHARACTERISTICS AND RETENTION CAPABILITIES

Bacterial populations With the unique construction of the ABR various proles of microbial communities may develop within each compartment. The microbial ecology within each reactor chamber will depend on the type and amount of substrate present, as well as external parameters such as pH and temperature. In the acidication zone of the ABR (front compartment(s) of reactor) fast growing bacteria capable of growth at high substrate levels and reduced pH will dominate. A shift to slower growing scavenging bacteria that grow better at higher pH will occur towards the end of the reactor. Various techniques have been applied to describe the population dynamics within the ABR, and the results are summarised in Table 11. By far the most common observation involved the shift in population of the two acetoclastic methanogens Methanosarcina sp. and Methanosaeta sp. At high acetate concentrations Methanosarcina outgrows Methanosaeta due to faster growth kinetics (doubling time 1.5 d compared with 4 d for Methanosaeta), however, at low concentrations Methanosaeta is dominant due to its scavenging capability (Ks=30 mg/l compared with 400 mg/l for Methanosarcina (Gujer and Zehnder, 1983)).

Tilche and Yang (1987) and Yang et al. (1988) compared the performance and bacterial populations of an anaerobic lter and a Hybridised Baed Reactor (HABR) at pilot scale treating molasses wastewater with maximum loading rates of 10.5 and 5.5 kg COD/m3 d for the anaerobic lter and HABR respectively. The major ndings of the study were: a large concentration of Methanosarcina at the front of the baed reactor which shifted to Methanosaeta towards the rear, compared with a domination of Methanosaeta in the lter reactor, and, hydrogen scavenging Methanobacterium were observed at the front of the baed reactor using epiuorescence microscopy. Explanations were oered to describe the lack of Methanosarcina in the lter reactor. Firstly, the acetate loading in the rst chamber of the HABR was 1000 mg/l which might be close to the apparent Ks value for Methanosarcina (data not given) and therefore may have favoured its growth. In contrast, acetate levels were 10 times lower in the lter reactor and therefore Methanosaeta had a kinetic advantage and dominated in the reactor. Secondly, lower supercial gas production rates in the baed reactor (5 m/d in the rst compartment of the HABR compared with 9 m/d in the lter) resulted in lower gas turbulence, and therefore fewer washouts of bioocs compared with the anaerobic lter. Hydrogen levels were also measured, and the high-

1570

William P. Barber and David C. Stuckey

est concentrations (4 104 atm) were noted in the rst chamber of the baed reactor, and this may explain the presence of Methanobacterium. The results were subsequently supported by Polprasert et al. (1992) where acetate concentrations as low as 20 mg/l enabled the domination of Methanosaetalike bacteria throughout a four-compartment reactor. Biomass activity Tilche and Yang (1987) and Yang et al. (1988) also discovered that 70% of all methane produced in the HABR came from the rst compartment, despite having only 10% of the VSS present within the reactor, and these ndings supported previous work (Bachmann et al., 1985; Orozco, 1988). Bachmann used a procedure based on the Anaerobic Toxicity Assay (ATA, Owen et al. (1979)) and discovered that the active fraction of acetate utilising methanogens as a percentage of the total VSS varied from 5.7 to 1.8%, with the largest values obtained at the front of the reactor and the lowest at the rear. In a study involving an 11-compartment open top baed reactor treating 500 mg/l sucrose at low temperature (13168C), Orozco (1988) quoted activities of 1.43 g COD-CH4/m3 in the rst seven chambers and 0.72 in chambers 7 to 11. Xing et al. (1991), and Boopathy and Tilche (1992) used ATP analysis to determine the relative position of the most active bacteria. Samples were taken from the top, middle and bottom of all three chambers from a reactor with a working volume of 150 l treating molasses wastewater at a loading rate of 20 kg COD/m3 d. The results showed that at least 85% of the activity came from the bottom of each compartment with the highest activity (92%) measured at the base of the rst compartment. However, the opposite trend was found in a study treating slaughterhouse wastewater (Polprasert et al., 1992). The reasons for this may lie in the concentration of intermediates, especially acetate, at the front of the reactor. In studies where methane activity was higher in the front compartments (Bachmann et al., 1985; Tilche and Yang, 1987; and Yang et al., 1988), acetate concentrations were relatively high and therefore provided the best environmental conditions for the growth of Methanosarcina which can grow quickly and eciently even at pH values as low as 6 (Speece, 1996). Another source of methane would be from hydrogen scavenging bacteria such as Methanobacterium (Tilche and Yang, 1987) and Methanobrevibacter, which would be stimulated by the higher hydrogen concentrations; the net eect would be a high methanogenic activity. In contrast, with dilute wastewaters, where acetate levels are low in the front compartment (as in the work by Polprasertet al), the likely scenario is that Methanosaeta would dominate. However, this species grows at a far slower rate

compared to Methanosarcina and is also far more sensitive to environmental conditions such as a reduced pH. This would encourage the growth of acid producing bacteria that would inevitably lead to a reduction in methane potential. Hassouna and Stuckey (1998) showed a shift in the activity of acid producing bacteria down the length of an eight-compartment baed reactor. Using the method of Owen et al. (1979), aliquots were removed from each compartment of ABRs treating a range of substrate concentrations. In the foremost compartments a glucose spike was readily converted to volatile acids within a few hours and this contrasted with the results from subsequent compartments which showed virtually no degradation of the spike. Granulation (and oc sizes) Although granulation is not necessary in the ABR for optimum performance, unlike suspended systems such as the UASB, various reports have noted the appearance of granules in the reactor. Boopathy and Tilche (1991) started up HABRs (the inoculum contained 4.01 g VSS/l) with a low initial loading rate (0.97 kg COD/kg VSS d) and liquid upow velocities below 0.46 m/h, in order to encourage the growth of occulent and granular biomass. Subsequently, stable granules of 0.5 mm appeared after one month in all chambers of the reactor and they were reported to be growing although no data was given; microscope studies subsequently showed that the granules were comprised primarily of acetoclastic methanogens. Similarly, Tilche and Yang (1987) found Methanosarcina coated ocs held together by brous bacteria resembling Methanosaeta. The ocs, which were formed after one month, were small with diameters less than 1.5 mm and were weak. Under the same loading conditions the authors also found densely packed granules typical of a UASB (d < 3 mm) formed after 3 months in an anaerobic lter. Boopathy and Tilche (1992) noticed similar particles of both types described above, which grew from 0.5 mm after one month to 3.5 mm after three months in a hybrid reactor. Granules, which were made from Methanosarcina clusters, were of low density and full of gas cavities and therefore lifted to the surface of the reactor due to high gas and liquid velocities during high loading. The particle size appeared to be partially dependent on substrate type. There was little dierence in particle size throughout the reactor when molasses alcohol stillage wastewater was treated. However, two weeks after the substrate was altered to raw molasses with a ten-fold increase in inlet COD a prole emerged which showed a steady decrease in particle size down the reactor. In addition, the sludge weight increased from <600 to 900 g in the rst compartment within the same time period (Xing et al., 1991). Orozco (1988) reported a similar decrease in

The anaerobic baed reactor: a review

1571

granule size from 5.4 mm in the rst chamber down to 1.5 mm in the last chamber of a reactor treating dilute carbohydrate waste. However, on a laboratory scale, (Barber and Stuckey, 1997) oc size seemed to grow to a maximum near the centre of an eight-compartment reactor and then reduce towards the rear. Typical oc sizes were 100, 230 and 175 mm in the front, middle and rear compartments respectively. These authors postulated that the oc size was a function of both gas production and COD concentration, with the largest particles growing when COD concentrations were suciently high to support growth, and gas production low enough to avoid oc breakage. Solids retention capability By using a chromic oxide sesqui tracer in a high solids swine wastewater (51 g/l), Boopathy and Sievers (1991) managed to measure the solids retention time for two hybrid reactors running at a hydraulic retention time of 15 d. A three-compartment reactor resulted in a solids retention time of 25 d compared with 22 d for a two-compartment unit. The two-compartment reactor had a larger initial chamber, and this provided a natural ltering action that enabled it to lose fewer solids to the euent. Despite this, the three-compartment reactor was found to be more ecient at converting the trapped material into methane on the basis of cellulose, lipid and protein measurements. In a comparative study, Orozco (1988) calculated the minimum solids retention time required to achieve certain removal eciencies in baed and UASB reactors under the same loading conditions, and concluded that the solid residence time in the UASB would have to be approximately 40% higher than the ABR in order to achieve the same removal rate. By assuming a series of perfectly mixed reactors, Grobicki and Stuckey (1991), calculated the solids retention times, biomass yield, and washout of biomass under several experimental conditions. Solids retention times varied from 7 to over 700 d (5 < 80 h) and large deviations in the results were attributed to varying degrees of granulation. Although a strong correlation was found to exist between the solids retention time and HRT, the

authors suggested that caution should be exercised when using the calculated gures due to the assumptions of perfectly mixed behaviour. Solids retention times of 300 d were reported by Garuti et al. (1992) using a 350 l reactor with a 15 h retention time and this gure is far higher than those calculated by Grobicki and Stuckey (1991) under similar conditions. These authors also calculated from theory and a mass balance, that the observed yields were very low (approximately 0.03 kg VSS/kg COD), which implies constant biomass concentration proles over time, but these ndings are in contrast to other researchers (Boopathy and Tilche, 1991; Xing et al, 1991). Boopathy et al. (1988) discovered that increasing the loading rate from 2.2 to 3.5 kg COD/m3 d made no signicant dierence to the amount of solids lost to the euent, with a maximum of 0.5 g/l occurring during start-up. These results were further supported in a hybrid reactor (Boopathy and Tilche, 1991) where virtually negligible euent VSS was found with loading rates between 6 and 12.5 kg COD/m3 d. However, a linear increase up to 17 g VSS/l at high loading rates (28 kg COD/m3 d) was observed. A similar correlation was also found to exist between the Sludge Volume Index (SVI) and the total solids lost from a pilot scale reactor (Garuti et al., 1992). Finally, in a recent study, Barber and Stuckey (1997) found that twice as many solids were lost during start-up by a reactor running at a low HRT of 20 h compared with one which was run on the same feed at long retention times (80 4 40 4 20 h), and this was linked to inferior COD removal since biomass accumulated faster in the reactor run at longer retention times.

MODELLING

Bachmann et al. (1983) found similar treatment behaviour under identical conditions in an ABR, anaerobic lter and a rotating biological disc reactor. In order to predict reactor performance, an attempt was made to develop a unied model for the xed lm reactors and also for the ABR. The authors considered the sludge particles found within the sludge bed of the ABR to be uidised spheres

Table 12. Model equations for ABR systems No. 1 2 3a 3b 4 Substrate model equations dS/dt = aCS +QS0QS, S = S0(a/Q)CS Df(d2Sf/dz2) = (kSfXf)/(Ks+Sf) Sn=S0/[(1 + k1W1/Q)(1 + k2W2/Q) F F F (1 + knWn/Q)] Sn=[S0(1 + R)n 1]/[(1 + R + k1W1/Q)(1 + R + k2W2/Q) F F F (1 + R + knWn/Q) (1 + R)n 1R] Df[(d2Sf/dr2) + (2/r)(dSf/dr)] = (kXfSf)/(Ks+Sf)
q q

Ref. Bachmann et al., 1983 Bachmann et al., 1985 Xing et al., 1991 Xing et al., 1991 Nachaiyasit, 1995

Nomenclature: a = surface area per unit reactor volume (L1), C = variable-order reaction coecient, Df=molecular diusivity in biolm (L2t1), k = maximum specic rate of substrate utilisation (MsMxt1), Ks=half-velocity constant (ML3), Q = specic ow rate (T1), q = variable-order reaction order, r = radius of a three-dimensional spherical particle (L), R = recycle ratio, S = substrate concentration (ML3), S0=inuent concentration (ML3), Sf=substrate concentration in biolm (ML3), Sn=euent substrate concentration (ML3), W = mass of sludge = volume/[Xf] (M), Xf=bacterial density (ML3), z = distance normal to biolm surface (L). Numerical subscripts refer to compartment number.

1572

William P. Barber and David C. Stuckey

with a surface area through which the solute must diuse for bacterial consumption. Therefore, they used a combination of a xed lm model (Williamson and McCarty, 1976) along with a variable order model (Rittmann and McCarty, 1978) which incorporated the concepts of liquid-layer mass transfer, Monod kinetics, and molecular diusion to accurately describe the process (Table 12). Two dierent approaches were employed; the rst was based on the concept of a rate limiting substrate (assumed to be acetate and propionate) diffusing into a ``deep'' xed bacterial lm. Application of the model was made possible by estimating the specic surface area in each of the reactor chambers from a data set, and then applying the results to simulate behaviour at dierent loadings. Although initial predictions were good, the model underestimated the level of COD removal at higher loading rates. The reasons for the discrepancies were given to be an unrealistic assumption of a constant diusion layer depth which would decrease at higher loading rates due to increased gas production, thereby improving substrate/biomass contact and ultimately reactor performance. The second evaluation was made by assuming a series of completely mixed dispersed growth reactors using Monod kinetics. Here values of the active micro-organism concentration were determined within each compartment for one loading, and the data applied to the same loading rates as with the ``xed lm'' model. The results of the second model termed ``the dispersed growth model'', did not give a realistic interpretation of the data since diusional limitations were not considered. Further work using the xed lm model was carried out by Bachmann et all. (1985) on baed reactors with an inuent substrate concentration of 8 g COD/l. The model predicted the following behaviour: a decrease in treatment eciency with (a) decreasing inuent substrate concentration at constant loading rates, (b) an increase in organic loading at constant inuent substrate concentration, and (c) an increase in recycle ratio at constant HRT since the reactor approaches completely mixed behaviour. Reactor eciency improved with reducing substrate concentration at constant HRT. Some of these ndings were mirrored in the work of Xing and Tilche (1992) on the modelling of a hybridised form of the baed reactor which had a working volume of 150 l, and treated 20 kg COD/ m3 d of molasses wastewater. The model focused on the ndings of ATP testing which concluded that virtually all of the active biomass was held within the base of each compartment, so the biomass weight and not concentration was used in the model. The main assumptions of the model were: all substrate consumption occurred within a granular sludge bed, and, the sludge bed was perfectly mixed due to gas evolution. The following predictions were made from the model: at constant or-

ganic loading the treatment eciency increased with increasing inuent substrate concentration; as HRT was reduced the performance of the reactor decreased; performance deteriorated with increasing loading (1116 kg COD/m3 d) with a constant sludge weight; an improvement in COD removal eciency was observed with increasing sludge weight until a certain concentration was reached, above which reactor performance becomes independent of biomass concentration; and nally an increase in recycle ratio coincided with a subsequent decrease in COD removal. Bachmann et al. (1983, 1985) assumed that the oc diameter was very large relative to the active biolm depth. However, this seems to be an unwarranted assumption since in anaerobic biolms the electron donor and acceptor are often the same organic, and since the three main microbial groups are symbiotic, oc particles may have active cores even with 3 mm diameters. In addition, many lm supports are not perfectly at, but can be considered suciently at if the biologically active thickness of the biolm is less than about 1% of the radius of curvature (i.e. the radius of a sphere plus a diusion layer, Rittmann and McCarty (1978)). At high loading rates this is not the case in the ABR, since sludge particles within the reactor act as uidised spheres with a surface area through which the substrate must diuse for consumption (Bachmann et al., 1985). These facts imply that a spherical model would provide a better t than a simple planar one. Nachaiyasit (1995) derived a spherical model using Monod kinetics combined with molecular diffusion of the substrate into the biomass aggregates based on the assumptions that: (i) substrate concentration could be described by a single parameter, COD, (ii) biomass concentration can be adequately described by a single parameter, VSS, (iii) the biomass composition is constant during balanced growth and (iv) the biological reactions of importance occur at constant temperature and pH. The calculation of important model parameters such as diusion layer thickness, and liquid phase mass transfer coecient followed the techniques proposed by Bachmann et al. (1985). In general, the model predicted better COD removal than was experimentally measured, and was most accurate for high loading rates (8 and 15 g COD/l at 20 h) than at short retention times (10, 5 h HRT with feed concentration of 4 g COD/l), but showed large deviations for the rst couple of compartments for some of the simulations. As with previous models certain trends appeared with the results, namely a decrease in removal eciency with increasing recycle ratios, decreasing HRTs (with xed substrate concentration), and increasing substrate concentrations (with xed HRT). However, the spherical model did provide a closer t than the earlier planar xed lm equations. Based on the ndings of a

The anaerobic baed reactor: a review

1573

sensitivity analysis that showed oc surface area and owrate had the greatest inuence on model predictions, the model was modied by making the surface area a tting parameter. Nachaiyasit then compared the results obtained from dierent models based on the same assumptions and with the same experimental data, and found the closest t with the spherical model. It was concluded that while the predictive capacity of the spherical model was not always good, it was useful as a tool for understanding the interaction between the various system parameters, and therefore could be used as a basis for the development of better predictive models. It seems that a combination of theoretical considerations and experimental ndings can be used together in order to generate models with a more realistic t. Since the accuracy of any model depends critically on the wastewater and substrate used, kinetic data should be experimentally determined for each compartment once the reactor is at steady state (Bachmann et al., 1985) by using simple bioassays. Such an approach may have enabled Nachaiyasit's spherical model to give a more realistic t at the front of the reactor. All modelling so far has used acetoclastic methane production as the rate-limiting step. However, it is evident from Section 5, that the structure of the reactor will cause a shift in the population dynamics of the two species (Methanosarcina and Methanosaeta) responsible for acetate consumption. Since both archea dier widely in kinetic ability, acetate loadings and pH will have an eect on reactor performance in each compartment. For reactors treating medium to high strength wastes acetate consumption at the front of the reactor will be higher than for a low strength waste. This will result in most of the COD being removed in the front of the reactor. For low strength wastes, acetate loadings will be low and this will encourage growth of Methanosaeta with little COD removed in the acidication zone. Channelling has been shown to be an important phenomenon in the ABR (see Section 3) and will aect the accuracy of any model. In order to take channelling into account it is necessary to calculate the number N of ideally mixed reactors in series using tracer studies (see Section 3.1). The results of these experiments could then be input as the number of real compartments into a reactor in series model. Also, correlations are available which show the eect of hydrodynamic dispersion on the substrate diusion coecient (Bear, 1972). Furthermore, by calculating the minimum solids retention time (Orozco, 1988), it should be possible to determine the correct compartment size for a given treatment eciency. On the basis of the literature it seems that for most cases only 24 compartments are necessary for adequate COD removal. However, reactors with more compartments will be far more resistant to hydraulic and organic shocks, since they will protect against the shift in acid pro-

duction towards the rear. Therefore a compromise will exist between optimal (required) compartment number, ``safe'' compartment number, and also upow liquid velocity. Despite the less than perfect predictive capabilities of the models described above, there is an urgent need to generate models for larger scale reactors. Boopathy and Tilche (1991) pointed out that at larger scale a greater evolution of gas per compartment cross sectional area can be expected, and this would cause an increase in mixing which would subsequently improve mass transfer rates leading to greater eciencies, but perhaps increased solids loss. Finally, it is also necessary to model reactor behaviour when hydrolysis is the rate-limiting step, as is the case with high solid inuents and lipid containing wastewaters, since by assuming rst-order rate kinetics it is possible to calculate the minimum solids retention time to achieve a given eciency (Pavlostathis and Giraldo-Gomez, 1991). With wastewaters containing a large amount of particulate material, it seems likely that COD removal will be low at the front end, and that the VFA prole formed will be shifted down the reactor, unless the reactor is modied, (Boopathy and Sievers, 1991) or extra compartments added, a drop of eciency may result.
FULL-SCALE EXPERIENCE

The performance data of a full-scale plant, treating domestic waste from a small town in Columbia (Tenjo, population <2500 inhabitants, Orozco (1997)), is presented in Table 13. The Tenjo reactors were designed to give a liquid upow velocity of 3 m/h based on laboratory ndings. However, despite following strict guidelines for start-up, the
Table 13. Full-scale data for an anaerobically operated baed reactor Tenjo, Colombia Wastewater 1 Composition 2 Strength (g BOD5/l) 3 Total solids (g/l) 4 Volatile solids (g/l) Performance 1 OLR (kg/m3d) 2 Removal eciency (%) 3 HRT (h) 4 Euent BOD5 (g BOD5/l) Reactor design 1 Reactor conguration 2 Reactor number 3 Reactor volume (m3) 4 Reactor dimensionsb (m) 5 Compartment number 6 Liquid upow velocity (m/h) 7 Packing material 8 Settling chamber Economics 1 See text Miscellaneous 1 Temperature (8C)
a

domestic/industrial mixa 0.314 0.90 0.25 0.85 H70 10.3 H0.1 open top reactor 2 394 (197 each) 2.7:17:4.3 8 3.00 plastic boxes internal gas/solid separation

15

Industrial dairy waste.bReactor dimensions: height:length:width.

1574

William P. Barber and David C. Stuckey


Table 14. Treatment eciencies for various reactor congurations

Feedstock Carbohydrate

Reactor type ABR ABR ABR UASB UASB AF USSB AAFEB ABR UASB UASB UASB UASB AF HABR HABR HABR UASB AF ABR AF ABR UASB ABR UASB UASB AF ABR 2-phase UASB AF

Reactor volume (l) 6.3 6 75 4.8 0.4a 4.2 H0.4 5.16 30 2 2 150 150 150 H85 125 8 8a 15 7 69,000 10 2.7 (2.5)e 5.75 1

Inlet COD (g/l) 7.1 110 0.440.47 110 0.490.55 8 6080b 0.050.60 0.480.73 0.73 1.502.20 1.502.20 8 8 115990 115990 510 100 510 0.48 0.48 58.5 5.5 2.23.2 0.9c 5.2 H1.8 20 45.2 0.72 49.8

Loading rate (kg/m3 d) 1 220 0.961.66 220 12.2 1 75 0.84.8 0.94.7 2.7 7 610 16.5 16.5 20 28 5.5 24 10.5 0.4 0.4 48 2.83 1.672.5 2 5.67 20 H5 1118.6

COD removal (%) 79 7299 8493 <5097 7786 92 7589 4095 75 77 85 8791 90 <90 77 50 98 75 98 6384 6489 6269 6080 8394 98 90 54 50 85 9095 2936

Ref. Bachmann et al., 1983 Bae et al., 1997 Orozco, 1988 Bae et al., 1997 Orozco, 1988 Bachmann et al., 1983 van Lier et al., 1996 Switzenbaum and Jewell, 1980 Polprasert et al., 1992 Zheng and Wu, 1985 Sayed et al., 1987 Lettinga et al., 1982 Ruiz et al., 1997 Ruiz et al., 1997 Boopathy and Tilche, 1991 Boopathy and Tilche, 1992 Tilche and Yang, 1987 Sanchez Riera et al., 1985 Tilche and Yang, 1987 Witthauer and Stuckey, 1982 Witthauer and Stuckey, 1982 Boopathy and Sievers, 1991 Cintoli et al., 1995 Holt et al., 1997 Zhou and Fang, 1997 Chang et al., 1995 Kanekar et al., 1996 Fox and Venkatasubbiah, 1996 Reis et al., 1995 Visser et al., 1992 Hilton and Archer, 1988

Slaughterhouse

Molasses

Greywater Piggery Phenol

Sulphate

a Liquid volume.bSucrose with VFA mixture, thermophilic treatment.cPhenol concentration.dCOD:SO4 ratio 8:1 for ABR, 10:1 for 2phase, 2:1 4 0.5:1 for UASB (thermophilic), 8:1 4 4:1 for AF.eAcidogenic stage made up of two reactors with a total volume of 2.7 l. The number in parentheses refers to a single methanogenic stage.

reactors experienced several practical problems during early operation. Hydraulic shocks increased solids washout, and poor screening of solid material caused the plastic packing media to oat with gas production. These problems were overcome by using a by-pass pipe during the rainy season and improved screening facilities. The reactors performed well with approximately 70% COD reduction and 80% removal of suspended solids over a two-month period. Varying the volumetric load between 0.4 and 2 kg/m3 d had no eect on removal eciency. However, the author concluded that a polishing lagoon was necessary to achieve discharge quality euent. Work is currently being undertaken to provide a wastewater treatment plant for a larger town. Although a detailed economic study was not presented, construction costs for the baed reactor were 20% less than those for UASB reactors in Columbia running at ambient temperature, and ve times less than a conventional activated sludge plant for a small town.
FUTURE PROSPECTS FOR THE ABR

toxic materials due to its inherent two-phase behaviour. Despite comparable performance with other well established technologies (Table 14), its future use will depend on exploiting its structure in order to treat wastewaters which cannot be readily treated. Outlined below is a list of possible processes that are feasible in the ABR. In situ aerobic polishing Unpublished work in this laboratory has shown that an aerobic polishing step can be inserted within an ABR with no detrimental eect on reactor performance. This is due to the fact that ``aerophobic'' methanogens can remain active even when oxygen is present, and whilst inside immobilised aggregates methanogenic archea are well shielded from oxygen by layers of facultative bacteria (Lettinga et al., 1997). Also, processes which inherently require both anaerobic and aerobic treatment (or detoxication) can be dealt with within a single reactor unit, such as black hemp liquors, wood extractives, coal processing industry, petrochemical, and textile dye wastewaters (Lettinga, 1995) thus signicantly reducing capital costs. Total nitrogen removal. Work is currently being undertaken to treat ammonia containing wastewaters with an anaerobic/aerobic baed reactor for total nitrogen removal. Ammonia present in the wastewater passes through the anaerobic compart-

The ABR shows promise for industrial wastewater treatment since it can withstand severe hydraulic and organic shock loads, intermittent feeding, temperature changes, and tolerate certain

The anaerobic baed reactor: a review


Table 15. Recommendations based on literature ndings Recommendations Start-up

1575

Recycle Low strength wastewater High strength wastewater High solids wastewater Temperature

low initial loading rates will encourage granule/oc growth pulses of methane precursors (e.g. acetate) have been successfully used to encourage methanogenic growth and dampen the eects of increases in loading rate start-up with long retention times reduces solids loss due to low liquid upow velocities and, promotes higher methanogen populations in every compartment recycle is benecial with respect to diluting toxicants in feed stream, increasing front pH and reducing production of foam and SMPs, but has several disadvantages outlined in Table 4 low retention time enables better mass transport due to improved hydraulic mixing and reduces biomass starvation in latter compartments methane production will originate from scavenging bacteria (Methanosaeta) long retention times reduce solids washout caused by high gas production, otherwise the reactor may be modied (by adding packing) to decrease biomass loss methane production will be mainly due to Methanosarcina, and hydrogen scavenging methanogens a larger front compartment has proved to be eective in treating wastewater with a high solids content reducing temperature to 258C from 358C has no eect on easily degradable waste, further decreases in temperature are detrimental on reactor performance, this may be due to potential toxicity, nutrient bioavailability and slower kinetic rates reactors started-up and kept at lower temperatures perform consistently well

ments largely unmetabolised, and is then oxidised to form nitrates and nitrites at the rear of the reactor. These can then be recycled to the anaerobic section where they act as alternative electron acceptors and are reduced to nitrogen. Complete sulphur removal. Sulphate is reduced at higher redox potentials than that at which methanogenesis begins (Henze and Harremoe s, 1983), and will therefore be converted to hydrogen sulphide at the front of a baed reactor at the expense of methanogenesis (Fox and Venkatasubbiah, 1996). Micro-aerobic polishing could be achieved within an aerobic stage to produce elemental sulphur, which could be recovered eliminating the need of a separate trickling lter unit. Enhancement of two-phase properties (better pH and temperature control) The optimum pH for a two-phase system has been widely quoted to be approximately 5 (Ghosh et al., 1975; Aivasidis et al., 1988; Speece, 1996). This implies that less buering would be required in a baed reactor since the pH is routinely above 6 in the rst compartment. Alternatively, buering and/or nutrients could be added separately in latter compartments to provide optimal conditions for scavenging methanogens.
CONCLUSIONS AND RECOMMENDATIONS

able resistance to toxic materials, shields syntrophic bacteria from elevated hydrogen levels, and results in high removal eciencies even at low hydraulic retention times (26 h). The physical structure of the reactor allows various modications to be made, such as an in situ aerobic polishing stage, resulting in providing the capability to treat wastewaters that currently require at least two separate units, therefore substantially reducing capital costs. However, in order to enhance the commercial potential of the ABR, more work still remains to be done in the following areas: modelling the fate of SMPs, solids, intermediate products, and COD removal; nutrient requirements; treatment of toxic wastewaters (e.g. polychlorinated aliphatics, nitrated organics, xenobiotics, haloaromatics, surfactants) which have been treated with success anaerobically; and an improved understanding of the factors controlling bacterial ecology. Finally, Table 15 shows a list of recommendations based on this review of the literature.

Acknowledgement The authors would like to thank Professor Chynoweth for his generous help with providing material for this review and the BBSRC for nancial support.
REFERENCES

Laboratory, pilot and full-scale work has shown that the ABR is capable of treating a variety of wastewaters of varying strength (0.45 < 1000 g/l), over a large range of loading rates (0.4 < 28 kg/ m3 d), and with high solids concentrations with satisfactory results (Table 6). Long biomass retention times are possible without granulation and solids/ liquid separation devices, and a selective pressure exists which enhances the development of appropriate bacterial populations in various parts of the reactor. This reactor conguration confers consider-

Aivasidis A., Bastin K. H. and Wandrey C. (1988) Optimisation of selection stress in a chemostat. Anaerobic Digestion, IAWPRC, 3546. Alves M. M., Pereira M. A., Mota M., Novais J. M. and Colleran E. (1997) Staged and non staged anaerobic lters: microbial selection, hydrodynamic aspects and performance. Proceedings of the 8th International Conference on Anaerobic Digestion, Vol. 2, Sendai, Japan, pp. 5663. Bachmann A., Beard V. L. and McCarty P. L. (1983) Comparison of Fixed Film Reactors with a Modied Sludge Blanket Reactor, Fixed Film Biological Processes for Wastewater Treatment, ed. Y. C. Wu and E. D. Smith. Noyes Data, NJ.

1576

William P. Barber and David C. Stuckey aerobic digester system development. Annual Report for General Electric Company, IGT Project 65044, Institute of Gas Technology, IIT Centre, 3424 S. State Street, Chicago, IL 60616. Fannin K. F., Srivastra V. J., Mensinger J., Conrad J. R. and Chynoweth D. P. (1982) Marine biomass program: anaerobic digester process development. Annual Report for General Electric Company, IGT Project 65044 and 30547, Institute of Gas Technology, IIT Centre, 3424 S. State Street, Chicago, IL 60616. Fox P. Venkatasubbiah V. (1996) Coupled anaerobic/ aerobic treatment of high-sulphate wastewater with sulphate reduction and biological sulphide oxidation. Wat. Sci. Technol. 34(56), 359366. Garuti G., Dohanyos M. and Tilche A. (1992) Anaerobic aerobic combined process for the treatment of sewage with nutrient removal: the Ananox1 process. Wat. Sci. Technol. 25(7), 383394. Ghosh S., Conrad J. R. and Klass D. L. (1975) Anaerobic acidogenesis of sewage sludge. J. WPCF 47, 3045. Grobicki A. M. W. (1989) Hydrodynamic characteristics and performance of the anaerobic baed reactor. Ph.D. dissertation, Department of Chemical Engineering, Imperial College, London, U.K. Grobicki A. M. W. and Stuckey D. C. (1989) The role of formate in the anaerobic baed reactor. Wat. Res. 23(12), 15991602. Grobicki A. M. W. and Stuckey D. C. (1991) Performance of the anaerobic baed reactor under steady state and shock loading conditions. Biotechnol. Bioeng. 37, 344 355. Grobicki A. M. W. and Stuckey D. C. (1992) Hydrodynamic characteristics of the anaerobic baed reactor. Wat. Res. 26, 371378. Guiot S. R., Sa B., Frigon J. C., Mercier P., Mulligan C., Tremblay R. and Samson R. (1995) Performances of a full-Scale novel multiplate anaerobic reactor treating cheese whey euent. Biotechnol. Bioeng. 45, 398495. Gujer W. and Zehnder A. J. B. (1983) Conversion processes in anaerobic digestion. Wat. Sci. Tech. 15, 127 167. Habets L. (1996) Overview of industrial anaerobic waste water treatment. Industrial Anaerobic WasteWater Treatment Conference, 18th September, SCI, London. Hassouna S. and Stuckey D. C. (1998), in preparation. Henze M. and Harremoe s P. (1983) Anaerobic treatment of wastewater in xed lm reactors: a literature review. Wat. Sci. Technol. 15(8/9), 1101. Hickey R. F., Wu W.-M., Viega M. C. and Jones R. (1991) Start-up, operation, monitoring and control of high-rate anaerobic treatment systems. Wat. Sci. Technol. 24, 207255. Hickey R. F., Vanderwielen J. and Switzenbaum M. S. (1987) Eects of organic toxicants on methane production and hydrogen gas levels during the anaerobic digestion of waste activated sludge. Wat. Res. 21, 1417 1427. Hilton M. G. and Archer D. B. (1988) Anaerobic digestion of a sulphate-rich molasses wastewater: inhibition of hydrogen sulphide production. Biotechnol. Bioeng. 31, 885888. Holt C. J., Matthew R. G. S. and Terzis E. (1997) A comparative study using the anaerobic baed reactor to treat a phenolic wastewater. Proceedings of the 8th International Conference on Anaerobic Digestion, Vol. 2, Sendai, Japan, pp. 4047. Iza J., Colleran E., Paris J. M. and Wu W.-M. (1991) International workshop on anaerobic treatment technology for municipal and industrial wastewaters: summary paper. Wat. Sci. Technol. 24(8), 116. Kanekar P., Sarnaik S. and Kelkar A. (1996) Microbial technology for management of phenol bearing dyestu wastewater. Wat. Sci. Technol. 33(8), 4751.

Bachmann A., Beard V. L. and McCarty P. L. (1985) Performance characteristics of the anaerobic baed reactor. Wat. Res. 19(1), 99106. Bae J.-H., Song K.-B. and Cho K.-M. (1997) Comparison of operational characteristics of UASB and ABR: organic removal eciency and the variations of PH2 and PCO. Proceedings of the 8th International Conference on Anaerobic Digestion, Vol. 1, Sendai, Japan, pp. 164 171. Barber W. P. and Stuckey D. C. (1997) Start-up strategies for anaerobic baed reactors treating a synthetic sucrose feed. Proceedings of the 8th International Conference on Anaerobic Digestion, Vol. 2, Sendai, Japan, pp. 3239. Barber W. P. and Stuckey D. C. (1998) Inuence of startup strategies on the performance of an anaerobic baed reactor. Environ. Technol. 19, 489501. Bear J. (1972) Dynamics of Fluids in Porous Media. Elsevier, New York. Boopathy R. and Sievers D. M. (1991) Performance of a modied anaerobic baed reactor to treat swine waste. Trans. ASAE 34(6), 25732578. Boopathy R. and Tilche A. (1991) Anaerobic-digestion of high-strength molasses waste-water using a hybrid anaerobic baed reactor. Wat. Res. 25(7), 785790. Boopathy R. and Tilche A. (1992) Pelletization of biomass in a hybrid anaerobic baed reactor (HABR) treating acidied waste-water. Bioresource Technol. 40(2), 101 107. Boopathy R., Larsen V. F. and Senior E. (1988) Performance of anaerobic baed reactor (ABR) in treating distillery waste-water from a Scotch Whisky factory. Biomass 16(2), 133143. Borja R., Banks C. J. and Wang Z. (1994) Stability and performance of an anaerobic downow lter treating slaughterhouse wastewater under transient changes in process parameters. Biotechnol. Appl. Biochem. 20, 371 383. Cayless S. M., da Motta Marques D. M. L. and Lester J. N. (1989) The eect of transient loading, pH and temperature shocks on anaerobic lters and uidised beds Environ. Technol. Lett. 10(11) 951968. Chang Y. J., Nishio N. and Nagai S. (1995) Characteristics of granular methanogenic sludge growth on phenol synthetic medium and methanogenic fermentation of phenolic wastewater in a UASB reactor. J. Ferm. Bioeng. 79(4), 348353. Chynoweth D. P., Srivastra V. J. and Conrad J. R. (1980) Research study to determine the feasibility of producing methane gas from sea kelp. Annual Report for General Electric Company, IGT Project 30502, Institute of Gas Technology, IIT Centre, 3424 S. State Street, Chicago, IL 60616. Cintoli R., Disabatino B., Galeotti L. and Bruno G. (1995) Ammonium uptake by zeolite and treatment in UASB reactor of piggery wastewater. Wat. Sci. Technol. 32(12), 7381. Cohen A., Breure A. M., van Andel J. G. and van Deursen A. (1980) Inuence of phase separation on the anaerobic digestion of glucose, I. Maximum COD-turnover-rate during continuous operation. Wat. Res. 14, 14391448. Cohen A., Breure A. M., van Andel J. G. and van Deursen A. (1982) Inuence of phase separation on the anaerobic digestion of glucose, II. Stability and kinetic responses to shock loadings. Wat. Res. 16, 449455. El-Mamouni R., Rouleau D., Mayer R., Guiot S. R. and Samson R. (1992) Comparison of the novel multiplate anaerobic reactor with the upow anaerobic sludge blanket reactor. 46th Purdue Industrial Waste Conference Proceedings. Lewis, Chelsea, MI 48118. Fannin K. F., Srivastra V. J., Conrad J. R. and Chynoweth D. P. (1981) Marine biomass program: an-

The anaerobic baed reactor: a review Kato M. T., Field J. A. and Lettinga G. (1997) The anaerobic treatment of low strength wastewaters. Proceedings of the 8th International Conference on Anaerobic Digestion, Vol. 1, Sendai, Japan, pp. 356 363. Kotsyurbenko O. R., Nozhevhikova A. N. and Zavarzin G. A. (1993) Methanogenic degradation of organic matter by anaerobic bacteria at low temperature. Chemosphere 27, 17451761. Lawrence A. W. and McCarty P. L. (1969) Kinetics of methane fermentation in anaerobic treatment. J. WPCF 41, R1R17. Lettinga G. (1995) Anaerobic digestion and wastewater treatment systems. Antonie van Leeuwenhoek 67, 328. Lettinga G., Field J., van Lier J., Zeeman G. and Hulsho Pol L. W. (1997) Advanced anaerobic wastewater treatment in the near future. Wat. Sci. Technol. 35(10), 512. Lettinga G., Hobma S. W., Hulsho Pol L. W. and de Zeeuw W. (1982) Design operation and economy of anaerobic treatment. Wat. Sci. Technol. 15(8), 175195. Lettinga G., van Velsen A. F. M., Hobma S. W., de Zeeuw W. and Klapwijk A. (1980) Use of the Upow sludge blanket (USB) reactor concept for biological wastewater treatment especially for anaerobic treatment. Biotechnol. Bioeng. 22, 699734. Levenspiel O. (1972) Non ideal ow. In Chemical Reaction Engineering, 2nd edition. Wiley, New York, pp. 253 308. Matsushige K., Inamori Y., Mizuochi M., Hosomi M. and Sudo R. (1990) Eects of recirculation ratio on the treatment of articial domestic wastewater using the anaerobicaerobic circulation lter process. Environ. Technol. 11, 989998. McCarty P. L. (1981) One Hundred Years of Anaerobic Treatment in Anaerobic Digestion 1981, ed. Hughes et al. Anaerobic Digestion 1981. Elsevier Biomedical Press B. V., pp. 321. McCarty P. L. and Bachmann A. (1992) United States Patent No. 5,091,315. Nachaiyasit S. (1995) The eect of process parameters on reactor performance in an anaerobic baed reactor. Ph.D. Dissertation, Department of Chemical Engineering, Imperial College, London, U.K. Nachaiyasit S. and Stuckey D. C. (1995) Microbial response to environmental changes in an anaerobic baed reactor (ABR). Antonie van Leeuwenhoek 67, 111123. Nachaiyasit S. and Stuckey D. C. (1997a) The eect of low temperature on the performance of an anaerobic baed reactor (ABR). J. Chem. Tech. Biotechnol. 69, 276284. Nachaiyasit S. and Stuckey D. C. (1997b) The eect of shock loads on the performance of an anaerobic baed reactor (ABR), 1. Step changes in feed concentration at constant retention time. Wat. Res. 31, 27372747. Nachaiyasit S. and Stuckey D. C. (1997c) The eect of shock loads on the performance of an anaerobic baed reactor (ABR), 2. Step and transient hydraulic shocks at constant feed strength. Wat. Res. 31, 27472755. Orozco A. (1988) Anaerobic wastewater treatment using an open plug ow baed reactor at low temperature. 5th International Symposium on Anaerobic Digestion, Bologna, Italy, pp. 759762. Orozco A. (1997) Pilot and full-scale anaerobic treatment of low-strength wastewater at sub-optimal temperature (158C) with a hybrid plug ow reactor. Proceedings of the 8th International Conference on Anaerobic Digestion, Vol. 2, Sendai, Japan, pp. 183191. Owen W. F., Stuckey D. C., Healy Jr J. B., Young L. Y. and McCarthy P. L. (1979) Bioassay for monitoring biochemical methane potential and anaerobic toxicity. Wat. Res. 13, 485492.

1577

Pavlostathis S. G. and Giraldo-Gomez E. (1991) Kinetics of anaerobic treatment: a critical review. Crit. Rev. Environ. Control 21(5,6), 411490. Pohland F. G. and Ghosh (1971) Developments in anaerobic treatment processes. In Biological Waste Treatment, ed. R. P. Canale. Interscience, New York, pp. 85106. Polprasert C., Kemmadamrong P. and Tran F. T. (1992) Anaerobic baed reactor (ABR) process for treating a slaughterhouse wastewater. Environ. Technol. 13, 857 865. Reis M. A. M., Lemos P. C. and Carrondo M. J. T. (1995) Biological sulphate removal of industrial euents using the anaerobic digestion. 9th Forum for Applied Biotechnology, Med. Fac. Landbouww. Univ. Gent, 60/ 4b, pp. 27012707. Rittmann B. E. and McCarty P. L. (1978) Variable-order model of bacterial-lm kinetics. J. Environ. Eng. Division 104(EE5), 889900. Rittmann B. E., Bae W., Namkung E. and Lu C.-J. (1987) A critical evaluation of microbial product formation in biological processes. Wat. Sci. Technol. 19, 517528. Ruiz I., Veiga M. C., de Santiago P. and Blazquez R. (1997) Treatment of slaughterhouse wastewater in a UASB reactor and an anaerobic lter. Bioresource Technol. 60(3), 251258. Sanchez Riera F., Cordoba P. and Sineriz F. (1985) Use of the UASB reactor for the anaerobic treatment of stillage from sugar cane molasses. Biotechnol. Bioeng. 27, 17101716. Sawyer C. N., McCarty P. L. and Parkin G. F. (1994) Chemistry for Environmental Engineering, 4th edition. McGraw-Hill. Sayed S., Campen L. and Lettinga G. (1987) Anaerobic treatment of slaughterhouse waste using a occulant sludge UASB reactor. Biol. Wastes 21, 1128. Sievers D. M. (1988) Particle trapping in an anaerobic baed reactor. ASAE paper No. 88-6606. ASAE, St. Joseph, MI. Speece R. E. (1996). Anaerobic Biotechnology for Industrial Wastewaters. Archae, Nashville, TN. Stronach S. M., Rudd T. and Lester J. N. (1986) Start-Up of anaerobic bioreactors. In Anaerobic Digestion Processes in Industrial Wastewater Treatment. Springer, Berlin. Stuckey D. C. (1983) Anaerobic digestion in developing countries: advances in fermentation. SCI meeting at Chelsea College, London. Switzenbaum M. S. and Jewell W. J. (1980) Anaerobic attached-lm expanded-bed reactor treatment. J. WPCF 52(7), 19531965. Tait S. J. and Freidman A. A. (1980) Anaerobic rotating biological contactor for carbonaceous wastewaters. J. WPCF 52(8), 22572269. Tilche A. and Vieira S. M. M. (1991) Discussion on reactor design of anaerobic lters and sludge bed reactors. Wat. Sci. Technol. 24(8), 193206. Tilche A. and Yang X. (1987) Light and scanning electron microscope observations on the granular biomass of experimental SBAF and HABR reactors. Proceedings of Gasmat Workshop, Netherlands, pp. 170178. van Lier J. B., Boersma F., Debets M. M. W. H. and Lettinga G. (1994) High rate thermophilic anaerobic wastewater treatment in compartmentalized upow reactors. Wat. Sci. Technol. 30(12), 251261. van Lier J. B., Groeneveld N. and Lettinga G. (1996) Development of thermophilic methanogenic sludge in compartmentalized upow reactors. Biotechnol. Bioeng. 50, 115124. Visser A., Gao Y. and Lettinga G. (1992) Anaerobic treatment of synthetic sulphate-containing wastewater under thermophilic conditions. Wat. Sci. Technol. 25(7), 193 202.

1578

William P. Barber and David C. Stuckey diluted Seine wastewater in the tropics. Trans. ASAE 30(4), 11051110. Yang X., Garuti G., Farina R., Parisi V. and Tilche A. (1988) Process dierences between a sludge bed lter and an anaerobic baed reactor treating soluble wastes. 5th International Symposium on Anaerobic Digestion, Bologna, Italy, pp. 355360. Young J. C. and McCarty P. L. (1969) The anaerobic lter for wastewater treatment. J. WPCF 4, 160173. Young H. W. and Young J. C. (1988) Hydraulic characteristics of upow anaerobic lters. J. Environ. Eng. 114(3), 621638. Zheng Y. J. and Wu W. N. (1985) A study of meat packing plant wastewater treatment with upow anaerobic sludge blanket process. Proceedings of 4th Int. Symp. on Anaerobic Digestion, China, pp. 327337. Zhou G. M. and Fang H. H. P. (1997) Co-degradation of phenol and m-cresol in a UASB reactor. Bioresource Technol. 61(1), 4752.

Weiland P. and Rozzi A. (1991) The start-up, operation and monitoring of high-rate anaerobic treatment systems: discussers report. Wat. Sci. Technol. 24(8), 257 277. Williamson K. and McCarty P. L. (1976) A model of substrate utilization by bacterial lms. J. WPCF 48(1), 9 24. Witthauer D. and Stuckey D. C. (1982) Laboratory studies on anaerobic processes to treat dilute organic waster in developing countries. Study by IRCWD, EAWAG Du bendorf, Switzerland. Xing J. and Tilche A. (1992) The eect of hydraulic retention time on the hybrid anaerobic baed reactor performance at constant loading. Biomass Bioenergy 3(1), 2529. Xing J., Boopathy R. and Tilche A. (1991) Model evaluation of hybrid anaerobic baed reactor treating molasses waste-water. Biomass and Bioenergy 1(5), 267274. Yang P. Y. and Moengangongo T. H. (1987) Operational stability of a horizontally baed anaerobic reactor for

Você também pode gostar