Você está na página 1de 8

Chemical Engineering Science 137 (2015) 677–684

Contents lists available at ScienceDirect

Chemical Engineering Science


journal homepage: www.elsevier.com/locate/ces

Mathematical modeling of upflow anaerobic sludge blanket (UASB)


reactors: Simultaneous accounting for hydrodynamics
and bio-dynamics
Yun Chen a,1, Jia He a,1, Yang Mu a,n, Ying-Chao Huo a, Ze Zhang a, Thomas A. Kotsopoulos b,
Raymond J. Zeng a,nn
a
CAS Key Laboratory of Urban Pollutant Conversion, Department of Chemistry, University of Science and Technology of China, Hefei 230026 China
b
Department of Hydraulics, Soil Science and Agricultural Engineering, School of Agriculture, Aristotle University of Thessaloniki, 54124 Thessaloniki, Greece

H I G H L I G H T S G R A P H I C A L A B S T R A C T

 Dispersive model and multi-CSTR


model simulated the hydraulics of
UASB reactor.
 Dispersive model showed great
advantage in giving tracer trends of
various heights.
 Dispersive model was integrated
with ADM1 to describe the bio-
dynamics.
 Good stimulation agreements with
experimental data during the
organic shocks.

art ic l e i nf o a b s t r a c t

Article history: Hydrodynamics and bio-dynamics are equally important in upflow anaerobic sludge blanket (UASB)
Received 8 April 2015 reactors, but few studies have been conducted to consider both issues when developing a mathematical
Received in revised form model. Therefore a novel ADM1-based dispersive model was developed in this study to provide a better
23 June 2015
description of UASB reactor performance by simultaneously taking into account the hydrodynamics and
Accepted 11 July 2015
bio-dynamics of the reactor. Organic overloading shock experiments were carried out in a laboratory-
Available online 26 July 2015
scale UASB reactor to validate the developed model. The results of the simulation model were in good
Keywords: agreement with experimental data. Although considerable calculation time was required, the ADM1-
Anaerobic digestion model No. 1 (ADM1) based dispersive model showed obvious advantages in predicting reactor status, which may give an early
Dispersive model
warning of reactor abnormalities. The models developed here are expected to simulate UASB reactors
Hydrodynamics
more effectively and to provide useful information for their design and operation.
Bio-dynamics
Upflow anaerobic sludge blanket (UASB) & 2015 Elsevier Ltd. All rights reserved.
reactor

1. Introduction several decades (Li et al., 2013; Schmidt and Ahring, 1996). The UASB
reactor has a special advantage in maintaining high biomass con-
Due to their obvious advantages for minimizing pollutants and centrations, which are usually developed at the bottom of the reactor
recovering energy, upflow anaerobic sludge blanket (UASB) reactors in the form of aggregated granular sludge (Ahn et al., 2002; Schmidt
have become widely used for wastewater treatment in the last and Ahring, 1996). Formation of granular sludge significantly
improves the ability of UASB reactors to treat high-concentration,
n
possibly refractory organic matter from wastewater (Li et al., 2013;
Corresponding author. Fax: þ 86 551 63607907.
nn Liu et al., 2015; Martínez et al., 2013; Mirzoyan and Gross, 2013).
Corresponding author. Fax: þ 86 551 63600203.
E-mail addresses: yangmu@ustc.edu.cn (Y. Mu), rzeng@ustc.edu.cn (R.J. Zeng). The Anaerobic Digestion Model 1 (ADM1) has been widely used
1
These authors contributed equally to this work. to simulate anaerobic digestion processes since its first release by

http://dx.doi.org/10.1016/j.ces.2015.07.016
0009-2509/& 2015 Elsevier Ltd. All rights reserved.
678 Y. Chen et al. / Chemical Engineering Science 137 (2015) 677–684

the International Water Association (IWA) task group (Batstone


et al., 2002). It consists of multiple steps to describe biochemical
processes, including hydrolysis, acidogenesis, acetogenesis, and
methanogenesis, as well as physico-chemical reactions such as
liquid–gas transfer and ion equilibrium in anaerobic digestion.
The ADM1 has been used to simulate UASB reactor performance
at different scales for treating a large variety of wastewaters such as
opium alkaloid effluents, grass silage digestion, cane-molasses
vinasse digestion and 3,4-dichloronitrobenzene (Barrera et al.,
2015; Dereli et al., 2010; Thamsiriroj et al., 2012; Xiao et al.,
2013). However, these studies assumed the flow pattern of UASB
reactors to be near-ideal mixing for simulation because ADM1 was
initially developed for continuous stirred-tank reactor (CSTR) sys-
tems (López and Borzacconi, 2009; Yu et al., 2012; Zhao et al., 2010).
The hydrodynamic characteristics of UASB reactors have been
extensively explored under different conditions (Batstone et al.,
2005; Peña et al., 2006; Ren et al., 2009). Batstone et al. (2005)
indicated that the flow pattern in a UASB could be between plug
flow and ideal mixing flow instead of ideal mixing. Some papers in
the literature have pointed out that the hydrodynamics of UASB Fig. 1. Schematic of an UASB reactor by the ADM1-based dispersive model.

reactors could be described by a dispersive model with a proper


diffusion parameter (Mu et al., 2008; Peña et al., 2006; Ren et al., respectively:
2009), whereas others showed that the multi-CSTR model could be  
∂Si ∂ ∂S ∂S
adjusted to simulate UASB reactor hydrodynamics (Bolle, 1985; ¼ D1 n i  un i þ r S;i ðz; t Þ ð4Þ
∂t ∂z ∂z ∂z
Heertjes et al., 1982; Rodríguez-Gómez et al., 2014). These results
imply that the flow pattern of UASB reactors is very important but  
∂X i ∂ ∂X ∂X
complex, and that therefore it cannot be simply regarded as an ideal ¼ D2 n i  un i þ r X;i ðz; t Þ ð5Þ
∂t ∂z ∂z ∂z
mixing pattern. Up to now, very limited work has been done to
consider complex reactor hydrodynamic characteristics simulta- The left-hand side of these two equations shows the variation
neously when developing an ADM1-based model to describe UASB in the balanced component. On the right-hand side of the
reactors (Mu et al., 2007; Tartakovsky et al., 2008). However, these equations, the first two parts represent the dispersive and con-
models are not validated in organic loading shocks which are highly vective components respectively, and the last part represents the
sensitive to UASB reactors and can cause process instability. bio-dynamic reaction of the substrate or biomass. The bio-
Therefore, this work has aimed to develop a mathematical dynamic reactions in these two equations were adopted from
model for simulating UASB reactors by simultaneously taking into the ADM1. In the hydrodynamic modeling part, another widely
account the hydrodynamics and bio-dynamics of the reactor. A accepted multi-CSTR (continuous stirred-tank reactor) model
laboratory-scale UASB reactor was set up and organic loading (Martin, 2000) was used to stimulate tracer variation.
shocks were used to calibrate and validate the developed model. The distribution of insoluble biomass is similar to that of
The hydrodynamic characteristics of the UASB reactor were soluble biomass, but with a comparatively smaller dispersion
investigated using tracers. The model developed here is expected coefficient D2, which could be deduced from the simulation results
to simulate UASB reactors more effectively and to provide useful based on experimental data. Furthermore, the biomass concentra-
information for their design and operation. tion above the sludge blanket was assumed to be zero due to its
sparse presence in this area. In all these partial differential
equations (PDEs), two different upflow velocities were used, with
a smaller one in the upper separator part due to its apparently
2. Materials and methods
greater cross-sectional area in a three-phase separator. Therefore,
a boundary was assumed between the reaction zone and the
2.1. Model development
three-phase separator in the UASB reactor. When 0 o s o h1 , then
u ¼ u1 , whereas when h1 o s o h1 þ h2 , then u ¼ u2 .
An axial dispersion model was developed to simulate the axial
soluble component and biomass species distributions in a UASB
2.2. Model implementation
reactor, as shown in Fig. 1. The tracer distribution is given in Eq. (1)
with the Danckwerts boundary conditions described in Eqs.
Numerical solutions of these equations in the two models were
(2) and (3):
generated using the MATLAB software (R2010a). The PDEs were
 
∂C ∂ ∂C ∂C calculated by the finite-difference method (FDM). The central
¼ D1 u ð1Þ
∂t ∂z ∂z ∂z finite-difference approximation was used for internal finite-
difference points.
∂C
D1 ¼ uðC l  C 0 Þ z ¼ 0 ð2Þ
∂z 2.3. Reactor operation

∂C
¼0z¼H ð3Þ Experiments were performed in a Plexiglas UASB reactor which
∂z
was equipped with a water jacket to maintain the reactor under
These equations show the hydrodynamic transport of the mesophilic conditions (37 1C). The internal diameter was 0.7 dm,
components. When combined with the bio-kinetic part, the and the height of the working zone was 7.5 dm, with a volume of
distributions of the soluble substrate and insoluble biomass in 2.9 L. In the three-phase separator, the internal diameter and the
the ADM1-based dispersive model are given in Eqs. (4) and (5) height were 1.2 dm and 2.5 dm respectively, with a volume of
Y. Chen et al. / Chemical Engineering Science 137 (2015) 677–684 679

Table 1 The injector and detector temperatures were 250 1C and 300 1C
Experimental conditions of organic overloading shocks conducted on the UASB respectively. Produced gas volume was recorded by a gas meter,
reactor.
and the gas contents were analyzed by a gas chromatograph
Shock run R1 R2 R3 R4 R5 R6 (Lunan model SP7890, CN) equipped with a thermal conductivity
detector and a 1.5-m stainless steel column packed with a 5 Å
Baseline (g-COD/L) 3.3 3.3 3.3 5 5 5 molecular sieve. The temperatures of the injector, detector, and
Shock run concentration (g-COD/L) 4.95 9.9 19.8 15 25 30 column were kept at 170 1C, 170 1C, and 150 1C, respectively. H2
Duration (h) 24 24 24 24 24 24
was used as the carrier gas.

2.7 L. Therefore, the total volume was 5.6 L. A gas meter was 3. Results and discussion
installed at the outlet to measure volumetric gas production.
The reactor was fed at a flow rate of 0.2 L/h with solutions A 3.1. Reactor performances
and B (1:1). The composition of solution A was as follows (in mg/
L): KCl 5.2, MgCl2  6H2O 7.2, CaCl2  2H2O 1, NH4Cl 0.4, Na2SO4 0.08, At steady state, the effluent pH of the UASB reactor was
MnCl2  4H2O 0.8, CoCl2  6H2O 1.2, H3BO3 0.2, CuCl2  2H2O 1.1, maintained at 7.27 0.3, and the gas produced by the reactor was
Na2MoO4  2H2O 0.1, ZnSO4  7H2O 3.2, FeSO4  7H2O 3.2, evaluated at a rate of 0.30–0.37 L CH4/g COD with CH4 concentra-
NiCl2  6H2O 0.5, NaHCO3 3.5. Solution B was glucose, and its tion in the 53–59% range.
concentration is given in Table 1. In R1 and R2 runs, only acetate was detected at the effluent.
The UASB reactor was inoculated with granule sludge from an Other VFAs were either below the detection baseline or within the
expanded granular sludge-bed reactor treating starch wastewater ability of the organisms to degrade them. Propionate, butyrate, and
(Huayi Ltd, China). Seed sludge was inoculated at 37 1C and iso-butyric acid were detected only when the reactor endured
sparged with nitrogen gas (4 99.99%) for 20 min before use. several organic shocks. Under shock loads, the effluent propionate
Following start-up, reactor operation was switched from batch to and acetate concentrations increased instantly. In R3, R4, R5 and R6,
continuous mode at an HRT (Hydraulic Retention Time) of 28 h propionate accumulated more heavily than acetate, whereas buty-
and an organic loading concentration of 3.3 g COD/L. The SRT rate and iso-butyric acid increased only slightly (Figs. 2–4). A
(Solids Retention Time) of the reactor was kept at one month. significant increase in discharged gas was observed, but the CH4
Six runs of organic-overloading shock tests were conducted on percentage decreased in the gas phase. Bubbles were generated in
the UASB reactor and are summarized in Table 1. Both HRT and greater amounts during shock phases. A fraction of the granules
SRT remained the same during the shocks. The reactor was were driven upwards along with the ascending bubbles, making the
operated for at least four weeks to restore it to baseline conditions sludge blanket expand slightly. In the first four runs, the reactor
before the next run. The duration of all shock tests was set to 24 h. managed to recover to steady state within three days. In Runs 5 and
Samples were collected at the outlet and analyzed for concentra- 6, the reactor took more than 10 days to restore itself. High loading
tions of volatile fatty acids (VFAs). shocks in this study demonstrated that propionate accumulates
more easily than acetate and butyrate (Figs. 2 and 3). This is the case
2.4. Tracer study because propionate is much more difficult to degrade due to its
positive Gibbs free energy change for conversion (Stams, 1994).
A pulse-response study was carried out to elucidate the Others have also pointed out that a pH over 7 selects a microbial
hydrodynamics of the UASB reactor. Li þ was chosen as the tracer community that produces mainly propionate rather than acetate
because of its ease of analysis and its lack of chemical reactivity and butyrate (Angenent et al., 2002). Without instant removal of
with the substances in the reactor. A solution containing 480 mg VFAs, propionate also accumulates quickly, which subsequently
Li2SO4 (10 mg/L Li þ on average in the whole reactor) was injected inhibits bacterial activity and prevents propionate from degrading
at time t¼0 in the influent at the bottom of the reactor. Sample to acetate (Björnsson et al., 1997; Pullammanappallil et al., 2001).
(1 mL) from the top, middle, and bottom sampling ports was The propionate/acetate ratio was below 1 when suffering in the
collected every two hours and was analyzed for the Li þ concen- low organic loading shocks (such as R1, R2). However, the ratio
tration. The dead volume of the reactor was estimated using an was over 1 when suffering in the high organic loading shocks
equation provided in a previous study (Ren et al., 2008). (such as R3 to R6). The maximum ratio was 7 in R4. Li et al. (2014)
suggested that the ratio of propionate to acetate higher than
2.5. Analyses 1.4 indicated impending digester failure, it was the similar results
to our study. So, it was concluded that the ratio of propionate to
Samples were first centrifuged at 10,000 rpm for 5 min and acetate was an early warning indicator.
then filtered with a 0.45 μm micro-filter membrane before analy-
sis. Li þ was determined by inductively coupled plasma optical 3.2. Hydrodynamic characteristics of the UASB reactor
emission spectroscopy (ICP OES Optima 7300 DV) at a wavelength
of 670 nm. The ICP operational parameters were as follows: Based on experimental data from tracer studies (Ren et al.,
40 MHz generator frequency; 1.1 kW radio-frequency power; 2008; Zeng et al., 2005), the estimated dead volume could be
15 L/min argon plasma flow rate; 0.5 L/min auxiliary argon flow ignored because it occupied only 2.22% of the whole reactor. The
rate; 0.8 L/min nebulizer argon flow rate; and 1.4 mL/min sample simulation results of the dispersive model of UASB reactor hydro-
flow rate. A 2.4 mm central tube internal-diameter torch was used. dynamics are shown in Fig. 5. The simulation results matched the
Concentrations of VFAs were measured by a gas chromatograph experimental data very well, with a standard deviation of 0.26,
(Agilent 7890, CA) equipped with a flame ionization detector and a and the dispersion coefficient derived from this model was
10 m  0.53 mm HP-FFAP fused-silica capillary column after being 1.2 dm2/h. Two upflow velocities, with the smaller one in the
acidified with 3% (v/v) formic acid. The configurations of the separator part, instead of one uniform velocity in the whole
column operating temperature was first 70 1C for 3 min, then reactor might contribute to the quality of the curve fitting.
rising at a rate of 10 1C/min to 180 1C, and then holding for The multi-CSTR model, another common model for describing
4.5 min; this procedure was then repeated for the same sample. hydraulic phenomena (Ren et al., 2009, 2008), was also developed
680 Y. Chen et al. / Chemical Engineering Science 137 (2015) 677–684

Fig. 2. Effluent acetate variations of the UASB reactor in six runs.

Fig. 3. Effluent propionate variations of the UASB reactor in the last four runs.
Y. Chen et al. / Chemical Engineering Science 137 (2015) 677–684 681

Fig. 4. Effluent butyrate and isobutyric acid variations of the UASB reactor in the last four runs.

to compare with the dispersive model. The number of CSTRs was


2.44, which was calculated based on the tracer study. The hydro-
dynamic curve generated by this model also gave a good descrip-
tion of the reactor's flow pattern at the height of the effluent port,
but was unable to reflect the tracer trajectories at other heights.
Simulation results for these three sampling ports from the dis-
persive model are shown in Fig. 6. Moreover, the multi-CSTR
model also failed to predict the experimental data trend during
the descending phase. The reason may have been that each tank in
the model is continuously stirred and the tracer once enters the
reactor, it will cause tracer to be discharged with the liquid, so it
leads to the model curve decline faster than the actual data.
Furthermore, the dispersive model has particular advantages in
simulating tracer concentrations at different heights (Fig. 6).
Theoretically, the closer the sampling port is to the injection port,
the more quickly the peak value will be reached (Zeng et al., 2005). Fig. 5. Simulation of effluent tracer concentration variation in the reactor using the
multi-CSTRs model and the dispersive model, respectively.
Experimental data have also shown this trend. The peak time
interval between the lower and the upper ports was approxi-
mately five hours, with values of 20 h and 15 h respectively, and
the peak values for the lower and upper ports were 6.28 mg/L and
6.93 mg/L respectively. This observation suggested that the flow
pattern in this UASB reactor was between mixing flow and plug
flow (Batstone et al., 2005). Therefore, it was more accurate to use
the combined dispersive ADM1 to demonstrate both the hydraulic
and bio-kinetic dynamics of the UASB reactor.

3.3. Sensitivity analysis

Although all the parameters influence the model output, many


of them have limited impact. Sensitivity analysis helps to identify
the important parameters and reduces the complexity of parameter
tuning. This was done using the absolute-relative sensitivity func-
tion. In the study, acetate and propionate were accumulated when
suffering from the organic loading shocks. So, the kinetic para- Fig. 6. Simulation result of the three sampling ports by the dispersive model.
meters about the uptake of acetate and propionate were recognized
as sensitive ones. The results of sensitivity analysis are shown in relevance of km_ac for the prediction of acetate concentration than
Fig. 7. S_ac sensitivity of the kinetic parameter km_ac was much ks_ac. Similarly, Fig. 7B indicates that the kinetic parameter km_pro
higher than that of the kinetic parameter ks_ac, indicating a higher also had a higher relevance with the propionate concentration
682 Y. Chen et al. / Chemical Engineering Science 137 (2015) 677–684

Fig. 8. Simulation results of R3 (A) and R4 (B) by the ADM1-based


dispersive model.
Fig. 7. Parameter sensitivity analysis for acetate and propionate concentrations in
the ADM1-based dispersive model.
based dispersive model could describe accurately the changes in
acetate and propionate concentrations in the reactor. Use of two
Table 2 upflow velocities in the ADM1-based dispersive model, with a
Estimated parameter values of km_pro and km_ac in the model (units: g-COD/ smaller one in the separator phase, may have contributed to good
(L  d)). agreement. In these experiments, increasing the loading concen-
tration reduced biomass activity, which led to acetate and propio-
Stable state Overloading shock
nate accumulation. In the model as developed, km_pro and km_ac
km_pro km_ac km_pro km_ac must be estimated for simulation. The parameter values during the
shocks were much lower than during steady state, which also
R3 15 12 8.6 8.8 demonstrated that biomass activity was inhibited. The same
R4 13 10 7 7.35
observation was also found in R3. As a reference, the simulated
acetate and propionate concentrations by ADM1 only were much
compared with ks_pro for S_pro sensitivity. Therefore, km_ac and higher than the experimental results (Fig. S1).
km_pro were chosen for parameter tuning. The fairly good modeling results for organic shocks give
credence to the dispersive model combined with bio-dynamics.
3.4. Model validation However, solving the partial differential equations should include
proper least internal finite-difference points to ensure conver-
Two cases, R3 and R4, were used to validate the developed gence of the equations, which required more than 30 min of
model. In R3, the estimated values of km_pro and km_ac are shown computation time to reach their final values.
in Table 2. These parameters were tuned manually. Values of the
other parameters were adopted from the original ADM1. 3.5. Implications of this study
In R3, the variations of effluent propionate and acetate con-
centrations in the UASB reactor were simulated using the ADM1- In the first two runs, a slight increase in acids did occur in the
based dispersive model, as shown in Fig. 8. The experimental and bottom part of the reactor (data not shown), although no obvious
simulated data matched very well. In this model, the particle acid accumulation was detected at the effluent, except of acetate at
dispersion coefficient is affected by gravitational force and is not as low concentrations. The last two runs experienced severe acid
free as the soluble components. Estimation gave a value of inhibition, especially propionate. Propionate accumulation has been
0.85 dm2/h for D2. Furthermore, because biomass aggregates commonly used to indicate organic loading abnormalities in
mainly at the bottom of the reactor and scarcely exists in the top advance (Hill et al., 1987; Marchaim and Krause, 1993; Nordstedt
part, s¼ 4 dm was selected as the boundary above which biomass and Thomas, 1985). The model as developed could successfully
concentration was considered to be zero. predict propionate fluctuations of lower-reactor samples, as shown
The UASB reactor took 2.5 days to recover from overloading in Fig. 9. The data in this figure were developed under the same
shock in R4, and experimental data for this case were used to conditions as Fig. 8B. This figure demonstrated that the lower the
validate the developed models again. Fig. 8 shows that the ADM1- height, the higher was the peak value and the more quickly it was
Y. Chen et al. / Chemical Engineering Science 137 (2015) 677–684 683

D2 Dispersive coefficient of particle concentration (dm2/d)


H Whole height of the reactor (dm)
HRT Hydraulic retention time
i Index of the components in the reactor
j Index of biodynamic reactions
N Number of the reactors
q Inflow of the reactor (L/d)
r S;i ðs; t Þ Net soluble transformation rate of the ith component
r X;i ðs; t Þ Net insoluble transformation rate of the ith component
z Axial position (dm)
t res;X Solid retention time that exceeds hydraulic time
Vd Dead volume of the reactor
Vliq Active volume of the reactor (L)
Xi Biomass concentration (g-COD/L)
Xi,in Inflow concentration of Xi (g-COD/L)
Fig. 9. Prediction of propionate variations at different reactor heights during the u Upflow velocity (dm/h)
organic overloading shock using the ADM1-based dispersive model. ρj Kinetic rate of reaction j
νi;j Rate coefficient of component i in reaction j
reached. This result is consistent with the dispersive hydraulic km_pro The maximum specific consumption rate of propionate
model, which indicates that the lowest port always experiences (g-COD/g-COD∙d)
the fastest and largest fluctuations (Fig. 6). If only effluent data were km_ac The maximum specific consumption rate of acetate (g-
used to describe reactor status, one might rush to the conclusion COD/g-COD∙d)
that the reactor is undergoing few or no interruptions, but detailed, ks_pro Half saturation value of propionate (g-COD/L)
more accurate information from the lower ports can indicate that ks_ac Half saturation value of acetate (g-COD/L)
the reactor is experiencing abnormal conditions. Because propio-
nate can serve as an indicator, fast, early propionate accumulation
in the lower section can demonstrate the instability of the inflow
constituents. Acknowledgments
Thousands of UASB reactors are functioning all over the world.
Strict monitoring is required to ensure that all these UASB reactors The authors would like to acknowledge the financial support
are in steady state. Therefore, development of models is essential. from National Hi-Technology Development 863 Program of China
The present research has demonstrated that the flow pattern of a (2011AA060901), the Hundred-Talent Program of Chinese Acad-
UASB reactor should not be regarded as ideal mixing. Therefore, emy of Sciences, National Natural Science Foundation of China
when modeling pilot-scale or full-scale UASB reactors, the hydro- (51222812), the Program for Changjiang Scholars and Innovative
dynamics should first be investigated. Based on the results, the Research Team in University, and the Fundamental Research Funds
bio-kinetics should be considered further when developing mod- for the Central Universities (wk2060190040).
els. However, extensive calibration of the hydraulic and kinetic
parts of the model is needed before a pilot- or full-scale system
can be simulated. Appendix A. Supporting information

Supplementary data associated with this article can be found in


4. Conclusions
the online version at http://dx.doi.org/10.1016/j.ces.2015.07.016.

The dispersive model and the multi-CSTR model were developed


to simulate the hydraulics of a UASB reactor. The former one showed References
great advantage in giving tracer trends of various heights while the
latter one failed to predict the experimental data trend during the Ahn, Y., Song, Y., Lee, Y., Park, S., 2002. Physicochemical characterization of UASB
descending phase. Then the dispersive model was integrated with sludge with different size distributions. Environ. Technol. 23, 889–897.
Angenent, L.T., Abel, S.J., Sung, S., 2002. Effect of an organic shock load on the
ADM1 to describe the bio-dynamics. It was found that the ADM1- stability of an anaerobic migrating blanket reactor. J. Environ. Eng. 128,
based dispersive model demonstrated good stimulation agreements 1109–1120.
with experimental data during the organic overloading shocks. Barrera, E.L., Spanjers, H., Solon, K., Amerlinck, Y., Nopens, I., Dewulf, J., 2015.
Modeling the anaerobic digestion of cane-molasses vinasse: extension of the
Although the ADM1-based dispersive model required almost half Anaerobic Digestion Model No. 1 (ADM1) with sulfate reduction for a very high
an hour to reach the final calculation result, it showed obvious strength and sulfate rich wastewater. Water Res. 71, 42–54.
advantages in predicting reactor status at different heights, which Batstone, D.J., Hernandez, J.L.A., Schmidt, J.E., 2005. Hydraulics of laboratory and
full-scale upflow anaerobic sludge blanket (UASB) reactors. Biotechnol. Bioeng.
may give an early warning of reactor abnormalities.
91, 387–391.
Batstone, D.J., Keller, J., Angelidaki, I., Kalyuzhnyi, S., Pavlostathis, S., Rozzi, A.,
Sanders, W., Siegrist, H., Vavilin, V., 2002. The IWA anaerobic digestion model
Nomenclature no 1(ADM 1). Water Sci. Technol. 45, 65–73.
Björnsson, L., Mattiasson, B., Henrysson, T., 1997. Effects of support material on the
pattern of volatile fatty acid accumulation at overload in anaerobic digestion of
t Time (h in tracer study, d in organic shock study) semi-solid waste. Appl. Microbiol. Biotechnol. 47, 640–644.
C Concentration of tracer Li þ in the reactor (mg/L) Bolle, W.L., 1985. An integral dynamic model for the UASB reactor. Biotechnol.
Bioeng. 28, 1621–1636.
C0 Tracer concentration of the influent (mg/L) Dereli, R.K., Ersahin, M.E., Ozgun, H., Ozturk, I., Aydin, A.F., 2010. Applicability of
Si Soluble concentration (g-COD/L) Anaerobic Digestion Model No. 1 (ADM1) for a specific industrial wastewater:
Ci,in Inflow concentration of Ci (g-COD/L) opium alkaloid effluents. Chem. Eng. J. 165, 89–94.
Heertjes, P., Kujivenhoven, L., Van der Meer, R., 1982. Fluid flow pattern in upflow
Cl Left boundary concentration of the tracer (mg/L) reactors for anaerobic treatment of beet sugar factory wastewater. Biotechnol.
D1 Dispersive coefficient of soluble concentration (dm2/h) Bioeng. 24, 443–459.
684 Y. Chen et al. / Chemical Engineering Science 137 (2015) 677–684

Hill, D., Cobb, S., Bolte, J., 1987. Using volatile fatty acid relationships to predict Peña, M.R., Mara, D.D., Avella, G.P., 2006. Dispersion and treatment performance
anaerobic digester failure. Trans. ASAE 30, 496–501. analysis of an UASB reactor under different hydraulic loading rates. Water Res.
López, I., Borzacconi, L., 2009. Modelling a full scale UASB reactor using a COD 40, 445–452.
global balance approach and state observers. Chem. Eng. J. 146, 1–5. Pullammanappallil, P.C., Chynoweth, D.P., Lyberatos, G., Svoronos, S.A., 2001. Stable
Li, J., Yu, L., Yu, D., Wang, D., Zhang, P., Ji, Z., 2014. Performance and granulation in performance of anaerobic digestion in the presence of a high concentration of
an upflow anaerobic sludge blanket (UASB) reactor treating saline sulfate propionic acid. Bioresour. Technol. 78, 165–169.
wastewater. Biodegradation 25, 127–136. Ren, T.-T., Mu, Y., Ni, B.-J., Yu, H.-Q., 2009. Hydrodynamics of upflow anaerobic
Li, L., He, Q., Wei, Y., He, Q., Peng, X., 2014. Early warning indicators for monitoring sludge blanket reactors. AIChE J. 55, 516–528.
the process failure of anaerobic digestion system of food waste. Bioresour. Ren, T.-T., Mu, Y., Yu, H.-Q., Harada, H., Li, Y.-Y., 2008. Dispersion analysis of an
Technol. 171, 491–494. acidogenic UASB reactor. Chem. Eng. J. 142, 182–189.
Liu, Y., Zhang, Y., Ni, B.-J., 2015. Zero valent iron simultaneously enhances methane Rodríguez-Gómez, R., Renman, G., Moreno, L., Liu, L., 2014. A model to describe the
production and sulfate reduction in anaerobic granular sludge reactors. Water performance of the UASB reactor. Biodegradation 25, 239–251.
Schmidt, J.E., Ahring, B.K., 1996. Granular sludge formation in upflow anaerobic
Res. 75, 292–300.
sludge blanket (UASB) reactors. Biotechnol. Bioeng. 49, 229–246.
Marchaim, U., Krause, C., 1993. Propionic to acetic acid ratios in overloaded
Stams, A.J., 1994. Metabolic interactions between anaerobic bacteria in methano-
anaerobic digestion. Bioresour. Technol. 43, 195–203.
genic environments. Antonie van Leeuwenhoek 66, 271–294.
Martínez, C., Celis, L., Cervantes, F., 2013. Immobilized humic substances as redox
Tartakovsky, B., Mu, S.J., Zeng, Y., Lou, S.J., Guiot, S.R., Wu, P., 2008. Anaerobic
mediator for the simultaneous removal of phenol and Reactive Red 2 in a UASB
digestion model no. 1-based distributed parameter model of an anaerobic
reactor. Appl. Microbiol. Biotechnol. 97, 9897–9905. reactor: II. Model validation. Bioresour. Technol. 99, 3676–3684.
Martin, A., 2000. Interpretation of residence time distribution data. Chem. Eng. Sci. Thamsiriroj, T., Nizami, A.S., Murphy, J.D., 2012. Use of modeling to aid design of a
55, 5907–5917. two-phase grass digestion system. Bioresour. Technol. 110, 379–389.
Mirzoyan, N., Gross, A., 2013. Use of UASB reactors for brackish aquaculture sludge Xiao, X., Sheng, G.-P., Mu, Y., Yu, H.-Q., 2013. A modeling approach to describe ZVI-
digestion under different conditions. Water Res. 47, 2843–2850. based anaerobic system. Water Res. 47, 6007–6013.
Mu, S.J., Zeng, Y., Tartakovsky, B., Wu, P., 2007. Simulation and control of an upflow Yu, L., Zhao, Q., Ma, J., Frear, C., Chen, S., 2012. Experimental and modeling study of
anaerobic sludge blanket (UASB) reactor using an ADM1-based distributed a two-stage pilot scale high solid anaerobic digester system. Bioresour. Technol.
parameter model. Ind. Eng. Chem. Res. 46, 1519–1526. 124, 8–17.
Mu, S.J., Zeng, Y., Wu, P., Lou, S.J., Tartakovsky, B., 2008. Anaerobic digestion model Zeng, Y., Mu, S.J., Lou, S.J., Tartakovsky, B., Guiot, S.R., Wu, P., 2005. Hydraulic modeling
no. 1-based distributed parameter model of an anaerobic reactor: I. Model and axial dispersion analysis of UASB reactor. Biochem. Eng. J. 25, 113–123.
development. Bioresour. Technol. 99, 3665–3675. Zhao, B.-H., Mu, Y., Dong, F., Ni, B.-J., Zhao, J.-B., Sheng, G.-P., Yu, H.-Q., Li, Y.-Y.,
Nordstedt, R., Thomas, M., 1985. Startup characteristics of anaerobic fixed bed Harada, H., 2010. Dynamic modeling the anaerobic reactor startup process. Ind.
reactors. Trans. ASAE 28, 1242–1247 1252. Eng. Chem. Res. 49, 7193–7200.

Você também pode gostar