Sie sind auf Seite 1von 44

Accepted Manuscript

Optimal operating conditions for maximum biogas production in anaerobic bioreactors

W. Balmant, B.H. Oliveira, D.A. Mitchell, J.V.C. Vargas, J.C. Ordonez

PII: S1359-4311(13)00667-4
DOI: 10.1016/j.applthermaleng.2013.09.033
Reference: ATE 5049

To appear in: Applied Thermal Engineering

Received Date: 9 July 2013


Revised Date: 15 September 2013
Accepted Date: 19 September 2013

Please cite this article as: W. Balmant, B.H. Oliveira, D.A. Mitchell, J.V.C. Vargas, J.C. Ordonez,
Optimal operating conditions for maximum biogas production in anaerobic bioreactors, Applied Thermal
Engineering (2013), doi: 10.1016/j.applthermaleng.2013.09.033.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT
Optimal operating conditions for maximum biogas

production in anaerobic bioreactors

W. Balmant1, B. H. Oliveira1, D. A. Mitchell1, J. V. C. Vargas2,*,J. C. Ordonez3


1
Departamento de Bioquímica e Biologia Molecular, Universidade Federal do

PT
Paraná, C. P. 19011, Curitiba, PR 81531-980, Brasil
2
Departamento de Engenharia Mecânica, Núcleo de P&D em Energia

RI
Autossustentável, Universidade Federal do Paraná, C. P. 19011, Curitiba, PR 81531-

SC
980, Brasil
3
Department of Mechanical Engineering, Energy & Sustainability Center, and Center

U
for Advanced Power Systems, Florida State University, Tallahassee, Florida 32310-
AN
6046, USA

E-mails: wbalmant@gmail.com, behaddad14@yahoo.com.br,


M
davidmitchell@ufpr.br, jvargas@demec.ufpr.br, ordonez@caps.fsu.edu
D
TE

Abstract

The objective of this paper is to demonstrate the existence of optimal


EP

residence time and substrate inlet mass flow rate for maximum methane production

through numerical simulations performed with a general transient mathematical


C

model of an anaerobic biodigester introduced in this study. It is herein suggested a


AC

*
Corresponding author: Tel. +55 41 33613307; fax: +55 41 33613129
E-mail address: jvargas@demec.ufpr.br
ACCEPTED MANUSCRIPT
simplified model with only the most important reaction steps which are carried out

by a single type of microorganisms following Monod kinetics. The mathematical

model was developed for a well mixed reactor (CSTR – Continuous Stirred-Tank

Reactor), considering three main reaction steps: acidogenesis, with a µmax of 8.64

day-1 and a KS of 250 mg/L, acetogenesis, with a µmax of 2.64 day-1 and a KS of 32

PT
mg/L, and methanogenesis, with a µmax of 1.392 day-1 and a KS of 100 mg/L. The

yield coefficients were 0.1-g-dry-cells/g-pollymeric compound for acidogenesis, 0.1-

RI
g-dry-cells/g-propionic acid and 0.1-g-dry-cells/g-butyric acid for acetogenesis and

SC
0.1 g-dry-cells/g-acetic acid for methanogenesis. The model describes both the

transient and the steady-state regime for several different biodigester design and

U
operating conditions. After model experimental validation, a parametric analysis was
AN
performed. It was found that biogas production is strongly dependent on the input

polymeric substrate and fermentable monomer concentrations, but fairly independent


M
of the input propionic, acetic and butyric acid concentrations. An optimisation study
D

was then conducted and optimal residence time and substrate inlet mass flow rate
TE

were found for maximum methane production. The optima found were very sharp,

showing a sudden drop of methane mass flow rate variation from the observed
EP

maximum to zero, within a 20% range around the optimal operating parameters,

which stresses the importance of their identification, no matter how complex the
C

actual bioreactor design may be. The model is therefore expected to be a useful tool
AC

for simulation, design, control and optimisation of anaerobic biodigesters.


ACCEPTED MANUSCRIPT
Keywords: Specific growth rates; performance coefficients; anaerobic bioreactor

Nomenclature

CSTR Continuous stirred-tank reactor

F Total molar flow rate of gas out of the gas phase, mol day-1

PT
G3 H2 partial pressure, atm

G4 CO2 partial pressure, atm

RI
G7 Methane partial pressure, atm

SC
H3 Henry’s constant for H2, g L-1.atm-1

H4 Henry’s constant for CO2, g L-1.atm-1

U
H7 Henry’s constant for methane, g L-1.atm-1

HRT
AN
Hydraulic retention time, V/Q, day

Kh Constant for first order polymer hydrolysis, day-1


M
KS1 Saturation constant for acidogens, g L-1
D

KS2 Saturation constant for syntrophs A, g L-1


TE

KS3 Saturation constant for hydrogenotrophic methanogens

regarding H2, g L-1


EP

KS4 Saturation constant for hydrogenotrophic methanogens

regarding CO2, g L-1


C

KS5 Saturation constant for acetoclastic methanogens, g L-1


AC

KS6 Saturation constant for syntrophs B, g L-1


ACCEPTED MANUSCRIPT
Kla3 Gas-liquid mass transfer coefficient for H2, day-1

Kla4 Gas-liquid mass transfer coefficient for CO2, day-1/

Kla7 Gas-liquid mass transfer coefficient for methane, day-1

MH2 Molar mass of H2, g mol-1

M CO 2 Molar mass of CO2, g mol-1

PT
M CH 4 Molar mass of methane, g mol-1

RI
pT Total pressure (gas phase), atm

Q Substrates input volumetric flow rate, m3 day-1

SC
Qg Biogas output volumetric flow rate, 22.4 × F , L day-1

r Reaction rate, day-1

U
R Universal gas constant, atm.L.mol-1 K-1

Sj
AN
Concentration of substance j, g L-1

Sj* Liquid phase saturation concentration of substance j, g L-1


M
Sj,ent Input concentration, g L-1
D

t Time, days
TE

T Temperature, K

V Liquid phase volume, L


EP

Vg Gas phase volume, L

xi Molar fraction of component i, G i / p T


C

X1 Acetogen concentration, g L-1


AC

X2 Syntroph A concentration, g L-1


ACCEPTED MANUSCRIPT
X3 Hydrogenotrophic methanogen concentration, g L-1

X4 Acetoclastic methanogen concentration, g L-1

X5 Syntroph B concentration, g L-1

YSi/Sp Yield of Si from Sp, g g-1

YSi/Xp Yield of Si from Xp, g g-1

PT
Greek symbols

µ Specific growth rate, day -1

RI
SC
Subscript

ent Inlet

U
j 0 polymeric phase

1 fermentable monomer
AN
2 propionic acid
M
3 liquid phase H2
D

4 liquid phase CO2


TE

5 acetic acid

6 butyric acid
EP

7 liquid phase methane

k 1 acidogens
C

2 syntrophs A
AC

3 hydrogenotrophic methanogens
ACCEPTED MANUSCRIPT
4 acetoclastic methanogens

5 syntrophs B

max maximum

opt optimal

PT
1. Introduction

RI
One of the goals of anaerobic digestion is the production of methane, which

can be converted into electricity by its combustion. The production of biogas, which

SC
is rich in methane, by such a process involves a consort of microorganisms that

U
degrade organic substrates present in biological wastes. Not only is biogas a clean-

AN
burning fuel but it also has a lower heating value close to the lower heating value of

methane, i.e., 50.28 MJ kg-1 [1]. Such value depends strongly on the proportion of
M
methane present in the mixture of gases, which is composed mostly by methane and

carbon dioxide.
D

Hence the importance of biodigestion could be summarized by: while being


TE

an efficient waste treatment method, reducing the organic load of the waste stream, it

also produces an environmentally friendly fuel [2]. Moreover, the remaining


EP

residues, both liquid and solid, can be used as biofertilisers [3].


C

Whenever anaerobic digestion is selected as the process to generate


AC

electricity by the combustion of biogas, the production rate of methane should be

maximized. In that direction, Marcos et al. [4] experimentally found optimal load
ACCEPTED MANUSCRIPT
rates for maximal biodegradation rates and methane production obtained from the

anaerobic co-digestion of solid (e.g., fat, intestines, rumen, bowels, whiskers) and

liquid (e.g., blood, washing water, manure) wastes of the meat industry, particularly

the ones rising from the municipal slaughterhouse of Badajoz, Spain.

Although mathematical models are efficient tools used to optimise a process,

PT
it should be noted that to obtain reliable parameters of an anaerobic digestion is very

RI
challenging. The complexity is mainly due to the large number of microorganisms

and compounds (i.e. large number of parameters) [5], which are summarized in Fig.

SC
1, therefore, determining accurate values for a specific bioreactor and substrate

would be highly onerous. Therefore, for the purpose of model development and

U
concision, the organic material can be classified into carbohydrates, proteins, fats
AN
and inert compounds. Figure 1 sketches the hydrolysis reactions for the first three,
M
producing sugars, amino acids and long chain fatty acids, respectively.

Several mathematical models based on each biochemical step involved in the


D

process and address variables that affect biogas production (e.g., temperature) have
TE

been proposed. Valle-Guadarrama [6] proposed a model, based on thermodynamic

principles, that predicts temperature changes in a pilot plant thermophilic anaerobic


EP

digester. Fixed and variable overall heat transfer coefficient values were used, in
C

good agreement with experimental data, concluding that temperature variation was
AC

affected by the heterogeneity of the feeding and extraction processes, by the

heterogeneity of the digestate recirculation through the heating system and by the
ACCEPTED MANUSCRIPT
lack of a perfect mixing inside the biodigester tank. However, the use of a searching

routine based on a minimal square optimization criterion for parameters

determination, model complexity and computational time are issues to be considered

when control and optimization are the objectives of the model.

In general, the published models assume uniform temperature within the

PT
reactor defining a class of models that the technical literature is rich of. One of the

RI
most complete models that have been proposed is the so called Anaerobic Digestion

Model No. 1 (ADM1), which is a structured model that includes multiple steps

SC
describing biochemical as well as physicochemical processes, but includes 26

dynamic state concentration variables, and 8 implicit algebraic variables per reactor

U
vessel or element, and implemented as differential equations only, there are 32
AN
dynamic concentration state variables [5].
M
The work of Vavilin et al. [7] is considered representative, and was published

as a first user-friendly application made available for general use, which was a new
D

version of the simulation model <METHANE> developed earlier [8, 9]. The model
TE

was experimentally validated by direct comparison to experimental data published by

Salminen et al. [10]. However, the model contained 31 ordinary differential


EP

equations (ODE) with respect to time with a number of kinetic and physical
C

parameters to be adjusted according to the particular system under analysis, i.e.,


AC

computational time could still be considered an issue if multiple cases are to be run

such as in control and optimization procedures. More recently, Vavilin et al. [11]
ACCEPTED MANUSCRIPT
added complexity to their previous models by presenting a new multidimensional (3

and 2-dimensional) anaerobic digestion model for a cylindrical reactor with non-

uniform influent concentration distributions, and successfully studied the way in

which mixing intensity affects the efficiency of continuous-flow anaerobic digestion.

Simpler models have also been published, such as seen in the work of

PT
Martinez et al. [12], in which the biodigestion process is described by only seven

RI
ODEs, but the authors stated they used a high time consuming genetic algorithm

followed by a gradient descendent procedure to adjust the model parameters in order

SC
to match the results to experimental data. One of the possible reasons of the need for

such fine parameter adjustments is that the model was dramatically simplified, not

U
considering for example the two-step kinetics of polymer hydrolysis, which is known
AN
to be important since it accounts for the balance between the rates of polymer
M
hydrolysis and methanogenesis.

Based on the bibliographic review, it is recognized that there are plenty of


D

published models that have been experimentally validated and are very successful in
TE

representing in detail the physical phenomena that occur in biodigestion processes.

Also, in the other extreme, mathematically very simple models were found, but that
EP

require complex refinement algorithms to be reliable.


C

Regarding biodigester optimization studies, very few studies have been


AC

published, and most of them are experimental. As pointed out earlier in the text, the

work of Marcos et al. [4] is one example. However, within the knowledge of the
ACCEPTED MANUSCRIPT
authors, there are no published studies to find general optimal residence time and

substrate inlet mass flow rate for maximum biogas production in the technical

literature.

Therefore, the aim of the present work is to demonstrate the existence of

optimal residence time and substrate inlet mass flow rate for maximum biogas

PT
production in biodigesters, through a simple enough mathematical model which lies

RI
between the two extremes discussed previously, i.e., that is neither too complex nor

too simple, so that practically meaningful results could be obtained directly. The

SC
model should involve as few parameters as possible, and still be representative of the

basic physical phenomena driving the process.

U
AN
2. Mathematical model
M
The reactions and processes occurring in the anaerobic digestion system are

simplified into the following general steps, which are shown in Fig. 2:
D

1. The polymeric substrate (S0) is hydrolyzed enzymatically, producing


TE

fermentable monomers (S1);

2. The fermentable monomers (S1) are transformed into propionic acid (S2),
EP

soluble hydrogen (S3), soluble carbon dioxide (S4), acetic acid (S5) and
C

butyric acid (S6) by acidogenic bacteria (X1);


AC

3. The propionic acid (S2) is transformed into H2 (S3), CO2 (S4) and acetic acid

(S5) by syntrophic bacteria A (X2);


ACCEPTED MANUSCRIPT
4. Acetic acid (S5) is transformed into methane (S7) and CO2 (S4) by

acetoclastic methanogenic bacteria (X4);

5. The butyric acid (S6) is transformed into H2 (S3) and acetic acid (S5) by

syntrophic bacteria B (X5);

6. CO2 and H2 are used by the hydrogenotrophic-methanogenic bacteria (X3) to

PT
generate methane (S7), and

7. Transfer of CO2, H2 and methane between the gas and liquid phases of the

RI
bioreactor.

SC
The model considers the following assumptions:

1. Gases behave as ideal gases;

U
2. The gas phase is well mixed;
AN
3. The total pressure in the gas phase is constant;

4. The temperature of the systems is uniform and constant;


M
5. The biogas only contains CO2, H2 and methane, and
D

6. The vapour pressure of water in the gas phase is negligible.


TE

2.1. Rate expressions


EP

The enzymatic hydrolysis of the polymeric substrate is considered to be a

first order process, as follows:


C

r1 = K h .S0 (1)
AC

The growth of each population of bacteria is directly proportional to their

concentration, according to the following equations:


ACCEPTED MANUSCRIPT
r2 = µ1.X1 (acidogens) (2)

r8 = µ 2 .X 2 (syntrophs A) (3)

r10 = µ3.X 3 (hydrogenotrophic methanogens) (4)

r11 = µ4.X4 (acetoclastic methanogens) (5)

r14 = µ5 .X5

PT
(syntrophs B) (6)

RI
In each case the specific growth rate depends on the relevant substrate

SC
concentration according to the Monod equation [13], as follows:

µ max 1.S1
µ1 = (acidogens) (7)
K S1 + S1

U
µ max 2 .S2
µ2 =
K S 2 + S2
AN (syntrophs A) (8)

µ max 3 .S3 S4
µ3 = ⋅
M
(hydrogenotrophic methanogens) (9)
K S3 + S3 K S4 +S4

µ max 4 .S5
D

µ4 = (acetoclastic methanogens) (10)


K S5 + S5
TE

µ max 5 .S6
µ5 = (syntrophs B) (11)
K S6 + S6
EP

The various products are treated as being growth-associated, including production of


C

propionic acid by acidogenic bacteria,


AC

r3 = YS2 / X1.µ1.X1 (12)


ACCEPTED MANUSCRIPT
production of H2 by acidogenic bacteria,

r4 = YS3 / X1.µ1.X1 (13)

production of CO2 by acidogenic bacteria,

r5 = YS4 / X1.µ1.X1 (14)

production of acetic acid by acidogenic bacteria,

PT
r6 = YS5 / X1.µ1.X1 (15)

RI
production of butyric acid by acidogenic bacteria,

r7 = YS6 / X1.µ1.X1 (16)

SC
production of H2 by syntrophic bacteria A,

U
r9 = YS3 / X 2 .µ 2 .X 2 (17)

AN
production of CO2, by the acetoclastic methanogens,

r12 = YS4 / X 4 .µ 4 .X 4 (18)


M
production of acetic acid by syntrophic bacteria B,

r13 = YS5 / X5 .µ5 .X5


D

(19)
TE

production of H2 by syntrophic bacteria B,

r15 = YS3 / X5 .µ5 .X5 (20)


EP

production of CO2 by syntrophic bacteria A,

r16 = YS4 / X 2 .µ 2 .X 2 (21)


C

production of acetic acid by syntrophic bacteria A,


AC

r17 = YS5 / X 2 .µ 2 .X 2 (22)


ACCEPTED MANUSCRIPT
production of methane by the hydrogenotrophic methanogens,

r18 = YS7 / X3.µ3.X3 (23)

and production of methane by the acetoclastic methanogens,

r19 = YS7 / X 4 .µ 4 .X 4 (24)

The rate of H2 transfer from the liquid to the gas phase is given by,

PT
r20 = Kla 3 .(S3 − S*3 ) (25)

RI
where

S*3 = H 3.G 3 (26)

SC
The rate of CO2 transfer from the liquid to the gas phase is given by,

U
r21 = Kla 4 .(S4 − S*4 ) (27)

where AN
S*4 = H 4 .G 4 (28)
M
The rate of CH4 transfer from the liquid to the gas phase is given by,
D

r22 = Kla 7 .(S7 − S*7 ) (29)


TE

where

S*7 = H 7 .G 7 (30)
C EP

2.2. Species conservation equations


AC

Concentrations balances are written for the species present in the bioreactor

as follows:
ACCEPTED MANUSCRIPT
a) polymeric substrate in the liquid phase of the bioreactor,

= ⋅ (S0 ent − S0 ) + r1
dS0 Q
(31)
dt V

b) fermentable monomers in the liquid phase of the bioreactor,

= ⋅ (S1ent − S1 ) + YS1/ S0 ⋅ r1 − YS1/ X1 ⋅ r2 − YS1/ S2 ⋅ r3 − YS1/ S3 ⋅ r4


dS1 Q
dt V (32)

PT
− YS1/ S4 ⋅ r5 − YS1/ S5 ⋅ r6 − YS1/ S6 ⋅ r7

c) acidogens in the liquid phase of the bioreactor,

RI
= ⋅ (X1ent − X1 ) + r2
dX1 Q
(33)
dt V

SC
d) propionic acid in the liquid phase of the bioreactor,

U
= ⋅ (S2ent − S2 ) + r3 − YS2 / X 2 ⋅ r8 − YS2 / S3 ⋅ r9 − YS2 / S4 ⋅ r16 − YS2 / S5 ⋅ r17
dS2 Q
(34)
dt V
AN
e) syntrophic bacteria A in the liquid phase of the bioreactor,

= ⋅ (X 2 ent − X 2 ) + r8
dX 2 Q
M
(35)
dt V

f) H2 in the liquid phase of the bioreactor,


D

= ⋅ (S3ent − S3 ) + r4 + r9 + r15 − YS3 / X 3 ⋅ r10 − YS3 / S7 ⋅ r18 − r20 (36)


dS3 Q
TE

dt V

g) CO2 in the liquid phase of the bioreactor,


EP

= ⋅ (S4 ent − S4 ) + r5 + r12 + r16 − YS4 / X 3 ⋅ r10 − YS4 / S7 ⋅ r18 − r21 (37)
dS4 Q
dt V
C

h) hydrogenotrophic methanogens in the liquid phase of the bioreactor,


AC

= ⋅ (X 3ent − X 3 ) + r10
dX 3 Q
(38)
dt V
ACCEPTED MANUSCRIPT
i) acetic acid in the liquid phase of the bioreactor,

= ⋅ (S5ent − S5 ) + r6 + r13 + r17 − YS5 / X 4 ⋅ r11


dS5 Q
dt V (39)
− YS5 / S4 ⋅ r12 − YS5 / S7 ⋅ r19

j) acetoclastic methanogens in the liquid phase of the bioreactor,

= ⋅ (X 4 ent − X 4 ) + r11
dX 4 Q

PT
(40)
dt V

k) butyric acid in the liquid phase of the bioreactor,

RI
= ⋅ (S6 ent − S6 ) + r7 − YS6 / S5 ⋅ r13 − YS6 / X 5 ⋅ r14 − YS6 / S3 ⋅ r15
dS6 Q
(41)
dt V

SC
l) syntrophic bacteria B in the liquid phase of the bioreactor,

U
= ⋅ (X 5ent − X 5 ) + r14
dX 5 Q
(42)
dt V
AN
m) methane in the liquid phase of the bioreactor,

= ⋅ (S7 ent − S7 ) + r18 + r19 − r22


dS7 Q
M
(43)
dt V

Since the overall pressure in the headspace is assumed to be constant, the


D

total molar flow rate of gas out of the gas phase is equal to the amount of gas
TE

transferred from the liquid to the gas phase, shown by the following expression:

( ) ( )
F = Kla 3. S3 − S*3 .V / M H 2 + Kla 4 . S4 − S*4 .V / M CO 2
EP

( )
(44)
+ Kla 7 . S7 − S*7 .V / M CH 4
C

n) H2 in the gas phase of the bioreactor, rearranged to be expressed in terms of the


AC

partial pressure of H2 in the gas phase,


ACCEPTED MANUSCRIPT
dG 3 R.T 
dt
= ( ) G 
⋅ Kla3. S3 − S*3 .V / M H 2 − F.V. 3 
Vg  PT 
(45)

o) CO2 in the gas phase of the bioreactor, rearranged to be expressed in terms of the

partial pressure of CO2 in the gas phase,

dG 4 R.T 
= ( ) G 
⋅ Kla 4 . S4 − S*4 .V / M CO2 − F.V. 4  (46)

PT
dt Vg  PT 

p) methane in the gas phase of the bioreactor, rearranged to be expressed in terms of

RI
the partial pressure of methane in the gas phase,

SC
dG 7 R.T 
dt
= ( ) G 
⋅ Kla 7 . S7 − S*7 V / M CH 4 − F.V. 7 
Vg  PT 
(47)

U
At this point, it is important to stress that the utilized hypotheses and model

AN
equations grant flexibility to the model, in the sense that it could be used to describe

a biodigester operating either as a continuous stirred-tank reactor (CSTR) or a batch


M
reactor. For describing a batch reactor, it is only necessary to set Q = 0 m3 day-1. In

fact, in this work, for performing the model experimental validation, biodigester
D

batch reactor experimental data available in the literature [10] are utilized, therefore
TE

the present model produces results for Q = 0 m3 day-1. After model experimental

validation, Q is allowed to vary, therefore the optimization study is conducted for


EP

biodigesters operating as continuous stirred-tank reactors (CSTR).


C
AC

3. Numerical method
ACCEPTED MANUSCRIPT
Equations (31) – (43), (45) – (47), for all species in the bioreactor in liquid

and gas phases, and the specified initial conditions form a system of 16 ordinary

differential equations (ODEs) which need to be solved together with 31 auxiliary

algebraic equations, i.e., Eqs. (1) – (30) and (44). The unknowns are the

concentrations (liquid phase) and partial pressures (gas phase) of the species in the

PT
bioreactor.

The solution to the ODEs depict the transient behavior of the system, starting

RI
from a set of initial conditions, and marching in time (checking for accuracy) until a

SC
steady state is achieved. The equations were integrated in time implicitly using a

Backward Euler method [14], which was implemented in a program written in

U
FORTRAN language, using the subroutine DASSL (differential algebraic system

solver) [15].
AN
Table 1 lists the initial values of variables and the values of parameters that
M
were used in the simulation. The kinetic parameters were obtained from the technical
D

literature [16]. The yield coefficients required by the mathematical model were
TE

determined by undertaking a stoichometric balance on the overall reaction for each

step. For that, the substrate efficiency coefficient was assumed as 10%, and all other
EP

stoichiometric coefficients were calculated from the balanced stoichiometric

equations [17]. Table 2 lists the calculated yield values.


C
AC

4. Results and discussion


ACCEPTED MANUSCRIPT
The numerical results obtained with the mathematical model were compared

to available batch reactor experimental data obtained by Salminen et al. [10] in

Figures 3 – 7. Table 1 displays the list of parameters and initial values used in the

simulations, which were selected according to the experimental data [10].

Figure 3 depicts the transient evolution of the polymeric substrate with

PT
respect to time. It is seen that the model was capable of capturing the descending

trend in time of the polymeric substrate concentration, S0, but the behaviour was

RI
monotonic whereas the experimental measurements showed that the concentration

SC
initially decreased, then on the 10th day stabilized up to the 20th day, and eventually

decreased to reach steady state at complete depletion around the 35th day, therefore

U
only qualitative agreement was observed in the polymeric substrate concentration
AN
transient evolution. On the other hand, the numerical results showed good

quantitative and qualitative agreement with the experimentally determined steady


M
state, i.e., around the 35th day in both cases. Therefore, the simulated profile for the
D

variation of concentration of polymeric substrate in time represents an average


TE

consumption rate as compared to the experimental data.

The unsteady simulation results for acetic and butyric acid show good
EP

qualitative and quantitative agreement with the experimental data reported by

Salminen et al. [10] as shown in Figs. 4 and 5, respectively. Figure 4 also shows that
C

the model could not predict accurately the time to achieve steady state for the acetic
AC

acid, but for butyric acid Fig. 5 demonstrates that the model predicted steady

conditions around the 20th day, therefore in good agreement with the experiments.
ACCEPTED MANUSCRIPT
For propionic acid, Figure 6 illustrates that the model fails to accurately

represent the process from the 10th to the 30th day. The transient numerical results

showed good agreement with the experiments until the 10th day. Experimentally it is

seen that the compound concentration was still increasing on the 10th day, in which

the simulation results started to drop, and continued to increase. The experimentally

PT
measured concentration peak (~ 3 g L-1) occurred around the 25th day and was more

than 20% greater than the numerically simulated one (~ 2.4 g L-1). Steady state was

RI
achieved experimentally around the 40th day whereas the numerical results stabilized

SC
around the 35th day, with the compound concentration dropping to values close to

zero.

U
In Figure 7, it is noted that the numerically calculated molar fraction of
AN
methane in the gas phase exhibits good qualitative agreement with the experimental

measurements both for the transient and steady state evolution of the process.
M
However, quantitatively, the model underestimates the final methane molar fraction
D

obtained in steady state conditions, which was close to 70% in the experimental data
TE

and 50% in the simulated results.

The disagreement between the numerical results and some of the


EP

experimental data is understandable, due to the model assumptions. For example, it

is well known that syntrophic bacteria, which are responsible for propionic and
C

butyric acid consumption, are inhibited by hydrogen, but this effect was not taken
AC

into account in the model. Future improvements of the model could include the

incorporation of the hydrogen inhibitory effect.


ACCEPTED MANUSCRIPT
Due to the qualitative and quantitative agreement of the numerical results and

experimental data in most of the transient and steady state compounds concentrations

and partial pressures, it is reasonable to state that the model has been experimentally

validated. This is mainly true for optimization purposes, since qualitative agreement

was observed in all comparisons, i.e., the actual system trends have been captured by

PT
the model, so that in spite of quantitative discrepancies that might occur in

numerically predicting maximum performance values (e.g., maximum methane

RI
molar fraction output), the locations of the predicted optima are expected to be

SC
accurate.

Additionally, the simplified reaction model of the anaerobic digestion process

U
that was developed in the present work is suitable for incorporation into models that
AN
describe more complex systems in which the biodigester contents are not

homogeneous. For that, in order to account for reactor non-uniformities, spatial


M
dependence could be introduced in the model without further increase in complexity,
D

by the use of a volume element model [18] which does not require the use of partial
TE

differential equations. In such case, the reaction model will be only slightly

modified, but the expressions describing the hydrodynamic aspects of the system
EP

will be more complex.

After model experimental validation, the analysis proceeded to evaluate the


C

response of the reactor to the variation of several operating conditions, i.e., by


AC

performing a parametric analysis. Although the model experimental validation has

been conducted for a batch reactor, generally, biodigesters operate continuously, so


ACCEPTED MANUSCRIPT
that the effect of the continuous variation of system variables on bioreactor

production is important to be assessed.

Figures 8 and 9 show the effects of the substrates concentrations on the

output total biogas volumetric flow rate. The input substrates concentrations S2ent,

S5,ent and S6,ent had no impact on the biogas volumetric flow rate, as it is seen in Figs.

PT
8a, 8b, and 8c, respectively. However, the input substrates concentrations S0ent and

S1ent had a significant effect on increasing biogas production, as it is shown in Figs.

RI
9a and 9b, respectively. What is more striking is the effect of S1ent on biogas

SC
production, which increases monotonically and linearly as S1ent increases. Such

effects demonstrate how much the hydrolysis step of the polymeric substrate limits

U
system efficiency. Note that when the system is fed with hydrolysed substrate (S1 is
AN
a product of the S0 hydrolysis), the biogas production increases considerably. The

analysis could be extended for higher values of S1ent, what is done in Fig. 9c, in
M
which S1ent was increased up to 2000 g L-1, and the data were fitted through linear
D

regression, producing the following correlation:


TE

Qg = 4.383 × S1ent with R 2 = 0.9997 (48)

The correlation given by Eq. (48) is valid only for the set of data used in this
EP

work. However, the output total biogas volumetric flow rate, Qg, is expected to

behave similarly for any continuous stirred-tank reactor (CSTR) biodigester, i.e.,
C

linearly with respect to the variation of S1ent.


AC

Another important variable that affects continuous biodigesters’ performance

is the substrates input volumetric flow rate, Q, which determines the system retention
ACCEPTED MANUSCRIPT
time, HRT (hydraulic retention time). Figure 10a shows that small or large HRT lead

to poor performance, so that there is an optimal hydraulic retention time (HRTopt ~ 5

days) for maximum biogas production, which corresponds to Qopt = 175 L day-1 for

the set of data used in the simulations conducted in this work, as it is seen in Fig.

10b. This optimal condition is physically explained by analyzing two extremes with

PT
fixed V: i) for small HRT, Q is high, therefore reaction kinetics is not fast enough to

process the substrates and biogas production is low, and ii) for large HRT, Q is

RI
small, therefore substrates concentrations are low, and biogas production is also low.

SC
Furthermore, the maximum is quite sharp, i.e., there is a significant drop in biogas

production that results from biodigester operation with HRT even slightly different

U
from HRTopt. For example, if a narrow range HRTopt ± 5% is considered, a 50%
AN
drop is observed in Qg with respect to Qg,max for the set of data used in this study.

The optimal values for HRT and Q found for the set of data used in the
M
numerical simulations conducted in this work are within what is expected for a
D

continuous stirred-tank reactor (CSTR) biodigester. However, depending on the


TE

input polymeric solids concentration, S0ent, the optimal substrates input volumetric

flow rate, Qopt, could change. As it is observed in Fig. 11, for high values of S0ent it is
EP

better to reduce Q, using HRT = 30 days instead of HRT = 6 days. The reason is that

with higher HRT there is more time to hydrolyze the polymeric substrate which in
C

turn generates more biogas.


AC

At this point it is important to generalize the analysis. Fundamentally, the

message is that no matter how complex the actual biodigester design might be, there
ACCEPTED MANUSCRIPT
will be an optimal retention time, HRTopt, and a corresponding optimal substrates

input volumetric flow rate, Qopt, worth to be found such that maximum biogas

production, Qg,max, will be achieved.

5. Conclusions

PT
In this paper, optimal hydraulic retention time, HRTopt (or input volumetric

flow rate, Qopt) was found that lead to sharp maximum biogas production within a

RI
narrow range, stressing the importance of its identification in any actual biodigester

SC
operation, i.e., for HRTopt ± 5% , a 50% drop was observed in Q with respect to

Qg,max, for the biodigester configuration tested in this study. For that, a general

U
transient mathematical model for the species management of batch and continuous
AN
stirred-tank reactor (CSTR) biodigesters has been developed. The model was

validated experimentally by direct comparison of transient and steady state


M
simulations results with previously published experimental data for a batch reactor
D

biodigester. Also, a sensitivity analysis was conducted for CSTR biodigesters


TE

identifying critical variables that affect output total biogas volumetric flow rate. This

was the case of the input fermentable monomer substrate concentration, S1, which
EP

was shown to affect significantly biogas production that increases monotonically as

S1 increases.
C

The key conclusion of this study is that the simplified model presented for the
AC

reactions that occur in anaerobic digestion has the ability to describe both transient

and steady state behaviour. Therefore, the model could be applied in practice for the
ACCEPTED MANUSCRIPT
design of small anaerobic bioreactors or large anaerobic bioreactors with induced

homogeneous mixtures.

The present model is also suitable for incorporation into mathematical

models that describe the operation of more complex anaerobic bioreactors, in which

the bioreactor contents are not homogeneously mixed. Therefore it is expected that

PT
such application could be used as an efficient tool for batch and CSTR biodigester

design, control and optimization.

RI
SC
Acknowledgements

The authors acknowledge with gratitude the support of the Brazilian National

U
Council of Scientific and Technological Development, CNPq (projects
AN
401117/2004-9, 552867/2007-1, 574759/2008-5 and 558835/2010–4), and of

NILKO Tecnologia Ltda.


M
D

References
TE

[1] A. Bejan, Advanced Engineering Thermodynamics, second ed, Chapter 7,

Wiley, New York, 1997.


EP

[2] H. Bouallagui, Mesophilic biogas production from fruit and vegetable waste

in a tubular digester, Bioresource Technology 86 (2003) 85-89.


C

[3] J.M. Lomas, C. Urbano, L.M. Camarero, Evaluation of a pilot scale


AC

downflow stationary fixed film anaerobic reactor treating piggery slurry in

the mesophilic range, Biomass & Bioenergy 17 (1999) 49-58.


ACCEPTED MANUSCRIPT
[4] A. Marcos, A. Al-Kassir, A.A. Mohamad, F. Cuadros, F. López-Rodríguez,

Combustible gas production (methane) and biodegradation of solid and liquid

mixtures of meat industry wastes, Applied Energy 87 (2010) 1729–1735.

[5] D.J. Batstone, J. Keller, I. Angelidaki, S.V. Kalyuzhnyi, S.G. Pavlostathis, A.

Rozzi, W.T.M. Sanders, H. Siegrist, V.A. Vavilin, The IWA Anaerobic

PT
Digestion Model No 1 (ADM1), Water Science and Technology 45:10 (2002)

65–73.

RI
[6] S. Valle-Guadarrama, T. Espinosa-Solares, I.L. Lopez-Cruz, M. Domaschko,

SC
Modeling temperature variations in a pilot plant thermophilic anaerobic

digester, Bioprocess and Biosystems Engineering 34:4 (2011) 459-470.

U
[7] V.A. Vavilin, L.Y. Lokshina, S.V. Rytov, The <METHANE> simulation
AN
model as the first generic user-friend model of anaerobic digestion, Moscow

University Chemistry Bulletin 41:6 (2000) 22-26, Supplement.


M
[8] V.B. Vasiliev, V.A. Vavilin, S.V. Rytov, A.V. Ponomarev, Simulation model
D

of anaerobic digestion of organic matter by a microorganism consortium:


TE

basic equations, Water Resources 20 (1993) 633-643.

[9] V.A. Vavilin, V.B. Vasiliev, A.V. Ponomarev, S.V. Rytov, Simulation-model
EP

methane as a tool for effective biogas production during anaerobic conversion

of complex organic-matter, Biores. Technol. 48:1 (1994) 1-8.


C

[10] E. Salminen, J. Rintala, L.Y. Lokshina, V.A. Vavilin, Anaerobic batch


AC

degradation of solid poultry slaughterhouse waste, Water Science and

Technology 41:3 (2000) 33-41.


ACCEPTED MANUSCRIPT
[11] V.A. Vavilin, L.Y. Lokshina, X. Flotats, I. Angelidaki, Anaerobic digestion

of solid material: multidimensional modeling of continuous-flow reactor with

non-uniform influent concentration distributions, Biotechnology and

Bioengineering 97:2 (2007) 354-366.

[12] E. Martinez, A. Marcos, A. Al-Kassir, M.A. Jaramillo, A.A. Mohamad,

PT
Mathematical model of a laboratory-scale plant for slaughterhouse effluents

biodigestion for biogas production, Applied Energy 95 (2012) 210–219.

RI
[13] J. Monod, The growth of bacterial cultures, Annual Review of Microbiology

SC
3 (1949) 371.

[14] D. Kincaid, W. Cheney, Numerical analysis, Wadsworth, Belmont, CA,

U
1991.
AN
[15] K.E. Brenan, S.L. Campbell, L.R. Petzold, Numerical solution of initial-value

problems in differential-algebraic equations, Classics in Applied


M
Mathematics, Philadelphia, PA, 1996.
D

[16] G.B. Ryhiner, E. Heinzle, J.I. Dunn, Modeling and simulation of anaerobic
TE

wastewater treatment and its application to control design: case whey,

Biotechnol. Prog. 9 (1993) 332-343.


EP

[17] M.L. Shuler, F. Kargi, Bioprocess engineering: basic concepts, second ed.,

Prentice Hall, 2001.


C

[18] J.V.C. Vargas, G. Stanescu, R. Florea, M.C. Campos, A numerical model to


AC

predict the thermal and psychrometric response of electronic packages,

ASME Journal of Electronic Packaging 123:3 (2001) 200-210.


ACCEPTED MANUSCRIPT

Figure Captions

Figure 1 The complex web of reactions within the anaerobic digestion system.

Steps of the biodigestion process: I) Hydrolysis; II) Acidogenesis; III)

Acetogenesis; IV) Methanogenesis – Microbial groups: 1) hydrolytics;

PT
2) acidogens; 3) acetogens; 4) methanogens. VFA = volatile fatty acids;

LCFA = long chain fatty acids.

RI
Figure 2 Simplified diagram of reactions and processes occurring in the anaerobic

SC
digestion system.

Figure 3 Transient evolution of the polymeric substrate concentration, S0 (solid

U
line = model prediction; squares = experimental data [10]).
AN
Figure 4 Transient evolution of acetic acid concentration, S5 (solid line = model
M
prediction; squares = experimental data [10]).
D

Figure 5 Transient evolution of butyric acid concentration, S6 (solid line = model


TE

prediction; squares = experimental data [10]).

Figure 6 Transient evolution of propionic acid concentration (solid line = model


EP

prediction; squares = experimental data [10]).


C

Figure 7 Transient evolution of methane molar fraction, x7 (solid line = model


AC

prediction; squares = experimental data [10]).


ACCEPTED MANUSCRIPT
Figure 8 The variation of the total biogas output volumetric flow rate, Qg, with

respect to: a) Input polymeric substrate concentration, S0ent; b) Input

fermentable monomers concentration, S1ent, up to 200 g L-1, and c) Input

fermentable monomers concentration, S1ent, up to 2000 g L-1.

Figure 9 The variation of the total biogas output volumetric flow rate, Qg, with

PT
respect to: a) Input propionic acid concentration, S2ent; b) Input acetic

RI
acid concentration, S5ent, and c) Input butyric acid concentration, S6ent.

Figure 10 The maximization of the total biogas output volumetric flow rate, Qg,

SC
with respect to: a) Hydraulic retention time, HRT, and b) Substrates

U
input volumetric flow rate, Q.

Figure 11 AN
The variation of the total biogas output volumetric flow rate, Qg, with

respect to input polymeric substrate concentration, S0ent, and hydraulic


M
retention time, HRT.
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

Table 1. Values of parameters [7, 10] and initial values of variables.

Symbol Value Unit Symbol Value Unit

Q 250 day-1 KS2 0.97 g L-1

V 2000 L KS3 0.00005 g L-1

PT
Vg 300 L KS4 0.0019 g L-1

R 0.082 atm.L.mol-1 K-1 KS5 0.019 g L-1

RI
T 305 K KS6 0.59 g L-1

g L-1

SC
G3,4,7 0 at t = 0 atm X1 0.1 at t = 0

PT 1 at t = 0 atm X2 0.1 at t = 0 g L-1

U
S0 60 g L-1 X3 0.01 at t = 0 g L-1

S1 1 at t = 0 g L-1 AN X4 0.01 at t = 0 g L-1

S2,3,4,5,6,7 0 at t = 0 g L-1 X5 0.1 at t=0 g L-1


M
S0ent 60 g L-1 X1,2,3,4,5ent 0 g L-1

S1,2,3,4,5,6,7ent 0 g L-1 Kla3,4,7 48 day-1


D

Kh 0.5 day-1 S3* Eq. (26) g L-1


TE

µ max1 0.2 day -1 S4* Eq. (28) g L-1

day -1 g L-1
EP

µ max2 0.00185 S7* Eq. (30)

µ max3 2 day -1 H3 0.01 g L-1.atm-1


C

µ max4 0.0225 day -1 H4 0.04 g L-1.atm-1


AC

µ max5 0.01 day -1 H7 0.07 g L-1.atm-1


ACCEPTED MANUSCRIPT
KS1 0.67 g L-1 MH2 2 g mol-1

Table 2. Yield values (Ya/b, yield of a from b), g g-1.

YS1/S0 1.11 YS2/S3 12.3 YS3/X2 0.811 YS4/X4 7.33 YS6/S3 22.0

YS1/S2 12.2 YS2/S4 1.68 YS3/X3 1.54 YS5/S4 1.36 YS6/X1 1.83

PT
YS1/S3 40.0 YS2/S5 1.23 YS3/X5 0.455 YS5/S7 3.75 YS6/X5 10.0

YS1/S4 2.56 YS2/X1 0.822 YS4/S7 2.75 YS5/X1 2.83 YS7/X3 3.08

RI
YS1/S5 3.53 YS2/X2 10.0 YS4/X1 3.91 YS5/X2 8.11 YS7/X4 2.67

YS1/S6 5.45 YS3/S7 0.5 YS4/X2 5.95 YS5/X4 10.0

SC
YS1/X1 10.0 YS3/X1 0.25 YS4/X3 8.46 YS5/X5 13.6

U
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

Highlights

>We introduce a general transient mathematical model for anaerobic biodigesters > The model
was experimentally validated > A simulation and optimization study was conducted with the
model > The existence of optimal hydraulic retention time was investigated > The model could
be a tool for simulation, design, control and optimization of biodigesters.

PT
RI
U SC
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

GAS PHASE
IV H2+CO2 CH4 CO2

PT
4 4
LIQUID PHASE
2
Simple 2 3
Lactic Ac.

RI
Carbohidrates Sugars A
2 3
Butyric Ac. c
2 e
H2+CO2
2 t

Acetic Acid
SC
VFA
2 3 i
Complex Propionic Ac.
2
Organic 1 2 3 c
Material Proteins Aminoacids Valeric Ac.

U
NH3
A

Lipids 1
AN
LCFA
2
c

I II III
M
D
TE
C EP
AC

Figure 1
ACCEPTED MANUSCRIPT

Gas

outlet
Gas Phase
G3 G7 G4

PT
r22
S7

RI
Liquid
r20 r18 r21
r19 outlet

SC
r17
r16
r15

U
XX22 r S2 r9 S3 S4 r12 S5 r13 S6
8

r3 r4
r10

XX33
AN
r5 X4
r11
r6 r7
r14

X5
M
S1
r2
r1 X1
D

S0
TE

Liquid Phase
EP

Liquid inlet
C
AC

Figure 2
ACCEPTED MANUSCRIPT

PT
Polymeric Substrate (S0)

RI
25
S0 concentration (g/L)

SC
-1
20
S0 (g L )
15

U
10
5
AN
0
M
0 20 40 60
t (day)
D

time (days)
TE
C EP
AC

Figure 3
ACCEPTED MANUSCRIPT

Acetic Acid (S5)

PT
6
S5 concentration (g/L)

RI
S5 (g L-1) 4

SC
3
2
1

U
0
0 10 20 AN 30 40 50 60
time (days)
t (day)
M
D
TE
C EP
AC

Figure 4
ACCEPTED MANUSCRIPT

Butyric Acid (S6)

6
S6 concentration (g/L)

PT
-1
5
S6 (g L )
4

RI
3
2

SC
1
0

U
0 20 40 60
t (day)
AN
time (days)
M
D
TE
C EP
AC

Figure 5
ACCEPTED MANUSCRIPT

Propionic Acid (S2)

3,5

PT
S2 concentration (g/L)

3
-1
S2 (g L )2,5

RI
2
1,5

SC
1
0,5
0

U
0 10 20 30 40 50 60
time (days)
AN
t (day)
M
D
TE
C EP
AC

Figure 6
ACCEPTED MANUSCRIPT

Methane (G7)

80
70

PT
60
x7 (%)
50
G7 (%)

40

RI
30
20

SC
10
0
0 10 20 30 40 50 60

U
time (days)
t (day)
AN
M
D
TE
EP

Figure 7
C
AC
ACCEPTED MANUSCRIPT

Qg (L day-1)

PT
RI
S0ent (g L-1)
(a)

SC
Qg (L day-1)

U
AN
M

S1ent (g L-1)
D

(b)
TE

Vaz ão de B iog ás

V az ão de B iogás = 4,383*C onc entraç ão de S 1


10000
EP
V az ão de B iog ás (L /dia)

R 2 = 0,9997
8000
Qg (L day-1)
6000
C

4000

2000
AC

0
0 500 1000 1500 2000 2500

S1ent (g L-1)
C onc e ntra ç ã o de E ntra da de S 1 (g /L )

(c)

Figure 8
ACCEPTED MANUSCRIPT

Qg (L day-1)

PT
S2ent (g L-1)

RI
(a)

SC
Qg (L day-1)

U
AN
M
S5ent (g L-1)
D

(b)
TE

Qg (L day-1)
C EP
AC

S6ent (g L-1)
(c)

Figure 9
ACCEPTED MANUSCRIPT

Qg (L day-1)

PT
RI
SC
HRT (day)

U
(a)
AN
M
Vaz ão de B iog ás
160
D

140
Vaz ão de B iog ás

Qg (L day-1) 120
(L /dia)

TE

100
80
60
40
EP

20
0
0 50 100 150 200 250
V a z ã o de E ntra da (L /dia )
C

Q (L day-1)
AC

(b)

Figure 10
ACCEPTED MANUSCRIPT

V az ão de B iog ás (L /dia)
Vaz ão de B iog ás
35
30
Qg (L day-1)
25

PT
20
HRT = 30 days
15 HR T 6 dias

RI
10 HR T 30 dias
6 days
5
0

SC
0 20 40 60 80
C onc e ntra ç ã oS0ent L-1/l))
de S(g0 (g

U
AN
M
D
TE
C EP
AC

Figure 11

Das könnte Ihnen auch gefallen