You are on page 1of 16

Journal of Environmental Management 241 (2019) 587–602

Contents lists available at ScienceDirect

Journal of Environmental Management


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

Research article

ADM1-based mechanistic model for the role of trace elements in anaerobic T


digestion processes
L. Frunzoa,∗, F.G. Fermosob, V. Luongoa, M.R. Matteia, G. Espositoc
a
Department of Mathematics and Applications “Renato Caccioppoli”, University of Naples Federico II, via Cintia I, Monte S. Angelo, 80126, Naples, Italy
b
Instituto de la Grasa (C.S.I.C.), Campus Universidad Pablo de Olavide, Edificio 46, Ctra. de Utrera km. 1, 41013, Sevilla, Spain
c
Department of Civil, Architectural and Environmental Engineering, University of Naples Federico II, via Claudio 21, 80125, Naples. Italy

ARTICLE INFO ABSTRACT

Keywords: An original mechanistic model able to describe the fate of trace elements (TE) in anaerobic digestion systems has
Trace elements been synthetized from mass balance equations. The model takes into account the main biochemical and physico-
Anaerobic digestion chemical processes affecting TE bioavailability and it is aimed at evaluating the effect that the combination of
Mathematical modelling such processes exerts on the system performance. Five main modules have been introduced: biochemistry,
Ordinary differential equations
physico-chemistry, sorption, complexation and precipitation. The model is based on mass conservation princi-
ples and is formulated as a set of ordinary differential equations for the soluble and particulate components
constituting the system. Model applications of two illustrative cases are provided. The first case is based on
experimental results and examines the effect of TE depletion in an AD process of food waste (FW). The second
case shows the effects of different metal supplements on methane production and biogas composition. The
simulation results confirm that the model can fairly be used to predict the effect of TE dynamics and bioa-
vailability, by considering biological, chemical and physicochemical processes in AD environments.

1. Introduction Fermoso et al., 2015; Ortner et al., 2014).


The definition of a mathematical model able to consider the com-
Anaerobic Digestion (AD) has been widely used to treat high volatile plex dynamics and bioavailability of TE in anaerobic environment is
solids (VS) content substrates and results in substrate stabilization and highly relevant for both academia and industries. It can allow the un-
high net energy recovery (Esposito et al., 2012). Optimal supply of derstanding of the main processes affecting TE in AD systems while
Trace Elements (TE) is a prerequisite for microbial growth and AD providing an user-friendly tool for TE requirements in full-scale appli-
microbial metabolism. Consequently, the deficiency of TE leads to cations.
significant limitation of microbial activity that can result in the com- In 2002 the International Water Association (IWA) Task Group for
plete failure of AD process (Van Hullebush et al., 2016). TE such as iron Mathematical Modelling of Anaerobic Digestion Processes developed a
(Fe), cobalt (Co), nickel (Ni), selenium (Se), molybdenum (Mo) and/or comprehensive mathematical model known as Anaerobic Digestion
tungsten (W), play a key role in the growth of anaerobic microorgan- Model 1 (ADM1) (Batstone et al., 2002). The ADM1 is a well-known and
isms, being part of essential enzymatic co-factors such as methyl- widely studied mathematical model, which can describe the conversion
coenzyme M, carbon monoxide dehydrogenase, or coenzyme M methyl- of complex organic compounds in methane (CH4). ADM1 simulates the
transferase, involved in the AD processes (Zhang et al., 2015a). main biochemical (related to the microbial community) and physico-
TE dynamics and bioavailability in AD systems are affected by chemical (related to the liquid or gaseous state of the variables) pro-
biological, chemical and physical processes. TE are subjected to com- cesses for a reasonable modelling of several intermediate compounds
plex interactions, with the biomass and the liquid phase (e.g., pre- formation leading to the final production of CH4. Although the ADM1
cipitation, organic complexation, bio-sorption, adsorption or inorganic model predicts the dynamic concentrations of biochemical components
complexation), which affect the TE bio-uptake and the process stability. fairly, its predictions are limited due to the insufficient consideration of
Predicting the fate of TE in anaerobic processes has received increased physicochemical and biochemical processes (Zhang et al., 2015a).
attention across the research community (Van Hullebush et al., 2016; During the last decade, many modifications/manipulations of the

Corresponding author.

E-mail addresses: luigi.frunzo@unina.it (L. Frunzo), fgfermoso@ig.csic.es (F.G. Fermoso), vincenzo.luongo@unina.it (V. Luongo),
mariarosaria.mattei@unina.it (M.R. Mattei), giovanni.esposito@unicas.it (G. Esposito).

https://doi.org/10.1016/j.jenvman.2018.11.058
Received 16 July 2018; Received in revised form 3 November 2018; Accepted 15 November 2018
Available online 22 April 2019
0301-4797/ © 2018 Elsevier Ltd. All rights reserved.
L. Frunzo, et al. Journal of Environmental Management 241 (2019) 587–602

original mathematical model have been developed in order to introduce is available on the supplementation of single elements such as Fe, Ni
specific process affecting the anaerobic conversion of organic sub- and Co (Fermoso et al., 2009). However, recent experimental activities
strates, such as sulphur (S) degradation and kinetics (Fedorovich et al., highlighted that a correct combination of TE supplementation can be
2003), or CaCO3 precipitation (Batstone and Keller, 2003). ADM1 ex- more beneficial than single element addition strategies (Moestedt et al.,
tensions have also been proposed to remove the discrepancies in both 2016). According to Thanh et al. (2016), TE can be supplemented
carbon and nitrogen balances (Blumensaat and Keller, 2005), to con- continuously, pre-loaded or pulsed into anaerobic reactors. Moreover,
sider the effects of particle size distribution in anaerobic digestion of the dosing strategy of the AD systems should take into account the
complex organics (Esposito et al., 2011a), to improve the physico- operational parameters of the process, such pH, Hydraulic Retention
chemical ADM1 framework by incorporating more inorganic compo- Time (HRT) and Organic Loading Rate (OLR), the temperature regime,
nents (metal ions and phosphates), or to include precipitation processes the digestion operating mode (single or double stage digestion), the
(Zhang et al., 2009; Flores-Alsina et al., 2016). type of microbial biomass operating the process, and the reactor con-
Up to date, an overall model, or ADM1 extension, able to take into figuration. Indeed, all these parameters can directly or indirectly in-
account the several processes affecting TE in AD systems is missing. In fluence TE speciation and thus their availability to microorganisms.
order to include TE dynamics and their effects on AD systems, ADM1
should be extended accounting for the main biological, chemical and 2.2. Effect of TE on biomass growth
physical processes where TE are involved. More precisely, it is note-
worthy to include the following dynamic processes, divided in three The microbial demands for TE in anaerobic environments are as
main categories: diverse as the involved microorganisms and their functions (Wintsche
Biochemical processes: et al., 2016). Indeed, the anaerobic microorganisms have different op-
timal growth conditions with different physiologies, growth kinetics,
− release of metal ions, sulphur (S) and phosphorus (P) during hy- nutrient requirements, and sensitivity levels to environmental changes
drolysis of complex organic matter, and their uptake during biolo- (Mudhoo and Kumar, 2013). Several studies highlighted that TE affect
gical conversion processes; all the trophic groups due to their functions in enzyme complexes
− inhibition and stimulation effects of TE on methanogenesis reac- (Wintsche et al., 2016). In AD systems, many TE, such as cobalt, nickel,
tions; iron, zinc, molybdenum and/or tungsten, are part, in the form of central
ions conferring catalythic functions, of the essential enzymes driving
Chemical processes: methanogenesis reactions (Zandvoort et al., 2006). In particular, Ni, Co
and Fe are essential cofactors of carbon monoxide dehydrogenase,
− precipitation processes of metal ions taking into account the asso- acetyl-CoA decarbonalyse, methyl-H4SPT:HS-CoM methyltransferase,
ciation/dissociation processes of carbonate, phosphate and sulphide methyl-CoM reductase and other enzymes involved in the acetoclastic
species; methanogenesis pathway. In experimental studies, some TE have been
− complexation reactions of metal and soluble components; found to affect the microbial activity of propionate oxidizing bacteria as
− metal sorption on particulate components, e.g. inert and microbial they utilize the hydrogenase enzyme (Romero-Güiza et al., 2016). It is
biomasses; assumed that methanogens groups, both hydrogenotrophic and acet-
oclastic, are usually characterized by slower growth rates compared to
Physical processes: the other microbial groups and could be considered as the most sensi-
tive microbial groups to TE limitations (Thanh et al., 2016). In addition,
− liquid-gas transfer of hydrogen sulphide. many TE can have an inhibitory effect on the microbial metabolism
depending on the concentration level as they can deactivate enzymes by
The aim of this study is to propose and apply a mechanistic model reacting with their functional groups, denature proteins and compete
based on ADM1 able to simulate the TE dynamics in AD systems. The with essential cations (Zhang et al., 2009).
model has been synthesized to mechanistically describe the main bio- The stimulatory and inhibitory concentration ranges of TE on
chemical and physicochemical processes affecting TE availability and to anaerobic microorganisms are disparate; for a general overview refer to
analyse the impact that the combination of such processes exerts on the the following recent studies (Thanh et al., 2016; Choong et al., 2016;
AD process stability and efficiency. Romero-Güiza et al., 2016; Glass and Orphan, 2012). Such incon-
sistency in the beneficial and inhibitory levels of TE is mainly related to
2. Most relevant processes affecting TE in AD systems the difference in the background values and TE distributions in the
feedstocks, in the diversity of the digestate texture as well as in the
2.1. Input of TE in AD environment characteristics of the specific microflora operating the biochemical
conversion (Cao et al., 2018). Moreover, the use of TE as AD additives,
TE in an AD environment usually come from two different sources, may smooth the inhibition of the anaerobic biomass by reducing the
i.e. TE content in the feedstock and TE supplementation as external concentration of H2S in the biogas through precipitation processes
addition. In many cases, the release of TE from the feedstock due to the (Romero-Güiza et al., 2016).
decomposition of the organic matter and associated compounds fulfil
the micronutrient requirements for a successful process performance. A 2.3. Fate of sulfur and phosphorous compounds
significant example has been recently published by Lee et al. (2018),
who highlighted that the degradation of TE containing biomasses sup- Sulfur and Phosphorous compounds are crucial factors in AD sys-
ports microbial activities during the AD process. If TE are released over tems as they are extremely reactive towards TE and they contextually
their toxic level on microorganisms, they contribute to the deterioration constitute fundamental nutrients for the anaerobic microbial consortia.
of the AD performance. High TE concentrations may also limit the It has been reported that the dry weight of bacteria may be constituted
proper use of digestate as fertilizer and cause environmental pollution by 1–3% and around 1% of phosphorous and sulphur, respectively
(Ortner et al., 2015). If TE are released below a threshold value for (Gerardi, 2006). In addition, sulfur and phosphorous are involved in
microbial requirement, external addition is needed to provide the various physicochemical processes, such as precipitation, affecting TE
minimum concentrations of such essential elements. This strategy is availability and dynamics in AD systems. To account for the fate of TE
highly recommended for AD of organic substrates with a low TE content in AD systems, the understanding of sulfur and phosphorous dynamics
such as food waste and stillage (Choong et al., 2016). A wide literature is a fundamental step as inorganic ligands, including phosphate and

588
L. Frunzo, et al. Journal of Environmental Management 241 (2019) 587–602

sulphide, are assumed to control the chemical speciation of TE continuously at a rate of 0.05–0.2 μmol h−1, it was observed an inter-
(Callander and Barford, 1983a,b). mittent decrease in methane production (Gonzalez-Gil et al., 1999).
P may exist in two different forms, i.e. inorganic and organic, and These rates of TE addition corresponded to TE concentrations of 1–4 μM
usually can be found in wastewaters in the prevalent form of ortho- and are in accordance with TE requirements based on the biomass yield
phosphate (PO43−), which represents the readily bioavailable form and and TE content of microorganisms. The authors thus concluded that
the product of poly-phosphate (H2P2O7) and organic phosphorous hy- when the amount of substrate is not limiting, metal dissolution can be
drolysis and mineralization (Sanchez et al., 2000, Gerardi, 2006). It is the rate-limiting step. It was argued that readily dissolved precipitates
an essential nutrient for a wide range of living organisms (De-Bashan provide the required nutrients for an exponential increase in methane
and Bashan, 2004; Kodera et al., 2013). It is often stated that anaerobic production rate by the methylotrophic microorganisms. However, if the
conversion of organic substrates promotes the mobility and availability metal uptake rate is higher than the related dissolution rate, then a TE
of phosphorous during AD (Möller and Stinner, 2010; Bachmann et al., limited environment is created. Conversely, it was found that growth
2011). Moreover, the bioaccumulation/sequestration process of phos- limitation due to slow dissolution rates of sulfides is unlikely to happen
phorous compounds in anaerobic environments by microbial consortia (Jansen et al., 2007). Indeed, it was found that the methanogenic ac-
usually competes with the precipitation process, due to the presence of tivity of Methanosarcina barkeri decreased with increase in presence of
binding compounds such as TE (Keating et al., 2016; Möller and Müller, total sulfide precipitates in the reactor. The authors argued that the
2012). For this reason, AD has been used as an efficient strategy for decrease in methane production may not be due to an increase in sul-
phosphorous removal leading to bioaccumulation and/or precipitation fide in the system as the consumption rates of Ni and Co did not slowed
in the form of metallic compounds, e.g. struvite crystallization, al- down with increase in sulfide. Recently, it was shown by sequential
lowing phosphorous recovery from digestate (De-Bashan and Bashan, extraction that in an AD system (for slaughterhouse waste) a significant
2004). portion of dosed TEs is not bioavailable (Ortner et al., 2015). However,
S compounds in a digester can be found in many different forms, as the study could not set a clear trend for carbonate bound and sulfide/
in the gas, liquid and solid phase. Many industrial processes generate organic bound fractions in the sequential extraction experiments.
high COD and high sulphur content (S-SO42-) wastewaters, which can Whereas, the dissolved and exchangeable fractions increased with in-
be used as feedstock to AD systems. Under anaerobic conditions, SRB crease in dose of TEs which also corresponds to increase methane
oxidize organic acids by using S-SO42- as final electron acceptor, thus production and decrease in VFA accumulation.
generating H2S and other reduced sulfur compounds (Fedorovich et al., Although mineral sulphides represent the major precipitate in
2003, Möller and Müller, 2012; Yekta et al., 2014). Moreover, it has anaerobic environments at the pH range usually adopted for anaerobic
been reported that proteins contain at least one of the sulphur amino digestion (Gustavsson, 2012, Liu et al., 2015; Roussel and Carliell-
acids, i.e. cysteine and methionine, releasing thiol groups (-SH) during Marquet, 2016), TE precipitates of carbonate and phosphate can be
anaerobic degradation (Gerardi, 2006). The presence of H2S, and its present in AD systems in significant amounts and have to be taken into
anionic species HS− and S2−, affects the availability of the TE account for a correct prediction of TE bioavailability. Chemical P pre-
(Gustavsson et al., 2013) and has inhibitory effects on hydro- cipitation is based on the formation of insoluble phosphate salts (e.g.
genotrophic methanogens (Batstone et al., 2002), acetate utilization struvite, hydroxyapatite, vivianite) and it is considered as a valuable
and, more generally, on overall microbial activities as well as methane method for the recovery of P from digested phosphate-rich sludge
generation (Hilton and Oleszkiewicz, 1988; Lin, 1993). (Cheng et al., 2017). Iron can be added to promote phosphorous re-
covery in AD and vivianite crystallization has been recognized as more
2.4. Precipitation competitive for ferrous ions over siderite (FeCO3) formation, which
reduces the demand of iron source to the system or omits a degassing
Typically, AD environments are characterized by the presence of step for the removal of bicarbonates as an undesired ferrous sink
abundant inorganic ligands, such as carbonate, phosphate and/or sul- (Saidou et al., 2009).
phide, which are assumed to control the speciation of TE by the for-
mation of poorly soluble metal precipitates (Yekta et al., 2017, Thanh 2.5. Complexation
et al., 2017). Carbonate, phosphate and sulphide interact also with the
light metals, usually contained in the digestion substrates (e.g. calcium, Different forms of TE are able to perform a large number of bio-
magnesium, potassium), leading to the formation of significant quan- chemical processes in the liquid phase, such as formation of inorganic
tities of struvite, hydroxyapatite and calcite. Sulphide concentration has and organic complexes, prior to make contact with the microbial bio-
been recognized as a primary factor in the definition of TE bioavail- mass (Thanh et al., 2016). Chelation of TE occurs due to their capability
ability as it regulates the formation of metal sulphides, which are of interacting with many functional groups, i.e. carboxyl, hydroxyl and
characterized by a very low solubility product resulting in a remarkably amino groups, available in AD environments. It should be noted that the
low free metal ion concentration (Barrera et al., 2015; Thanh et al., formation of soluble metal organic and inorganic complexes might have
2016). Sulphide chemistry has been claimed as the main regulator of a beneficial effect on TE bioavailability as it increases the level of total
the trace elements bioavailability in the AD system (Fermoso et al., soluble metals in an anaerobic digester competing with precipitation
2009). In presence of sulphide, TE bioavailability has been found to kinetics (Callander and Barford, 1983a,b; Fermoso et al., 2008). The
decrease in the order Fe > Ni > Co, leading to a suboptimal biogas addition of a chelating agent leads to the formation of specific TE
formation. Fe is often supplied in larger quantities than Ni and Co to compounds that could be differently up-taken by the microbial biomass
promote the precipitation of FeS and the consequent consumption of depending on their nature and physical characteristics. The following
the metabolite biosulphide produced by sulfate reducing bacteria, as chelating agents have been investigated: nitrilotriacetic acid (NTA) (Hu
well as the adsorption and co-precipitation of Ni and Co on FeS. Indeed, et al., 2008), ethylenediaminetetraacetic acid (EDTA) (Vintiloiu et al.,
it is a common practice in full-scale anaerobic digestion to supply Fe in 2013), and ethylenediamine-N,N′-disuccinic acid ([S,S]-EDDS) (Zhang
the form of FeCl2 or FeCl3 in order to reduce the H2S content in the et al., 2015c). In all these works, the addition of the chelating agents
biogas composition and thus lower the cost for H2S removal (Choong resulted in an increased methane production due to the improvement in
et al., 2016). The precipitation estimation in the context of AD is largely the element bioavailability (Choong et al., 2016). However, the effect of
limited to sulfide species. Sequential extraction has largely been used to such chelating agents to humans and the environment must be further
assess the different fractions of TEs in AD systems and in some studies assessed. Other chelating agents are represented by the soluble-micro-
the effect of precipitation/dissolution on methane production has been bial products and extracellular polymeric substances which enhance TE
explicitly investigated. For example, when metals were added bioavailability through metal binding and complexation (Choong et al.,

589
L. Frunzo, et al. Journal of Environmental Management 241 (2019) 587–602

Fig. 1. Schematic representation of the conversion pathways in the mathematical model.

2016). soluble and particulate components constituting the system. The model
has been synthesized to mechanistically include the most relevant
2.6. Sorption processes affecting TE in AD systems and to analyse the impact that the
combination of such processes exerts on the AD process stability and
Sorption is one of the main mechanisms affecting TE bioavailability. conversion efficiency. The processes have been implemented as sys-
It involves a series of processes, such as complex formation, chelation of tematic additions to the original ADM1. In general form, the model is
metals, ion exchange, adsorption, inorganic micro-precipitation and formulated in terms of four groups of state variables:
translocation of metal into the cell, which lead to TE immobilization
and sequestration on inorganic and/or organic particulate compounds, 1) n1 soluble components in liquid phase Si , including the compounds
which reduce the bioavailability of TE for the microbial species. deriving from the hydrolysis of the complex organic matter, all the
Biosorption has been recognized as a promising technology for the re- forms (ionic, free and total) of the bio-products from the biochem-
moval of several TE from contaminated wastewaters (Frunzo, 2017) ical reactions carried out by the diverse microbial groups con-
due to the high efficiency, cost effectiveness and affinity with metals, stituting the anaerobic consortium, the TE affecting the AD system
metalloids and other pollutants from many biological materials (Gadd, dynamics and, inorganic cations and anions;
2009). Indeed, the potential of many living organisms and agricultural 2) n2 n1 particulate components Xi , representing the concentration of
wastes as biosorbents has been largely studied and reviewed (Wang and the microbial groups carrying out the biochemical reactions, the
Chen, 2009). Since most microorganisms live in the form of biofilms, complex organic matter fed to the AD system, and the particulate
the different nature of the cell agglomerates and the presence/compo- material deriving from the hydrolysis and disintegration steps;
sition of EPS further contribute to biosorption (Flemming and 3) n1 gas components Sgas, i (i.e. hydrogen, carbon dioxide, methane,
Wingender, 2010; Comte et al., 2006; Papirio et al., 2017). Weakly hydrogen sulphide), which are in equilibrium with the corre-
acidic functional groups present in the cell wall and EPS are involved in sponding components in the liquid phase;
the mechanism proposed above. In addition to those groups, poly- 4) m3 particulate components Xp, i , representing the mineral pre-
saccharides and amino acids constituting the basic structure of the cell cipitates deriving from the mineral precipitation process involving
wall contain polarizable groups (sites), such as phosphate, carboxyl, specific soluble compounds Si .
hydroxyl and amino-groups, which are capable of interaction with ca-
tions. These sites may contribute to the reversible TE-binding capacity 3.1. Mass balance equations
of the biomass of granular sludge (Van Hullebush et al., 2003). In
systems involving bacterial colonies forming biofilms, the higher con- The differential equations governing TE, substrates, products and
centration of EPS directly results in increasing biosorption yields. The bacterial groups dynamics involved in the AD processes take the fol-
distribution, immobilization, and remobilization of TE depend on the lowing form:
sorption properties of the biofilms, determined by many factors in-
cluding pH, redox potential, and concentration of ligands present dVliq Si
= qin Sin, i qout Si + Vliq i A, i (t , S) Vliq T , i (t , S , Sgas)
dt
within the biofilm matrix (Thanh et al., 2016). m1 m
+ Vliq j = 1 i, j j
(t , S, X) Vliq j =21 S, j (t , S , X)
+ Vliq
m3 m
(t , S, Xp) Vliq j =41 C , j (t , S), i = 1, …, n1, t > 0, (1)
3. Mathematical model j = 1 i, j P , j

The proposed extended anaerobic digestion model accounting for dVliq Xi


= qin Xin, i qout Xi + Vliq
m1
j=1 i, j j (t , S , X),
TE dynamics in AD systems is based on mass conservation principles dt
i = n1 + 1, …, n2 , t > 0, (2)
and is formulated as a set of ordinary differential equations for the

590
Table 1
Biochemical rate coefficients ( i, j ) and kinetic rate equations ( j ) for the soluble components of the modified ADM1 including TEs dynamics.

Component→ 1 2 3 4 5 6 7 8
L. Frunzo, et al.

Process ↓
Ssu Saa Sfa Sva Sbu Spro Sac SH2

1 Disintegration of OFMSW
2 Hydrolysis of RB 1
Carbohydrates
3 Hydrolysis of SB 1
Carbohydrates
4 Hydrolysis of RB Proteins 1
5 Hydrolysis of SB Proteins 1
6 Hydrolysis of RB Lipids 1-ffa,li ffa,li
7 Hydrolysis of SB Lipids 1-ffa,li ffa,li
8 Uptake of Sugars −1 (1-Ysu) fbu,su (1-Ysu) fpro,su (1-Ysu) fac,su (1-Ysu) fH2,su
9 Uptake of AminoAcids −1 (1-Yaa) fva,aa (1-Yaa) fbu,aa (1-Yaa) fpro,aa (1-Yaa) fac,aa (1-Yaa) fH2,aa
10 Uptake of LCFA −1 (1-Yfa) 0.7 (1-Yfa) 0.3
11 Uptake of Valerate −1 (1-YC4) 0.54 (1-YC4) 0.31 (1-YC4) 0.15
12 Uptake of Butyrate −1 (1-YC4) 0.8 (1-YC4) 0.2
13 Uptake of Propionate −1 (1-Ypro) 0.57 (1-Ypro) 0.43
14 Uptake of Propionate by −1 (1-YpSRB) 0.57
pSRB
15 Uptake of Acetate −1
16 Uptake of Acetate by aSRB −1
17 Uptake of Hydrogen −1
18 Uptake of Hydrogen by −1
hSRB
19 Decay of Xsu

591
20 Decay of Xaa
21 Decay of Xfa
22 Decay of XC4
23 Decay of Xpro
24 Decay of XpSRB
25 Decay of Xac
26 Decay of XaSRB
27 Decay of XH2
28 Decay of XhSRB
Monosaccarides Amino Acids Long Chain Fatty Acids Total Valerate Total Butyrate Total Propionate Total Acetate Liquid Hydrogen
(kg COD m−3) (kg COD m−3) (kg COD m−3) (kg COD m−3) (kg COD m−3) (kg COD m−3) (kg COD m−3) (kg COD m−3)

9 10 11 12 13 14 15 16 Rate rj
−3−1
[kgCODmd /
−3−1
SCH4 SIC SIN SIP SIS SSO SI SMe kmol md ]

1 fSI,Xc 1
2 2
3 3
4 f_IS f_Maa 4
5 f_IS f_Maa 5
6 f_IP 6
7 f_IP 7
8 i = 1 9,11 16,38 55 Ci i,8
-(Ysu)Nbac

9 i = 1 9,11 16,38 55 Ci i,9


Naa-(Ysu)Nbac

(continued on next page)


Journal of Environmental Management 241 (2019) 587–602
Table 1 (continued)

9 10 11 12 13 14 15 16 Rate rj
−3−1
[kgCODmd /
L. Frunzo, et al.

−3−1
SCH4 SIC SIN SIP SIS SSO SI SMe kmol md ]

10 -(Yfa)Nbac 10
11 -(YC4)Nbac 11
12 -(YC4)Nbac 12
13 i = 1 9,11 16,38 55 Ci i,13
-(Ypro)Nbac

14 i = 1 9,11 16,38 55 Ci i,14


-(YpSRB)Nbac (1-YpSRB)0.43/64 -(1-YpSRB)0.43/64

15 (1-Yac) i = 1 9,11 16,38 55 Ci i,15


-(Yac)Nbac
-(Yac)Mbac
16 i = 1 9,11 16,38 55 Ci i,16
-(YaSRB)Nbac (1-YaSRB)/64 -(1-YaSRB)/64
-(YaSRB)Mbac
17 (1-Yac) i = 1 9,11 16,38 55 Ci i,17
-(YH2)Nbac
-(YH2)Mbac
18 i = 1 9,11 16,38 55 Ci i,18
-(YhSRB)Nbac (1-YhSRB)/64 -(1-YhSRB)/64
-(YhSRB)Mbac
19 i = 1 9,11 16,38 55 Ci i,19 i = 11,39 54 Ni i,19

20 i = 1 9,11 16,38 55 Ci i,20 i = 11,39 54 Ni i,20

21 i = 1 9,11 16,38 55 Ci i,21 i = 11,39 54 Ni i,21

592
22 i = 1 9,11 16,38 55 Ci i,22 i = 11,39 54 Ni i,22

23 i = 1 9,11 16,38 55 Ci i,23 i = 11,39 54 Ni i,23

24 i = 1 9,11 16,38 55 Ci i,24 i = 11,39 54 Ni i,24

25 i = 1 9,11 16,38 55 Ci i,25 i = 11,39 54 Ni i,25

26 i = 1 9,11 16,38 55 Ci i,26 i = 11,39 54 Ni i,26

27 i = 1 9,11 16,38 55 Ci i,27 i = 11,39 54 Ni i,27

28 i = 1 9,11 16,38 55 Ci i,28 i = 11,39 54 Ni i,28

Liquid Methane Inorganic Carbon Inorganic Nitrogen Inorganic Phosphorous Total sulphides Total sulphates Soluble Inerts Soluble Metal
(kg COD m−3) (kmol C m−3) (kmol N m−3) (kmol P m−3) (kmol S m−3) (kmol S m−3) (kg COD m−3) (kmol Me m−3)
Journal of Environmental Management 241 (2019) 587–602
L. Frunzo, et al. Journal of Environmental Management 241 (2019) 587–602

dVgas Sgas, i
= qgas Sgas, i + Vliq reaction rates adopted in the proposed model will be detailed in the
T , i (t , S, Sgas),
dt
following sections.
i = 1, …, n1, t > 0, (3)
dVliq Xp, i 3.2. Biochemical reaction rates
dt
= qin X pin, i qout Xp, i + Vliq P , i (t , S, Xp),
i = 1, …, m3, t > 0, (4) The large applicability and high versatility of the original ADM1 is
where: due to the simple structure of the mathematical model and the rea-
sonable prediction of the main products deriving from the anaerobic
n1 denotes the number of soluble components; conversion of complex organic matter. Several extensions have been
n2 n1 denotes the number of particulate components; proposed over the last decade to specifically focus on relevant processes
m1 denotes the number of biochemical processes taken into account; and biological interactions occurring in AD due to specific substrate
m2 denotes the number of sorption processes taken into account; compositions or reactor configurations. In agreement with the sche-
m3 denotes the number of precipitation processes taken into ac- matic approach of ADM1, five groups of biochemical reactions have
count; been individuated to account for TE dynamics in AD (Fig. 1): i) the
m4 denotes the number of complexation processes taken into ac- disintegration of complex organic compounds XC in carbohydrates Xch ,
count; proteins Xp , lipids Xli and, particulate inerts XI , ii) the hydrolysis of
these compounds in monosaccharides Xsu , amino acids Xaa , and fatty
i, j is the stoichiometric coefficient of species i on biochemical
process j; acids Xfa and XC4 , and the contextual release of inorganic carbon XIC ,
nitrogen XIN , phosphorous XIP , sulfur XIS and TEs SMe2 + , iii) the acid-
i, j is the stoichiometric coefficient of species i on precipitation
process j; ogenesis, iv) the acetogenesis and v) the methanogenesis steps.
The disintegration step has been modelled by using a surface based
i is the stoichiometric coefficient for the acid base reaction invol-
ving the ith soluble component; kinetic approach Eq. (10) which explicitly takes into account the spe-
Si denotes the ith soluble component, S = (S1, …, Sn1) ; cific surface area of organic compounds (Esposito et al., 2011b). The
Xi denotes the ith particulate component, X = (Xn1+ 1, …, Xn2 ) ; biodegradability characteristics of the hydrolysable substrates (carbo-
Sgas, i denotes the ith component in gas form, Sgas = (Sgas,1, …, Sgas, n1) ; hydrates, proteins and lipids) have been taken into account by the
Xp, i denotes the ith precipitate, Xp = (Xp,1, …, Xp, m3 ) ; addition of supplementary state variables accounting for the rapid
j (t , S , X) represents the rate of the jth biochemical process;
( Xch R , Xpr R and Xli R ) or slow ( Xch S , Xpr S and Xli S ) biodegrad-
S , j (t , S , X) represents the rate of the jth sorption process;
ability of these compounds (Esposito et al., 2011a,b). First order ki-
netics have been used to describe the hydrolysis of these compounds in
P , j (t , S , X p) represents the rate of the jth precipitation process;
anaerobic environments (Eqs. (11)–(16)).
C , j (t , S) represents the rate of the jth complexation process;
According to (De Gracia et al., 2006; Flores-Alsina et al., 2016), the
A, i (t , S) represents the acid base kinetic rate equation for the ith
soluble component; fate of sulphur and phosphorous have been included into the original
model by the addition of new state variables, namely total sulphates
T , i (t , S , S gas) represents the gas transfer rate for the ith soluble
component. SSO , total sulphides SIS and total phosphorous SIP , as reported in the
modified biochemical matrix of the proposed model (Tables 1 and 2).
Note that although the total amount of TE in anaerobic effluents is The oxidized sulphur is usually supplemented to the system through
normally used as a practical index of TE bioavailability, the effective industrial wastewaters. The last two compounds instead originate from
biouptakable amount of TE mainly depends on the concentrations of the the hydrolysis of proteins and amino acids (Shi and Chance, 2011) and
free metal ions within the reactor. For this reason, new dynamic state then assume different forms according to the acid base reactions re-
variables, which account for the concentration of the generic free metal ported in the following section. Similarly, TE are supposed to be sup-
ion in liquid phase, have been introduced for the mathematical mod- plemented to the anaerobic system or contained in the digestion sub-
elling of all the biological and physicochemical phenomena affecting TE strates. The supplementation is modelled as an inlet flux to the reactor
bioavailability in anaerobic systems. Moreover, a charge balance ac- and it is expressed as a function of the inlet flow rate and the con-
counting for all the ionic species is needed to completely define the centration of the supplemented metal ion. The TE contained in the
model as the pH of the system represents a major biological inhibition feedstock are supposed to be included mainly in the protein fraction of
factor. Such equation can be written in general form as follows: the digestion substrate. The main reason relies on the predominance of
p q
metal ions within the functional groups characterizing these com-
Qi+ Qi = 0, p + q < n1, pounds. TE release into the liquid phase derives from the hydrolysis of
i=1 i=1 (5) proteins and the amino acids bioconversion. Therefore, a specific pro-
duction rate has been considered in the mass balance equation related
where
to the generic metal ion. According to other studies (Fedorovich et al.,
p defines the number of cationic components;
2003; Batstone, 2006; Barrera et al., 2015), sulphate reduction has been
q defines the number of anionic components;
incorporated in the original ADM1 including the inhibitory effect of the
Qi+ represents the cationic equivalent concentration of species ith;
produced sulphide on methanogenic groups causing potential toxicity
Qi represents the anionic equivalent concentration of species ith.
and corrosion (Flores-Alsina et al., 2016). The behaviour of sulphate
The differential algebraic system obtained by coupling Eq. (3.5) to
reducing bacteria (SRB) competing for different intermediates, i.e.
the mass balance equations (3.1)–(3.4) can be solved after introducing
propionate, acetate and hydrogen, has been included by adding the new
the following initial conditions:
state variables XpSRB , XaSRB and XhSRB (Barrera et al., 2015). According
Si (0) = Si0, i = 1, …, n1, (6) to Fedorovich et al. (2003), dual term Monod type kinetics have been
used to describe the reduction of SSO by SRB using the above-mentioned
Xi (0) = Xi0 , i = n1 + 1, …, n2, (7) electron donors. In addition, the biomass decay processes related to SRB
0 have been included in the model and described as first order kinetics.
Sgas, i (0) = Sgas , i, i = 1, …, n1, (8)
To cover the discrepancy between the nitrogen content in the biomasses
Xp, i (0) = X p0, i , and the feeding composite material in the original ADM1, additional
i = 1, …, m3. (9)
terms related to the decay processes have been included for the in-
The biochemical, acid/base, gas transfer, sorption, precipitation organic nitrogen as in Blumensaat and Keller (2005). The inhibition

593
Table 2
Biochemical rate coefficients ( i, j ) and kinetic rate equations ( j ) for the particulate components of the modified ADM1 including TEs dynamics.

Component→ 38 39 40 41 42 43 44 45 46
L. Frunzo, et al.

Process ↓
XC Xch_R Xch_S Xpr_R Xpr_S Xli_R Xli_S Xsu Xaa

1 Disintegration of OFMSW −1 fCh,Xc fr fCh,Xc fs fPr,Xc fr fPr,Xc fs fLi,Xc fr fLi,Xc fs


2 Hydrolysis of RB −1
Carbohydrates
3 Hydrolysis of SB −1
Carbohydrates
4 Hydrolysis of RB Proteins −1
5 Hydrolysis of SB Proteins −1
6 Hydrolysis of RB Lipids −1
7 Hydrolysis of SB Lipids −1
8 Uptake of Sugars Ysu
9 Uptake of AminoAcids Yaa
10 Uptake of LCFA
11 Uptake of Valerate
12 Uptake of Butyrate
13 Uptake of Propionate
14 Uptake of Propionate by pSRB
15 Uptake of Acetate
16 Uptake of Acetate by aSRB
17 Uptake of Hydrogen
18 Uptake of Hydrogen by hSRB
19 Decay of Xsu 1 −1
20 Decay of Xaa 1 −1
21 Decay of Xfa 1

594
22 Decay of XC4 1
23 Decay of Xpro 1
24 Decay of XpSRB 1
25 Decay of Xac 1
26 Decay of XaSRB 1
27 Decay of XH2 1
28 Decay of XhSRB 1
Organic Composites R.Biodegradable S.Biodegradable R.Biodegradable S.Biodegradable R.Biodegradable S.Biodegradable Sugars Amino Acids
(kg COD m−3) Carbohydrates Carbohydrates Proteins Proteins Lipids Lipids Degraders Degraders
(kg COD m−3) (kg COD m−3) (kg COD m−3) (kg COD m−3) (kg COD m−3) (kg COD m−3) (kg COD m−3) (kg COD m−3)

47 48 49 50 51 52 53 54 55 Rate rj
−3
[kgCODm-
−1
Xfa XC4 Xpro XpSRB Xac XaSRB Xh2 XhSRB XI d /
−3−1
kmol md ]

1 fXi,Xc 1
2 2
3 3
4 4
5 5
6 6
7 7
8 8
9 9
10 Yfa 10
11 YC4 11
12 YC4 12
Journal of Environmental Management 241 (2019) 587–602
L. Frunzo, et al. Journal of Environmental Management 241 (2019) 587–602

terms for the uptake rates of the diverse microbial groups have been

−3−1
]
adopted from Barrera et al. (2015). TE have been proposed to affect
−3
[kgCODm-

kmol md
Rate rj

/
mainly the methanogenic activity, both hydrogenotrophic and acet-
−1

21
20
14

17

24
13

15
16

18
19

22
23

25
26
27
28
oclastic as methanogens have been proved to be strongly affected by TE
d

deprivation (Wintsche et al., 2016). Based on these reasonings, the


microbial kinetics of hydrogenotrophic and acetoclastic methanogens

Particulate Inerts
have been modified by including an inhibition function which depends

(kg COD m−3)


on the concentration of the free metal ions in the liquid phase (IMe2 +).
Such function accounts for the inhibition of methanogenic activity due
to TE deprivation and it has been modelled as a secondary substrate
55

XI

inhibition term. The functional forms of the inhibition terms Ii are re-
ported in Table 3. Furthermore, a consumption term has been in-
troduced in the mass balance for the generic free metal ion in liquid
phase to take into account the microorganisms' uptake of TEs. All the

(kg COD m−3)


SRB Hydrogen
kinetic rate equations i have been following listed according to the

Degraders
ADM1 and its further extensions (Esposito et al., 2011b; Frunzo et al.,
XhSRB

YhSRB

2012; Barrera et al., 2015):


−1
54

1 = Ksbk Ca , (10)
Hydrogen Degraders
(kg COD m−3) 2 = Khyd _ ch _ R X ch _ R , (11)

3 = Khyd _ ch _ S Xch _ S , (12)

4 = Khyd _ pr _ R Xpr _ R , (13)


−1
Xh2

Yh2
53

5 = Khyd _ pr _ S Xpr _ S , (14)


SRB Acetate Degraders

6 = Khyd _ li _ R Xli _ R , (15)


(kg COD m−3)

7 = Khyd _ li _ S Xli _ S , (16)

Ssu
XaSRB

YaSRB

= Km, su Xsu I1,


−1

8
(17)
52

K ssu + Ssu

Saa
9 = Km, aa Xaa I1,
(18)
Acetate Degraders

K saa + Saa
(kg COD m−3)

Sfa
10 = Km, fa Xfa I2,
K sfa + Sfa (19)
−1
Xac

Yac
51

Sva 1
11 = Km, c 4 XC 4 I3,
K sc4 + Sva 1 + Sbu /Sva (20)

Sbu 1
SRB Propionate

(kg COD m−3)

12 = Km, C 4 XC 4 I3,
K sc4 + Sbu 1 + Sva/ Sbu (21)
Degraders

Spro
XpSRB

YpSRB

= Km, pro Xpro I3,


−1

13
50

K spro + Spro (22)


Propionate Degraders

Spro SSO
14 = Km, pSRB XpSRB I6,
K spro + Spro K sSO4 pSRB + SSO (23)
(kg COD m−3)

Sac
15 = Km, ac Xac I4 ,
K sac + Sac (24)
Xpro

Ypro

−1
49

Sac SSO
16 = Km, aSRB XaSRB I6,
K sac + Sac K sSO4 aSRB + SSO (25)
Valerate and Butyrate

Sh2
(kg COD m−3)

17 = Km, h2 Xh2 I5,


K sh2 + Sh2 (26)
Degraders

Sh2 SSO
= Km, hSRB XhSRB I6,
−1
XC4

18
48

K sh2 + Sh2 K sSO4 hSRB + SSO (27)


(kg COD m−3)

= K dec, Xsu Xsu , (28)


Table 2 (continued)

19
Long Chain
Fatty Acids
Degraders

20 = K dec, Xaa Xaa , (29)


−1
Xfa
47

21 = K dec, Xfa Xfa , (30)


13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28

22 = K dec, Xc4 X c 4 , (31)

595
L. Frunzo, et al. Journal of Environmental Management 241 (2019) 587–602

Table 3
Functional form of the inhibition terms adopted in the proposed model.

Inhibition Term Expression Used for Reference

IpH Uptake of sugars, amino acids, LCFA, VFAs, hydrogen and sulphate reduction. Batstone et al. (2002)
( )
2
pH pHUL
exp 3 , pH < pHUL
pHUL pHLL

1, pH > pHUL
IIN , lim 1 Uptake of sugars, amino acids, LCFA, VFAs and hydrogen. Batstone et al. (2002)
1 + K S , IN / SIN
IH2 1 Uptake of LCFA and VFAs. Batstone et al. (2002)
1 + S H2 / KI , H2
IH2 S 1 Uptake of VFAs, hydrogen and sulphate reduction. Barrera et al. (2015)
1 + S H2 S / KI , H2 S
INH3 1 Uptake of acetate. Batstone et al. (2002)
1 + SNH3 / KI , NH3
IMe2 + 1 Uptake of acetate and hydrogen. This Study
1+K /S
I , Me2 + Me2 +

23 = K dec, Xpro Xpro , (32) H2 S H+ + HS , pka = 7.05, (46)

24 = K dec, XpSRB XpSRB , (33) HS H+ + S 2 , pka = 14, (47)

25 = K dec, Xac Xac , (34) H2 PO4 H+ + HPO42 , pka = 7.21, (48)

26 = K dec , XaSRB XaSRB , (35) HPO42 H+ + PO43 , pka = 12.32. (49)

27 = K dec, Xh2 Xh2 , (36) The rate coefficients and the kinetic rate equations for all the acid
base reactions implemented in the proposed model have been for-
28 = K dec, XhSRB XhSRB . (37) mulated as suggested by Rosen and Jeppsson (2006) and are reported in
Table 4. Note that for each acid base pair Si/ Si + 1, where Si corresponds
The inhibition terms in Eqs. (17)–(27) have been modified from
to the acidic component, the kinetic rate expressions are related as
Barrera et al. (2015) to consider the effect of TE on the microbial ki-
follows:
netics. They are formulated as follows:
A, i + 1 = A, i , (50)
I1 = IpH · IIN , lim, (38)
= KA / Bva (Sva (S H + + K a, va) Sva) , (51)
I2 = IpH · IIN , lim·IH2, (39) A,17

I3 = IpH · IIN , lim·IH2 ·IH2 S , (40) A,19 = KA / Bbu (Sbu (S H + + K a, bu) Sbu ), (52)

I4 = IpH · IIN , lim·INH3·IH2 S · IMe2 +, (41) A,21 = KA / Bpro (Spro (S H + + K a,pro) Spro), (53)
I5 = IpH ·IIN , lim· IH2 S ·IMe2 +, (42) A,23 = KA / Bac (Sac (S H + + Ka,ac) Sac ) , (54)
I6 = IpH ·IH2 S . (43) = KA / Bco2 (Shco3
A,25 (S H + + K a,co2) + K a, co2 Sco32 SIC ), (55)

3.3. Acid-base process rate A,26 = KA / Bhco3 (Sco32 (S H + + Ka,hco3) + K a,hco3 Sco2 SIC ) , (56)

= KA / Bnh4 (Snh3 (S H + + K a, nh3) Snh4), (57)


Additional acid-base reactions have been implemented in the pro- A,28

posed model in order to take into account specific ionic forms of the
= KA / h2po4 (Shpo42 (S H + + K a,h2po4) + K a,h2po4 Spo43 SIP ) , (58)
soluble components, reacting with the bioavailable forms of TE in A,30

physicochemical processes such as precipitation and complexation. For = KA / hpo4 (Spo43


what concerns inorganic carbon, the acid-base pair HCO3 /CO32 has A,31 (S H + + Ka,hpo4) + K a,hpo4 Sh2po4 SIP ) , (59)
been introduced according to Eq. (44), due to the relevant percentage of = KA / h2s (Shs
A,33 (S H + + K a, h2s) + K a,h2s S s2 SIS ) , (60)
carbonate precipitates usually found in AD effluents (Langerak et al.,
1999). According to Barrera et al. (2015), the acid-base reaction in- = KA / hs (S s2 (S H + + K a,hs) + K a,hs Sh2s SIS ) ,
A,34 (61)
volving the undissociated sulphuric acid H2 SO4 has been neglected as it
can be considered completely dissociated in water (pKa < −2). To A,36 = KA / Bso (Sso42 (S H + + Ka,hso4) Sso4 ) , (62)
account for SRB metabolism and sulphur rich wastewaters, the che-
mical equilibrium related to the pair HSO4 / SO42 has been included
according to Eq. (45). In relation to reduced sulphur compounds, the 3.4. Gas-transfer process rate
acid-base pairs H2 S /HS and HS / S 2 have been considered according
to Eqs. (46) and (47). The sulphide ions S 2 have been introduced to The liquid-gas transfer processes for the variables
effectively model the formation of sulphide precipitates. Similarly, the SH 2, SCH 4, SIC , SIS in the liquid phase have been considered. The for-
acid-base reactions accounting for the equilibrium pairs H2 PO4 /HPO42 mulation of the kinetic rate expressions follows the approach of Rosen
and HPO42 / PO43 have been implemented according to Eqs. (48) and and Jeppsson (2006) for SH 2, SCH 4, SIC and Barrera et al. (2015) for SIS
(49). The dissociation of H3 PO4 has not been considered due to the low (see Table 5). For all the other components in liquid phase T , i has been
pKa characterizing such reaction (pKa = 2.14). considered equal to zero.

HCO3 H+ + CO32 , pka = 10.33, (44) T ,8 = KLa (Sh2 16KH , h2 pgas, h2 ), (63)

HSO4 H+ + SO42 , pka = 1.99, (45) T ,9 = KLa (Sch4 64KH , ch4 pgas, ch4 ), (64)

596
L. Frunzo, et al. Journal of Environmental Management 241 (2019) 587–602

Table 5
Liquid phase yield coefficients and rate equations ( T ,i ) for the liquid-gas
[kmol m−3d−1] transfer processes implemented in the proposed model.

Component→ 8 9 10 13 Rate rj
Process ↓ [kgCODm−3d−1]
Rate rj

SH2 SCH4 SIC SIS


A,17
A,19
A,21
A,23
A,25
A,26
A,28
A,30
A,31
A,33
A,34
A,36
T8 H2 Transfer −1 T,8
2-

T9 CH4 Transfer −1
SSO4

−1
T,9
37

T10 CO2 Transfer −1 T,10


T12 H2S Transfer −1

T,12
SHSO4
36

1
for the acid base reactions in the differential implementation of the modified ADM1 including TEs dynamics.

T ,10 = KLa (Sco2 KH , co2 pgas, co2 ), (65)


2-

−1
35

SS

= KLa (Sh2s 48KH , h2s pgas, h2s ), (66)


T ,12
−1
SHS
34

1
SH2S

3.5. Sorption process rate


33

1
3-

The sorption processes have been implemented in the proposed


SPO4

−1
32

model by following the approach introduced in D'Acunto et al. (2015)


and D'Acunto et al. (2018). A non-reversible mechanism for TE sorption
SHPO42-

has been considered. The generic TE SMe2 + has been supposed to adsorb
−1
31

on the various particulate components, including the microbial bio-


masses constituting the anaerobic consortium and the inert biomass.
SH2PO4−

Each of these components is characterized by a specific concentration of


30

binding sites, which can be found in two states: occupied or free. The
1

fraction of free binding sites evolves over time due to the production of
Snh3

−1
29

new biomass and the progressive sorption of TE. Conversely, the frac-
tion of occupied binding sites can only increase over time as the sorp-
+
Snh4

tion phenomenon has been modelled as a non-reversible process. The


28

rate characterizing the jth sorption process has been formulated as


2-

follows:
Sco3

−1
27

¯X
S, j = K s, Xn1+ 1 + j SMe2 + (1 n1+ 1 + j ) N Xn1+ 1 + j Xn1+ 1 + j , j = 1, …, m2 , (67)
Shco3-

−1
26

where
1

Ks, Xn1+1 + j represents the sorption constant on the particulate com-


SCO2
25

ponent Xn1+ 1+ j ;
1

n1+ 1 + j denotes the fraction of occupied binding sites on the com-


¯X
−1
-
Sac
24

ponent Xn1+ 1+ j ;
N Xn1+ 1 +j is the site density of the component Xn1+ 1+ j .
Shac
23

Note that the sorption process on the complex organic matter


1

XC = Xn1+ 1 has not been included in the proposed model. The number
Spro-

−1
22

of sorption processes implemented in the present model coincides with


the number of particulate components (except the complex organic
Shpro

matter) taken into account in the biochemical reactions


21

1
A, i )

m2 = n2 (n1 + 1) = 17 .
Sbu-

−1
Biochemical rate coefficients ( i ) and kinetic rate equations (

20

3.6. Precipitation process rate


Shbu
19

Diverse precipitation processes affect TE availability in AD systems


−1
-
Sva
18

(Zhang et al., 2015a). The precipitation is supposed to occur when the


solution is oversaturated for the soluble components involved in the
Shva
17

specific precipitation process. It depends on the specific solubility


product (Ksp), representing the maximum number of undissociated ions
Dihydrogen sulphide acid base
Hydrogen Phosphate acid base
Inorganic nitrogen acid base

remaining in equilibrium with the undissolved phase. According to


Carbon dioxide acid base

Zhang et al. (2015a), the precipitation rates have been formulated as


Bicarbonate acid base
Propionate acid-base

kinetic based rate equations, whose expression is reported as in the


Phosphate acid base

Sulphide acid base


Sulphate acid base
Butyrate acid-base
Valerate acid-base

Acetate acid base

following equation:
Component→

1 v
Process ↓

v
+ +
1
prec = kr , m + A v
v
[([M m ]v ) [Aa ]v ] v ksp, m ,Av
v+

(68)
Table 4

A17
A19
A21
A23
A25
A26
A28
A31
A32
A34
A35
A37

where:

597
L. Frunzo, et al. Journal of Environmental Management 241 (2019) 587–602

Table 6
Precipitation rate coefficients ( i, j ) and rate equations ( P, j ) for the precipitation processes implemented in the proposed model.

Component→ 8 10 12 13 P,1 P,2 P,3 Rate P, j


Process ↓ [kmol m−3d−1]
2+
SMe SIC SIP SIS XMeCO3 XMe3(PO4)2 XMeS

P1 Carbonate precipitation −1 −1 1 P,1


P2 Phosphate precipitation −3 −2 1 P,2
P3 Sulphide precipitation −1 −1 1 P,3

pH 10-3 Sulfide Speciation


9 1
S 2-
8 H 2S

[Mol L -1]
HS -
pH

7 0.5

5 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Time [d] Time [d]
Cumulative biogas production 10-5 Iron
1.5 4
CH 4gas freeFe
CH 4 [LL -1d -1]

[Mol L -1]

1 3

0.5 2

0 1
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Time [d] Time [d]
10-8 IronSulfide Pr & AA
3 8
FeS Pr
S 2- 6 AA
[gCOD L -1]
[Mol L ]

2
-1

4
1
2

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Time [d] Time [d]
Fig. 2. Experimental values (square) and model predictions (continuous lines).

m v+ A v Concentrations of cations and anions coefficients related to the precipitation processes implemented in this
kr , m + A v is the precipitation rate constant model are reported in Table 6.
v
ksp, m +, A v is the solubility product 1/2 2
v
P ,1 = KP,1 ((SMe2 + SCO32 )1/2 kspk ,1 ) , (69)
v+ and v are the total number of cationic and anionic species and
v = v+ + v . 3 2 1/5 1/5 2
= KP,2 ((SMe 2+ S 3 ) kspk ,2 ) ,
P ,2 PO 4 (70)
For each precipitating compound, the free cations and anions, ty-
1/2 2
pically inorganic carbon, inorganic sulphur and inorganic phosphorous, P ,3 = KP,3 ((SMe2 + S S2 )1/2 kspk ,3 ) . (71)
dynamically react following specific kinetics, which are regulated by
the charge balance and the precipitation and solubility constant,
namely kr , m + A v .and ksp, m +, A v , respectively. In the proposed model, 4. Application and numerical results
v v
three precipitation rates have been included for the generic trace ele-
ment SMe2 + . The latter is supposed to precipitate in the form of carbo- FW is one of the most promising waste organic substrates to be used
nate, sulphide and phosphate according to Eqs. (69)–(71). The rate as feedstock for AD and renewable energy production. It has been es-
timated that FW production in China could approximately reach

598
L. Frunzo, et al. Journal of Environmental Management 241 (2019) 587–602

Table 7
Operational parameters used for the numerical simulations.

Parameter Unit Simulation Set Simulation Set 2


1

Digester volume mL 150 150


Headspace volume mL 30 30
Influent FW rate mL d−1 5 5
Total FW Carbohydrates fraction – 0.65 0.65
Total FW Proteins fraction – 0.20 0.20
Total FW Lipids fraction – 0.05 0.05
Total FW Inert fraction – 0.10 0.10
Temperature °C 37 37
Sulfur content of Proteins mol gCOD−1 0.075 0.0075
Sulfur content of Lipids mol gCOD−1 0.015 0.015
Iron content of AA mol gCOD−1 0.0001 0.0001
Iron content of Lipid mol gCOD−1 0.00075 0.00075
Iron content of inoculum mol gCOD−1 0.255 0
Iron supply mol L−1 0 5*10−3-1*10−7
Henry's law coefficient for H2S mol bar−1 0.1 0.1
Gas-Liquid H2S transfer coefficient d−1 200 200
FeS solubility product (Pr_FeS) mol L−1 104 104
FeS precipitation rate constant (Ksp_FeS) d−1 18 18
Stimulation constant of Fe2+ (KI,Fe) mol L−1 0.0724 0.0724

1.4 × 108 tons per year in 2020, which is equivalent to 10 million tons parameter value for mesophilic AD (Batstone et al., 2002). In addition,
of coal based on energy conversion by electricity production (Ye et al., the kinetic disintegration constant Ksbk and the hydrolysis kinetic
2018). Despite the advantages deriving from the conversion of carbon- constants for all simulations have been derived from literature (Esposito
rich substrate via AD, nowadays the key issues of FW bio-conversion lie et al., 2012). The organic substrate has been fractionated as 50%
on the instability and unefficiency of AD reactors in long-term appli- readily biodegradable and 50% slowly biodegradable. The kLa coeffi-
cations (Zhang and Jahng, 2012). This is mainly due to the lack of some cients for liquid-to-gas transfer of methane and hydrogen have been
nutrients (i.e. TE) that are limiting for the anaerobic microorganisms adopted according to Esposito et al. (2011a) and (2011b), respectively,
metabolism and induce VFA accumulation and poor process perfor- whereas the kLa value of 200 d−1 (Barrera et al., 2015) has been applied
mance for FW anaerobic conversion (Romero-Güiza et al., 2016). To for hydrogen sulfide.
overcome this problem, co-digestion with TE-rich substrates (such as In accordance to the experimental evidence, iron represents the
piggery wastewater) and TE external supplementation have been limiting factor leading to the reactor failure after 80 days of operation.
identified as promising strategies for the optimization of FW anaerobic To properly model the effect of iron deprivation on AD process, a
conversion (Zhang et al., 2015b, Romero-Güiza et al., 2016). specific inhibition term has been introduced as reported in Table 3. The
The availability of a mathematical model able to predict TE defi- expression for IFe2 + has been calibrated by setting a kinetic constant
ciency and the optimal supplementation strategy is a fundamental re- KI , Fe2 + to the value of 0.0724 mol L−1.
quirement for the management of AD bio-reactors at full scale. For this The model outputs have been reported in Fig. 2. Numerical simu-
reason, the proposed mathematical model has been applied to two il- lations demonstrated the model capability of predicting pH, sulphide
lustrative cases: the first one is based on the experimental results speciation, methane production, free iron, iron sulphide, protein and
achieved by Zhang and Jahng (2012), who examine the effect of trace amino acid concentrations. Notably, the pH evolution over time and the
metals limitation in an AD process of FW. The second case investigates methane production predicted by the model have been compared with
the effects of different iron dosing on methane production and it is the experimental data. The model reproduces in a reasonable way both
aimed at individuating the optimal external iron supplementation in the methane and pH profiles and it was able to accurately predict the
long-term operations. In both cases, iron has been considered as the process failure after about 80 operation days.
only limiting TE due to its recognized high requirements in AD systems Due to the semi-continuous operation, the iron concentration de-
(Choong et al., 2016). creases and the low iron content of the FW gradually determines the
Application 1: Effect of iron limitation in an AD process of FW. failure of the reactor. Indeed, the inhibition of methanogenesis results
The first numerical application has been based on the experimental in VFA accumulation and fast decrease of the reactor pH. The pH value
results of Zhang and Jahng (2012). In their work, the authors carried- drops from 7.5 to 5; the sulphide speciation changes and the amino acid
out different semi-continuous AD experiments for FW conversion aimed hydrolysis efficiency decreases. Indeed, when the pH reaches the value
at identifying the role of the addition of an individual trace element or a of 5, the sulphide is mainly present in the form of H2S, which leads to
solution containing different TE in AD bio-reactors. the increase of H2S gas production (data not shown) and the consequent
To this aim, seven reactors were operated in parallel. Six reactors inhibition of amino acid hydrolysis. The latter causes an additional
were supplied with individual (or combined) TE solutions while one decrease of free iron content into the reactor.
rector was operated for 95 days without any trace element addition. By Application 2: Effect of iron supply in an AD process of FW.
analysing the behaviours of the six reactors, the authors highlighted the Figs. 3 and 4 show the results of model simulations (Simulation Set
remarkable effect of the low iron content in the FW that differently 2, Table 7) performed to assess the effect of different external iron
from Co, Mo and Ni, plays a crucial role for the stability of the process. supply on AD performance. This model application accounts for the
The experiment related to the reactor without TE addition (cfr. effect of increasing free iron concentration on sulfide speciation and
Reactor R1) has been qualitatively reproduced in silico (Fig. 2). The biogas composition during the biological anaerobic conversion of FW.
operational parameters used for the numerical simulations have been The external addition of iron in the range of
reported in Table 7. 0.005–0.000005 mol L−1 has been investigated to explore the dynamic
The values of the kinetic and stoichiometric parameters as well as effects of iron addition and optimize the TE dosing strategy. To reach a
the Henry's constants have been selected according to the suggested stable initial condition (day 0), 100 simulation days have been

599
L. Frunzo, et al. Journal of Environmental Management 241 (2019) 587–602

CH 4 production Free Iron concentration


2 10-2

1.5

Fe2+ [Mol L -1]


CH 4 [LL -1d -1]

10-3

10-4
0.5

Increasing Iron supply


0 10-5
-100 0 100 200 -100 0 100 200
Time [d] Time [d]

H 2S production
10-4 10-6 Iron Sulfide concentration
5 3
Increasing Iron supply
2.5
4
H 2S [LL -1 d-1]

FeS [Mol L -1]

2
3
1.5
2
1

1
0.5
Increasing Iron supply
0 0
-100 0 100 200 -100 0 100 200
Time [d] Time [d]
0.005M Fe In 0.0005M Fe In 0.00005M Fe In 0.000005M Fe In

Fig. 3. Methane production, free iron concentration, hydrogen sulphide gas production and iron sulphide concentration with increasing iron supplementation.

considered for all the simulation sets. prediction of real scale reactor failures under very low TE concentration
As it is possible to notice (Fig. 3), in all the experiments the methane regime, sophisticated experimental techniques for the data collection
production profile denotes the same trend as increasing iron supple- and ad-hoc experimental activities are still required.
mentation does not affect methanogenic activity. Indeed, the minimum
free iron requirement for methanogenic bacteria is available during all
the simulation tests. On the other hand, the increase of iron supply 5. Conclusion
directly results in the increase of iron sulphide concentration, with
consequent precipitation of sulfur compounds and a significant de- A mathematical model able to consider the complex dynamics of
crease of H2S gas production during AD of FW. TEs in anaerobic systems has been presented. The model was im-
The potential use of the presented model as a predictive tool for TE plemented as an original extension of ADM1 and it is able to account for
deficiency in AD has been further highlighted by investigating the effect the main processes affecting TE evolution in AD bio-reactors. Two
of lower external iron additions (in the range of different applications have been presented: the first relates to a real
0.000005–0.0000001 mol L−1) to the simulated bioreactor (Fig. 4). experimental case where TE starvation occurs during the AD of FW. The
Due to extremely low free iron concentrations, methane production second application has been aimed at evaluating the biogas composi-
profiles collapsed after 85, 100 and 130 days when the initial iron tion and hydrogen sulphide production due to different TE external
concentration was set to 0.000001, 0.0000005 and 0.0000001 mol L−1, additions. Due to the complex interactions among the biological, che-
respectively. According to this evidence, iron sulphide concentration mical and physicochemical processes involved in TE speciation, the
and precipitation was not significant and very low H2S gas production model can be used as a predictive tool to obtain a deeper understanding
were observed before the AD reactor failure. Noteworthy, to definitely of TE bioavailability and to develop an optimal supplementation
reach an appropriate calibration and validation of the model for the strategy for full scale applications.

600
L. Frunzo, et al. Journal of Environmental Management 241 (2019) 587–602

CH 4 production Free Iron concentration


1.5 10-3

Fe2+ [Mol L -1]


CH 4 [LL -1d -1]
1 10-4

0.5 10-5

0 10-6
-100 0 100 200 -100 0 100 200
Time [d] Time [d]

H 2S production
10-3 10-7 Iron Sulfide concentration
2 1.5

1.5
H 2S [LL -1 d-1]

FeS [Mol L -1]


1

0.5
0.5

0 0
-100 0 100 200 -100 0 100 200
Time [d] Time [d]
0.000005M Fe In 0.000001M Fe In 0.0000005M Fe In 0.0000001M Fe In

Fig. 4. Methane production, free iron concentration, hydrogen sulphide gas production and iron sulphide concentration with lowering iron supplementation.

Acknowledgments 1947–1957.
Callander, I.J., Barford, J.P., 1983b. Precipitation, chelation, and the availability of me-
tals as nutrients in anaerobic digestion. II. Applications. Biotechnol. Bioeng. 25 (8),
This work has been developed in the framework of the COST Action 1959–1972.
1302 (“European Network on Ecological Roles of Trace Metals in Cao, W., Wang, M., Liu, M., Zhang, Z., Sun, Z., Miao, Y., Hu, C., 2018. The chemical and
Anaerobic Biotechnologies”) supported by COST (European dynamic distribution characteristics of iron, cobalt and nickel in three different
anaerobic digestates: effect of pH and trace elements dosage. Bioresour. Technol.
Cooperation in Science and Technology). L. Frunzo and M.R. Mattei 269, 363–374.
acknowledge the support from the Progetto Giovani G.N.F.M. 2017 Cheng, X., Wang, J., Chen, B., Wang, Y., Liu, J., Liu, L., 2017. Effectiveness of phosphate
"Analisi di Sistemi Biologici Complessi". L. Frunzo also acknowledges removal during anaerobic digestion of waste activated sludge by dosing iron (III). J.
Environ. Manag. 193, 32–39.
the support from the project VOLAC - Valorization of OLive oil wastes Choong, Y.Y., Norli, I., Abdullah, A.Z., Yhaya, M.F., 2016. Impacts of trace element
for sustainable production of biocide-free Antibiofilm Compounds, supplementation on the performance of anaerobic digestion process: a critical review.
funded by CARIPLO foundation. Bioresour. Technol. 209, 369–379.
Comte, S., Guibaud, G., Baudu, M., 2006. Biosorption properties of extracellular poly-
meric substances (EPS) resulting from activated sludge according to their type: so-
References luble or bound. Process Biochem. 41 (4), 815–823.
D'Acunto, B., Esposito, G., Frunzo, L., Mattei, M.R., Pirozzi, F., 2015. Mathematical
Bachmann, S., Wentzel, S., Eichler-Löbermann, B., 2011. Codigested dairy slurry as a modeling of heavy metal biosorption in multispecies biofilms. J. Environ. Eng. 142
phosphorus and nitrogen source for Zea mays L. and Amaranthus cruentus L. J. Plant (9), C4015020.
Nutr. Soil Sci. 174 (6), 908–915. D'Acunto, B., Frunzo, L., Mattei, M.R., 2018. On a free boundary problem for biosorption
Barrera, E.L., Spanjers, H., Solon, K., Amerlinck, Y., Nopens, I., Dewulf, J., 2015. in biofilms. Nonlinear Anal. Real World Appl. 39, 120–141.
Modeling the anaerobic digestion of cane-molasses vinasse: extension of the De Gracia, M., Sancho, L., García-Heras, J.L., Vanrolleghem, P., Ayesa, E., 2006. Mass and
Anaerobic Digestion Model No. 1 (ADM1) with sulfate reduction for a very high charge conservation check in dynamic models: application to the new ADM1 model.
strength and sulfate rich wastewater. Water Res. 71, 42–54. Water Sci. Technol. 53 (1), 225–240.
Batstone, D.J., 2006. Mathematical modelling of anaerobic reactors treating domestic De-Bashan, L.E., Bashan, Y., 2004. Recent advances in removing phosphorus from was-
wastewater: rational criteria for model use. Rev. Environ. Sci. Bio/Technol. 5 (1), tewater and its future use as fertilizer (1997–2003). Water Res. 38 (19), 4222–4246.
57–71. Esposito, G., Frunzo, L., Panico, A., Pirozzi, F., 2011a. Modelling the effect of the OLR and
Batstone, D.J., Keller, J., 2003. Industrial applications of the IWA anaerobic digestion OFMSW particle size on the performances of an anaerobic co-digestion reactor.
model No. 1 (ADM1). Water Sci. Technol. 47 (12), 199–206. Process Biochem. 46 (2), 557–565.
Batstone, D.J., et al., 2002. The IWA anaerobic digestion model no 1 (ADM1). Water Sci. Esposito, G., Frunzo, L., Panico, A., Pirozzi, F., 2011b. Model calibration and validation
Technol. 45, 65–73. for OFMSW and sewage sludge co-digestion reactors. Waste Manag. 31 (12),
Blumensaat, F., Keller, J., 2005. Modelling of two-stage anaerobic digestion using the 2527–2535.
IWA Anaerobic Digestion Model No. 1 (ADM1). Water Res. 39 (1), 171–183. Esposito, G., Frunzo, L., Giordano, A., Liotta, F., Panico, A., Pirozzi, F., 2012. Anaerobic
Callander, I.J., Barford, J.P., 1983a. Precipitation, chelation, and the availability of me- co-digestion of organic wastes. Rev. Environ. Sci. Bio/Technol. 11 (4), 325–341.
tals as nutrients in anaerobic digestion. I. Methodology. Biotechnol. Bioeng. 25 (8), Fedorovich, V., Lens, P., Kalyuzhnyi, S., 2003. Extension of enaerobic digestion model No.

601
L. Frunzo, et al. Journal of Environmental Management 241 (2019) 587–602

1 with processes of sulfate reduction. Appl. Biochem. Biotechnol. 109 (1–3), 33–45. Ortner, M., Rameder, M., Rachbauer, L., Bochmann, G., Fuchs, W., 2015. Bioavailability
Fermoso, F.G., Bartacek, J., Jansen, S., Lens, P.N.L., 2009. Metal supplementation to of essential trace elements and their impact on anaerobic digestion of slaughterhouse
UASB bioreactors: from cell-metal interactions to full-scale application. Sci. Total. waste. Biochem. Eng. J. 99, 107–113.
Environ. 407 (12), 3652–3667. Papirio, S., Frunzo, L., Mattei, M.R., Ferraro, A., Race, M., D'Acunto, B., Esposito, G.,
Fermoso, F.G., Collins, G., Bartacek, J., O'Flaherty, V., Lens, P., 2008. Acidification of 2017. Heavy metal removal from wastewaters by biosorption: mechanisms and
methanol-fed anaerobic granular sludge bioreactors by cobalt deprivation: induction modeling. In: Sustainable Heavy Metal Remediation. Springer, Cham, pp. 25–63.
and microbial community dynamics. Biotechnol. Bioeng. 99 (1), 49–58. Romero-Güiza, M.S., Vila, J., Mata-Alvarez, J., Chimenos, J.M., Astals, S., 2016. The role
Fermoso, F.G., Van Hullebusch, E.D., Guibaud, G., Collins, G., Svensson, B.H., Carliell- of additives on anaerobic digestion: a review. Renew. Sustain. Energy Rev. 58,
Marquet, C., Frunzo, L., 2015. Fate of trace metals in anaerobic digestion. In: Biogas 1486–1499.
Science and Technology. Springer International Publishing, pp. 171–195. Rosén, C., Jeppsson, U., 2006. Aspects on ADM1 Implementation within the BSM2
Flemming, H.C., Wingender, J., 2010. The biofilm matrix. Nat. Rev. Microbiol. 8 (9), 623. Framework. Department of Industrial Electrical Engineering and Automation, Lund
Flores-Alsina, X., Solon, K., Mbamba, C.K., Tait, S., Gernaey, K.V., Jeppsson, U., Batstone, University, Lund,, Sweden, pp. 1–35.
D.J., 2016. Modelling phosphorus (P), sulfur (S) and iron (Fe) interactions for dy- Roussel, J., Carliell-Marquet, C., 2016. Significance of vivianite precipitation on the
namic simulations of anaerobic digestion processes. Water Res. 95, 370–382. mobility of iron in anaerobically digested sludge. Front. Environ. Sci. 4, 1–12.
Frunzo, L., 2017. Modeling Sorption of Emerging Contaminants in Biofilms. arXiv pre- https://doi.org/10.3389/fenvs.2016.00060.
print arXiv:1706.04541. Saidou, H., Korchef, A., Moussa, S.B., Amor, M.B., 2009. Struvite precipitation by the
Frunzo, L., Esposito, G., Pirozzi, F., Lens, P., 2012. Dynamic mathematical modeling of dissolved CO2 degasification technique: impact of the airflow rate and pH.
sulfate reducing gas-lift reactors. Process Biochem. 47 (12), 2172–2181. Chemosphere 74 (2), 338–343.
Gadd, G.M., 2009. Biosorption: critical review of scientific rationale, environmental im- Sanchez, E., Borja, R., Weiland, P., Travieso, L., Martin, A., 2000. Effect of temperature
portance and significance for pollution treatment. J. Chem. Technol. Biotechnol. 84 and pH on the kinetics of methane production, organic nitrogen and phosphorus
(1), 13–28. removal in the batch anaerobic digestion process of cattle manure. Bioprocess Eng. 22
Gerardi, M.H., 2006. Wastewater Bacteria, vol. 5 John Wiley & Sons. (3), 247–252.
Glass, J., Orphan, V.J., 2012. Trace metal requirements for microbial enzymes involved in Shi, W., Chance, M.R., 2011. Metalloproteomics: forward and reverse approaches in
the production and consumption of methane and nitrous oxide. Front. Microbiol. metalloprotein structural and functional characterization. Curr. Opin. Chem. Biol. 15
3, 61. (1), 144–148.
Gonzalez-Gil, G., Kleerebezem, R., Lettinga, G., 1999. Effects of nickel and cobalt on Thanh, P.M., Ketheesan, B., Yan, Z., Stuckey, D., 2016. Trace metal speciation and
kinetics of methanol conversion by methanogenic sludge as assessed by on-line CH4 bioavailability in anaerobic digestion: a review. Biotechnol. Adv. 34 (2), 122–136.
monitoring. Appl. Environ. Microbiol. 65 (4), 1789–1793. Thanh, P.M., Ketheesan, B., Yan, Z., Stuckey, D., 2017. Effect of Ethylenediamine-N, N′-
Gustavsson, J., 2012. Cobalt and Nickel Bioavailability for Biogas Formation. PhD disuccinic acid (EDDS) on the speciation and bioavailability of Fe2+ in the presence
Thesis. University of Linkoping, Sweden. of sulfide in anaerobic digestion. Bioresour. Technol. 229, 169–179.
Gustavsson, J., Yekta, S.S., Sundberg, C., Karlsson, A., Ejlertsson, J., Skyllberg, U., van Hullebusch, E.D., Guibaud, G., Simon, S., Lenz, M., Yekta, S.S., Fermoso, F.G., Jain,
Svensson, B.H., 2013. Bioavailability of cobalt and nickel during anaerobic digestion R., Duester, L., Roussel, J., Guillon, E., Skyllberg, U., Almeida, C.M.R., Pechaud, Y.,
of sulfur-rich stillage for biogas formation. Appl. Energy 112, 473–477. Garuti, M., Frunzo, L., Esposito, G., Carliell-Marquet, C., Ortner, M., Collins, G., 2016.
Hilton, B.L., Oleszkiewicz, J.A., 1988. Sulfide-induced inhibition of anaerobic digestion. Methodological approaches for fractionation and speciation to estimate trace element
J. Environ. Eng. 114 (6), 1377–1391. bioavailability in engineered anaerobic digestion ecosystems: an overview. Critic.
Hu, Q.H., Li, X.F., Du, G.C., Chen, J., 2008. Effect of nitrilotriacetic acid on bioavailability Rev. Eviron. Sci. Technol. 46 (16), 1324–1366.
of nickel during methane fermentation. Chem. Eng. J. 143 (1–3), 111–116. Van Hullebusch, E.D., Zandvoort, M.H., Lens, P.N., 2003. Metal immobilisation by bio-
Jansen, S., Gonzalez-Gil, G., van Leeuwen, H.P., 2007. The impact of Co and Ni speciation films: mechanisms and analytical tools. Rev. Environ. Sci. Biotechnol. 2 (1), 9–33.
on methanogenesis in sulfidic media-biouptake versus metal dissolution. Enzym. Vintiloiu, A., Boxriker, M., Lemmer, A., Oechsner, H., Jungbluth, T., Mathies, E.,
Microb. Technol. 40 (4), 823–830. Ramhold, D., 2013. Effect of ethylenediaminetetraacetic acid (EDTA) on the bioa-
Keating, C., Chin, J.P., Hughes, D., Manesiotis, P., Cysneiros, D., Mahony, T., Smith, C.J., vailability of trace elements during anaerobic digestion. Chem. Eng. J. 223, 436–441.
Mc Grath, J.W., O’Flaherty, V., 2016. Biological phosphorus removal during high- Wang, J., Chen, C., 2009. Biosorbents for heavy metals removal and their future.
rate, low-temperature, anaerobic digestion of wastewater. Front. Microbiol. 7, 226. Biotechnol. Adv. 27 (2), 195–226.
Kodera, H., Hatamoto, M., Abe, K., Kindaichi, T., Ozaki, N., Ohashi, A., 2013. Phosphate Wintsche, B., Glaser, K., Sträuber, H., Centler, F., Liebetrau, J., Harms, H., Kleinsteuber,
recovery as concentrated solution from treated wastewater by a PAO-enriched bio- S., 2016. Trace elements induce predominance among methanogenic activity in
film reactor. Water Res. 47 (6), 2025–2032. anaerobic digestion. Front. Microbiol. 7, 2034.
Langerak, E.P.A.V., Beekmans, M.M.H., Beun, J.J., Hamelers, H.V.M., Lettinga, G., 1999. Ye, M., Liu, J., Ma, C., Li, Y.Y., Zou, L., Qian, G., Xu, Z.P., 2018. Improving the stability
Influence of phosphate and iron on the extent of calcium carbonate precipitation and efficiency of anaerobic digestion of food waste using additives: a critical review.
during anaerobic digestion. J. Chem. Technol. Biotechnol. Int. Res. Process Environ. J. Clean. Prod. 192, 316–326.
Clean Technol. 74 (11), 1030–1036. Yekta, S.S., Svensson, B.H., Björn, A., Skyllberg, U., 2014. Thermodynamic modeling of
Lee, J., Park, K.Y., Cho, J., Kim, J.Y., 2018. Releasing characteristics and fate of heavy iron and trace metal solubility and speciation under sulfidic and ferruginous condi-
metals from phytoremediation crop residues during anaerobic digestion. tions in full scale continuous stirred tank biogas reactors. Appl. Geochem. 47, 61–73.
Chemosphere 191, 520–526. Yekta, S.S., Skyllberg, U., Danielsson, Å., Björn, A., Svensson, B.H., 2017. Chemical
Lin, C.Y., 1993. Effect of heavy metals on acidogenesis in anaerobic digestion. Water Res. speciation of sulfur and metals in biogas reactors–Implications for cobalt and nickel
27 (1), 147–152. bio-uptake processes. J. Hazard. Mater. 324, 110–116.
Liu, Y., Wang, Q., Zhang, Y., Ni, B.J., 2015. Zero valent iron significantly enhances me- Zandvoort, M.H., Van Hullebusch, E.D., Fermoso, F.G., Lens, P.N.L., 2006. Trace metals in
thane production from waste activated sludge by improving biochemical methane anaerobic granular sludge reactors: bioavailability and dosing strategies. Eng. Life
potential rather than hydrolysis rate. Sci. Rep. 5, 1–6. https://doi.org/10.1038/ Sci. 6 (3), 293–301.
srep08263. Zhang, L., Jahng, D., 2012. Long-term anaerobic digestion of food waste stabilized by
Moestedt, J., Nordell, E., Yekta, S.S., Lundgren, J., Martí, M., Sundberg, C., Björn, A., trace elements. Waste Manag. 32 (8), 1509–1515.
2016. Effects of trace element addition on process stability during anaerobic co-di- Zhang, L., Keller, J., Yuan, Z., 2009. Inhibition of sulfate-reducing and methanogenic
gestion of OFMSW and slaughterhouse waste. Waste Manag. 47, 11–20. activities of anaerobic sewer biofilms by ferric iron dosing. Water Res. 43 (17),
Möller, K., Müller, T., 2012. Effects of anaerobic digestion on digestate nutrient avail- 4123–4132.
ability and crop growth: a review. Eng. Life Sci. 12 (3), 242–257. Zhang, Y., Piccard, S., Zhou, W., 2015a. Improved ADM1 model for anaerobic digestion
Möller, K., Stinner, W., 2010. Effects of organic wastes digestion for biogas production on process considering physico-chemical reactions. Bioresour. Technol. 196, 279–289.
mineral nutrient availability of biogas effluents. Nutrient Cycl. Agroecosyst. 87 (3), Zhang, W., Wu, S., Guo, J., Zhou, J., Dong, R., 2015b. Performance and kinetic evaluation
395–413. of semi-continuously fed anaerobic digesters treating food waste: role of trace ele-
Mudhoo, A., Kumar, S., 2013. Effects of heavy metals as stress factors on anaerobic di- ments. Bioresour. Technol. 178, 297–305.
gestion processes and biogas production from biomass. Int. J. Environ. Sci. Technol. Zhang, W., Zhang, L., Li, A., 2015c. Enhanced anaerobic digestion of food waste by trace
10 (6), 1383–1398. metal elements supplementation and reduced metals dosage by green chelating agent
Ortner, M., Rachbauer, L., Somitsch, W., Fuchs, W., 2014. Can bioavailability of trace [S, S]-EDDS via improving metals bioavailability. Water Res. 84, 266–277.
nutrients be measured in anaerobic digestion? Appl. Energy 126, 190–198.

602

You might also like