Vous êtes sur la page 1sur 13

w a t e r r e s e a r c h 7 1 ( 2 0 1 5 ) 4 2 e5 4

Available online at www.sciencedirect.com

ScienceDirect

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

Modeling the anaerobic digestion of cane-


molasses vinasse: Extension of the Anaerobic
Digestion Model No. 1 (ADM1) with sulfate
reduction for a very high strength and sulfate rich
wastewater

Ernesto L. Barrera a,e,*, Henri Spanjers b, Kimberly Solon c,


Youri Amerlinck d, Ingmar Nopens d, Jo Dewulf e
a
Study Center of Energy and Industrial Processes, Sancti Spiritus University, Ave de los Martires 360, 60100 Sancti
Spiritus, Cuba
b
Faculty of Civil Engineering and Geosciences, Department of Water Management, Section Sanitary Engineering,
Delft University of Technology, Stevinweg 1, 2628 CN Delft, The Netherlands
c
Division of Industrial Electrical Engineering and Automation (IEA), Department of Biomedical Engineering (BME),
Lund University, Box 118, SE-221 00 Lund, Sweden
d
BIOMATH, Department of Mathematical Modeling, Statistics and Bioinformatics, Ghent University, Coupure Links
653, 9000 Gent, Belgium
e
Research Group ENVOC, Faculty of Bioscience Engineering, Ghent University, Coupure Links 653, 9000 Ghent,
Belgium

article info abstract

Article history: This research presents the modeling of the anaerobic digestion of cane-molasses vinasse,
Received 7 July 2014 hereby extending the Anaerobic Digestion Model No. 1 with sulfate reduction for a very
Received in revised form high strength and sulfate rich wastewater. Based on a sensitivity analysis, four parameters
22 October 2014 of the original ADM1 and all sulfate reduction parameters were calibrated. Although some
Accepted 15 December 2014 deviations were observed between model predictions and experimental values, it was
Available online 24 December 2014 shown that sulfates, total aqueous sulfide, free sulfides, methane, carbon dioxide and
sulfide in the gas phase, gas flow, propionic and acetic acids, chemical oxygen demand
Keywords: (COD), and pH were accurately predicted during model validation. The model showed high
ADM1 (±10%) to medium (10%e30%) accuracy predictions with a mean absolute relative error
Mathematical modeling ranging from 1% to 26%, and was able to predict failure of methanogenesis and sulfido-
Simulation genesis when the sulfate loading rate increased. Therefore, the kinetic parameters and the

Abbreviations: ADM1, Anaerobic Digestion Model No. 1; aSRB, Acetate sulfate reducing bacteria; COD, Chemical oxygen demand; CI,
Confidence interval; COV, Covariance matrix; FIM, Fisher information matrix; hMA, Hydrogenotrophic methanogenic archea; hSRB,
Hydrogenotrophic sulfate reducing bacteria; LCFA, Long chain fatty acids; ODE, Ordinary differential equation; OLR, Organic loading rate;
pSRB, Propionate sulfate reducing bacteria; SCOD, Soluble COD concentration; SLR, Sulfate loading rate; SRB, Sulfate reducing bacteria;
UASB, Upflow anaerobic sludge bed; XCOD, Particulate COD concentration.
* Corresponding author. Study Center of Energy and Industrial Processes, Sancti Spiritus University, Ave de los Martires 360, 60100 Sancti
Spiritus, Cuba. Tel.: þ53 41336118, þ32 92645948 (Belgium).
E-mail addresses: ernestol@uniss.edu.cu, ErnestoLuis.BarreraCardoso@UGent.be (E.L. Barrera), h.l.f.m.Spanjers@tudelft.nl (H.
Spanjers), kimberly.solon@iea.lth.se (K. Solon), Youri.Amerlinck@UGent.be (Y. Amerlinck), ingmar.nopens@UGent.be (I. Nopens), Jo.
Dewulf@UGent.be (J. Dewulf).
http://dx.doi.org/10.1016/j.watres.2014.12.026
0043-1354/© 2014 Elsevier Ltd. All rights reserved.
w a t e r r e s e a r c h 7 1 ( 2 0 1 5 ) 4 2 e5 4 43

Sulfate reduction model structure proposed in this work can be considered as valid for the sulfate reduction
Vinasse process in the anaerobic digestion of cane-molasses vinasse when sulfate and organic
loading rates range from 0.36 to 1.57 kg SO4 2 m3 d1 and from 7.66 to 12 kg COD m3 d1,
respectively.
© 2014 Elsevier Ltd. All rights reserved.

Nomenclature Sgas,i Gas phase concentration of the component i (for


H2S units are in kmol m3) Fraction
Ci Carbon content of the component
TCOD Total COD concentration kg COD m3
i kmol C kg COD1
ta,df t-values obtained from the Student-t distribution
eff_COD Effluent COD concentration kg COD m3

I1e4 Inhibition functions e
Xi Concentration of the particulate component
Ih2s,j Sulfide inhibition function for process j e
i kg COD m3
IpH,j pH inhibition function for process j e
Yi Yield of biomass on substrate
KA/B,i Acid-base kinetic parameter for component
kg COD_Xi kg COD_S1
i m3 kmol1 d1
i

Ka,i Acid-base equilibrium coefficient for component Greek letter


i kmol m3 Ґn,m Sensitivity value of the nth process variable (y)
kdec,j First order decay rate for process j d1 with respect to the mth model parameter (q) e
KH,i Henri's law coefficient for component ni,j Biochemical rate (or liquid phase yield) coefficient
i kmol m3 bar1 for component i on process j e
KI,h2s,j 50% inhibitory concentration of free H2S on the rj Kinetic rate of process j (for acid-base and
process j kmol m3 gaseliquid equations units are in kmol m3 d1)
kLa Gaseliquid transfer coefficient d1 kg COD_Si m3 d1
km,j Monod maximum specific uptake rate for process d Perturbation factor Fraction
j kg COD_Si kg COD_X1 i d1 s Standard deviations Vary with q
KS,j Half saturation coefficient of component i on a Significance level %
process j (for SO4 2 units are in kmol m3)
Subscript
kg COD m3
bac Pertaining to bacteria e
Ni Nitrogen content of the component
dec Pertaining to decay processes e
i kmol N kg COD1
df Degree of freedom
Pgas,i Pressure of gas component i bar
free Undissociated form of the species e
Qgas Gas flow rate m3 d1
gas Pertaining to the gas phase e
Si Concentration of the soluble components i (for
i Pertaining to soluble or particulate component e
hydrogen ions, sulfates, sulfides and its ionized
j Pertaining to processes e
forms units are in kmol m3) kg COD_Si m3
I Pertaining to the inert fraction e

by considering the oxidation (by SRB) of the available hydrogen


1. Introduction only (Batstone, 2006). However, in systems with high sulfate
concentrations volatile fatty acids (butyric, propionic and acetic)
Many industrial processes, especially in food and fermenta- have, to be included as electron donors in the sulfate degradation
tion industries, generate wastewaters with high levels of reactions in addition to hydrogen (Batstone, 2006). Barrera et al.
chemical oxygen demand (COD) and sulfate (Zub et al., 2008). (2014) provided a characterization of the sulfate reduction pro-
Vinasse obtained from ethanol distillation in the sugar cane cess in the anaerobic digestion of a very high strength and sulfate
industry (cane-molasses vinasse) is a typical example of very rich vinasse. The authors demonstrated that propionate sulfate
high strength and sulfate rich liquid substrate (Barrera et al., reducing bacteria considerably contribute to propionic acid
2013). Hence, the anaerobic digestion of vinasse promotes degradation at SO4 2 /COD ratios 0.10 as a result of hydrogen
the activity of sulfate reducing bacteria (SRB) producing sul- limitation. This suggests that reactions involving volatile fatty
fide. The latter is distributed among aqueous sulfide (H2S free, acids need to be included to properly model sulfate reduction in
HS and S2), hydrogen sulfide in the biogas and insoluble the anaerobic digestion of such vinasses. The Anaerobic Diges-
metallic sulfides. tion Model No. 1 (ADM1), developed by the IWA Task Group for
Modeling has proven to be an important tool for under- Mathematical Modeling of Anaerobic Digestion Processes, is one
standing, design and control of the sulfate reduction process of the most sophisticated and complex anaerobic digestion
(Batstone, 2006). A simple approach to model sulfate reduction is models, involving 19 biochemical processes and two types of
44 w a t e r r e s e a r c h 7 1 ( 2 0 1 5 ) 4 2 e5 4

physiochemical processes (Batstone et al., 2002). The simple  E3 to E4: the influent SO4 2 was increased whereas the
approach of Batstone (2006) to model sulfate reduction as an influent COD concentration was decreased to increase the
extension of ADM1 has been used to model the anaerobic SO4 2 /COD ratio to 0.1.
digestion of vinasse under dynamic conditions without success,  E4 to E6: the concentration of influent COD and SO4 2
exhibiting under prediction of H2S and over prediction of volatile was increased while keeping the SO4 2 /COD ratio at 0.1.
fatty acids (Hinken et al., 2013). In order to extend ADM1,  E7: the concentration of influent COD and SO4 2 was
Fedorovich et al. (2003) included the sulfate reduction process reduced to control toxicity, keeping the SO4 2 /COD ratio at
starting from previously reported work (Kalyuzhnyi et al., 1998; 0.1.
Kalyuzhnyi and Fedorovich, 1998; Knobel and Lewis, 2002;  E8 and E9: the influent SO4 2 was increased while
Ristow et al., 2002). The approach of Fedorovich et al. (2003) can keeping a constant influent COD concentration to increase
be considered as complex because of the inclusion of valerate/ the SO4 2 /COD ratio to 0.15 and 0.20, respectively.
butyrate, propionate, acetate and hydrogen in the sulfate
degradation reactions (Batstone, 2006). This model (Fedorovich Operating conditions were grouped in data set D1 (oper-
et al., 2003) was calibrated for organic deficient (SO4 2 /COD ating conditions E1, E2, E3 and E4 in Barrera et al. (2014))
ratios  1.5) synthetic wastewaters (Omil et al., 1996, 1997), for calibration and direct validation, and data set D2 (oper-
hereby focusing on volatile fatty acids, sulfates and methane gas ating conditions E5, E6, E7, E8 and E9 in Barrera et al.
phase concentrations. Furthermore, the agreement between (2014)) for cross validation.
model and experimental values for the concentrations of total
aqueous sulfide (Sh2s), free sulfides (Sh2s,free) and gas phase sul-
fides (Sgas,h2s) was not reported. Likely because of these limita- 2.2. Model description and implementation
tions, the extension of Fedorovich et al. (2003) is not commonly
used (Lauwers et al., 2013). The original ADM1 is described in a scientific and technical
Consequently, an extension of ADM1 with sulfate reduc- report prepared by an IWA Task Group (Batstone et al., 2002).
tion to model the anaerobic digestion of a very high strength This model takes into account seven bacterial groups. The
and sulfate rich vinasse may overcome the current limitation biological degradation processes are described using Monod
of models by (1) describing the sulfate reduction process in the kinetics, while the extracellular processes (disintegration and
anaerobic digestion of vinasse, (2) predicting the sulfur com- hydrolysis) and the biomass decay are described using first-
pounds in both the gas and liquid phases, (3) increasing order kinetics.
applicability of ADM1 to specific industrial wastewaters The ADM1 extension with sulfate reduction for high
(vinasse), and (4) simplifying the existing approach to reduce strength and sulfate rich wastewater was implemented in
complexity and to support further implementations. MatLab/Simulink 2008b following the original ADM1 (Batstone
Therefore, the work presented here attempts to model the et al., 2002) and the approaches discussed by Barrera et al.
anaerobic digestion of real cane-molasses vinasse by (2013). The model was implemented as a set of ordinary dif-
extending ADM1 with sulfate reduction for a very high ferential equations using the ODE 15s as numerical solver.
strength and sulfate-rich complex wastewater, including Based on experimental observations previously discussed
volatile fatty acids (propionic and acetic acids) in the sulfate in Barrera et al. (2014), butyric acid was neglected as organic
degradation reactions, hereby including an accurate predic- matter for SRB in the model structure (5% of the total volatile
tion of Sh2s, Sh2s,free and Sgas,h2s. fatty acids concentration), whereas propionic (considered
incompletely oxidized by propionate SRB) and acetic acids, as
well as hydrogen, were considered as the electron donors for
the sulfate reduction processes following the biochemical
2. Materials and methods degradation reactions (1), (2) and (3) below:

C2 H5 COOHþ0:75H2 SO4 /CH3 COOHþCO2 þH2 Oþ0:75H2 S (1)


2.1. Experimental data

CH3 COOH þ H2 SO4 /2CO2 þ 2H2 O þ H2 S (2)


Experimental observations from a characterization study of
the sulfate reduction process in the anaerobic digestion of a
4H2 þ H2 SO4 /H2 S þ 4H2 O (3)
very high strength and sulfate-rich vinasse (Barrera et al.,
2014) were used for model calibration and validation. During Consequently, three SRB groups were considered to be
these experiments a 3.5 L UASB reactor was operated under active inside the reactor; i.e. propionate sulfate reducing
dynamic conditions for a period of 75 days (following a 55 day bacteria (pSRB), acetate sulfate reducing bacteria (aSRB) and
start-up period). The experimental set-up, analytical methods hydrogenotrophic sulfate reducing bacteria (hSRB). A dual
and operating conditions are described in detail in Barrera term Monod type kinetics was used to describe the uptake rate
et al. (2014). They can be briefly described as follows, where of these substrates (Fedorovich et al., 2003). The biochemical
the E-codes indicate successive experiments conducted under rate coefficients (ni,j) and kinetic rate equation (rj) for soluble
different operating conditions: and particulate components are listed in Table 1. Similar to
decay of other microbial species, first order kinetics was used
 E1 to E3: the concentration of influent COD and SO4 2 to describe the decay of SRB. Additionally, rate coefficients
was gradually increased while keeping the SO4 2 /COD ratio and kinetic rate equations for acid-base reactions for sulfides
at 0.05. and sulfates (in the form recommended by Rosen and
Table 1 e Biochemical rate coefficients (ni,j) and kinetic rate equation (rj) for soluble and particulate components added to ADM1 to model the sulfate reduction process in
the anaerobic digestion of cane-molasses vinasse.
Component i 6 7 8 8a 9a 10 11 13 21a 22a 23a Rate (r, kg COD m3 d1)
/
j Process Y Spro Sac Sh2 Sso4 Sh2s SIC SIN Xc XpSRB XaSRB XhSRB
P S Sso4
10a Uptake of 1 (1  YpSRB)0.57 (1  YpSRB) (1  YpSRB)  Ci ni;10a YpSRB$Nbac YpSRB km;pSRB $KS;propro
þSpro $KS;so4;pSRB þSso4 $XpSRB $I4
Propionate 0.43/64 0.43/64 i¼19a

w a t e r r e s e a r c h 7 1 ( 2 0 1 5 ) 4 2 e5 4
i¼1124
by pSRB
P
11a Uptake of 1 (1  YaSRB)/64 (1  YaSRB)/64  Ci ni;11a YaSRB$Nbac YaSRB km;aSRB $KS;acSacþSac $KS;so4;aSRB
Sso4
þSso4 $XaSRB $I4
Acetate by i¼19a
i¼1124
aSRB
P
12a Uptake of 1 (1  YhSRB)/64 (1  YhSRB)/64  Ci ni;12a YhSRB$Nbac YhSRB km;hSRB $KS;h2Sh2þSh2 $KS;so4;hSRB
Sso4
þSso4 $XhSRB $I4
Hydrogen i¼19a
i¼1124
by hSRB
17a Decay of Cbac þ Cxc Nbac þ Nxc 1 1 kdec,XpSRB$XpSRB
XpSRB
18a Decay of Cbac þ Cxc Nbac þ Nxc 1 1 kdec,XaSRB$XaSRB
XaSRB
19a Decay of Cbac þ Cxc Nbac þ Nxc 1 1 kdec,XhSRB$XhSRB
XhSRB
Total Total acetate Total Total Hydrogen Inorganic Inorganic Composites Propionate Acetate SRB Hydrogeno- Inhibition term for Xc4 & Xpro
propionate (kg COD m3) hydrogen sulfates sulfide carbon nitrogen (kg COD m3)1 SRB (kg COD m3) trophic (See I2 in Batstone et al. (2002))
(kg COD m3) gas (kmol m3) (kmol m3) (kmol m3) (kmol m3) (kg COD m3) SRB I2$Ih2s,810
3
(kg COD m ) (kg COD m3) Inhibition term for Xac
(See I3 in Batstone et al. (2002))
I3$Ih2s,11
Inhibition term for Xh2
(See I1 in Batstone et al. (2002))
I1$Ih2s,12
Inhibition term for pSRB,
aSRB & hSRB
I4 ¼ IpH $Ih2s;ð10a; 11a;12aÞ

45
46 w a t e r r e s e a r c h 7 1 ( 2 0 1 5 ) 4 2 e5 4

Table 2 e Rate coefficients (ni,j) and kinetic rate equation (rj) for acid-base reactions in the differential equation
implementation added to ADM1 to model the sulfate reduction process in the anaerobic digestion of cane-molasses
vinasse.
Component i / 8a.1 8a.2 9a.1 9a.2 Rate (rj , kmol m3 d1)
j Process Y Shso4 Sso42 Sh2s,free Shs
A12 Sulfide acidebase 1 1 KA=B;h2s $ðShs $ðSHþ þ Ka;h2s Þ  Ka;h2s $Sh2s;total Þ
A13 Sulfate acidebase 1 1 KA=B;so4 $ðSso4;total $SHþ  ðKa;so4 þ SHþ ÞÞ

Jeppsson (2006)) were considered (Table 2). Un-dissociated inhibition terms in Table 1). All pH inhibitions were based on
sulfuric acid (H2SO4) and sulfide ions (S2) were considered the Hill function as suggested in Rosen and Jeppsson (2006).
negligible and were not included in the acid-base reactions. As
sulfuric acid is a strong acid (pKa < 2), it can be considered
2.3. Model inputs and initial conditions
completely dissociated, whereas sulfide ions S2 exist in small
amounts (pKa z 14) in the liquid phase of anaerobic reactors
The influent characterization of cane-molasses vinasse is
in which a pH between 6.5 and 8 is required. The dissociation
shown in Table 4. Sugar, protein and lipid contents were
Eqs (4) and (5) were included in the model.
experimentally determined in the filtered and unfiltered
HSO þ 2 vinasse and used to calculate the soluble sugars (Ssu) and
4 %H þ SO4 pKa ¼ 1:99 (4)
particulate carbohydrates (Xch), the soluble amino acids (Saa)
and particulate proteins (Xpr) as well as the long chain fatty
H2 S %Hþ þ HS pKa ¼ 7:01 (5)
acids (Sfa) and particulate lipids (Xli), respectively. The total
To model the stripping of H2S, the liquid phase yield coef- cation concentration was determined as the sum of Naþ, Kþ,
ficient (ni,j) and the rate equations (rj) for liquidegas transfer Ca2þ, Mg2þ, Mn2þ, and Zn2þ species concentrations whereas
process were also included (Table 3). Nomenclature in Tables NO2  , NO3  , PO4 3 , and Cl concentrations were used to
1e3 was adopted from ADM1 (Batstone et al., 2002). determine the total anion concentration of vinasse.
Despite the fact that Sh2s has been found to inhibit anaer- The difference between the soluble COD (SCOD) and the
obic digestion (Visser et al., 1996), Sh2s,free was assumed to be total COD of Ssu, Saa, Sfa and Sac was assumed to be the soluble
inhibitory for modeling purposes (Kalyuzhnyi et al., 1998; inert concentration of vinasse (SI). Similarly, the difference
Kalyuzhnyi and Fedorovich, 1998; Knobel and Lewis, 2002; between the COD concentration of the particulate matter
Ristow et al., 2002). A non-competitive inhibition function (XCOD) and the total COD of Xch, Xpr and Xli was assumed to be
for sulfides (Ih2s,j) was considered in all cases (Knobel and the particulate inert concentration of vinasse (XI) (See Table 4).
Lewis, 2002). The inhibition terms I1 and I2 were adopted The concentrations of these known input variables (Ssu, Saa,
from the original ADM1 for the uptake of sugars, amino acids Sfa, Sac, SI, Xch, Xpr, Xli, and XI) under specific operating con-
and long chain fatty acids (processes not shown in Table 1). ditions (E1 to E9) were calculated from the total COD
However, the inhibition term for valerate, butyrate and pro- (TCOD) of the diluted vinasse at these operating conditions
pionate degraders (I2 in Batstone et al. (2002)) as well as the and the compositions of raw vinasse as given in Table 4. This
inhibition term for acetotrophic methanogens (I3 in Batstone is illustrated for sugars in Eq. (6).

TCOD_diluted_vinasse_operating conditionsðE  1 to E  9Þ
Ssu operating conditionsðE1 to E9Þ ¼ $Ssu_raw vinasse_Table 4 (6)
TCOD_raw_vinasse_Table 4

et al. (2002)) and the inhibition term for hydrogenotrophic ADM1 requires a large number of input variables. Reason-
methanogens (I1 in Batstone et al. (2002)) was multiplied by able assumptions were made for the concentration of the
Ih2s,j in order to include the free sulfide inhibition in this model unknown input variables Sh2, Sch4, Xsu, Xaa, Xfa, Xc4, Xpro, Xac,
extension. The inhibition term I4 was added to account for pH Xh2, XpSRB, XaSRB and XhSRB. Their default concentrations in
inhibition (IpH,j) and Ih2s,j of pSRB, aSRB and hSRB (see ADM1 were set for the operating condition E1, whereas
concentrations for the cases E2 to E9 were calculated
similar to Eq. (6).
The initial conditions for the dynamic simulation were
Table 3 e Liquid phase yield coefficient (ni,j) and rate estimated as recommended by Rieger et al. (2012). Steady state
equations (rj) for the liquidegas transfer process added to
simulations were run and the values of the state variables at
ADM1 to model the sulfate reduction process in the
anaerobic digestion of cane-molasses vinasse. the end of this simulation period were used as initial condi-
tions for the dynamic simulation. Since this procedure as-
Component i / 9a Rate (rj, kmol m3 d1)
sumes that the reactor is operated in a typical way for an
j Process Y Sh2s,free extended period prior to the dynamic simulation (similar to
T9a H2S Transfer 1 kL a$ðSh2s;free  KH;h2s $Pgas;h2s Þ the experiments used for calibration and validation), this was
w a t e r r e s e a r c h 7 1 ( 2 0 1 5 ) 4 2 e5 4 47

Table 4 e Model based influent characterization of cane-molasses vinasse.


Components Names Units Values
Solubles
Ssu Sugar concentration kg COD m3 33.73
Saa Amino acid concentration kg COD m3 5.82
Sfa LCFA concentration kg COD m3 0.09
Sac Acetic acid concentration kg COD m3 1.36
SSI Inert concentration kg COD m3 16.97
SCOD Soluble COD concentration kg COD m3 57.97
Particulates
Xch Carbohydrate concentration kg COD m3 6.91
Xpr Protein concentration kg COD m3 0.09
Xli Lipid concentration kg COD m3 0.14
XI Inert concentration kg COD m3 0.00
Xc Composite concentration kg COD m3 0.00
XCOD Particulate COD concentration kg COD m3 7.15
TCOD Total COD kg COD m3 65.12
Total cation Total cation concentration kmol m3 0.315
Total anion Total anion concentration kmol m3 0.073

considered sufficient to establish the initial conditions for this Sensitivity analysis has been widely applied to reduce model
work (Rieger et al., 2012). complexity, to determine the significance of model parame-
ters and to identify dominant parameters (Dereli et al., 2010;
Silva et al., 2009; Tartakovsky et al., 2008).
2.4. Sensitivity analysis
The local relative sensitivity analysis method (Dochain
and Vanrolleghem, 2001) is employed here in order to
Despite the fact that all parameters affect the model output,
calculate sensitivity functions for the dynamic simulations.
the output sensitivity differs from one parameter to another.

Fig. 1 e Most sensitive model parameters arranged in descending order, for the process variables Spro, Sac, pH, Qgas, Sgas,ch4,
Sgas,co2, Sso4, Sh2s, and Sgas,h2s.
48 w a t e r r e s e a r c h 7 1 ( 2 0 1 5 ) 4 2 e5 4

The numerical calculation of sensitivity functions uses the Once a set of estimated parameters has been obtained, it
finite difference approximation (Dochain and is necessary to question the predictive quality of the
Vanrolleghem, 2001). The sensitivities are quantified in resulting model through validation (Donoso-Bravo et al.,
terms of the variation of measurable process variables 2011). Direct and cross validation are usually considered
under the perturbation of model parameters in their as steps of the model validation procedure (Donoso-Bravo
neighborhood domain (Eq. (7)). The value of the perturba- et al., 2011). Therefore, the data was divided into two sub-
tion factor d was chosen such that the differences between sets as recommended by Donoso-Bravo et al. (2011): (1) data
the resulting sensitivity values of different parameters can used during model calibration (data set D1) for direct vali-
be detected. dation, and (2) unseen data (data set D2) for cross validation.
       The accuracy of the predictions for direct and cross vali-
vyn t yn t y ðt; qm þ d$qm Þ  yn t; qm =yn ðt; qm Þ dation were determined by using the mean absolute relative
Ґn;m ðtÞ ¼ ¼ n
vqm =qm d $qm =qm error and they were classified as high (±10%) or medium
(7) (10%e30%) accurate quantitative prediction (Batstone and
Keller, 2003).
where Ґn,m is the dimensionless sensitivity value of the nth
process variable with respect to the mth model parameter; yn
(n ¼ 1, … 9) denotes the nth process variable (i.e. Spro, Sac, pH,
Qgas (gas flow), Sgas,ch4, Sgas,co2, Sso4, Sh2s, and Sgas,h2s); qm is
the mth model parameter, m ¼ 1, … 40 (see parameters in 3. Results and discussion
Appendix A); and qmþd$qm is the perturbated parameter
value. The sensitivity values for each process variable to 3.1. Sensitivity analysis
each model parameter for data set D1 (days 0e36) and data
P Steady state simulations using the ADM1 benchmark param-
set D2 (days 37e75), were computed as Ґn;m ðtÞ (expressed as
PP eter values (Rosen and Jeppsson, 2006) and values given by
% in respect to the total Ґn;m ðtÞ) and arranged in
descending order. Fedorovich et al. (2003) for sulfate reduction showed discrep-
ancies greater than 50% between the experimental results and
the model predictions.
2.5. Model calibration, parameter uncertainties and To further improve the dynamic predictions, a sensitivity
validation procedure analysis was performed under dynamic conditions in order
to determine the most important parameters to be used in
Calibration of the more sensitive model parameters is the dynamic calibration. The resulting local sensitivity
now required. Model calibration was performed on an P PP
values ( Ґn;m = Ґn;m ; expressed as %) for each process
expert ebasis by a trial and error approach, driven by variable (Spro, Sac, pH, Qgas, Sgas,ch4, Sgas,co2, Sso4, Sh2s, and
knowledge from the sensitivity analysis and using the Sgas,h2s) are shown in Fig. 1. The perturbation factor d was
parameter ranges reported in the literature as constraints. set as 1% for all the calculations as in Tartakovsky et al.
The iterative procedure reported by Dereli et al. (2010) was (2008). It is noteworthy that negative values indicated a
applied. decrease of the process variable when the parameter was
In order to provide information about the uncertainty of perturbed.
the calibrated parameters, confidence intervals (CI) for the Fig. 1 allows the identification of the most sensitive
resulting set of parameters were calculated based on the model parameters for each process variable. For example,
Fisher information matrix (FIM) (Eq. (8)) (Dochain and the process variable Spro is highly sensitive to parameters
Vanrolleghem, 2001). km,hSRB, KS,hSRB and Ysu (see nomenclature of the parame-
Xt  vy T X
1 
vyn
 ters in Appendix A), whereas Sac is highly sensitive to km,ac
FIM ¼ i¼1 vq
n
ðtÞ $ ðtÞ (8) and Yac. The fact that some model parameters affected
m vqm
several process variables at the same time (e.g. the model
P1 parameter Ysu affected the process variables Spro, pH,
where, vyn/vqm(t) are the absolute sensitivity values and
is calculated as the inverse of the covariance matrix of the Sgas,ch4 and Sgas,co2) was also useful for model calibration.
measurement error (Dochain and Vanrolleghem, 2001). Sub- Additionally, it was observed from the sensitivity analysis
sequently, the covariance matrix (COV) can be approximated that the effect of the model parameters from days 0 to 36
by the inverse of the FIM matrix (COV ¼ FIM1) and the stan- (data set D1) varied in comparison to those from days 37 to
dard deviations (s) for the parameters (qm) can be obtained by 75 (data set D2) (Fig. 1). In that sense, an increased sensi-
using Eq. (9) (Dochain and Vanrolleghem, 2001). tivity towards km,hSRB on Spro was observed near the end of
the experiments (data set D2) because of the increase of
  pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
s qm ¼ COVm;m (9) influent sulfates (see experimental conditions in Barrera
et al. (2014)), which made the sulfate reduction process
Confidence intervals for the parameters (Eq. (10)) were predominant leading to a higher sensitivity (with respect to
calculated for a confidence level of 95% (a ¼ 0.05) and the t- data set D1).
value was obtained from the Student-t distribution. Therefore, the sensitivity analysis enabled ranking the
effect of the model parameters on each process variable,
 
qm ±ta;df $s qm (10) which yields information useful for model calibration.
w a t e r r e s e a r c h 7 1 ( 2 0 1 5 ) 4 2 e5 4 49

3.2. Model calibration In contrast, fitting of KS,so4,pSRB, KS,so4,aSRB, and KS,so4,hSRB


was required to predict Sso4 as these parameters solely impact
The model was calibrated using 36 days of dynamic data (data this process variable (results not shown). Values up to 10
set D1). During these days, the organic loading rate (OLR) of times higher than those from the literature (Barrera et al.,
the upflow anaerobic sludge bed (UASB) reactor was gradually 2013) were retrieved for KS,so4,pSRB, KS,so4,aSRB, and KS,so4,hSRB.
increased from 7.66 to 12.00 kg COD m3 d1 and later reduced This observation was attributed to the use (by previous mod-
to 7.9 kg COD m3 d1. At the same time, the sulfate loading elers) of experimental observations based on organic deficient
rate (SLR) was increased from 0.36 to 0.76 kg SO4 2 m3 d1 at a substrates (SO4 2 /COD ratios  1.5) (Alphenaar et al., 1993;
constant hydraulic retention time of 4.86 days, resulting in an Omil et al., 1996, 1997) to fit models (Fedorovich et al., 2003;
increase of the SO4 2 /COD ratio from 0.05 to 0.10 (Barrera Kalyuzhnyi et al., 1998), by increasing the maximum specific
et al., 2014). uptake rate and decreasing the half saturation coefficient of
Initial values of the model state variables were taken from SRB. The half saturation coefficient for hSRB (KS,hSRB) agreed
steady state simulations at the operating conditions of E1 well with values reported (Batstone et al., 2006) and was 86% of
and the dynamic input variables were calculated from the the half saturation coefficient of hydrogenotrophic meth-
influent characterization of cane-molasses vinasse (Table 4), anogenic archea (hMA), showing that hSRB can outcompete
as illustrated in Eq. (6). hMA for hydrogen (Omil et al., 1997; Rinzema and Lettinga,
An iterative method (Dereli et al., 2010) was applied for the 1988).
calibration of the most sensitive parameters by fitting the
model to the experimental results for the process variables 3.3. Parameter uncertainty estimation
Sso4, Sh2s, Sh2s,free, Sgas,h2s, Qgas, Sgas,ch4, Sgas,co2, Spro, Sac, eff_-
COD (effluent COD) and pH. Although a larger number of The confidence intervals (CI) for the calibrated parameters are
ADM1 parameters were sensitive to the process variables shown in Appendix A. They were found to be below 20% in all
(Fig. 1), only km,pro, km,ac, km,h2, and Yh2 were used for cali- cases which yields a satisfactory confidence in the determined
bration, as the sensitivity analysis revealed them to be among set of parameters (confidence level 95%). The correlation be-
the most sensitive model parameters (Fig. 1). In this way, the tween the calibrated parameters was also calculated based on
number of calibrated ADM1 parameters was kept to a strict the covariance matrix (COV), rendering the following results:
minimum. In addition, all sulfate reduction parameters (70%
among the most sensitive parameters of Fig. 1) were cali-  Strong correlation (0.7) between the parameter pairs
brated. The estimated parameter values providing the best fit [km,aSRB, YaSRB]; [km,pro, KI,h2s,pro]; and [km,ac, KI,h2s,ac].
(based on the mean absolute relative error) between model  Moderate correlation (0.4e0.7) between the parameter
predictions and experimental results are reported in column 7 pairs [Yh2, YhSRB]; [km,pSRB, YpSRB]; [km,hSRB, KS,hSRB];
(Calibration this work) of Appendix A. All other parameters [KS,so4,aSRB, KS,so4,hSRB]; [KS,so4,hSRB, KI,h2s,h2]; [KS,so4,pSRB,
were adopted from Rosen and Jeppsson (2006). KI,h2s,aSRB]; [KS,so4,pSRB, KI,h2s,hSRB]; and [km,hSRB, KI,h2s,hSRB].
During calibration, the values obtained for km,pro, km,ac, and
km,h2 were in agreement with values used to calibrate the
anaerobic digestion of cane-molasses (Romli et al., 1995) 3.4. Direct validation
(Column 5, Appendix A). The fact that parameter values used
for calibration in Romli et al. (1995) were used for calibration in The deviation between model predictions and experimental
this work, was likely because of the similar characteristics of observations was used for direct validation. The results are
both substrates (cane-molasses and cane-molasses vinasse, shown in Fig. 2. It can be seen that the process variables
respectively), which favored the uptake rate of propionate, (except for Spro) were predicted quite well after the model
acetate and hydrogen leading to required modification (in calibration (Fig. 2AeH), exhibiting a mean absolute relative
respect to ADM1 parameter values) of km,pro, km,ac, and km,h2 error below 10% (1%e9.7%), which is considered as a high
during the calibration in this work (Appendix A). accuracy quantitative prediction (Batstone and Keller, 2003).
Concerning the calibration of the sulfate reduction pa- However, deviations between model predictions and
rameters, the yield coefficients, the Monod maximum spe- experimental values for Spro (Fig. 2G) led to a mean absolute
cific uptake rates and the half saturation coefficients were relative error higher than 30%, which can be considered as a
found in the range of values found in the literature (Barrera qualitative prediction that can demonstrate the overall
et al., 2013). However, the 50% inhibitory concentrations of qualitative response of the system (Batstone and Keller,
free H2S (Appendix A) were lower than the values used to 2003).
calibrate the sulfate reduction processes (Barrera et al., 2013), The increase of the OLR and the influent COD concentra-
but similar to experimental values reported as inhibitory tion in a UASB reactor fed with vinasse caused an increase of
(150 mg S L1 (0.0047 kmol m3)) for methanogens and SRB the propionic acid concentration (Harada et al., 1996). How-
(except for propionate degraders, which is 70 mg S L1 ever, in the experimental values used for calibration in this
(0.0022 kmol m3)) (Rinzema and Lettinga, 1988). These work, Spro remained constant (see experimental values in
values agreed well with the experimental observations used Fig. 2G) when the OLR was increased on day 8 (Barrera et al.,
for calibration in this work (Barrera et al., 2014) and therefore 2014). This was likely because sludge in the UASB reactor
they can be considered a better approximation for the real assimilated the increase of the OLR by degrading the excess
phenomena. propionate. Despite the fact that the model could not predict
50 w a t e r r e s e a r c h 7 1 ( 2 0 1 5 ) 4 2 e5 4

Fig. 2 e Comparison between experimental values and model predictions after the model calibration: (A) Sso4, (B) Sh2s, (C)
Sh2s,free, (D) Sgas,h2s, (E) Qgas, (F) Sgas,ch4 & Sgas,co2, (G) Spro & Sac, and (H) eff_COD & pH.

this observation (Fig. 2G), the over prediction observed for Spro 3.5. Cross validation
during days 8e15 when the OLR increased was in agreement
with the phenomenon described by Harada et al. (1996). A cross validation study was performed to assess the quality
Moreover, the under prediction of Spro during days 23e25 can and applicability of the calibrated model. The model outputs
be attributed to the slight under prediction of Sh2s,free (con- were compared with data set D2 (days 37e75) under the
stant at the concentration of 0.0018 kmol m3) that reduces operating conditions E5, E6, E7, E8 and E9 (Barrera
the inhibitory effect of sulfide on propionate degrading bac- et al., 2014) without changing the previously optimized
teria during the simulation (Fig. 2C). parameter set. During these periods, the OLR and SLR of the
Additionally, the over prediction of Sh2s during days 21e27 UASB reactor were in the range of 7.72e10.69 kg COD m3 d1
(Fig. 2B) can be attributed to hydrogen sulfide loss during the and 0.76e1.57 kg SO4 2 m3 d1, respectively (Barrera et al.,
experiments used for calibration as the sulfur recovery in the 2014). SO4 2 /COD ratios of 0.10, 0.15 and 0.20 were applied in
reactor outlet streams decreases from 100% to 90% (Barrera the periods covering the validation study unlike the periods
et al., 2014). used for the calibration study.
w a t e r r e s e a r c h 7 1 ( 2 0 1 5 ) 4 2 e5 4 51

Fig. 3 e Validation of model predictions with experimental values: (A) Sso4, (B) Sh2s, (C) Sh2s,free, (D) Sgas,h2s, (E) Qgas, (F) Sgas,ch4
& Sgas,co2, (G) Spro & Sac, and (H) eff_COD & pH.

Fig. 3 (B, C, F, G & H) presents the comparison of model pre- during these periods (Fig. 3D and E). This was likely due to
dictions and experimental values for the process variables dur- the assumption of a constant gaseliquid transfer coefficient
ing the validation study. As can be seen, Sh2s, Sh2s,free, Sgas,ch4, (200 d1) for H2S even when the biogas production rate, and
Sgas,co2, Sac, and pH were well predicted by the model showing a consequently its stripping effect, decreased after day 60 in
mean absolute relative error below 10% (1%e10%), which is the model predictions (Fig. 3E). The over and under pre-
considered as a highly accurate quantitative prediction (±10%) dictions of Sh2s,free from days 40 to 45 and 48 to 55, respec-
(Batstone and Keller, 2003). A medium accurate quantitative tively (Fig. 3C), were attributed to slight deviations (±3.1%)
prediction (10%e30%) was achieved for Spro, Sso4, Sgas,h2s, Qgas, in the pH prediction (Fig. 3H).
and eff_COD (Fig. 3 A, E, G & H), as the mean absolute relative During the validation period the model was also able to
error ranged from 12% to 26% (Batstone and Keller, 2003). predict the reactor failure (for methanogenesis and sulfi-
The underestimation observed for Sso4 during days 51e62 dogenesis) from days 70 to 75. Methanogenesis failure in
and 73 to 75 (Fig. 3A), was in agreement with the lower the model predictions was evidenced by a pH decrease due
Sh2s,free predicted during these days (Fig. 3C), which reduced to Sac increase, which led to a decrease of Qgas and Sgas,ch4
the inhibitory effect of Sh2s,free on SRB during the simula- (Fig. 3E, F, G & H). At the same time, sulfidogenesis failure in
tion. The excess consumption of sulfate was accumulated in the model prediction was evidenced by the increase of Sso4
the gas phase since higher Sgas,h2s and Qgas were predicted while Sh2s remained constant, showing that the increase in
52 w a t e r r e s e a r c h 7 1 ( 2 0 1 5 ) 4 2 e5 4

the SLR resulted in accumulation of sulfates in the effluent 26%. Moreover, the model was able to predict failure of meth-
rather than conversion to hydrogen sulfide (Fig. 3A and B). anogenesis and sulfidogenesis when the sulfate loading rate
In addition, the model predicted an Sh2s,free increase as a increased. Therefore, the kinetic parameters and the model
result of a pH decrease showing a severe sulfide inhibition structure proposed in this work can be considered as valuable to
during these days (Fig. 3C and H). A discussion on the effect describe the sulfate reduction process in the anaerobic digestion
of the high sulfate concentrations on the methane pro- of cane-molasses vinasse, by predicting the sulfur compounds in
duction rates as well as on the inhibition effects of H2S to the gas and liquid phases, increasing the applicability of ADM1 to
the anaerobic consortia (methanogens and sulfate reducing specific industrial wastewaters (vinasse).
bacteria) can be found in Barrera et al. (2014).

Acknowledgments
4. Conclusions
The authors wish to express their gratitude to Dr. Jeppsson Ulf
An extension of ADM1 with sulfate reduction was proposed,
(Department of Industrial Electrical Engineering and Auto-
calibrated and validated for the description of the anaerobic
mation, Lund University, Sweden) for kindly providing the
digestion of cane-molasses vinasse (high strength and sulfate
codes and documents related to ADM1. This work has been
rich wastewater). Based on the results of a sensitivity analysis,
supported financially by the VLIR Program, Project
only four parameters of the original ADM1 (km,pro, km,ac, km,h2
ZEIN2009PR362 (Biogas production from waste from local
and Yh2) and all the sulfate reduction parameters were fitted
food, wood and sugar cane industries for increasing self-
during calibration. Despite the fact that some deviations were
sufficiency of energy in Sancti Spiritus, Cuba).
observed between model predictions and experimental values, it
was shown that the process variables Sso4, Sh2s, Sh2s,free, Sgas,h2s,
Qgas, Sgas,ch4, Sgas,co2, Spro, Sac, eff_COD, and pH were predicted Appendix A. Stoichiometric and kinetic
reasonably well during model validation. The model showed parameters selected for sensitivity analysis.
high (±10%) to medium (10%e30%) accurate quantitative pre- Values reported in literature and calibrated for
dictions with a mean absolute relative error ranging from 1 to this work.

Parameter Names Units Benchmark Cane- Sulfate Calibration CI


valuesa molassesb reductionc this work (95%)
Ysu Yield of sugar degraders kg COD_Xsu kg COD_S1
su 0.100 0.100
(Xsu)
Yaa Yield of amino acids kg COD_Xaa kg COD_S1
aa 0.080 0.080
degraders (Xaa)
Yfa Yield of LCFA degraders kg COD_Xfa kg COD_S1
fa 0.060 0.060
(Xfa)
1
Yc4 Yield of valerate and kg COD_Xc4 kg COD_Sva & bu 0.060 0.060
butyrate degraders (Xc4)
Ypro Yield of propionate kg COD_Xpro kg COD_S1
pro 0.040 0.040
degraders (Xpro)
Yac Yield of acetate degraders kg COD_Xac kg COD_S1
ac 0.050 0.050
(Xac)
Yh2 Yield of hydrogen degraders kg COD_Xh2 kg COD_S1
h2 0.060 0.070 ±0.0042
(Xh2)
YpSRB Yield of pSRB (XpSRB) kg COD_XpSRB kg COD_S1
pro 0.0329d 0.027e0.035 0.035 ±0.0040
YaSRB Yield of aSRB (XaSRB) kg COD_XaSRB kg COD_S1
ac 0.0342d 0.033e0.041 0.041 ±0.0060
YhSRB Yield of hSRB (XhSRB) kg COD_XhSRB kg COD_S1
h2 0.0366d 0.037e0.077 0.051 ±0.0047
km,su Monod maximum specific kg COD_Ssu kg COD_X1
su d
1
30 30
uptake rate of sugars by Xsu
km,fa Monod maximum specific kg COD_Sfa kg COD_X1
fa d
1
6 6
uptake rate of LCFA by Xfa
km,c4 Monod maximum specific kg COD_Sva & bu kg 20 20
uptake rate of HVa & HBu by COD_X1
c4 d
1

Xc4
km,pro Monod maximum specific kg COD_Spro kg COD_X1
pro d
1
13 15 16 ±1.21
uptake rate of HPr by Xpro
km,ac Monod maximum specific kg COD_Sac kg COD_X1
ac d
1
8 9.4 12 ±0.73
uptake rate of HAc by Xac
km,h2 Monod maximum specific kg COD_Sh2 kg COD_X1
h2 d
1
35 43 43 ±4.26
uptake rate of H2 by Xh2
km,pSRB Monod maximum specific kg COD_Spro kg COD_X1
pSRB d
1
12.6d 9.60e23.1 23 ±2.35
uptake rate of HPr by pSRB
w a t e r r e s e a r c h 7 1 ( 2 0 1 5 ) 4 2 e5 4 53

e (continued )
Parameter Names Units Benchmark Cane- Sulfate Calibration CI
valuesa molassesb reductionc this work (95%)
km,aSRB Monod maximum specific kg COD_Sac kg COD_X1
aSRB d
1
7.1d 4.19e18.5 18.5 ±2.32
uptake rate of HAc by aSRB
km,hSRB Monod maximum specific kg COD_Sh2 kg COD_X1
hSRB d
1
26.7d 26.7e64.9 63 ±7.81
uptake rate of H2 by hSRB
kdec,Xac First order decay rate for Xac d1 0.020 0.020
kdec,Xh2 First order decay rate for Xh2 d1 0.020 0.020
KS,su Half saturation coefficient kg COD_Ssu m3 0.500 0.500
for the uptake of sugars by
Xsu
KS,aa Half saturation coefficient kg COD_Saa m3 0.300 0.300
for the uptake of amino
acids by Xaa
KS,c4 Half saturation coefficient kg COD_Sva & bu m3 0.200 0.200
for the uptake of HVa & HBu
by Xc4
KS,pro Half saturation coefficient kg COD_Spro m3 0.100 0.100
for the uptake of HPr by Xpro
KS,ac Half saturation coefficient kg COD_Sac m3 0.150 0.150
for the uptake of HAc by Xac
KS,h2 Half saturation coefficient kg COD_Sh2 m3 7.0e-6 7.0e-6
for the uptake of H2 by Xh2
KS,pSRB Half saturation coefficient kg COD_Spro m3 0.110d 0.015e0.295 0.110 ±0.010
for the uptake of HPr by
pSRB
KS,aSRB Half saturation coefficient kg COD_Sac m3 0.220d 0.024e0.220 0.120 ±0.015
for the uptake of HAc by
aSRB
KS,hSRB Half saturation coefficient kg COD_Sh2 m3 0.000100d 4.0e-6e e 1.0e-4 6e-06 ±6.2e-7
for the uptake of H2 by hSRB
KS,so4,pSRB Half saturation coefficient kmol m3 0.000200d 7.7e-5 e 2.0e-4 0.00200 ±0.00039
for the uptake of SO4 2 by
pSRB
KS,so4,aSRB Half saturation coefficient kmol m3 0.000100d 1.0e-4 e 2.9e-4 0.00100 ±0.00023
for the uptake of SO4 2 by
aSRB
KS,so4,hSRB Half saturation coefficient kmol m3 0.000104d 9.0e-6 e 1.0e-4 0.00105 ±0.00017
for the uptake of SO4 2 by
hSRB
KI,h2s,c4 50% inhibitory kmol m3 0.00750d 0.0075d 0.00440 ±0.00065
concentration of free H2S
for Xc4
KI,h2s,pro 50% inhibitory kmol m3 0.00750d 0.0075d 0.00280 ±0.00048
concentration of free H2S
for Xpro
KI,h2s,ac 50% inhibitory kmol m3 0.00720d 0.0072d 0.00440 ±0.00053
concentration of free H2S
for Xac
KI,h2s,h2 50% inhibitory kmol m3 0.00630d 0.0063d 0.00440 ±0.00075
concentration of free H2S
for Xh2
KI,h2s,pSRB 50% inhibitory kmol m3 0.00813d 0.0058e0.0089 0.00480 ±0.00086
concentration of free H2S
for pSRB
KI,h2s,aSRB 50% inhibitory kmol m3 0.00780d 0.0051e0.018 0.00470 ±0.00028
concentration of free H2S
for aSRB
KI,h2s,hSRB 50% inhibitory kmol m3 0.00780d 0.0078e0.017 0.00470 ±0.00083
concentration of free H2S
for hSRB

CI: confidence interval.


a
Rosen and Jeppsson (2006).
b
Romli et al. (1995).
c
Barrera et al. (2013).
d
Fedorovich et al. (2003).
e
Batstone et al. (2006).
54 w a t e r r e s e a r c h 7 1 ( 2 0 1 5 ) 4 2 e5 4

references population dynamics between sulfate reducing and


methanogenic bacteria. Biodegradation 9 (3), 187e199.
Kalyuzhnyi, S.V., Fedorovich, V.V., 1998. Mathematical modelling
of competition between sulphate reduction and
Alphenaar, P., Visser, A., Lettinga, G., 1993. The effect of liquid
methanogenesis in anaerobic reactors. Bioresour. Technol. 65
upward velocity and hydraulic retention time on granulation
(3), 227e242.
in UASB reactors treating wastewater with a high sulphate
Knobel, A.N., Lewis, A.E., 2002. A mathematical model of a high
content. Bioresour. Technol. 43 (3), 249e258.
sulphate wastewater anaerobic treatment system. Water Res.
Barrera, E.L., Spanjers, H., Dewulf, J., Romero, O., Rosa, E., 2013.
36 (1), 257e265.
The sulfur chain in biogas production from sulfate-rich liquid ve, J., Impe, J.F.V.,
Lauwers, J., Appels, L., Thompson, I.P., Degre
substrates: a review on dynamic modeling with vinasse as
Dewil, R., 2013. Mathematical modelling of anaerobic
model substrate. J. Chem. Technol. Biotechnol. 88, 1405e1420.
digestion of biomass and waste: power and limitations. Prog.
Barrera, E.L., Spanjers, H., Romero, O., Rosa, E., Dewulf, J., 2014.
Energy Combust. Sci. 39, 383e402.
Characterization of the sulfate reduction process in the
Omil, F., Lens, P., Hulshoff Pol, L.W., Lettinga, G., 1996. Effect of
anaerobic digestion of a very high strength and sulfate rich
upward velocity and sulphide concentration on volatile fatty
vinasse. Chem. Eng. J. 248, 383e393.
acid degradation in a sulphidogenic granular sludge reactor.
Batstone, D., 2006. Mathematical modelling of anaerobic reactors
Process Biochem. 31, 699e710.
treating domestic wastewater: rational criteria for model use.
Omil, F., Lens, P., Hulshoff Pol, L.W., Lettinga, G., 1997.
Rev. Environ. Sci. Biotechnol. 5 (1), 57e71.
Characterization of biomass from a sulfidogenic, volatile fatty
Batstone, D.J., Keller, J., 2003. Industrial applications of the IWA
acid-degrading granular sludge reactor. Enzyme Microb.
anaerobic digestion model no. 1 (ADM1). Water Sci. Technol.
Technol. 20, 229e236.
47 (12), 199e206.
Rieger, L., Gillot, S., Langergraber, G., Shaw, A., 2012. Good
Batstone, D.J., Keller, J., Angelidaki, I., Kalyuzhnyi, S.V.,
Modelling Practice: Guidelines for Use of Activated Sludge
Pavlostathis, S.G., Rozzi, A., Sanders, W.T.M., Siegrist, H.,
Models. IWA Publishing.
Vavilin, V.A., 2002. The IWA Anaerobic Digestion Model No 1.
Rinzema, A., Lettinga, G., 1988. Anaerobic treatment of sulfate-
Scientific and Technical Report No. 13. IWA Publishing,
containing waste water. In: Wise, D.L. (Ed.), Biotreatment
London.
Systems, vol. III. CRC Press Inc, Boca Raton, USA, pp. 65e109.
Batstone, D.J., Keller, J., Steyer, J.P., 2006. A review of ADM1
Ristow, N.E., Whittington-Jones, K., Corbett, C., Rose, P.,
extensions, applications, and analysis: 2002e2005. Water Sci.
Hansford, G.S., 2002. Modelling of a recycling sludge bed
Technol. 54 (4), 1e10.
reactor using AQUASIM. Water S.A 28 (1), 111e120.
Dereli, R.K., Ersahin, M.E., Ozgun, H., Ozturk, I., Aydin, A.F., 2010.
Romli, M., Keller, J., Lee, P.J., Greenfield, P.F., 1995. Model
Applicability of anaerobic digestion model no. 1 (ADM1) for a
prediction and verification of a two stage high-rate anaerobic
specific industrial wastewater: opium alkaloid effluents.
wastewater treatment system subject to shock loads. Process
Chem. Eng. J. 165 (1), 89e94.
Saf. Environ. Prot. 73, 151e154.
Dochain, D., Vanrolleghem, P.A., 2001. Dynamical Modelling and
Rosen, C., Jeppsson, U., 2006. Aspects on ADM1 Implementation
Estimation in Wastewater Treatment Processes. IWA
within the BSM2 Framework. Department of Industrial
Publishing, London, UK.
Electrical Engineering and Automation (IEA). Lund University,
Donoso-Bravo, A., Mailier, J., Martin, C., Rodrı́guez, J., Aceves-
Lund, Sweden.
Lara, C.A., Wouwer, A.V., 2011. Model selection, identification
Silva, F., Nadais, H., Prates, A., Arroja, L., Capela, I., 2009.
and validation in anaerobic digestion: a review. Water Res. 45
Modelling of anaerobic treatment of evaporator condensate
(17), 5347e5364.
(EC) from a sulphite pulp mill using the IWA anaerobic
Fedorovich, V., Lens, P., Kalyuzhnyi, S., 2003. Extension of
digestion model no. 1 (ADM1). Chem. Eng. J. 148, 319e326.
anaerobic digestion model no. 1 with processes of sulfate
Tartakovsky, B., Mu, S.J., Zeng, Y., Lou, S.J., Guiot, S.R., Wu, P.,
reduction. Appl. Biochem. Biotechnol. 109, 33e45.
2008. Anaerobic digestion model no. 1-based distributed
Harada, H., Uemura, S., Chen, A.-C., Jayadevan, J., 1996. Anaerobic
parameter model of an anaerobic reactor: II. model validation.
treatment of a recalcitrant distillery wastewater by a
Bioresour. Technol. 99 (9), 3676e3684.
thermophilic UASB reactor. Bioresour. Technol. 55 (3),
Visser, A., Hulshoff Pol, L.W., Lettinga, G., 1996. Competition of
215e221.
methanogenic and sulfidogenic bacteria. Water Sci. Technol.
Hinken, L., Pato n Gasso  , M., Weichgrebe, D., Rosenwinkel, K.H.,
33 (3), 99e110.
2013. Implementation of Sulphate Reduction and Sulphide
Zub, S., Kurissoo, T., Menert, A., Blonskaja, V., 2008. Combined
Inhibition in ADM1 for Modelling of a Pilot Plant Treating
biological treatment of high-sulphate wastewater from yeast
Bioethanol Wastewater. Santiago de Compostela. Spain.
production. Water Environ. J. 22 (4), 274e286.
Kalyuzhnyi, S., Fedorovich, V., Lens, P., Hulshoff Pol, L.,
Lettinga, G., 1998. Mathematical modelling as a tool to study

Vous aimerez peut-être aussi