toxicants inhibiting anaerobic digestion: a review

52
Toxicants Inhibiting Anaerobic Digestion: A Review Jian Lin Chen, Raphael Ortiz, Terry W.J. Steele, David C. Stuckey PII: S0734-9750(14)00154-2 DOI: doi: 10.1016/j.biotechadv.2014.10.005 Reference: JBA 6849 To appear in: Biotechnology Advances Received date: 28 March 2014 Revised date: 8 October 2014 Accepted date: 8 October 2014 Please cite this article as: Chen Jian Lin, Ortiz Raphael, Steele Terry W.J., Stuckey David C., Toxicants Inhibiting Anaerobic Digestion: A Review, Biotechnology Advances (2014), doi: 10.1016/j.biotechadv.2014.10.005 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.

Upload: david-c

Post on 27-Feb-2017

221 views

Category:

Documents


7 download

TRANSCRIPT

Page 1: Toxicants inhibiting anaerobic digestion: A review

�������� ����� ��

Toxicants Inhibiting Anaerobic Digestion: A Review

Jian Lin Chen, Raphael Ortiz, Terry W.J. Steele, David C. Stuckey

PII: S0734-9750(14)00154-2DOI: doi: 10.1016/j.biotechadv.2014.10.005Reference: JBA 6849

To appear in: Biotechnology Advances

Received date: 28 March 2014Revised date: 8 October 2014Accepted date: 8 October 2014

Please cite this article as: Chen Jian Lin, Ortiz Raphael, Steele Terry W.J., StuckeyDavid C., Toxicants Inhibiting Anaerobic Digestion: A Review, Biotechnology Advances(2014), doi: 10.1016/j.biotechadv.2014.10.005

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 proofbefore it is published in its final form. Please note that during the production processerrors may be discovered which could affect the content, and all legal disclaimers thatapply to the journal pertain.

Page 2: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

1

Toxicants Inhibiting Anaerobic Digestion: A

Review

Jian Lin Chen a, Raphael Ortiz

b, Terry W.J. Steele

b *, and David C. Stuckey

a, c *

a Advanced Environmental Biotechnology Centre, Nanyang Environment & Water Research

Institute, Nanyang Technological University, Singapore 637141.

b School of Materials Science & Engineering, College of Engineering, Nanyang Technological

University, Singapore 637141.

c Department of Chemical Engineering, Imperial College London, London SW7 2AZ, and

Nanyang Environment & Water Research Institute, Advanced Environmental Biotechnology

Centre, Nanyang Technological University, Singapore 637141.

*Corresponding authors at: c

Department of Chemical Engineering, Imperial College

London, London SW7 2AZ, UK. Tel.: +44 207 5945591; fax: +32 15 317453. E-mail

address: [email protected] (David C. Stuckey); b

School of Materials Science &

Engineering, Nanyang Technological University N4.1-01-29, 50 Nanyang Drive,

Singapore 639798. Tel: +65-6592-7594, Fax: +65-6790-9081, Email address:

[email protected] (Terry W.J. Steele).

Page 3: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

2

Content

1 Introduction

2 Toxicants

2.1 Organic Toxicants

2.1.1 Chlorophenols

2.1.1.1 Toxicity of chlorophenols to anaerobic digestion

2.1.1.2 Mechanism of chlorophenols’ toxicity

2.1.2 Halogenated aliphatics

2.1.2.1 Toxicity of halogenated aliphatics to anaerobic digestion

2.1.2.2 Mechanism of halogenated aliphatics toxicity

2.1.3 Long chain fatty acids

2.1.3.1 Inhibition of long chain fatty acids to anaerobic digestion

2.1.3.2 Mechanism of long chain fatty acids inhibition

2.2 Inorganic Toxicants

2.2.1 Ammonia

2.2.1.1 Toxicity of ammonia to anaerobic digestion

2.2.1.2 Mechanism of ammonia toxicity

2.2.2 Sulfide

2.2.2.1 Toxicity/inhibition of sulfide to anaerobic digestion

2.2.2.2 Mechanisms of sulfide toxicity/inhibition

2.2.3 Heavy Metals

2.2.3.1 Toxicity of heavy metals to anaerobic digestion

2.2.3.2 Mechanisms of heavy metal toxicity

2.3 Nanomaterials

2.3.1 Toxicity/inhibition of nanoparticle/nanotubes to anaerobic

digestion

2.3.2 Mechanisms of toxicity/inhibition of nanoparticle/nanotubes

Page 4: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

3

3 Conclusions

Page 5: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

4

Abstract

Anaerobic digestion is increasingly being used to treat wastes from many sources because

of its manifold advantages over aerobic treatment, e.g. low sludge production and low

energy requirements. However, anaerobic digestion is sensitive to toxicants, and a wide

range of compounds can inhibit the process and cause upset or failure. Substantial

research has been carried out over the years to identify specific inhibitors/toxicants, and

their mechanism of toxicity in anaerobic digestion. In this review we present a detailed

and critical summary of research on the inhibition of anaerobic processes by specific

organic toxicants (e.g., chlorophenols, halogenated aliphatics and long chain fatty acids),

inorganic toxicants (e.g., ammonia, sulfide and heavy metals) and in particular,

nanomaterials, focusing on the mechanism of their inhibition/toxicity. A better

understanding of the fundamental mechanisms behind inhibition/toxicity will enhance the

wider application of anaerobic digestion.

Keywords: Anaerobic digestion; Inhibition; Toxicant; Chlorophenols, Halogenated

aliphatics, Long chain fatty acids, Nanomaterials, Ammonia, Sulfide, Heavy metals.

Page 6: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

5

List of Abbreviations

ADM1 Anaerobic Digestion Model No 1 AGS anaerobic granular sludge

CF chloroform CP monochlorophenol DCP dichlorophenol DCM dichloromethane DGGE denaturing gradient gel electrophoresis

DWCNT double walled carbon NTs

EGB expanded granular sludge bed EPS extracellular polymeric substances

FA free ammonia

FAN free ammonia nitrogen

IC50 half maximal inhibitory concentration

LCFA long chain fatty acid

MPB methane producing bacteria

NP nanoparticle

NT nanotube

NZVI nano zero valent iron

PCE perchlorethylene

PCP pentachlorophenol PEG polyethylene glycol chain

QSAR quantitative structure-activity relationship

ROS reactive oxygen species

SRB sulfate reducing bacteria

SWNT single-walled NT

TAN total ammonia nitrogen

TCE trichloroethylene TCP trichlorophenol

TeCP tetrachlorophenol TRFLP Terminal restriction fragment length

polymorphism UASB upflow anaerobic sludge bed

WWTP wastewater treatment plant

Page 7: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

6

1. Introduction

Anaerobic digestion can reduce organic pollution from the liquid outputs of homes,

industry and agriculture, while potentially offsetting the use of fossil fuels at the same

time. In addition, it offers numerous other significant advantages, such as lower energy

requirements, and less sludge production compared with traditional aerobic treatment

(Chen et al., 2008). Anaerobic digestion consists of a series of microbial processes that

convert organics to methane and carbon dioxide, and can take place under psychrophilic

(<20°C), mesophilic (25−40°C) or thermophilic (50−65°C) conditions, although

biodegradation under mesophilic conditions is most common. It also enables higher

loading rates than aerobic treatment, and a greater destruction of pathogens (Ravuri,

2013). Anaerobic digestion can be divided into three major microbial steps, i.e.

hydrolysis, acidogenesis/acetogenesis, and methanogenesis represented in Figure 1

(Amaya et al., 2013). During hydrolysis, a consortia of bacteria break down complex

organics (e.g. proteins, cellulose, lignin, and lipids) from the influent into soluble

monomers such as amino acids, simple sugars, glycerol, and fatty acids. Hydrolysis of

these complex polymers, some of which are insoluble, is catalyzed by extracellular

enzymes such as cellulases, proteases, and lipases (Batstone and Jensen, 2011).

Acidogenesis includes fermentation and anaerobic oxidation (β-oxidation), which are

carried out by fermentative acidogenic and acetogenic bacteria, respectively (Batstone

and Jensen, 2011). Fermentative acidogenic bacteria convert sugars, amino acids, and

fatty acids to organic acids (e.g. acetic, propionic, formic, lactic, butyric, or succinic

acids), alcohols and ketones (e.g. ethanol, methanol, glycerol, and acetone), acetate,

carbon dioxide, and hydrogen. Acetate is the major product of carbohydrate fermentation.

Page 8: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

7

Acetogenic bacteria convert fatty acids (e.g. long chain fatty acids) and alcohols into

acetate, hydrogen, and carbon dioxide, which are used by the methanogens. In the

methanogenesis step, acetate, hydrogen, and carbon dioxide are converted into methane

by methanogenic microorganisms, which are also classified as archaea composed of both

gram-positive and gram-negative bacteria with a wide variety of shapes, e.g., coccoid and

bacilli (Michael and Constantinos, 2006). Hydrolysis of insoluble polymers is generally

considered as rate-limiting among these successive steps, although with soluble feeds

methanogenesis is regarded as the key step in anaerobic digestion (Appels et al., 2008).

One of the main drawbacks to anaerobic digestion is its higher sensitivity to toxicants

than aerobic treatment. With the rapid development of nanotechnology, emerging

nanomaterials are starting to be used in some industrial products and will inevitably be

released into the environment; some nanomaterials have already been found in

wastewater treatment plants (WWTPs) and waste sludge (Yang et al., 2013). Hence, more

attention is being paid to their impact on the environment (Ju-Nam and Lead, 2008; Yang

et al., 2013), and in this review, we will summarize recent work on the effect of

nanoparticles and nanotubes on anaerobic digestion and the mechanisms by which they

may act. Besides, a wide range of organic chemicals can inhibit anaerobic digestion,

including halogenated benzenes (van Beelen and van Vlaardingen, 1994), halogenated

phenols (Armenante et al., 1999; Liu et al., 2008), phenol and alkyl phenols (Fedorak and

Hrudey, 1984; Levén et al., 2012), halogenated aliphatics (Adamson and Parkin, 1999;

Stuckey et al., 1980; van Hylckama Vlieg and Janssen, 2001), and long chain fatty acids

(LCFAs) (Hwu et al., 1996; Palatsi et al., 2012). In addition, many inorganic compounds,

such as ammonia (Ho and Ho, 2012; Liu and Sung, 2002), sulfide (Cai et al., 2008; Lopes

Page 9: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

8

and Lens, 2011) and heavy metals (Altaş, 2009; Oleszkiewicz and Sharma, 1990) are also

reported to be inhibitory, and the mechanisms behind inhibition are gradually becoming

understood. In this review we have decided to focus on the more common and typical

toxicants found in anaerobic membrane bioreactors (Casu et al., 2012; Feng et al., 2013;

Stuckey, 2012). This review will also present a detailed comparative summary of

research on the inhibition of anaerobic processes by nanomaterials, specific organic (i.e.,

chlorophenols, halogenated aliphatics and long chain fatty acids) and inorganic toxicants

(i.e., ammonia, sulfide and heavy metals), and critically analyze their mechanism of

inhibition.

2. Toxicants

2.1 Organic Toxicants

2.1.1 Chlorophenols

Chlorophenols are a group of chemicals produced by adding chlorine to phenol, and

include monochlorophenols (CPs), dichlorophenols (DCPs), trichlorophenols (TCP),

tetrachlorophenols (TeCPs), and pentachlorophenol (PCP). Chlorophenols are used

widely as pesticides, herbicides, antiseptics and fungicides as well as preservatives for

wood, glue, paint, vegetable fibers and leather. They are found to be highly persistent in

both aquatic and terrestrial environments, and are harmful to humans due to their

carcinogenicity (Muller and Caillard, 1986). Therefore, chlorophenols are listed as

priority pollutants by the U.S. Environmental Protection Agency (U.A. EPA).

2.1.1.1 Toxicity of chlorophenols to anaerobic digestion

Page 10: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

9

Although biodegradable, chlorophenols have been shown to be inhibitory to anaerobic

systems (Hernandez and Edyvean, 2008; Puyol et al., 2012; Wang et al., 1991). The

toxicity of chlorophenols depends on their degree of chlorination, the position of the

chlorine, and the purity of the sample; generally the level of inhibition increases with the

number of chlorine substitutions. For example, TCPs are all significantly more inhibitory

than DCPs, while PCP is considered to be the most toxic, e.g. 0.5-10 mg/L PCP can

cause inhibition to acidogens and methanogens (Chen et al., 2008). Adsorption tests

suggest that biological dechlorination is the main process for PCP removal in anaerobic

digestion, and due to the dechlorination high PCP removal efficiencies can be obtained

with the formation of DCP as the primary intermediate, followed by TCP and TeCP

(Damianovic et al., 2009). It was found that the partial dechlorination products of PCP,

such as the intermediate 3,4,5-TCP, is more toxic than PCP itself (Wu et al., 1989). For

chlorophenols with the same number of chlorine groups, the position of the substitution

influenced toxicity, and the sequence in decreasing order of toxicity was - meta para

ortho (Pera-Titus et al., 2004). By studying the biodegradation of chlorophenols in

anaerobic propionate-fed systems, Jin and Bhattacharya (1997) reported that the toxicity

of TCPs decreased in the following order: 2,4,5-TCP > 2,3,5-TCP > 2,4,6-TCP > 2,3,6-

TCP; while the six DCPs and three CPs showed slight toxicity to both propionate and

acetate degradation. In contrast, it has been reported that the chlorine position on CPs had

no significant effect on toxicity for either propionate or acetate degradation in sulfate

reducing anaerobic systems (Uberoi and Bhattacharya, 1997). Ennik-Maarsen et al.

(1998), however, found that during treatment of potato processing wastewater with a full

scale upflow anaerobic sludge blanket (UASB) reactor, the methanogenic activity of the

Page 11: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

10

sludge was inhibited more by 3-CP and 4-CP than by 2-CP; and this order can be related

to solvent/water partition coefficients (Kishino and Kobayashi, 1994) and these values

can be seen in Figure 2. In another study of chlorophenol’s toxicity to anaerobic sludge

from a sugar factory wastewater plant with high methanogenic activity, Vallecillo and

Vallecillo (1999) found that PCP showed the highest toxicity, and they concluded that

chlorophenol toxicity increased with the number of substituted chlorine atoms, except

2,4-DCP which was more toxic than 2,4,6-TCP. The toxicity of various chlorophenols

towards the syntrophic methanogenic reaction is summarized in Figure 2.

By now it is well known that substrate structure and concentration have an important

influence on methanogenic inhibition; however, the initial biomass concentration has not

always been taken into account. The initial biomass concentration is important since it

defines the initial substrate to microorganism ratio, S0/X0. This ratio determines whether,

in a given batch system, the cell will multiply or only grow without dividing (Chudoba et

al., 1991). For example, when studying the influence of S0/X0 on the inhibition of

methane production caused by 4-chlorophenol (4-CP), it was observed that inhibition

decreases as S0/X0 decreases (initial biomass concentration increases), and that for the

same value of S0/X0 there was an increase in inhibition when X0 decreases (Moreno

Andrade, 2003). In addition, based on the assumption that biomass growth is negligible

with respect to methane production, i.e. biomass concentration can be considered

constant during the biological tests, Puyol et al. (2012) proposed a kinetic model to

predict the inhibition of methanogenesis caused by chlorophenols and recommended it as

a useful tool to improve the recognized methanogenesis modeling by the IWA Anaerobic

Page 12: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

11

Digestion Model No 1 (ADM1) (Batstone et al., 2002) when treating inhibitory and/or

toxic compounds with anaerobic technologies.

2.1.1.2 Mechanism of chlorophenol toxicity

It is widely believed that only molecules of cyclic hydrocarbons dissolved in the aqueous

phase are available for intracellular metabolism (Wodzinsk.Rs and Bertolin.D, 1972).

Transfer of hydrophobic substrates, such as chlorophenols, proceeds via dissolving in the

aqueous phase and subsequently being taken up by the cell; however, direct contact

between the cell membrane’s hydrophobic components and hydrophobic compounds is

prevented by the cell wall and/or particularly, the cell membrane’s hydrophobic

components (Sikkema, 1995). Partitioning of these hydrophobic molecules into the lipid

bilayer of the cytoplasmic membrane is the most important mechanism in the uptake of

hydrophobic compounds, and may result in the accumulation of these compounds in the

lipid bilayer to enhance their availability to the cell and may cause toxicity problems as

well (Sikkema et al., 1994). Partition coefficients of hydrophobic compounds in

octanol/water, olive oil/water, diethyl ether/water, hexadecane/water, etc. have been used

to predict the effects of these compounds on intact cells (e.g., bioconcentration,

biodegradation and toxicity) (Osborne et al., 1990). Sierra-Alvarez and Lettinga (1991)

found that the logarithm of the partition coefficient in octanol/water (logP), an indicator

of hydrophobicity, was positively correlated with methanogenic inhibition, suggesting

that hydrophobicity is an important factor contributing to the toxicity of most inhibitory

aromatic compounds. Ennik-Maarsen et al. (1998) also obtained a high correlation

between logP for chloro-substituted benzenes and phenols and their methanogenic

toxicity. In addition, it has been proved that the protein-to-lipid ratios in the membrane

Page 13: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

12

and membrane functioning could be significantly influenced by phenols (Keweloh et al.,

1990). Finally, Heipieper et al. (1991) reported that addition of phenol and 4-

chlorophenol to suspensions of E. coli induced an efflux of potassium ions. Nevertheless,

the inhibitory action of chlorophenols seems to be directly related to these lipophilic

compounds’ partitioning behavior, which is expressed as logP. This suggests that the

(cytoplasmic) membrane is the primary site for the toxic action of chlorophenols by

disrupting the proton gradient through the membrane, and interfering with cellular energy

transduction, thereby decreasing cell growth due to an uncoupling of the catabolic and

anabolic reactions (Chen et al., 2008).

2.1.2 Halogenated aliphatics

Halogenated aliphatics (HAs) are organic chemicals in which one or more hydrogen

atoms has been replaced by a halogen. HAs are used in industry as solvents, chemical

intermediates, and fumigants and insecticides, and are found in the chemical, paint and

varnish, textile, rubber, plastics, dye-stuff, pharmaceutical and dry-cleaning industries

(Fishbein, 1986). Many of these synthetic compounds, especially the chlorinated

insecticides, due to their poor bioavailability, xenobiotic structure or high toxicity of the

compound itself or of intermediates, are recalcitrant and persist in the environment,

although some can be degraded under certain conditions (Cappelletti et al., 2012).

2.1.2.1 Toxicity of halogenated aliphatics to anaerobic digestion

Although microorganisms have been reported with the ability to remove halogens from

aliphatic compounds by the activity of enzymes known as dehalogenases (Slater, 1994),

most of the HAs are strong methanogenic inhibitors (Yu and Smith, 2000). Among HAs

Page 14: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

13

it has been suggested that brominated aliphatics, in general, are more inhibitory to

methanogens than their chlorinated analogs (Belay and Daniels, 1987; Freedman, 1991).

Among the most frequently encountered contaminating HAs in the environment are

polychlorinated aliphatic hydrocarbons, such as dichloromethane (DCM, CH2Cl2),

chloroform (CF, CHCl3), trichloroethylene (TCE, C2HCl3), and perchlorethylene (PCE,

C2Cl4), which are common priority pollutants found in wastewaters, soils, and aquifers

(Riley and Zachara, 1992).

Chloroform (CF) is the most widely used chloroaliphatic whose methanogenic toxicity

has been studied widely. By studying the inhibitory effect of DCM, CF, TCE, and PCE,

Yu and Smith (2000) found that CF was the most inhibitory, and at an aqueous

concentration at 0.09 mg/L CF inhibited methanogenesis completely; while DCM

inhibited methanogenesis at 3.9 mg/L, TCE at 18 mg/L, but PCE did not inhibit

methanogenesis at 14.5 mg/L. Furthermore, among six chlorinated hydrocarbon solvents

including 1,1,1-trichloroethane and carbon tetrachloride, CF was reported to be the most

toxic to the anaerobic digestion of sewage sludge (Swanwick and Foulkes, 1971).

Meanwhile, with the concentration inhibiting the mineralisation rate by 10% (IC10) at

0.04 mg/L, CF proved to be extremely toxic to anaerobic mineralisation of 4-

chlorophenol in microcosms with methanogenic sediment compared to benzene (IC10 at

150 mg/L), 1,2-dichloroethane (IC10 at 71 mg/L) and pentachlorophenol (6 mg/L) (van

Beelen and van Vlaardingen, 1994). The toxicity of CF to many obligate anaerobic

prokaryotes has been reported (Weathers and Parkin, 2000). One mg/L of CF can

completely inhibit dechlorination of PCE by a chlororespiring anaerobic isolate (Maymo-

Gatell et al., 2001), suggesting inhibition by CF for reductive dechlorination of

Page 15: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

14

chloroethenes is a general problem for sites co-contaminated with CF, which can only be

overcome by first removing the CF (Bagley et al., 2000). However, it has been suggested

that the high toxicity of CF to methanogenesis might be attributed to the formation of

very toxic and reactive intermediates during the slow anaerobic degradation of CF in

anaerobic sludge, although methanogenic toxicity data for these intermediates are lacking

(Trohalaki et al., 2003). Terminal restriction fragment length polymorphism (TRFLP)

analyses combined with a clone library showed that CF inhibited not only methanogenic

activity, but also the structure of methanogenic communities (Xu et al., 2010) because the

acetoclastic Methanosaetaceae were more sensitive to CF than hydrogenotrophic

Methanobacteriales and Methanomicrobiales. Xu et al. (2010) suggested that it is

probably some hydrogenotrophic methanogens, such as Methanobacteriaceae and

Methanomicrobiaceae, which can synthesize coenzyme M (CoM) which exhibit lower

rates of transport of external CoM into the cell, and are likely to be more resistant to CF

and other toxicants such as 2-bromoethanesulfonate.

2.1.2.2 Mechanism of halogenated aliphatics toxicity

Unlike chlorophenols, there is no identified relationship between the number of chloro-

substituents and the toxicity of chloroaliphatics. Wimley and White (1993) proposed that

polarity is an important factor, however, this is insufficient for predicting the toxicity of

different HAs. Furthermore, Gill and Ratledge (1972) demonstrated that, besides

halogenated substituents, the toxicity of HAs is related to their chain length, perfectly

correlating with their hydrophobicity and their solubility in water. Originally developed

for pharmaceutical applications (Leo, 1975), quantitative structure-activity relationships

(QSARs) have also been successfully used to predict environmental toxicity (Erturk et al.,

Page 16: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

15

2012; Jensen et al., 2008; Wu et al., 2013). Although QSARs have been helpful in

assessing the toxicity of specific groups of chemicals (e.g., phenols, chlorinated

aromatics), and also have been applied to toxicity prediction for HAs (Trohalaki and

Pachter, 2003; Trohalaki et al., 2003), they do not provide any information on the

mechanisms of these toxic effects, and this needs further investigation. By coordinating

the study on inhibition of methanogenesis by DCM, CF, TCE, and PCE, Yu and Smith

(2000) proposed a model to explain why corrinoids and porphinoids are not only the

dechlorination catalysts, but also the target moieties by which CF inhibits methanogens

(Figure 3). This model proposes that CF can be either a direct or indirect methanogenic

inhibitor due to its molecular structure, which is similar to some of the key C-1

intermediates of the methanogenesis pathway. In direct inhibition, due to its high redox

potential (10.56 V), CF would bind to the intracellular reduced corrinoid and porphinoid

enzymes (pool B, Figure 3) of the methanogenesis pathway, resulting in a higher affinity

for these reduced enzymes. This would prevent the normal substrate binding, such as a

methyl group, and channel electron flow away from methanogenesis towards

dechlorination (Figure 3), and hence CF’s binding to these enzymes leads to direct

inhibition of methanogenesis. Normal methanogenesis will resume when the

corrinoid/porphinoid enzymes are freed after CF is depleted. This model explains how

CF is dechlorinated reductively, and why the inhibition of methanogenesis is alleviated

once CF is dechlorinated. On the other hand, the binding of CF to the intracellular free

corrinoids/porphinoids lowers the free corrinoid/porphinoid concentrations (pool F, Fig.3)

in the cell, and could shift the equilibrium between the intracellular pools of free

corrinoids/porphinoids and protein-bound corrinoids/porphinoids (pool B, Fig.3) to the

Page 17: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

16

free-form side. This would result in the indirect inhibition of methanogenesis because

protein-bound corrinoids and porphinoids are essential for methanogenesis. In addition,

electron sinks of intermediates of CF and free corrinoids/porphinoids (pool F, Fig.3)

would drain the electron flow from normal methanogenesis to CF dechlorination, and this

in turn would indirectly inhibit methanogenesis.

2.1.3 Long chain fatty acids

Fatty acids are chains of carbons with attached hydrogen molecules at one end, and an

acid group at the other end. Long chain fatty acids (LCFAs) are fats that have several

carbons in their chain and include unsaturated and saturated fats. LCFAs are important

fractions of the organic matter in oil/fat wastewater (Stoll and Gupta, 1997). Anaerobic

digestion has been used for many decades to treat oily/fatty wastes, e.g., ice-cream wastes

(Hawkes et al., 1995), dairy wastewater (Perle et al., 1995), fish waste (Achour et al.,

2000), slaughterhouse wastewater (Cuetos et al., 2008) and vegetable waste (Li et al.,

2002) from the food processing industries. These lipids are neither easily treated by

conventional means, nor decomposed biologically, due to their formation of insoluble

aggregates and floating on the surface of the wastewater (Stoll and Gupta, 1997).

2.1.3.1 Inhibition of long chain fatty acids to anaerobic digestion

Although lipid rich wastewaters have high methane potential, one of its intermediate

products, LCFAs, can lead to inhibition (Palatsi et al., 2009). Hanaki et al. (1981) found

that the addition of LCFAs caused the appearance of a lag period in methane production

from acetate, and in the anaerobic digestion process. According to the Second Law of

Thermodynamics, energy must flow from a higher to a lower potential.

Page 18: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

17

Thermodynamically acido-acetogenesis (fermentation) is endothermic and non-

spontaneous (ΔH > 0 and ΔG > 0), and this implies that LCFA fermentation cannot

proceed spontaneously, and that LCFAs form insoluble hydrophobic aggregates in the

aqueous phase (Oh and Martin, 2010). However, acidogenesis together with acetogenesis

decomposes the long chain saturated fatty acids to acetic acid though shorter chain fatty

acids (β-oxidation) (Oh and Martin, 2010) and by electrochemical coupling with

methanogenesis, a strongly spontaneous process with a negative enthalpic driving force

(ΔG ≪ 0 and ΔH < 0) (Oh and Martin, 2007) occurs, and the non-spontaneous

fermentation is thermodynamically switched to a spontaneous process (ΔG < 0 and ΔH <

0) (Bartoschek et al., 2000). This implies that LCFA inhibition of methanogenesis could

cause the failure of LCFA fermentation and the whole anaerobic digestion bioprocess.

Severe LCFA inhibition was observed in an anaerobic membrane bioreactor for the

treatment of lipid rich corn-to-ethanol thin stillage, and Dereli et al. (2014) suggested that

the extensive dissolution of LCFAs in the reactor broth possibly caused inhibition of

methanogenesis. Inhibition of methanogens by LCFAs is reported as the limitation to

exploit biogas production from fat-rich wastewaters (Silva et al., 2014). Inhibition by

LCFAs of anaerobic processes depends on the type of LCFA, the microbial population,

and the temperature (Silvestre et al., 2014). For instance, oleic acid, followed by palmitic

and stearic acids, has been described as the LCFA with the greatest inhibitory effect on

thermophiles (Pereira et al., 2005).

2.1.3.2 Mechanism of long chain fatty acid inhibition

Although LCFA degradation was suggested as the “limiting step” in methanogenesis due

to its perceived limitations (Novak and Carlson, 1970), by combining hydrolytic,

Page 19: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

18

fermentative, syntrophic acetogenic (SAB) and methanogenic microorganisms, anaerobic

digestion of LCFAs has been achieved (Summers and Bousfield, 1980). The limiting step

was suggested to be closely related to the initial concentration of LCFAs, and therefore

high concentrations of LCFAs lead to the failure of anaerobic digesters (Chen and

Hashimoto, 1978; Tijero et al., 1989). It is believed that inhibition of anaerobic

metabolism by LCFAs is because of the adsorption of the LCFAs onto the cell wall and

membrane affecting metabolic transport (Alves et al., 2001; Masse et al., 2002; Neves et

al., 2006). This adsorption delays methane production, but can be prevented by providing

a competitive synthetic adsorbent (such as bentonite) (Figure 4) (Palatsi et al., 2009), and

the bio-physics of LCFA inhibition in anaerobic digestion has been mathematically

modeled by Zonta et al. (2013). In addition, Hwu et al. (1996) proved that the toxicity of

a model LCFA (oleate) to acetoclastic methanogens in anaerobic sludge was not

dependent on three biological factors (sludge origin, specific acetoclastic methanogenic

activity and sludge adaptation to lipids), but was closely correlated to the physical factor

of specific surface area of the sludge. Besides the inhibition of methanogenic bacteria,

Pereira et al. (2005) and Neves et al. (2009) proposed that LCFA adsorption and

accumulation on biomass can create a physical barrier and hinder the transfer of

substrates and products, inducing an initial delay in methane production, and even

causing sludge flotation and washout. Addition of calcium has been shown to reduce

LCFA inhibition, probably due to the formation of insoluble salts (Hanaki et al., 1981).

Dilution of the reactor’s content with inoculum, thus increasing the biomass/LCFA ratio,

or the addition of adsorbents, were found to be the best strategies to recover thermophilic

manure reactors subjected to LCFA inhibition (Palatsi et al., 2009).

Page 20: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

19

2.2 Inorganic Toxicants

2.2.1 Ammonia

2.2.1.1 Toxicity of ammonia to anaerobic digestion

Ammonia is an essential nutrient for the growth of microorganisms involved in anaerobic

digestion, as well as acting as an inhibitor at certain high concentrations (Koster and

Lettinga, 1984). The fermentation of nitrogen-containing materials such as urea and

proteins releases ammonia-nitrogen which exists largely as the ionized form (NH4+), but

this depends strongly on pH as its pKa is 9.3, and hence the toxic unionized form (NH3)

increases with increasing pH. The anaerobic digestion of livestock wastes releases high

levels of ammonia which raises the pH and forms higher levels of free ammonia (FA)

which is widely known to inhibit methanogenic microorganisms with low methane

production (Jin et al., 2012). Zhou and Qiu (2006) proved that concentrations of ammonia

nitrogen at 2.48 and 2.89 g/L can inhibit 50% of specific methanogenic activity in upflow

anaerobic sludge bed (UASBs) and expanded granular sludge bed (EGBs) reactors,

respectively. In terms of a 50% reduction in methane production, a wide range of

ammonia concentrations has been documented, with the inhibitory total ammonia

nitrogen concentration ranging from 1.7 to 14 NH3–N g/L (Chen et al., 2008). Some other

findings for the impact of ammonia on anaerobic digestion have been summarized in

Table 1. In a recent study on the impact of ammonia on thermophilic anaerobic digestion,

it was found that it was clearly affected by increasing concentrations of ammonia; the

threshold C/N ratio was found to be 4.40, corresponding to 620 mg free ammonia/L

(Siles et al., 2010). Calli et al. (2005) proved that acetogenic bacteria are more sensitive

Page 21: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

20

than methanogenic archaea to free ammonia, which has been suggested to be the active

component causing ammonia inhibition (Angelidaki and Ahring, 1993; He et al., 2011).

Meanwhile acetate-utilizing bacteria adapted to ammonia were shown to grow with a free

ammonia concentration of up to 800 mg-N/L, while many lower free ammonia

concentrations (100–150 mg-N/L) have been reported to initially inhibit an unadapted

process (Braun et al., 1981; De Baere et al., 1984).

The free ammonia concentration (CNH3) can be calculated according to the following

equation:

(1)

where TAN is the total ammonia nitrogen concentration, mg/L, Ka is the temperature

dependent dissociation constant (0.564 10-9

at 25 ºC, 1.097 10-9

at 35 ºC and 3.77

10-9

at 55 ºC), and CH is the concentration of hydrogen ions (Kayhanian, 1999).

Accordingly, the free ammonia concentration (CNH3) depends primarily on three

parameters; the total ammonia nitrogen concentration, temperature and pH (CH).

Ammonia concentrations below 200 mg/L are suggested to be beneficial to anaerobic

processes as this provides an essential nitrogen nutrient for anaerobic microorganisms

(Liu and Sung, 2002). Several studies found that the fermentation of high ammonia-

containing wastes is more easily inhibited at thermophilic temperatures than at

mesophilic temperatures (Angelidaki and Ahring, 1994; Bayr et al., 2012; Braun et al.,

1981), which is in agreement with the fact that the ratio of free ammonia to the total

ammonium will be much higher at higher temperatures as noted by Garcia and Angenent

Page 22: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

21

(2009). Therefore, wastewater rich in ammonium, urea and protein is more difficult to

treat under thermophilic conditions (i.e., 55–65 °C), even though the kinetics are more

favorable compared to mesophilic conditions (i.e., 25–37 °C) (Bocher et al., 2008; El-

Mashad et al., 2004). In aqueous solutions, there is a chemical balance between

ammonium ions (NH4+) and FA (NH3):

(2)

where obviously pH can affect the equilibrium of NH3/NH4+, and a high pH is conducive

to the formation of FA (Mosquera-Corral et al., 2005). Furthermore, Koster (1986)

reported that when the pH value increases, the biogas process becomes more sensitive

towards ammonia; from equation 2 an increase in pH results in more free ammonia. An

increase in pH from 7 to 8 will actually lead to an eight-fold increase in the free ammonia

concentration.

2.2.1.2 Mechanism of ammonia toxicity

Free ammonia is more toxic to methanogens than ionised ammonium (NH4+) because it

diffuses more rapidly through the cell membrane, causing proton imbalance, and/or

potassium (K+) deficiency, while ionised ammonium may just inhibit the methane

synthesizing enzyme directly (Gerardi, 2006; Kayhanian, 1999). When free ammonia

diffuses passively into methanogens, the difference in intracellular pH causes some of

them to convert to ammonium (NH4+), absorbing protons (H

+) in the process, while by

using a potassium antiporter, the cells then expend energy on proton balancing (Sprott et

al., 1984). However, the diffusion of free ammonia molecules into the cell also depends

on the physiology of the methanogens, e.g., on the basis of kilograms of NH3 entering per

Page 23: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

22

kilogram cell mass per hour, less free ammonia diffused into the cell of Methanosarcina,

which consist of larger spherical cells than into the smaller rod-shaped cell of

Methanothrix (Wiegant and Zeeman, 1986). Another reason why the ionized form of

ammonia is more beneficial than the free form is that the hydroxide, produced from

equation 2, can react with carbon dioxide, produced during the anaerobic digestion

process, to form bicarbonate (Kayhanian, 1999):

(3)

This will increase the buffering capacity of the anaerobic reactor, making the process less

susceptible to minor fluctuations in the relative production rates of the acetogenic and

methanogenic bacteria and, therefore, more stable.

A common approach to ammonia inhibition relies on dilution, and various types of

inhibition can be counteracted by increasing the biomass retention in the reactor.

Moreover, to mitigate the inhibition of thermophilic anaerobic treatment of digested

piggery wastewater by ammonia, Ho and Ho (2012) studied the effect of pH reduction,

zeolite, biomass and humic acid on methane production, and suggested that a reduction in

the pH from 8.3 to 6.5, and addition of 10–20 g/L zeolite, as more effective strategies

than addition of biomass and humic acid to mitigate ammonia inhibition.

2.2.2 Sulfide

Sulfide containing waste streams are generated by a number of industries such as

petrochemical plants, tanneries, viscose rayon factories, and coal gasification for

electricity production, or the anaerobic treatment of sulfate containing wastewater (Cai et

al., 2008).

Page 24: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

23

2.2.2.1 Toxicity/inhibition of sulfide to anaerobic digestion

Sulfidogens or sulfate reducing bacteria (SRB), which reduce sulfate to sulfide, play a

significant role in the anaerobic digestion of complex substrates. McCartney and

Oleszkiewicz (1991) found that SRB: i) generate sulfides which may be inhibitory and/or

toxic to SRB and methane producing bacteria (MPB); ii) reduce the rate of

methanogenesis; and iii) decrease the quantity of methane produced by competing for the

available carbon and/or hydrogen. In addition, Karhadkar et al. (1987) proposed two

stages of methanogenic inhibition due to sulfate reduction, viz. primary inhibition due to

competition for substrate from the SRBs, and secondary inhibition resulting from

methanogenic population decline due to inhibition of cellular function by soluble sulfides.

During the anaerobic treatment of wastewater containing sulfate, the competition

between SRB and MPB for acetate as their common primary substrate can affect

treatment efficiency. Although the primary competitive inhibition can be overcome in the

anaerobic treatment of sulfate-rich wastewaters, to maintain a low oxidation–reduction

potential in the reactors makes the presence of small amounts of sulfide advantageous. It

has been shown that sulfide ions can inhibit methane formation, suggesting non-

competitive inhibition of methanogenesis due to the sulfide resulting from SRB activity

may result in process failure (Paula Jr. and Foresti, 2009). It has been reported that high

levels of sulfide (IC50 ∼1300 mg/L H2S) did not affect SRB growing on acetate and

ethanol (Isa et al., 1986), and it has even been observed that sulfate removal rates

increased on ethanol and sugar with increasing total sulfide concentrations up to 1424

mg/L (Greben et al., 2005). However, sulfide was reported to be toxic to unacclimated

methanogens at concentrations of 50 only mg/L (Parkin et al., 1983), while

Page 25: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

24

Maillacheruvu and Parkin (1996) found that H2S inhibited acetotrophic methanogens

much more than hydrogenotrophic methanogens. Yamaguchi et al. (1999) reported that

the IC50 of H2S for acetotrophic and hydrogenotrophic methanogens was 160 and 220

mg/L, respectively.

2.2.3.2 Mechanisms of sulfide toxicity/inhibition

Theoretically, the reduction of sulfate to sulfide:

(∆G = −154 kJ) (4)

(∆G = − 43 kJ) (5)

yields more energy than methanogenesis:

(∆G = −135 kJ) (6)

(∆G = −28.5 kJ) (7)

which makes the latter noncompetitive and may reduce the rate of methanogenesis and

decrease the quantity of methane production (Karhadkar et al., 1987). In addition, sulfide

can denature proteins due to the formation of cross-links among polypeptide chains, and

can interfere with key metabolic enzymes in the cells (Madigan et al., 2003). Sulfide can

also interfere with the assimilation of sulfur and affect the intracellular pH (Visser, 1995).

Therefore, uncoupling growth from energy production and cell maintenance needs more

energy (Okabe et al., 1995). The toxicity of sulfide is often associated with its

undissociated form (H2S) due to the facilitated passage of neutral molecules across cell

membranes, and its high reactivity with cellular components (O'Flaherty et al., 1998).

Page 26: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

25

The chemical equilibrium of the sulfide species is pH dependent, with most of the total

sulfide (TS) in the HS− form at pH 8, while most is H2S at pH 6. The aqueous H2S

concentration can be calculated from the equilibrium equation (McCartney and

Oleszkiewicz, 1991):

(8)

Sulfide toxicity to methanogens is proportional to its concentration (on an added basis) in

the substrate and H2S concentration in the gas phase, which also incorporates the effect of

pH. Therefore, diluting the wastewater stream can prevent toxicity, although in general

this approach is considered undesirable due to the increase in the total volume of

wastewater that must be treated. Chen et al. (2008) suggested incorporating a sulfide

removal step in the overall process to reduce the sulfide concentration in an anaerobic

treatment system. In addition, adaptation of the MPB to free H2S, particularly in reactors

with biomass that cannot be washed out easily, eg biofilms and membrane systems, could

increase the tolerance of MPB to sulfide.

2.2.3 Heavy Metals

Heavy metals are often present in industrial wastewaters and municipal sludge in

significant concentrations, and the most frequently found are copper (Cu), zinc (Zn), lead

(Pb), mercury (Hg), chromium (Cr), cadmium (Cd), iron (Fe), nickel (Ni), cobalt (Co)

and molybdenum (Mo) (Altaş, 2009). However, many metals are required for the

activation or functioning of many enzymes and coenzymes in anaerobic digestion. Some

heavy metals, such as Ni, Co and Mo, are required at low concentrations, while the order

Page 27: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

26

of heavy metal composition in cells was found to be Fe >> Zn Ni > Co = Mo > Cu

when analyzing ten methanogenic strains (Takashima and Speece, 1989). However,

excessive amounts of heavy metals can lead to inhibition or toxicity (Li and Fang, 2007).

2.2.3.1 Toxicity of heavy metals to anaerobic digestion

The values of the half maximal inhibitory concentration (IC50) of different heavy metals

in other anaerobic digestion systems are summarized in Table 2. Unlike many other toxic

substances discussed above, heavy metals are non-biodegradable and can accumulate to

potentially toxic concentrations (Nayono, 2009; Sterritt and Lester, 1980). The potential

toxicity of heavy metals is controlled, to a large extent, by the physical and chemical

nature of the environment in which they are present, and this is correlated to different

ion-specific physicochemical parameters, e.g., standard reduction-oxidation potential,

electronegativity, the solubility product of the corresponding metal-sulfide complex, the

Pearson softness index, electron density and the covalent index (Workentine et al., 2008).

Heavy metals have been reported to be inhibitory to anaerobic microorganisms including

acetogens (Li and Fang, 2007), acidogens (Yu and Fang, 2001a; b; Zayed and Winter,

2000), methanogens (Karri et al., 2006; Mori et al., 2000) and SRB (Utgikar et al., 2003).

Evaluating the toxicity of heavy metals during the anaerobic digestion of sewage sludge

indicated severe inhibition at different concentration ranges for certain heavy metals,

such as from 70 to 400 mg/L for Cu, 200 to 600 mg/L for Zn and 10 to 2000 mg/L for Ni

(Ahring and Westermann, 1985). Moreover, lower concentrations of heavy metals

showing toxicity to anaerobic digestion were found in experiments under more defined

conditions. For example, the IC50 of acetate-degrading methanogens was reported to be

7.7, 12.5, 16 and 67.2 mg/L for Cd, Cu, Zn and Pb, respectively (Lin, 1992), and Cu, Zn

Page 28: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

27

and Ni caused 50% inhibition of mixed methanogens at 10, 40 and 60 mg/L, respectively

(Zayed and Winter, 2000). Acidogens are believed to be more resistant to heavy metal

toxicity than methanogens (Zayed and Winter, 2000). However, Hickey et al. (1989)

proposed that, compared to methanogenic populations, some trophic groups or organisms

within the anaerobic consortia might be more severely inhibited by a pulsed addition of

heavy metals. Studies have been carried out on the effect of heavy metals on a variety of

microbial species in anaerobic digestion. For example, with the addition of 5 mg Cd/L the

activity of Betaproteobacteria, which are involved in nutrient removal in activated sludge,

decreased significantly from 30.7% to 2.1% in an anaerobic–anoxic–oxic system

compared to the same system without the addition of cadmium (Tsai et al., 2005). In

another study on 2-CP degradation in anaerobic bioreactors, it was found the abundance

of the predominant archaeal species (e.g. Methanothrix soehngenii, Methanosaeta concilü

and uncultured euryarchaeota) decreased with Cd2+

and Cu2+

addition at high

concentration (shocked by 300 mg/L for 3 days), and the 2-CP anaerobic degradation

system was more sensitive to Cu2+

than Cd2+

(Huang, 2008). This observation is useful

with reference to the fundamentals, and the monitoring and control of anaerobic

membrane reactor responses to ramp/shock heavy metals loads.

2.2.3.2 Mechanisms of heavy metal toxicity

In an anaerobic environment, heavy metals may be: i) precipitated as sulfides (except Cr)

carbonates or hydroxides (Gonçalves et al., 2007; Gould and Genetelli, 1978); ii)

chelated and maintained in solution by compounds produced during digestion (Callander

and Barford, 1983; Walker et al., 2003), e.g. Soluble Microbial Products (SMPs); and iii)

adsorbed/bound to sludge ligands (Alibhai et al., 1985). At present it is believed that the

Page 29: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

28

failure of anaerobic digestion due to heavy metals occurs only when the concentration of

their free ions exceeds some threshold value (Leighton and Forster, 1997; Oleszkiewicz

and Sharma, 1990). Aluminum appears to be an exception as it becomes gradually

inhibitory only beyond 1500 mg/L as A12(SO4)3, and this high toxicity threshold could be

attributed to its low solubility under anaerobic conditions (Leighton and Forster, 1997). It

is believed that heavy metals show their toxicity due to their disruption of enzyme

function and structure by binding with thiol and other groups on protein molecules, or by

replacing natural metals in enzyme prosthetic groups (Soldatkin et al., 2012). In addition,

Oleszkiewicz and Sharma (1990) summarized various mechanisms of metal

toxicity/inhibition as: i) substitution of metallic enzyme cofactors (metal prosthetic

group); ii) combining with the outstanding sulfhydryl group (-SH) such as in cysteine; iii)

inactivation of the mercapto group in coenzyme M (2-mercaptoethanesulfonate) in

methanogens; and iv) tight binding to acid groups in the side chains of the amino acids in

the polypeptide chain.

Because heavy metals are usually present as mixtures in the influent of an anaerobic

digester, specific antagonistic and synergistic effects have been reported and are of great

interest in managing and ameliorating toxicity. Synergism means the enhanced toxic

effect of one metal in the presence of small amounts of another metal, while antagonism

refers to the effect of one metal which alleviates the toxic effect of another metal.

Oleszkiewicz and Sharma (1990) reviewed these phenomena in an early study on the

toxicity of Pb, Cu, Cd, Zn and Hg on anaerobic digestion, and showed that aluminum

silicate was antagonistic to the toxicity of other heavy metals. Another interesting heavy

Page 30: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

29

metal, nickel, has been proved to be synergistic in Ni-Cu, Ni-Mo-Co, and Ni-Hg systems,

and antagonistic in Ni-Hg, Ni-Zn, and Ni-Cd systems (Babich and Stotzky, 1983).

Heavy metal ions can be removed, or their concentrations reduced, by several

mechanisms such as precipitation, sorption and chelation by organic and inorganic

ligands, which are proposed as the most important methods for mitigating heavy metal

toxicity (Agrawal et al., 2011). For example, the addition of sulfide has been the main

method to precipitate heavy metals in anaerobic treatment (Kieu et al., 2011). Meanwhile,

Aquino and Stuckey (2007) proved that some insoluble sulfide salts were biologically

available under anaerobic condition. In addition, sorption of heavy metals onto activated

carbon, kaolin, bentonite, diatomite and waste materials such as compost and cellulose

pulp waste can also mitigate inhibition (Ulmanu et al., 2003). More recently, anaerobic

biological processes relying on the activity of SRB are being considered for the treatment

of heavy metal containing effluents (Kieu et al., 2011).

2.3 Nanomaterials

As the technological benefits of nanotechnology begin to move rapidly from laboratory to

large-scale industrial application (Abhilash, 2010; Schmid and Riediker, 2008), release of

nanomaterials to the environment is inevitable, and major environmental receptors will be

soil, sediment, and biosolids from wastewater treatment (Brar et al., 2010; Eduok et al.,

2013). Understanding the effect of nanoparticles on anaerobic microbial activity and

communities is important in order to enhance anaerobic process design.

2.3.1 Toxicity/inhibition of nanoparticle/nanotubes to anaerobic digestion

Page 31: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

30

The applications for nanoparticles (NPs, particles of any shape with dimensions in the

1×10−9

and 1×10−7

m range (Alemán et al., 2007)) are growing rapidly and outpacing

scientific investigations into the potential environmental impact as emerging contaminant.

The impact of metallic and metal oxide NPs (e.g. silver nanoparticles (Ag-NPs), nano-

ZnO, nano-TiO2, nano-Al2O3, nano-SiO2, nano-Au, nano-CeO2 and nano zero valent iron

(NZVI)) on sludge anaerobic digestion has been reviewed, suggesting that Ag-NPs, nano-

Al2O3, nano-SiO2 and nano-TiO2 are chemically stable NPs that have no adverse effects

on microbes under anaerobic conditions (Stasinakis, 2012; Yang et al., 2013), while

nano-Au presented no or low toxicity to anaerobic biomass and nano-CeO2 was the most

toxic to both mesophilic and thermophilic biomass (Stasinakis, 2012). When

investigating the behavior of Ag-NPs in a non-aerated tank, and in a pilot WWTP, Kaegi

et al. (2011) found that Ag-NPs transformed to Ag2S in less than 2 hours under anaerobic

conditions, and most of the Ag in both the sludge and effluent of the WWTP was present

in the form of Ag2S. In another batch anaerobic digestion system over less than two

weeks, Jin et al. (2012) observed that there was no significant difference in biogas or

methane production between the sludge treated with 40 mg Ag-NPs and the control. In

the same study, using quantitative PCR assays at moderate concentrations (≤ 40 mg/L),

Jin et al. (2012) found that Ag-NPs had no significant impact on methanogenic

assemblages and anaerobic production due to almost no release of Ag+ ions from Ag-NPs,

and this result was comparable to the findings by Kaegi et al. (2011). However, when

studied in activated sludge, it was proposed that the Ag-NPs apparently delivered Ag+ to

the bacteria more effectively, and result in pronounced microbial population shifts that

would hinder some wastewater treatment processes (Yang et al., 2014). ZnO-NPs are

Page 32: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

31

increasingly and commonly used in cosmetic products, sunscreen formulations, paints,

plastics, and packaging, and are indirectly released into the environment (Ju-Nam and

Lead, 2008). During the anaerobic digestion of cattle manure, the inhibition of methane

production by ZnO-NPs can be partially attributed to the soluble bioavailable fraction of

the metal found in the liquid phase of the reaction after a 14-day incubation period,

although the high toxicity of ZnO-NPs cannot only be explained by the release of toxic

Zn2+

ions (Luna-delRisco et al., 2011). Investigating the chemical transformation of two

ZnO-NPs and one hydrophobic ZnO-NP commercial formulation (used in personal care

products) in anaerobic digestion, Lombi et al. (2012) reported that both “native” Zn, and

Zn added either as a soluble salt or as NPs were rapidly converted to sulfides in all

treatments, which suggests that released Zn2+

ions cannot be the key mechanism for

inhibition of anaerobic digestion by ZnO-NPs. However, Mu and Chen (2011) proposed

that the toxic effect of ZnO-NPs on methane production was mainly due to the release of

Zn2+

from ZnO-NPs, which may inhibit the hydrolysis and methanation steps of digestion.

The negative influence of a shock load of ZnO-NPs on methane production in anaerobic

granular sludge (AGS) has also been reported (Mu et al., 2012). In the same study, a

decrease of proteins in the extracellular polymeric substances (EPS) released by AGS by

69.6% was observed when the dose of ZnO-NPs was greater than 100 mg/g-TSS,

suggesting that the decline of EPS induced by ZnO-NPs resulted in their deteriorating

protective role on the inner microorganisms of AGS, which corresponded with the lower

general physiological activity of AGS observed, and the death of microorganisms.

Besides nanoparticles, carbon nanotubes (NTs) are being investigated for medical

applications because of their theoretical capability to penetrate cell membranes, although

Page 33: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

32

their low solubility makes them challenging to work with in biological systems. Polar

functional groups, such as polyethylene glycol chains (PEG), are introduced into NTs to

increase their solubility and their bioavailability. Examining the effect of C60 fullerene

NTs on the structure and function of the microbial community in digester sludge, Nyberg

et al. (2008) proposed that C60 fullerenes NTs have no significant effect on the anaerobic

community over a few months exposure according to the community structure changes

monitored by denaturing gradient gel electrophoresis (DGGE), using primer sets

targeting the small subunit rRNA genes of bacteria, archaea, and eukarya. However, in an

anaerobic environment it was found that single-walled NTs (SWNT) functionalized with

polyethylene glycol chains (SWNT- PEG) showed a significant effect on microbial

community structure and function after exposure over a few months by monitoring

methanogenesis, functional gene primers for mcrA gene, and PEG diol dehydratase assay

(Nyberg et al., 2009). When studying the effect of double walled carbon NTs (DWCNT)

on a trichloroethylene (TCE)-dechlorinating culture, it was reported that the rate of

dechlorination first increased as the DWCNT loading was increased, but then decreased

as the DWCNT loading increased beyond 800 mg/L; meanwhile, it was observed that

methanogenesis decreased with increasing DWCNT loading, and no methane was

produced in cultures with 1600 mg DWCNT/L.

2.3.2 Mechanisms of toxicity/inhibition of nanoparticle/nanotubes

As nanomaterials often exhibit physical and chemical properties significantly different

from those of their molecular or macrosized analogs, concern has been growing regarding

their toxicity in the environment (both natural and manmade), and the detailed

mechanisms of how they inhibit anaerobic digestion is still being investigated. One

Page 34: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

33

important mechanism is the release of metal ions caused by corrosion and dissolution of

the NPs. By investigating algae, crustaceans, fish, bacteria, yeast, nematodes, protozoa

and mammalian cell lines, several studies proposed that the toxicity of CuO NPs is due to

the release of soluble Cu ions (Bondarenko et al., 2013). Apart from the release of toxic

Cu ions, CuO NPs have also been shown to cause toxicity by membrane disruption in

Escherichia coli as well (Zhao et al., 2013). Meanwhile, the toxicity of CuO NPs to

acetoclastic methanogenic activity is most likely caused by both the CuO NPs themselves

and copper ions released to the culture medium (Otero-González et al., 2014). However,

Gonzalez-Estrella et al. (2013) suggested that release of toxic metal species by NP-

dissolution was the principal mechanism of methanogenic inhibition caused by Cu0, CuO,

and ZnO NPs with IC50 values of 62–250 mg/L. Another main mechanism for NPs

toxicity is the generation of reactive oxygen species (ROS) which primarily damage cell

membranes; however, it is unlikely that ROS will be the cause of membrane damage in

anaerobic environments. It is suggested that Al2O3 NPs induced changes in the bacterial

membranes of the anaerobe Ruminococcus flavefaciens 007C by direct physical

interaction, however, under the same conditions nano-TiO2 did not show a significant

effect on the membrane profile of the same bacterium (Vodovnik et al., 2012). On the

other hand, although toxicity of many types of nanotubes to anaerobic microorganisms

have been reported, e.g. DWCNT’s effect on methanogenesis (Kannepalli et al., 2008),

the mechanism of their toxicity is still unknown. A study of the effective visible-light-

driven bismuth vanadate NT for inactivation of bacteria, especially under anaerobic

conditions, indicated that the destruction process of bacterial cells began from the cell

wall to other cellular components, possibly by the hydroxyl radical adsorbing onto the

Page 35: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

34

surface rather than by the free radical in bulk solution, although both were derived from

photogenerated h+ (Wang et al., 2012). Systematic exposition of NT and NP toxicity to

anaerobic digestion so far is not available. Physical and chemical transformations of

nanomaterials control their toxicity and bioavailability, and therefore must be considered

in future risk assessments.

3 Conclusions

As an efficient waste treatment technology that harnesses natural anaerobic

decomposition to treat waste, reduce waste volume and generate biogas as well,

anaerobic digestion has been widely used as a source of renewable energy. However,

anaerobic digestion can be inhibited to varying degrees by toxic materials present in the

system; these substances may be components of the influent waste stream, or byproducts

of the metabolic activities of the digester bacteria. Inhibitory toxic compounds include

organics, ammonia, sulfide, heavy metals, and the emerging nanomaterials, and are often

present in the processing of wastes from agricultural and industrial operations such as

molasses fermentation, petroleum refining and the tanning industries. These toxic

compounds principally obstruct the activities of the sensitive obligate hydrogen

producing acetogens and methanogenic portions of the digester population, as well as

cause retarded methane formation, a decrease in the methane content of biogas, or can

even cause complete failure of methanogenesis. However, because of the difference in

anaerobic microorganisms and waste composition, results from previous studies on

inhibition of anaerobic processes vary substantially. In this review, we have summarized

and highlighted the effect of specific organic toxicants (e.g., chlorophenols, halogenated

Page 36: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

35

aliphatics and long chain fatty acids), inorganic toxicants (e.g., ammonia, sulfide and

heavy metals) and the emerging nanomaterials on anaerobic digestion. In addition, better

understanding the mechanism(s) of inhibition or toxicity of different toxicants in an

anaerobic digester provides insights into overcoming these toxic effects and possible

solutions or strategies to properly cope with it, successfully apply anaerobic digestion and

significantly improve waste treatment efficiency. On the other hand, measuring the

toxicant concentration and monitoring them are an essential precautionary strategy. At

present for anaerobic waste treatment, no work has been carried out on the pre-

measurement of toxicity before the waste is introduced into the digester. Most research

has focused on detoxification after inhibition and not on stopping/ameliorating toxicity

before it happens. Hence, a rapid response method to determine toxic substances in the

feed stream, and toxic byproducts in the digester, needs to be developed to protect the

anaerobic microcosm from instability, and hence enable digesters to operate without toxic

perturbations. In addition, how to control toxicity once it has occurred is another

important question to be considered; we will review the state of the art on toxicity

monitoring and control for anaerobic digestion in our next article.

Page 37: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

36

REFERENCES

Abhilash M. Potential applications of nanoparticles. Int J Pharm Bio Sci. 2010;1:1-12.

Achour M, Khelifi O, Bouazizi I, Hamdi M. Design of an integrated bioprocess for the treatment

of tuna processing liquid effluents. Process Biochem. 2000;35:1013-7.

Adamson DT, Parkin GF. Biotransformation of mixtures of chlorinated aliphatic hydrocarbons by

an acetate-grown methanogenic enrichment culture. Water Res. 1999;33:1482-94.

Agrawal J, Sherameti I, Varma A. Detoxification of Heavy Metals: State of Art. In: Sherameti I,

Varma A, editors. Detoxification of Heavy Metals: Springer Berlin Heidelberg; 2011. p. 1-34.

Ahring BK, Westermann P. Sensitivity of thermophilic methanogenic bacteria to heavy metals.

Curr Microbiol. 1985;12:273-6.

Alemán J, Chadwick AV, He J, Hess M, Horie K, Jones RG, et al. Definitions of terms relating to

the structure and processing of sols, gels, networks, and inorganic-organic hybrid materials

(IUPAC recommendations 2007). Pure Appl Chem. 2007;79:1801-29.

Alibhai KRK, Mehrotra I, Forster CF. Heavy metal binding to digested sludge. Water Res.

1985;19:1483-8.

Altaş L. Inhibitory effect of heavy metals on methane-producing anaerobic granular sludge. J

Hazard Mater. 2009;162:1551-6.

Alves MM, Mota Vieira JA, Alvares Pereira RM, Pereira MA, Mota M. Effects of lipids and

oleic acid on biomass development in anaerobic fixed-bed reactors. Part II: Oleic acid toxicity

and biodegradability. Water Res. 2001;35:264-70.

Amaya OM, Barragán MTC, Tapia FJA. Microbial Biomass in Batch and Continuous System,

Biomass Now - Sustainable Growth and Use, Dr. Miodrag Darko Matovic (Ed.)2013.

Angelidaki I, Ahring BK. Thermophilic anaerobic digestion of livestock waste: the effect of

ammonia. Appl Microbiol Biot. 1993;38:560-4.

Angelidaki I, Ahring BK. Anaerobic thermophilic digestion of manure at different ammonia loads:

effect of temperature. Water Res. 1994;28:727-31.

Appels L, Baeyens J, Degrève J, Dewil R. Principles and potential of the anaerobic digestion of

waste-activated sludge. Prog Energ Combust. 2008;34:755-81.

Aquino SF, Stuckey DC. Bioavailability and toxicity of metal nutrients during anaerobic

digestion. Journal of Environmental Engineering. 2007;133:28-35.

Armenante PM, Kafkewitz D, Lewandowski GA, Jou C-J. Anaerobic–aerobic treatment of

halogenated phenolic compounds. Water Res. 1999;33:681-92.

Babich H, Stotzky G. Toxicity of nickel to microbes: Environmental aspects. In: Allen IL, editor.

Adv Appl Microbiol: Academic Press; 1983. p. 195-265.

Bagley DM, Lalonde M, Kaseros V, Stasiuk KE, Sleep BE. Acclimation of anaerobic systems to

biodegrade tetrachloroethene in the presence of carbon tetrachloride and chloroform. Water

Res. 2000;34:171-8.

Bartoschek S, Vorholt JA, Thauer RK, Geierstanger BH, Griesinger C. N-Carboxymethanofuran

(carbamate) formation from methanofuran and CO2 in methanogenic archaea. Eur J Biochem.

2000;267:3130-8.

Batstone DJ, Jensen PD. 4.17 - Anaerobic Processes. In: Wilderer P, editor. Treatise on Water

Science. Oxford: Elsevier; 2011. p. 615-39.

Batstone DJ, Keller J, Angelidaki I, Kalyuzhnyi SV, Pavlostathis SG, Rozzi A, et al. The IWA

anaerobic digestion model No. 1 (ADM1). Water Sci Technol. 2002;45:65-73.

Bayr S, Rantanen M, Kaparaju P, Rintala J. Mesophilic and thermophilic anaerobic co-digestion

of rendering plant and slaughterhouse wastes. Bioresource Technol. 2012;104:28-36.

Belay N, Daniels L. Production of ethane, ethylene, and acetylene from halogenated

hydrocarbons by methanogenic bacteria. Appl Environ Microb. 1987;53:1604-10.

Page 38: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

37

Bocher BT, Agler MT, Garcia ML, Beers AR, Angenent LT. Anaerobic digestion of secondary

residuals from an anaerobic bioreactor at a brewery to enhance bioenergy generation. J Ind

Microbiol Biot. 2008;35:321-9.

Bondarenko O, Juganson K, Ivask A, Kasemets K, Mortimer M, Kahru A. Toxicity of Ag, CuO

and ZnO nanoparticles to selected environmentally relevant test organisms and mammalian

cells in vitro: a critical review. Arch Toxicol. 2013;87:1181-200.

Borja R, Sánchez E, Weiland P. Influence of ammonia concentration on thermophilic anaerobic

digestion of cattle manure in upflow anaerobic sludge blanket (UASB) reactors. Process

Biochem. 1996;31:477-83.

Brar SK, Verma M, Tyagi RD, Surampalli RY. Engineered nanoparticles in wastewater and

wastewater sludge - Evidence and impacts. Waste Manage. 2010;30:504-20.

Braun R, Huber P, Meyrath J. Ammonia toxicity in liquid piggery manure digestion. Biotechnol

Lett. 1981;3:159-64.

Cai J, Zheng P, Mahmood Q. Effect of sulfide to nitrate ratios on the simultaneous anaerobic

sulfide and nitrate removal. Bioresource Technol. 2008;99:5520-7.

Callander I, Barford J. Precipitation, chelation, and the availability of metals as nutrients in

anaerobic digestion. I. Methodology. Biotechnol Bioeng. 1983;25:1947-57.

Calli B, Mertoglu B, Inanc B, Yenigun O. Effects of high free ammonia concentrations on the

performances of anaerobic bioreactors. Process Biochem. 2005;40:1285-92.

Cappelletti M, Frascari D, Zannoni D, Fedi S. Microbial degradation of chloroform. Appl

Microbiol Biot. 2012;96:1395-409.

Casu S, Crispino NA, Farina R, Mattioli D, Ferraris M, Spagni A. Wastewater treatment in a

submerged anaerobic membrane bioreactor. Journal of Environmental Science and Health Part

a-Toxic/Hazardous Substances & Environmental Engineering. 2012;47:204-9.

Chen Y, Cheng JJ, Creamer KS. Inhibition of anaerobic digestion process: a review. Bioresource

Technol. 2008;99:4044-64.

Chen YR, Hashimoto AG. Kinetics of Methane Fermentation. 1978. p. 269-82.

Chudoba P, Chevalier JJ, Chang J, Capdeville B. Effect of anaerobic stabilization of activated

sludge on its production under batch conditions at various S(o)/X(o) ratios. Water Sci Technol.

1991;23:917-26.

Cuetos MJ, Gómez X, Otero M, Morán A. Anaerobic digestion of solid slaughterhouse waste

(SHW) at laboratory scale: Influence of co-digestion with the organic fraction of municipal

solid waste (OFMSW). Biochem Eng J. 2008;40:99-106.

Damianovic MHRZ, Moraes EM, Zaiat M, Foresti E. Pentachlorophenol (PCP) dechlorination in

horizontal-flow anaerobic immobilized biomass (HAIB) reactors. Bioresource Technol.

2009;100:4361-7.

De Baere LA, Devocht M, Van Assche P, Verstraete W. Influence of high NaCl and NH4Cl salt

levels on methanogenic associations. Water Res. 1984;18:543-8.

Dereli RK, van der Zee FP, Heffernan B, Grelot A, van Lier JB. Effect of sludge retention time

on the biological performance of anaerobic membrane bioreactors treating corn-to-ethanol thin

stillage with high lipid content. Water Res. 2014;49:453-64.

Eduok S, Martin B, Villa R, Nocker A, Jefferson B, Coulon F. Evaluation of engineered

nanoparticle toxic effect on wastewater microorganisms: Current status and challenges.

Ecotox Environ Safe. 2013;95:1-9.

El-Mashad HM, Zeeman G, Van Loon WKP, Bot GPA, Lettinga G. Effect of temperature and

temperature fluctuation on thermophilic anaerobic digestion of cattle manure. Bioresource

Technol. 2004;95:191-201.

Ennik-Maarsen KA, Louwerse A, Roelofsen W, Stams AJM. Influence of monochlorophenols on

methanogenic activity in granular sludge. Water Res. 1998;32:2977-82.

Page 39: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

38

Erturk MD, Sacan MT, Novic M, Minovski N. Quantitative structure-activity relationships

(QSARs) using the novel marine algal toxicity data of phenols. J Mol Graph Model.

2012;38:90-100.

Fedorak PM, Hrudey SE. The effects of phenol and some alkyl phenolics on batch anaerobic

methanogenesis. Water Res. 1984;18:361-7.

Feng B, Fang Z, Hou J, Ma X, Huang Y, Huang L. Effects of heavy metal wastewater on the

anoxic/aerobic-membrane bioreactor bioprocess and membrane fouling. Bioresource

Technology. 2013;142:32-8.

Fishbein L. Halogenated aliphatic hydrocarbons: uses and environmental occurrence. In: Fishbein

L, O'Neill IK, editors. Environmental Carcinogens - Selected methods of analysis. Vol. 7.

Some volatile halogenated hydrocarbons. Lyon: International Agency for Research on Cancer;

1986. p. 47-67.

Freedman DL. Biodegradation of Polychlorinated Methanes in Methanogenic Systems: Annual

Report. 1 Dec 88-1 Mar 89: University of Illinois at Urbana-Champaign. Department of Civil

Engineering.; 1991.

Garcia ML, Angenent LT. Interaction between temperature and ammonia in mesophilic digesters

for animal waste treatment. Water Res. 2009;43:2373-82.

Gerardi MH. Methane-Forming Bacteria. Wastewater Bacteria: John Wiley & Sons, Inc.; 2006. p.

161-3.

Gill CO, Ratledge C. Toxicity of n-alkanes, n-alk-1-enes, n-alkan-1-ols and n-alkyl-1-bromides

towards yeasts. J Gen Microbiol. 1972;72:165-72.

Gonçalves MMM, da Costa ACA, Leite SGF, Sant’Anna Jr GL. Heavy metal removal from

synthetic wastewaters in an anaerobic bioreactor using stillage from ethanol distilleries as a

carbon source. Chemosphere. 2007;69:1815-20.

Gonzalez-Estrella J, Sierra-Alvarez R, Field JA. Toxicity assessment of inorganic nanoparticles

to acetoclastic and hydrogenotrophic methanogenic activity in anaerobic granular sludge. J

Hazard Mater. 2013;260:278-85.

Gould MS, Genetelli EI. Heavy metal complexation behavior in anaerobically digested sludges.

Water Res. 1978;12:505-12.

Greben HA, Maree JP, Eloff E, Murray K. Improved sulphate removal rates at increased sulphide

concentration in the sulphidogenic bioreactor. Water Sa. 2005;31:351-8.

Hanaki K, Matsuo T, Nagase M. Mechanism of inhibition caused by long-chain fatty acids in

anaerobic digestion process. Biotechnol Bioeng. 1981;23:1591-610.

Hawkes FR, Donnelly T, Anderson GK. Comparative performance of anaerobic digesters

operating on ice-cream wastewater. Water Res. 1995;29:525-33.

He P, Guan D, Wu D, Lü F, Shao L. Inhibitory effect of ammonia and lincomycin on anaerobic

digestion. Huagong Xuebao/CIESC Journal. 2011;62:1389-94.

Heipieper HJ, Keweloh H, Rehm HJ. Influence of phenols on growth and membrane-permeability

of free and immobilized Escherichia coli. Appl Environ Microb. 1991;57:1213-7.

Hernandez JE, Edyvean RGJ. Inhibition of biogas production and biodegradability by substituted

phenolic compounds in anaerobic sludge. J Hazard Mater. 2008;160:20-8.

Hickey RF, Vanderwielen J, Switzenbaum MS. The effect of heavy metals on methane

production and hydrogen and carbon monoxide levels during batch anaerobic sludge digestion.

Water Res. 1989;23:207-18.

Ho L, Ho G. Mitigating ammonia inhibition of thermophilic anaerobic treatment of digested

piggery wastewater: use of pH reduction, zeolite, biomass and humic acid. Water Res.

2012;46:4339-50.

Huang A. Effects of Cd(II) and Cu(II) on microbial characteristics in 2-chlorophenol-degradation

anaerobic bioreactors. J Environ Sci. 2008;20:745-52.

Hwu C-S, Donlon B, Lettinga G. Comparative toxicity of long-chain fatty acid to anaerobic

sludges from various origins. Water Sci Technol. 1996;34:351-8.

Page 40: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

39

Isa Z, Grusenmeyer S, Verstraete W. Sulfate reduction relative to methane production in high-rate

anaerobic digestion: technical aspects. Appl Environ Microb. 1986;51:572-9.

Jensen GE, Niemela JR, Wedebye EB, Nikolov NG. QSAR models for reproductive toxicity and

endocrine disruption in regulatory use - a preliminary investigation. Sar Qsar Environ Res.

2008;19:631-41.

Jin P, Bhattacharya SK. Toxicity and biodegradation of chlorophenols in anaerobic propionate

enrichment culture. Water Environ Res. 1997;69:938-47.

Jin R-C, Yang G-F, Yu J-J, Zheng P. The inhibition of the anammox process: a review. Chem

Eng J. 2012;197:67-79.

Ju-Nam Y, Lead JR. Manufactured nanoparticles: An overview of their chemistry, interactions

and potential environmental implications. Sci Total Environ. 2008;400:396-414.

Kaegi R, Voegelin A, Sinnet B, Zuleeg S, Hagendorfer H, Burkhardt M, et al. Behavior of

metallic silver nanoparticles in a pilot wastewater treatment plant. Environ Sci Technol.

2011;45:3902-8.

Kannepalli S, Fennell DE, Huang W. Effect of double-walled carbon nanotubes on a TCE-

dechlorinating culture. Prepr. Ext. Abstr. - ACS Natl. Meet., Am. Chem. Soc., Div. Environ.

Chem. 2008;48:369-72.

Karhadkar PP, Audic J-M, Faup GM, Khanna P. Sulfide and sulfate inhibition of methanogenesis.

Water Res. 1987;21:1061-6.

Karri S, Sierra-Alvarez R, Field JA. Toxicity of copper to acetoclastic and hydrogenotrophic

activities of methanogens and sulfate reducers in anaerobic sludge. Chemosphere.

2006;62:121-7.

Kayhanian M. Ammonia inhibition in high-solids biogasification: an overview and practical

solutions. Environ Technol. 1999;20:355-65.

Keweloh H, Weyrauch G, Rehm HJ. Phenol-induced membrane changes in free and immobilized

Escherichia coli. Appl Microbiol Biot. 1990;33:66-71.

Kieu HTQ, Müller E, Horn H. Heavy metal removal in anaerobic semi-continuous stirred tank

reactors by a consortium of sulfate-reducing bacteria. Water Res. 2011;45:3863-70.

Kishino T, Kobayashi K. Relation between the chemical structures of chlorophenols and their

dissociation constants and partition coefficients in several solvent-water systems. Water Res.

1994;28:1547-52.

Koster IW. Characteristics of the pH-influenced adaptation of methanogenic sludge to ammonium

toxicity. J Chem Technol Biot. 1986;36:445-55.

Koster IW, Lettinga G. The influence of ammonium-nitrogen on the specific activity on

pelletized methanogenic sludge. Agr Wastes. 1984;9:205-16.

Leighton IR, Forster CF. The effect of heavy metal ions on the performance of a two-phase

thermophilic anaerobic digester. Process Saf Environ. 1997;75:27-32.

Leo A. Quantitative structure-activity relationships in drug-design. In: Drews J, Hahn E, editors.

Drug Receptor Interactions in Antimicrobial Chemotherapy: Springer Vienna; 1975. p. 45-56.

Levén L, Nyberg K, Schnürer A. Conversion of phenols during anaerobic digestion of organic

solid waste – a review of important microorganisms and impact of temperature. J Environ

Manage. 2012;95, Supplement:S99-S103.

Li C, Fang HHP. Inhibition of heavy metals on fermentative hydrogen production by granular

sludge. Chemosphere. 2007;67:668-73.

Li YY, Sasaki H, Yamashita K, Saki K, Kamigochi I. High-rate methane fermentation of lipid-

rich food wastes by a high-solids co-digestion process. 2002. p. 143-50.

Lin CY. Effect of heavy metals on volatile fatty acid degradation in anaerobic digestion. Water

Res. 1992;26:177-83.

Lin CY, Chen CC. Effect of heavy metals on the methanogenic UASB granule. Water Res.

1999;33:409-16.

Page 41: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

40

Liu T, Sung S. Ammonia inhibition on thermophilic aceticlastic methanogens. Water Sci Technol.

2002;45:113-20.

Liu X-W, He R, Shen D-S. Studies on the toxic effects of pentachlorophenol on the biological

activity of anaerobic granular sludge. J Environ Manage. 2008;88:939-46.

Lombi E, Donner E, Tavakkoli E, Turney TW, Naidu R, Miller BW, et al. Fate of zinc oxide

nanoparticles during anaerobic digestion of wastewater and post-treatment processing of

sewage sludge. Environ Sci Technol. 2012;46:9089-96.

Lopes SIC, Lens PNL. 6.32 - Anaerobic Treatment of Organic Sulfate-Rich Wastewaters. In:

Editor-in-Chief: Murray M-Y, editor. Comprehensive Biotechnology (Second Edition).

Burlington: Academic Press; 2011. p. 399-418.

Luna-delRisco M, Orupõld K, Dubourguier H-C. Particle-size effect of CuO and ZnO on biogas

and methane production during anaerobic digestion. J Hazard Mater. 2011;189:603-8.

Madigan MT, Martinko JM, Parker J. Brock Biology of Microorganisms: Prentice Hall/Pearson

Education; 2003.

Maillacheruvu KY, Parkin GF. Kinetics of growth, substrate utilization and sulfide toxicity for

propionate, acetate, and hydrogen utilizers in anaerobic systems. Water Environ Res.

1996;68:1099-106.

Masse L, Massé DI, Kennedy KJ, Chou SP. Neutral fat hydrolysis and long-chain fatty acid

oxidation during anaerobic digestion of slaughterhouse wastewater. Biotechnol Bioeng.

2002;79:43-52.

Maymo-Gatell X, Nijenhuis I, Zinder SH. Reductive dechlorination of cis-1,2-dichloroethene and

vinyl chloride by "dehalococcoides ethenogenes". Environ Sci Technol. 2001;35:516-21.

McCartney DM, Oleszkiewicz JA. Sulfide inhibition of anaerobic degradation of lactate and

acetate. Water Res. 1991;25:203-9.

Michael N, Constantinos PH. Chapter 8 Biological processes. Waste Management Series:

Elsevier; 2006. p. 171-218.

Moreno Andrade I. Influence of the initial substrate to microorganisms concentration ratio on the

methanogenic inhibition test. Water Sci Technol. 2003;48:17-22.

Mori K, Hatsu M, Kimura R, Takamizawa K. Effect of heavy metals on the growth of a

methanogen in pure culture and coculture with a sulfate-reducing bacterium. J Biosci Bioeng.

2000;90:260-5.

Mosquera-Corral A, González F, Campos JL, Méndez R. Partial nitrification in a SHARON

reactor in the presence of salts and organic carbon compounds. Process Biochem.

2005;40:3109-18.

Mu H, Chen Y. Long-term effect of ZnO nanoparticles on waste activated sludge anaerobic

digestion. Water Res. 2011;45:5612-20.

Mu H, Zheng X, Chen Y, Chen H, Liu K. Response of anaerobic granular sludge to a shock load

of zinc oxide nanoparticles during biological wastewater treatment. Environ Sci Technol.

2012;46:5997-6003.

Muller F, Caillard L. Chlorophenols. In: Gerhartz W (ed) Ullmann's encyclopedia of industrial

chemistry, vol A7. VCH Weinheim, pp 1-8. 1986.

Nayono SE. Anaerobic digestion of organic solid waste for energy production: KIT Scientific

Publishing; 2009.

Neves L, Oliveira R, Alves MM. Anaerobic co-digestion of coffee waste and sewage sludge.

Waste Manage. 2006;26:176-81.

Neves L, Oliveira R, Alves MM. Co-digestion of cow manure, food waste and intermittent input

of fat. Bioresource Technol. 2009;100:1957-62.

Novak JT, Carlson DA. The kinetics of anaerobic long chain fatty acid degradation. J Water

Pollut Control Fed. 1970;42:1932-43.

Nyberg L, Turco RF, Nies L. Assessing the impact of nanomaterials on anaerobic microbial

communities. Environ Sci Technol. 2008;42:1938-43.

Page 42: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

41

Nyberg LM, Nies LF, Turco RF. Assessment of the impact of pegylated single-walled nanotubes

in an anaerobic environment. 237th ACS National Meeting, IEC-115. Salt Lake City, UT,

United States: American Chemical Society; 2009.

O'Flaherty V, Mahony T, O'Kennedy R, Colleran E. Effect of pH on growth kinetics and sulphide

toxicity thresholds of a range of methanogenic, syntrophic and sulphate-reducing bacteria.

Process Biochem. 1998;33:555-69.

Oh ST, Martin AD. Thermodynamic equilibrium model in anaerobic digestion process. Biochem

Eng J. 2007;34:256-66.

Oh ST, Martin AD. Long chain fatty acids degradation in anaerobic digester: thermodynamic

equilibrium consideration. Process Biochem. 2010;45:335-45.

Okabe S, Nielsen PH, Jones WL, Characklis WG. Sulfide product inhibition of Desulfovibrio

desulfuricans in batch and continuous cultures. Water Res. 1995;29:571-8.

Oleszkiewicz JA, Sharma VK. Stimulation and inhibition of anaerobic processes by heavy metals

- a review. Biol Wastes. 1990;31:45-67.

Osborne SJ, Leaver J, Turner MK, Dunnill P. Correlation of biocatalytic activity in an organic-

aqueous two-liquid phase system with solvent concentration in the cell membrane. Enzyme

Microb Tech. 1990;12:281-91.

Otero-González L, Field JA, Sierra-Alvarez R. Inhibition of anaerobic wastewater treatment after

long-term exposure to low levels of CuO nanoparticles. Water Res. 2014;58:160-8.

Palatsi J, Affes R, Fernandez B, Pereira MA, Alves MM, Flotats X. Influence of adsorption and

anaerobic granular sludge characteristics on long chain fatty acids inhibition process. Water

Res. 2012;46:5268-78.

Palatsi J, Laureni M, Andrés MV, Flotats X, Nielsen HB, Angelidaki I. Strategies for recovering

inhibition caused by long chain fatty acids on anaerobic thermophilic biogas reactors.

Bioresource Technol. 2009;100:4588-96.

Parkin GF, Speece RE, Yang CHJ, Kocher WM. Response of methane fermentation systems to

industrial toxicants. J Water Pollut Control Fed. 1983:44-53.

Paula Jr. DR, Foresti E. Sulfide toxicity kinetics of a uasb reactor. Braz J Chem Eng.

2009;26:669-75.

Pera-Titus M, Garc a-Molina V, Baños MA, Giménez J, Esplugas S. Degradation of

chlorophenols by means of advanced oxidation processes: a general review. Appl Catal B-

environ. 2004;47:219-56.

Pereira MA, Pires OC, Mota M, Alves MM. Anaerobic biodegradation of oleic and palmitic acids:

evidence of mass transfer limitations caused by long chain fatty acid accumulation onto the

anaerobic sludge. Biotechnol Bioeng. 2005;92:15-23.

Perle M, Kimchie S, Shelef G. Some biochemical aspects of the anaerobic degradation of dairy

wastewater. Water Res. 1995;29:1549-54.

Puyol D, Sanz JL, Rodriguez JJ, Mohedano AF. Inhibition of methanogenesis by chlorophenols:

a kinetic approach. New Biotechnol. 2012;30:51-61.

Ravuri Hk. Role of factors influencing on anaerobic process for production of bio hydrogen:

Future fuel 2013.

Riley RG, Zachara J. Chemical contaminants on DOE lands and selection of contaminant

mixtures for subsurface science research. Pacific Northwest Lab., Richland, WA (United

States); 1992.

Schmid K, Riediker M. Use of nanoparticles in Swiss industry: a targeted survey. Environ Sci

Technol. 2008;42:2253-60.

Sierra-Alvarez R, Lettinga G. The effect of aromatic structure on the inhibition of acetoclastic

methanogenesis in granular sludge. Appl Microbiol Biot. 1991;34:544-50.

Sikkema J. Mechanisms of membrane toxicity of hydrocarbons. Microbiol Rev. 1995;59:201-22.

Sikkema J, Debont JAM, Poolman B. Interactions of cyclic hydrocarbons with biological

membranes. J Biol Chem. 1994;269:8022-8.

Page 43: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

42

Siles JA, Brekelmans J, Martín MA, Chica AF, Martín A. Impact of ammonia and sulphate

concentration on thermophilic anaerobic digestion. Bioresource Technol. 2010;101:9040-8.

Silva SA, Cavaleiro AJ, Pereira MA, Stams AJM, Alves MM, Sousa DZ. Long-term acclimation

of anaerobic sludges for high-rate methanogenesis from LCFA. Biomass Bioenerg.

2014;67:297-303.

Silvestre G, Illa J, Fernández B, Bonmatí A. Thermophilic anaerobic co-digestion of sewage

sludge with grease waste: Effect of long chain fatty acids in the methane yield and its

dewatering properties. Appl Energ. 2014;117:87-94.

Slater JH. Microbial dehalogenation of haloaliphatic compounds. In: Ratledge C, editor.

Biochemistry of microbial degradation: Springer Netherlands; 1994. p. 379-421.

Soldatkin OO, Kucherenko IS, Pyeshkova VM, Kukla AL, Jaffrezic-Renault N, El'skaya AV, et

al. Novel conductometric biosensor based on three-enzyme system for selective determination

of heavy metal ions. Bioelectrochemistry. 2012;83:25-30.

Sprott G, Shaw KM, Jarrell KF. Ammonia/potassium exchange in methanogenic bacteria. J Biol

Chem. 1984;259:12602-8.

Stasinakis AS. Review on the fate of emerging contaminants during sludge anaerobic digestion.

Bioresource Technol. 2012;121:432-40.

Sterritt RM, Lester JN. Interactions of heavy metals with bacteria. Sci Total Environ. 1980;14:5-

17.

Stoll U, Gupta H. Management strategies for oil and grease residues. Waste Manage Res.

1997;15:23-32.

Stuckey DC. Recent developments in anaerobic membrane reactors. Bioresource Technol.

2012;122:137-48.

Stuckey DC, Owen WF, McCarty PL, Parkin GF. Anaerobic toxicity evaluation by batch and

semi-continuous assays. J Water Pollut Control Fed. 1980:720-9.

Summers R, Bousfield S. A detailed study of piggery-waste anaerobic digestion. Agr Wastes.

1980;2:61-78.

Swanwick JD, Foulkes M. Inhibition of anaerobic digestion of sewage sludge by chlorinated

hydrocarbons. Water Pollut Control. 1971;70:58-70.

Takashima M, Speece RE. Mineral nutrient requirements for high-rate methane fermentation of

acetate at low SRT. Res J Water Pollut C. 1989;61:1645-50.

Tijero J, Guardiola E, Cortijo M, Moreno L. Kinetic study of anaerobic digestion of glucose and

sucrose. J Environ Sci Heal A. 1989;24:297-319.

Trohalaki S, Pachter R. Quantum descriptors for predictive toxicology of halogenated aliphatic

hydrocarbons. Sar Qsar Environ Res. 2003;14:131-43.

Trohalaki S, Pachter R, Geiss KT, Frazier JM. Toxicity QSARS for halogenated aliphatics

derived using metabolite descriptors. Chem Res Toxicol. 2003;16:1663-.

Tsai Y-P, You S-J, Pai T-Y, Chen K-W. Effect of cadmium on composition and diversity of

bacterial communities in activated sludges. Int Biodeter Biodegr. 2005;55:285-91.

Uberoi V, Bhattacharya SK. Effects of chlorophenols and nitrophenols on the kinetics of

propionate degradation in sulfate reducing anaerobic systems. Environ Sci Technol.

1997;31:1607-14.

Ulmanu M, Marañón E, Fernández Y, Castrillón L, Anger I, Dumitriu D. Removal of copper and

cadmium ions from diluted aqueous solutions by low cost and waste material adsorbents.

Water, Air, and Soil Pollution. 2003;142:357-73.

Utgikar VP, Tabak HH, Haines JR, Govind R. Quantification of toxic and inhibitory impact of

copper and zinc on mixed cultures of sulfate-reducing bacteria. Biotechnol Bioeng.

2003;82:306-12.

Vallecillo A, Vallecillo. Anaerobic biodegradability and toxicity of chlorophenols. Water Sci

Technol. 1999;40:161-8.

Page 44: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

43

van Beelen P, van Vlaardingen P. Toxic effects of pollutants on the mineralization of 4-

chlorophenol and benzoate in methanogenic river sediment. Environ Toxicol Chem.

1994;13:1051-60.

van Hylckama Vlieg JET, Janssen DB. Formation and detoxification of reactive intermediates in

the metabolism of chlorinated ethenes. J Biotechnol. 2001;85:81-102.

Visser A. The anaerobic treatment of sulfate containing wastewater: Landbouwuniversiteit te

Wageningen; 1995.

Vodovnik M, Kostanjšek R, Zorec M, Marinšek Logar R. Exposure to Al2O3 nanoparticles

changes the fatty acid profile of the anaerobe Ruminococcus flavefaciens. Folia Microbiol.

2012;57:363-5.

Walker DJ, Clemente R, Roig A, Bernal MP. The effects of soil amendments on heavy metal

bioavailability in two contaminated Mediterranean soils. Environ Pollut. 2003;122:303-12.

Wang W, Yu Y, An T, Li G, Yip HY, Yu JC, et al. Visible-light-driven photocatalytic

inactivation of E. coli K-12 by bismuth vanadate nanotubes: Bactericidal performance and

mechanism. Environ Sci Technol. 2012;46:4599-606.

Wang YT, Gabbard HD, Pai PC. Inhibition of acetate methanogenesis by phenols. J Environ Eng-

asce. 1991;117:487-500.

Weathers LJ, Parkin GF. Toxicity of chloroform biotransformation to methanogenic bacteria.

Environ Sci Technol. 2000;34:2764-7.

Wiegant W, Zeeman G. The mechanism of ammonia inhibition in the thermophilic digestion of

livestock wastes. Agr Wastes. 1986;16:243-53.

Wimley WC, White SH. Membrane partitioning: distinguishing bilayer effects from the

hydrophobic effect. Biochemistry-us. 1993;32:6307-12.

Wodzinsk.Rs, Bertolin.D. Physical state in which naphthalene and bibenzyl are utilized by

bacteria. Appl Microbiol. 1972;23:1077-81.

Workentine ML, Harrison JJ, Stenroos PU, Ceri H, Turner RJ. Pseudomonas fluorescens' view of

the periodic table. Environ Microbiol. 2008;10:238-50.

Wu FC, Mu YS, Chang H, Zhao XL, Giesy JP, Wu KB. Predicting water quality criteria for

protecting aquatic life from physicochemical properties of metals or metalloids. Environ Sci

Technol. 2013;47:446-53.

Wu W-M, Hickey RF, Bhatnagar L, Zeikus JG. Fatty acid degradation as a tool to monitor

anaerobic sludge activity and toxicity. Proceedings of the Industrial Waste Conference 1989.

p. 225-33.

Xu K, Liu H, Chen J. Effect of classic methanogenic inhibitors on the quantity and diversity of

archaeal community and the reductive homoacetogenic activity during the process of

anaerobic sludge digestion. Bioresource Technol. 2010;101:2600-7.

Yamaguchi T, Harada H, Hisano T, Yamazaki S, Tseng IC. Process behavior of UASB reactor

treating a wastewater containing high strength sulfate. Water Res. 1999;33:3182-90.

Yang Y, Quensen J, Mathieu J, Wang Q, Wang J, Li M, et al. Pyrosequencing reveals higher

impact of silver nanoparticles than Ag+ on the microbial community structure of activated

sludge. Water Res. 2014;48:317-25.

Yang Y, Zhang C, Hu Z. Impact of metallic and metal oxide nanoparticles on wastewater

treatment and anaerobic digestion. Environ Sci: Processes Impacts. 2013;15:39-48.

Yu HQ, Fang HHP. Inhibition by chromium and cadmium of anaerobic acidogenesis. 2001a. p.

267-74.

Yu HQ, Fang HHP. Inhibition on acidogenesis of dairy wastewater by zinc and copper. Environ

Technol. 2001b;22:1459-65.

Yu ZT, Smith GB. Inhibition of methanogenesis by C-1- and C-2-polychlorinated aliphatic

hydrocarbons. Environ Toxicol Chem. 2000;19:2212-7.

Zayed G, Winter J. Inhibition of methane production from whey by heavy metals - protective

effect of sulfide. Appl Microbiol Biot. 2000;53:726-31.

Page 45: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

44

Zhao J, Wang Z, Dai Y, Xing B. Mitigation of CuO nanoparticle-induced bacterial membrane

damage by dissolved organic matter. Water Res. 2013;47:4169-78.

Zheng XJ, Yu HQ. Biological hydrogen production by enriched anaerobic cultures in the

presence of copper and zinc. J Environ Sci Heal A. 2004;39:89-101.

Zhou H-B, Qiu G-Z. Inhibitory effect of ammonia nitrogen on specific methanogenic activity of

anaerobic granular sludge. J Cent South Univ. 2006;13:63-7.

Zonta Ž, Alves MM, Flotats X, Palatsi J. Modelling inhibitory effects of long chain fatty acids in

the anaerobic digestion process. Water Res. 2013;47:1369-80.

Page 46: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

45

Table 1

A summary of research findings for the impact of ammonia on anaerobic digestion

Substrate Temperature

(°C)

pH Critical TAN conc. or as

specified (mg/L)

Critical FAN conc. or as

specified (mg/L)

Reference

Synthetic

wastewater

(glucose)

35 6.8-7.0 2480 (50% inhibition methane

producing) -

(Zhou and Qiu,

2006)

Cattle manure 55 7.2

4000 (50% inhibition of aceticlastic) 280 (50% inhibition of

aceticlastic)

(Borja et al., 1996) 7500 (50% inhibition of

hydrogenotrophic methanogens)

520 (50% inhibition of

hydrogenotrophic methanogens)

Synthetic

wastewater

(glucose)

52 7.8 7000 (75% inhibition methane

producing)

620 (100% inhibition methane

producing)

(Siles et al., 2010)

Synthetic

wastewater (yeast

extract)

35 7.7 1445 (50% inhibition methane

producing)

27 (50% inhibition methane

producing)

(He et al., 2011)

Synthetic

wastewater (yeast

extract)

35 8.1 - 800 (COD removal efficiencies

of 78–96%)

(Calli et al., 2005)

Slaughterhouse

by-products 55 7.5

5600 (50% inhibition methane

producing)

635 (50% inhibition methane

producing)

(Bayr et al., 2012)

Swine waste 35 7.6 > 5200 (100% inhibition methane

producing)

200 (100% inhibition methane

producing)

(Garcia and

Angenent, 2009)

Page 47: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

46

Table 2

IC50 values of six heavy metals and comparison

Bioactivity

parameter

Temperature

(°C) Product

Carbon

source

IC50 (mg/L)

Reference Cd Cr Cu Ni Zn Pb

Methane

production

potential

35 Methane Glucose 36 27 N.A. 35 7.5 - (Altaş, 2009)

Methane

production

potential

37 Methane Whey - - - - 19.2 27.7 (Zayed and Winter,

2000)

Methane

production

potential

35 Methane VFA 330 250 130 1600 270 8000 (Lin and Chen, 1999)

Sulfate-

Reducing

Bacteria

inhibition

36 Sulfide Yeast extract - - 1136 - 1648 - (Utgikar et al., 2003)

Hydrogen

production

potential

26 Hydrogen Sucrose 3300 3000 30 1300 1500 >5000 (Li and Fang, 2007)

Hydrogen

production

potential

35 Hydrogen Dairy

wastewater 170 72 65 - 120 -

(Yu and Fang, 2001a)

(Yu and Fang, 2001b)

Hydrogen

production

potential

37 Hydrogen Glucose - - 350 - >500 - (Zheng and Yu, 2004)

Page 48: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

48

Fig. 1. Schematic representation of the main conversion processes in anaerobic digestion

(Amaya et al., 2013) including: 1) hydrolysis, breakdown of complex organics into

soluble monomers; 2) acidogenesis, converting small organic molecules into volatile fatty

acids; 3) acetogenesis, converting volatile fatty acids into acetic acid, carbon dioxide, and

hydrogen; and 4) methanogenesis, consuming hydrogen and converting acetate into

methane and carbon dioxide.

Fig. 2. Toxicity sequence of various chlorophenols towards the syntrophic methanogenic

reaction. The logarithm of the partition coefficient of each chlorophenol in octanol/water

(logP) has been listed beside it according to www.guidechem.com.

Fig. 3. A working model showing the direct and indirect inhibition of methanogenesis by

chloroform, trichloroethylene, and perchloroethylene. Presumably, there is an equilibrium

between the protein-bound corrinoids/porphinoids (pool B) and the free

corrinoids/porphinoids (pool F) within methanogen cells to explain the direct and indirect

inhibition of methanogenesis (Yu and Smith, 2000).

Fig. 4. Inhibitory effect of LCFA adsorption over anaerobic granular biomass and

prevention by synthetic adsorbent (bentonite) addition.

Fig. 1.

Page 49: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

49

Fig. 2.

2.15 2.39 2.50

2.75 2.84 3.06 3.06 3.33 3.62

3.61 3.69 3.72 3.77 4.01 4.56 5.12

Page 50: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

50

Fig. 3.

Page 51: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

51

Fig. 4.

Page 52: Toxicants inhibiting anaerobic digestion: A review

ACC

EPTE

D M

ANU

SCR

IPT

ACCEPTED MANUSCRIPT

52

Highlights

critical review of anaerobic toxicity focussing on fundamental mechanisms

looks at specific organics -chlorophenols, halogenated aliphatics, long fatty acids

looks at specific inorganics -ammonia, sulfide and heavy metals

looks at novel nanomaterials and how they inhibit anaerobes

need to develop toxicity sensors with rapid response times