You are on page 1of 15

Bioresource Technology Reports 8 (2019) 100310

Contents lists available at ScienceDirect

Bioresource Technology Reports


journal homepage: www.journals.elsevier.com/bioresource-technology-reports

The anaerobic digestion process of biogas production from food waste: T


Prospects and constraints

Sagor Kumar Pramanika, Fatihah Binti Sujaa, Shahrom Md Zaina, Biplob Kumar Pramanikb,
a
Department of Civil and Structural Engineering, Faculty of Engineering, Universiti Kebangsaan Malaysia, 43600 UKM Bangi, Selangor, Malaysia
b
School of Engineering, RMIT University, Melbourne, VIC 3000, Australia

A R T I C LE I N FO A B S T R A C T

Keywords: The unrestrained release of huge quantities of food waste (FW) has become a significant concern because it
Anaerobic digestion causes intensive environmental pollution. However, FW is a proper substrate that can be treated by anaerobic
Biogas digestion (AD) due to its excellent biodegradability and high-water content. Studies have demonstrated that the
Co-digestion AD process of turning FW into biogas is an effective solution for FW treatment. This manuscript reviews the
Food waste
characteristics of FW, the biological process and biochemical reaction involved in the AD process, various op-
Pre-treatment
erational parameters and classification of the AD process, and the co-digestion and pre-treatment of the AD
process for biogas production. Both co-digestion and pre-treatment processes could improve the FW hydrolysis
rate and methane production. However, further improvement of this technology is required to assess its eco-
nomic feasibility. Challenges and future perspectives of biogas production from FW are also discussed to improve
the performance of AD technology.

1. Introduction 157, 154, 74.7, 51, and 44 kg per person in Australia, America, Japan,
Germany, United Kingdom, India, and China, respectively.
The quantity of food losses and waste has grown enormously in the Since FW creates harmful impacts at the environmental level, ap-
past few years because of the rapid growth of the world economy and propriate management and treatment of FW have become the major
population. It is estimated that approximately 33.3% of food produced objective of numerous countries across the world. One of the main
globally for human consumption is lost or wasted through the food environmental impacts of FW is associated with the embedded carbon
supply chain (i.e., 1.6 gigatons of food per year) which has a production from the earlier life cycle phases of food before it became waste.
value of $750 billion (Ma and Liu, 2019; Slorach et al., 2019). The Moreover, activities related to food production such as agriculture
waste of food is a non-productive use of scarce resources (land, water (including land-use change), processing, manufacturing, transportation,
and fertiliser) and leads to environmental degradation (Gokarn and storage, refrigeration, distribution and retail have an embedded GHG
Kuthambalayan, 2017; Slorach et al., 2019). Baroutian et al. (2018) effect (Papargyropoulou et al., 2014). The same study reported that
reported that the total amount of FW produced by a single person is agriculture is associated with approximately 22% of all GHG emissions
160–295 kg/year all over the world. Food is wasted during the food compared to livestock production (about 18%). The final disposal of FW
supply chain. The chain includes agricultural production, processing, in landfills has also given rise to major environmental pollutions. Clercq
distribution, consumption and post-harvest handling stages (Gustavsson et al. (2017) showed that considerable amounts of GHG including
et al., 2011; Papargyropoulou et al., 2014). It can be noted that de- methane (CH4) and carbon dioxide (CO2) are produced when FW is
veloped countries tend to have major losses (70 to 80%) associated with disposed of in landfills. They showed that the emission of GHG into the
the retail and consumer stages, whereas food wastage is higher at the atmosphere contributes to global warming; methane is a potent GHG
immediate post-harvest stages in developing countries (Gokarn and with a greenhouse effect that is 25 times more powerful than CO2.
Kuthambalayan, 2017). Papargyropoulou et al. (2014) revealed that the Slorach et al. (2019) reported that global food loss and waste generate
quantity of food losses and waste not only fluctuated between devel- annually 6.7% of total anthropogenic GHG emissions. Papargyropoulou
oped and developing countries but also varied in low-income countries et al. (2014) demonstrated another environmental effect associated
having poor producers and customers. A recent report developed by with FW, which is the disturbance of the biogenic phases of phosphorus
Magnet (2018) noted that annual FW generation reached 361, 278, and nitrogen, applied as fertilisers in agriculture.


Corresponding author.
E-mail address: biplob.pramanik@rmit.edu.au (B.K. Pramanik).

https://doi.org/10.1016/j.biteb.2019.100310
Received 9 July 2019; Received in revised form 19 August 2019; Accepted 20 August 2019
Available online 21 August 2019
2589-014X/ © 2019 Elsevier Ltd. All rights reserved.
S.K. Pramanik, et al. Bioresource Technology Reports 8 (2019) 100310

There are several old technologies for treating different types of methane yields. Bong et al. (2018) reported that fruit and vegetable
wastes (e.g., animal manure, household food waste, agricultural crop waste has low lipid content but relatively high cellulosic content,
residue, and organic industrial waste) around the world. Historical whereas FW and kitchen waste have high lipid content because of the
evidence indicates that, the anaerobic digestion process is one of the presence of animal fat and oil. Studies have reported that fruit and
oldest technologies. Granado et al. (2017) reported that the concept of vegetable waste had a lipid content of 11.8%, whereas FW and kitchen
anaerobic digestion has been introduced around 1870 through the de- waste were reported to have 33.22% and 21.6% of lipid content, re-
velopment of the septic tank system. In 1939, the first anaerobic di- spectively (Wang et al., 2014; Yong et al., 2015). Y. Li et al. (2017a)
gestion plant was constructed in the USA for treating organic fraction of reported that FW rich in lipids could produce a higher amount of me-
municipal solid waste, whereas several anaerobic digestion plants were thane compared to carbohydrates and proteins. However, high lipid
constructed over the last few decades in Europe (Karthikeyan et al., content can cause system failure due to the formation of long-chain
2018). The development of anaerobic digestion technology has im- fatty acids. This occurs when the mass transformation of soluble or-
proved very rapidly after the energy crisis in the 1970s (Deepanraj ganics into bacteria decreases due to the destruction of the cellular
et al., 2014). Currently, anaerobic digestion technology is being used membrane (Leung and Wang, 2016). Y. Li et al. (2017a) stated that FW
not only for the treatment of organic wastes but also for the treatment rich in carbohydrate content will affect the carbon and nitrogen ratio
of wastewater. For installing large-scale biogas plants, Germany and (C/N), and thus, nutrient restrictions and quick acidification could
Switzerland are the pioneer countries in the global biogas industry occur due to increased organic matter.
(Karthikeyan et al., 2018). The United States has about 2100 current The total solids and volatile solids of each type of FW fall in the
operational anaerobic digestion plants, whereas Asia has the largest ranges of 10.7%–41% and 10%–34.4%, respectively (Table 1), in-
number of small-scale household anaerobic digesters that are used in dicating that water accounts for 60%–90% in fruit and vegetable waste,
rural areas for lighting and cooking (Vasco-correa et al., 2018). They kitchen waste and FW. FW is considered to be a readily biodegradable
also pointed out that socio-economic hurdle, existing infrastructure, organic substrate because of its large quantity of moisture content
policymakers, technology availability and consistency are the main (Zhang et al., 2014). The characteristics of FW also define the relative
differences in the development of anaerobic digestion plants around the quantities of organic carbon and nitrogen present in the FW. The C/N
world. ratio for each type of FW was found to vary in the range of 12.7–28.84,
AD is an excellent alternative for FW treatment compared to waste and displayed an acidic pH of 4.1–6.5 (Table 1). The methane pro-
treatment, energy supply, and environmental protection (Leung and duction of every category of FW fall in the range of 346–551.4 mL/
Wang, 2016). There are numerous benefits related to the AD process g VSadded (Table 1), which is higher compared to cow manure (233 mL/
such as decreased GHG emission, digestate for application in agronomy, g VSadded), grass silage (306 mL/g VSadded), and oat straw (203 mL/
small footprint production, and the generation of high-quality renew- g VSadded) (Huttunen and Rintala, 2007).
able fuel (Ariunbaatar, 2014). However, the drawbacks of the AD
process such as relatively high capital costs, long retention time, and 3. Biological process and biochemical reaction involved in the AD
the required control of certain key parameters (e.g., pH, temperature, process
feed rate, alkalinity) prevents it from being widely implemented
(Ariunbaatar, 2014). Herein, the objective of this paper is to review the Anaerobic digestion is a biological process, which breaks down
characteristics of FW, the biological process and biochemical reaction complex organic matter into simpler chemicals components in the ab-
involved in the AD process, various operational parameters and clas- sence of oxygen. During this process, a gas that is mainly composed of
sification of the AD process, and the co-digestion and pre-treatment of CH4 and CO2, also referred to as biogas, is produced as the end products
the AD process for biogas production. This paper also discusses the under ideal conditions. A minor quantity of hydrogen sulphide (H2S),
challenges and future perspectives of enhancing biogas production from ammonia (NH3), and other gases are also present when biogas is pro-
FW and maintaining the AD process effectively. duced in the AD plant (Monnet, 2003). The AD process can be divided
into four stages i.e. hydrolysis, acidogenesis, acetogenesis, and metha-
2. Characteristics of FW nogenesis, as it is a multi-step biochemical process (Zhang et al., 2014).
In the AD procedure, various kinds of bacteria degrade the organic
FW is characterised by complex components and organic material. substance continuously in a multi-step method and via parallel reac-
There are several types of FW such as fruit and vegetable waste, tions. Microorganisms play an essential role in the AD process, and the
household and restaurant FW, brewery waste, and dairy waste (Xu bacterial groups are dissimilar among the phases of hydrolysis, acid-
et al., 2018). Studies have found that the composition of FW varies ification, and methane production (P. Wang et al., 2018). Li et al.
based on geographical changes, seasonal changes, cooking procedures, (2015) investigated the relationship between microbial community
and consumption patterns (Meng et al., 2015; Xu et al., 2018). They structure and process stability and compared the microbial community
reported that FW consists of various organic components such as pro- structure in both stable and deteriorative phase using 454-pyr-
teins, carbohydrate polymers (starch, cellulose, hemicelluloses, and osequencing. They reported that bacteria are responsible for the de-
lignin), lipids, and organic acids. Fisgativa et al. (2016) studied 102 gradation of FW to intermediate metabolites which can be later used by
different FW samples and reported that the characteristic of FW dis- methanogens. They concluded that acid-producing bacteria such as
played high coefficient of variance (CV). They indicated that the var- Acholeplasma and Actinomyces increased dramatically at deteriorative
iations of 24% of the studied characteristics were described by the phase compared with stable stage, which may be the failure indicator in
geographical change, seasonal change and the type of collection source. anaerobic digester treating FW. In hydrolysis, insoluble complex poly-
They observed that FW has an average pH of 5.1 (CV 13.9%), carbon mers comprising carbohydrates, proteins, lipids, and other organics are
and nitrogen ratio of 18.5% (CV 31.8%), 36% of carbohydrates (CV converted into smaller soluble molecules. It can be noted that hydro-
57.2%), 26% of protein (CV 62.2%), 15% of fats (CV 52.0%), and lysis is a comparatively slow stage and therefore can limit the rate of the
biomethane potential of 460.0 NL CH4/kg VS (CV 19%). entire AD process, especially when FW is used as the feedstock (Kothari
As mentioned in Table 1, the percentage range of degradable car- et al., 2014; Leung and Wang, 2016; Ostrem, 2004; A. Zhang et al.,
bohydrates, proteins, and lipids is (5.7%–53%), (2.3%–28.4%), and 2015; Zhang et al., 2014). In acidogenesis, monomers and dissolved
(1.3%–30.3%), respectively. Meng et al. (2015) and Xu et al. (2018) compounds such as sugars, amino acids, and fatty acids resulting from
indicated that carbohydrates and proteins have a higher hydrolysis rate hydrolysis are converted into simple molecules with a small molecular
due to its rapid degradability compared to lipid. Thus, quickly de- weight such as volatile fatty acids (i.e., propionic, butyric, acetic acid),
gradable carbohydrates and lipid-rich food wastes can produce high alcohols, and different kinds of gases (CO2, H2, and NH3) (A. Zhang

2
S.K. Pramanik, et al. Bioresource Technology Reports 8 (2019) 100310

Table 1
Characteristics and methane potential of some food waste reported in literatures.
Source TS (%) VS (%) VS/TS (%) C/N ratio pH Carbohydrates (%) Proteins (%) Lipids (%) Methane yield (mL/g VS) Reference

KW 24.9 23.1 92.8 18.24 – 49 17.3 23 501 J. Jiang et al. (2018)


FVW 13.8 12.88 93.4 4.5 7.74 3.28 2.87 516 Edwiges et al. (2018)
FW 20 19.26 96.3 15.5 – 47.6 24.1 28.3 548.1 Li et al. (2018)
19.2 18.45 96.1 12.7 – 41.3 28.4 30.3 541.1
20.9 20.02 95.8 14 – 38.4 26.3 25.3 545.3
FW 10.86 10.22 94 15.18 4.16 5.71 2.29 1.31 460 Xiao et al. (2019)
FW 25.94 24.59 94.7 17.5 – 48 15.1 10.6 346.2 Shi et al. (2018)
FW 10.69 10.06 94 – 4.18 5.69 2.29 1.30 477–459 Xiao et al. (2018)
FW 24.3 22.5 92.6 23.11 5.02 – 3.38 – 386.7–551.4 Liu et al. (2017)
FW 19.1 18.53 97 17.7 – 10.8 4 3.7 536.19 Y. Li et al. (2017a)
17.2 16.72 97.2 13.4 – 9.4 4.5 2.9 441.23
20.5 19.78 96.5 17.8 – 11 3.6 5.3 531.30
19.7 19.05 96.7 16.8 – 10.1 4.3 4.7 537.37
19.6 18.91 96.5 15.4 – 9.3 4.6 5 533.01
20.0 19.26 96.3 15.5 9.2 4.6 5.5 544.95
AFW 41 34.44 84 – – 52 12 25 – Naroznova et al. (2016)
VFW 24 22.32 93 – – 53 5 14 425
KW 19.1 17.80 93.2 14.41 4.5 11.8 2.5 3.5 372.1 Li et al. (2016a)
FW 23.2 21.7 93.5 4.4 13.7 2.9 6.5 425.2 W. Zhang et al. (2015)
FW 20.05 19.21 95.81 28.4 – 33.22 14.03 25.25 381 Yong et al. (2015)
FW 29.4 28.01 95.3 14.2 4.1 – 18.1 19 529 Browne and Murphy (2013)
FW 18.1 17.1 94 13.2 6.5 11.2 3.3 2.3 479.5 Zhang et al. (2011)

Note: FVW: fruit and vegetable waste, KW: kitchen waste, FW: food waste, AFW: animal food waste, VFW: vegetable food waste.

et al., 2015). In acetogenesis, acetogenic bacteria use volatile fatty acids utilisation rates, and bacterial development (Khalid et al., 2011). The
for their growth and the growth of these bacteria is slow with a dou- same study reported that cell energy fatigue and intracellular substance
bling time of 1.5 to 4 days (Kothari et al., 2014). The concentration of leakage could occur due to this lower temperature. They also indicated
products created in this phase differs according to the type of bacteria that AD operating under mesophilic conditions offers higher stability
as well as culture circumstances such as temperature and pH (Ostrem, and requires lower energy cost compared to the thermophilic condition.
2004). Methanogenesis is the final step in the anaerobic digester, where AD operating under thermophilic condition provides several benefits
the methanogens use acetic acid, hydrogen, and carbon dioxide to such as the higher growth of methanogenic bacteria at a higher tem-
produce methane gas. Most methanogen bacteria require an optimum perature, reduced retention time, destruction of pathogens, enhanced
pH range between 6.5 and 7.5 (Leung and Wang, 2016). The four steps digestibility and better degradability of solid substrates, besides dif-
of anaerobic biodegradation process such as hydrolysis, acidogenesis, ferentiating liquid and solid portions (Dobre et al., 2014; A. Zhang
acetogenesis, and methanogenesis, and the bacteria involved in each et al., 2015). However, the drawbacks of the thermophilic condition
stage of the AD process, are schematically displayed in Fig. 1. must be considered. There are several demerits of the thermophilic
condition such as a greater amount of disproportion and higher energy
4. Parameters affecting the AD process of biogas production requirement because of the associated high temperature (A. Zhang
et al., 2015).
Different categories of bacteria are engaged in the AD process. The
bacteria need an environment that has reached equilibrium to produce 4.2. pH and VFA
biogas from FW. Leung and Wang (2016) reported that a stable process
could be achieved if the bacteria stayed in an ideal condition. Fluc- pH is the most significant parameter that affects the performance
tuations in environmental conditions could interrupt the micro- and stability of an anaerobic digester. Microorganisms are sensitive to
organism's equilibrium, which could perhaps impede or even close pH. This is because every group of bacteria needs a different pH range
down the AD process (Ostrem, 2004). Therefore, operational conditions for their growth (Appels et al., 2008). The ideal pH range for hydrolysis,
(such as temperature, pH and VFA, the ratio of carbon and nitrogen, acetogenesis, and methanogenesis is almost 6.0, 6.0–7.0, and 6.5–7.5,
retention time, and organic loading rate) in the AD process need to be respectively (Leung and Wang, 2016). Gerardi (2003) reported that the
continuously observed and maintained within optimum ranges. The pH required for acid-forming bacteria and methane-forming bacteria
effect and optimised range of these parameters on biogas production is > 5.0 and 6.2, respectively, for acceptable enzymatic activity. Me-
are explained in the following section. thanogenic bacteria display better performance in a pH range of 6.8–7.2
(Yadvika et al., 2004). CH4 production was found to be 75% more ef-
4.1. Temperature ficient with a pH of > 5.0 (Yadvika et al., 2004). Krishna and
Kalamdhad (2014) indicated that other main factors contribute to the
Methane production is highly influenced by temperature, which fluctuation of pH such as alkalinity, volatile fatty acid (VFA), the
affects bacterial performance within an anaerobic digester. Both me- quantity of CO2 production, and the concentration of bicarbonate
thanogenic and volatile acid-forming microorganisms are affected by (HCO3) during the AD process. They reported that the relationship
temperature. Changes in temperature extensively influence the perfor- between VFA and HCO3 concentrations should be controlled, as it helps
mance of methanogenic microorganisms compared to the operating in adjusting the optimum pH during the AD process. For a stable and
temperature (Gerardi, 2003). The AD process can take place at various well-buffered digestion process, it is important to maintain at least
temperatures, which are normally classified into three types, i.e. psy- 1.4:1 (molar ratio) of HCO3/VFA or a buffering capacity of
chrophilic, mesophilic, and thermophilic temperatures. The mesophilic 70 meq CaCO3/L (Appels et al., 2008). Volatile fatty acids are short-
and thermophilic temperature ranges are between 20 and 40 °C (usually chain fatty acids (acetic acid, propionic acid, butyric acid, and valeric
35 °C) and 50 and 65 °C (typically 45 °C), respectively (Kothari et al., acid), which are the primary intermediate products produced from the
2014). A lower temperature reduces methane production, FW AD of FW (Zhang et al., 2014). Xu et al. (2014) and Shi et al. (2018)

3
S.K. Pramanik, et al. Bioresource Technology Reports 8 (2019) 100310

Food Waste

Complex polymers
pH= 5.5-6.0
Hydrolysis

Carbohydrates Proteins Lipids

Clostridium, Acetivibrio, Clostridium, Proteus, Clostridium, Fermentative


Staphylococcus, Bacteroides Peptococcus, Vibrio, Micrococcus,
Bacteroides, Bacillus Staphylococcus bacteria
Acidogenesis
pH= 6.0-7.0

Monomers

Sugars Amino acids LCFA

Clostridium, Lactobacillus, Escherichia, Staphylococcus, Acidogenic


Eubacterium limosum, Micrococcus, Bacillus, Sarcina, Veillonella,
Pseudomonas, Desulfovibrio, Desulfuromonas, bacteria
Streptococcus
Desulfobacter, Selenomonas, Streptococcus
Acetogenesis
pH= 6.0-7.0

VFAs and Clostridium, Syntrophobacter,


Syntrophomonas
alcohols

Acetogenic
Clostridium, Syntrophobacter, Syntrophomonas bacteria

Acetate oxidizing bacteria


Acetate H2, CO2
Homoacetogenic bacteria

Methanobacterium, Methanocalculus,
Methanogenesis

Methanosarcina, Methanospirillum, Methanobrevibacter, Methanoplanus, Methanogenic


pH= 6.5-7.5

Methanosaeta Methanoregula, Methanococcus, archaea


Methanoculleus

Acetotrophic Hydrogenotrophic
CH4, CO2
methanogens methanogens

Fig. 1. Four biological steps and the respective bacterial groups engaged in every phase of the AD process. Information collected from Deepanraj et al. (2014),
Gonzalez-Fernandez et al. (2015), Kothari et al. (2014), Leung and Wang (2016), and P. Wang et al. (2018).

reported that methane production was inhibited completely when the for an effective AD process (Kothari et al., 2014). Zhang et al. (2014)
concentrations of VFA fell in the range of 5800 to 6900 mg/L. reported that the C/N ratio greatly influences the stability of the AD
process. This is because the optimal C/N ratio not only helps to main-
4.3. Carbon and nitrogen ratio tain a suitable environment, but it also helps to control proper nutrient
balance through the development of microorganisms. The microbial
The C/N ratio represents the relationship between the quantity of population could increase gradually if the quantity of nitrogen is low in
carbon and nitrogen present in FW. An optimum C/N ratio is required the FW, and thus, more time will be required to decompose the existing

4
S.K. Pramanik, et al.

Table 2
Comparison of different anaerobic digestion process for food waste.
Process Reactor volume and type Inoculum Operating condition Methane yield (mL/ VS removal (%) Reference
g VS)

Single-stage 2 L semi-continuous CSTR, Anaerobic sludge OLR = 5 g VS/L day, HRT = 20 days, pH = 7.7 494 74.7 Jo et al. (2018)
mesophilica
Two-stage 0.5 L and 2 L semi-continuous CSTR, OLR = 4 g VS/L day, HRT = 25 days, pH = 7.5 511 78.9
mesophilic
Single-stage 230 L CSTR, thermophilicb Anaerobic sludge OLR = 3.5 kg VS/m3 day, HRT = 20 days 450 93.6 Micolucci et al. (2018)
Two-stage 200 and 760 L CSTR, thermophilic 550 96.2
Single-stage 4 L semi-continuous CSTR, Mesophilic anaerobic sludge HRT = 30 days, pH = 8.31, TVFA = 0.87 g HAc/L 477 83.22 Xiao et al. (2018)
thermophilic
Two-stage 2 L and 8 L semi-continuous CSTR, HRT = 30 days, pH = 3.83 and 8.17, TVFA = 1.64 and 459 82.02
thermophilic 1.08 g HAc/L
Single-stage 1 and 20 L semi-continuous, Anaerobic seed sludge OLR = 1.6–10 g VS/L, pH = 5.1–7.8 199 30.1 Zhang et al. (2017)
Two-stage mesophilic 249 44.2
Three-stage 307 83.5

5
Single-stage 5 m3 reactor, mesophilic Anaerobic sludge OLR = 3.79 kg VS/m3/day, pH = 7.32 380 96 Grimberg et al. (2015)
Two-stage OLR = 0.78 kg VS/m3/day, pH = 5.2, 8.4 446 93
Single-stage 6 L CSTR, mesophilic Mesophilic seed sludge OLR = 2.4 g VS/L/day, pH = 7.77, HRT = 30 days 440 74.1 Wu et al. (2015)
Temperature-phase two- 1.5 L thermophilic CSTR, 6 L OLR = 14.2 and 2.6 g VS/L/day, pH = 5.36 and 7.59, HRT = 3 440 80.1
stage mesophilic CSTR and 12 days
Mesophilic 3 L continuous CSTR Anaerobic seed sludge OLR = 7.75 g VS/L/day, HRT = 10 days, pH = 7.64 350 – Q. Li et al. (2017)
Thermophilic OLR = 5.19 g VS/L/day, HRT = 15 days, pH = 7.86. 407
Mesophilic 500 mL laboratory-scale bottle (semi- Mesophilic sewage sludge OLR = 1.5 g VS/L/day, HRT = 20 days 371 94.7 Liu et al. (2017)
Thermophilic continuous) Thermophilic sewage sludge OLR = 2.5 g VS/L/day, HRT = 20 days 541 93
Batch 75 L upflow anaerobic reactor, Seed sludge OLR = 6.1 kg COD/m3/day, HRT = 30 days, pH = 7.5 266a 80.9b Park et al. (2018)
Continuous mesophilic OLR = 7.9 kg COD/m3/day, HRT = 13 days, pH = 7.5 326a 71.2b
Batch 1 L glass digesters, mesophilic Anaerobic activated sludge OLR = 8 g VS/L, pH = 7.5 388 – Zhang et al. (2013)
Continuous OLR = 10 g VS/L, pH = 7.1–7.7 317
Wet 6.0 L continuous reactor, mesophilic Mesophilic seed sludge OLR = 2.35 kg VS/m3/day, SRT = 20 days, pH = 7.39 370 80.1 Yi et al. (2014)
Dry OLR = 9.41 kg VS/m3/day, SRT = 20 days, pH = 7.82 480 85.6

a
L/kg COD.
b
COD removal.
Bioresource Technology Reports 8 (2019) 100310
S.K. Pramanik, et al. Bioresource Technology Reports 8 (2019) 100310

FW, resulting in lower CH4 yield. In contrast, ammonia inhibition could On the other hand, higher OLR causes shorter HRT, which may result in
occur, preventing microbial growth, especially if the concentration of microorganism washout, and this could lead to lower biogas production
nitrogen is more than the microbial necessity (Kothari et al., 2014; (Leung and Wang, 2016). Hu et al. (2018) reported that the thermo-
Leung and Wang, 2016). It is found that microorganisms use carbon philic reactor can be adapt to a wide range of OLR (3.0–14.4 kg-COD/
25–30 (Yadvika et al., 2004), 25–35 (Krishna and Kalamdhad, 2014) or m3/day) compared to the mesophilic reactor (OLR of 3.0–7.3 kg-COD/
30–35 (Leung and Wang, 2016) times quicker than nitrogen during the m3/day) in anaerobic FW treatment. Liu et al. (2017) pointed out that
AD process. Therefore, the optimum ratio of C/N in the substrate should the optimal OLR on AD of FW under thermophilic and mesophilic
be almost 20–30:1 (Yadvika et al., 2004) 25–30:1(Krishna and condition was 2.5 and 1.5 g-VS/L/day, respectively.
Kalamdhad, 2014) or 30–35:1 (Leung and Wang, 2016). The co-diges-
tion of organic substrates with animal manure, lignocellulosic biomass, 5. Classification of the AD process of FW
and sewage sludge can be used to enhance the C/N ratios (Khalid et al.,
2011). The classification of the anaerobic reactor when treating FW de-
pends on the temperature, feeding type, and moisture content of the
4.4. Retention time FW. Single-stage and two/multi-stage processes have also influenced
the performance of the AD process (Table 2) (Deepanraj et al., 2014).
Retention time is one of the major parameters, which needs to be Mesophilic or thermophilic, batch or continuous process, or wet or dry
repeatedly observed in the anaerobic digesters. Retention time is the digestion are the various categories of the AD process, which are further
time needed for the complete degradation of organic matter or the described in the following sections.
average time organic matter remains in a digester (Deepanraj et al.,
2014; Mao et al., 2015). There are two important types of retention 5.1. Wet and dry digestion
times involved in the AD system: solid retention time (SRT) and hy-
draulic retention time (HRT). SRT is the average time that the bacteria The AD process can be categorised as wet or dry digestion based on
spend in the digester, whereas HRT is the average time that the liquid the total solid concentration in FW. The anaerobic digestion of food
sludge spends in the digester (Deepanraj et al., 2014). According to Mao waste is termed as a dry process if the total solid concentration of the
et al. (2015), bacterial development rate related to retention time de- food waste stays between 20 and 40%, whereas the anaerobic digestion
pends on process temperature, substrate composition, and organic of food waste is considered as a wet process when the total solid con-
loading rate; the shorter the HRT, the higher the value of organic centration of the food waste is < 15% (Kothari et al., 2014; Deepanraj
loading rate. A retention time of 10–40 days is necessary to treat or- et al., 2014). Kothari et al. (2014) noted that most of the AD plants
ganic waste in mesophilic temperature, while a lesser retention time constructed during the 1980s depended on the wet system, whereas
could be used in thermophilic temperature (Kothari et al., 2014). new plants constructed in the last decade are mostly based on the dry
Yadvika et al. (2004) described that high capital cost and large reactor process. They also noted that the HRT, OLR, and volatile solid removal
volume are the main requirements for a longer HRT. However, shorter rate of the dry process were 14–60 days, 12–15 kg VS/m3/day, and
HRT offers insufficient time for the optimal degradation of the sub- 40%–70%, respectively, whereas the HRT, OLR, and volatile solid de-
strate. Yadvika et al. (2004) noted that HRT varies with climate change. struction rate of the wet process were 25–60 days, < 5 kg VS/m3/day,
For example, for tropical countries and in cold weather, HRT fluctuates and 40%–75%, respectively. It is important to note that a higher cost for
from 30 to 50 days and 100 days, respectively. SRT maintains the digestate dewatering post-treatment and higher reactor volume are
bacterial population in the reactor, which could result in waste stabi- required for the wet process compared to the dry process. Kothari et al.
lisation (Dobre et al., 2014). A portion of the microbial community is (2014) noted a higher biogas production rate in the dry process com-
removed every time sludge is withdrawn, resulting in a stable condition pared to the wet process. Yi et al. (2014) examined dry (total
and the prevention of process failure (Appels et al., 2008). A long SRT solid = 20%) and wet (total solid = 5%) processes of AD during food
helps permit biological acclimation to toxic compounds, and this can be waste treatment, and found that the CH4 production (0.48 L/g VS) and
achieved by increasing both the volume of the digester and the con- volatile solid reduction (85.6%) were higher in the dry process com-
centration of the bacteria (solids) (Gerardi, 2003). Chen et al. (2018) pared to the wet process. They also found that the dry process allowed a
observed that the CH4 yield reduced with higher SRT and the highest higher VFA concentration and OLR compared to the wet process, re-
CH4 yield was achieved at an SRT of 6 days, compared to 7.5 days and sulting in a decreased possibility of inhibition of the AD technique. The
10 days. Fernández-Rodríguez et al. (2014) reported that the maximum comparison of dry versus wet digestion process has shown that the dry
CH4 yield was achieved at an SRT of 5–8 days compared to SRTs of process had a more advantageous energy balance and economic per-
4 days and 3 days. They also indicated that SRTs lower than 4 days were formance compared to the wet process. However, complete mixing of
inappropriate for a single-stage dry AD of organic fraction of municipal the waste is not possible in dry process, and, thus, the ideal contact of
solid waste, which would render the process unbalanced. microorganisms and substrate cannot be guaranteed. Conversely, wet
process offered several important benefits including greater flexibility
4.5. Organic loading rate over the type of feedstock accepted, dilution of inhibitory substances by
process water and necessity of less sophisticated mechanical equip-
Organic loading rate (OLR) is a significant operational parameter ment. Overall, dry digestion process was more cost-effective, and it had
that affects the CH4 yield. OLR is defined as the amount of dry organic greater biogas productivity than wet process.
solids, which can be fed into the digester per day per unit volume of
digester capacity (Kothari et al., 2014). If the reactor is overfed beyond 5.2. Mesophilic or thermophilic digestion
the suitable OLR, inhibitory substrates such as fatty acids could be ac-
cumulated and CH4 production could be reduced. This is because micro The AD process can take place at various temperatures, usually
bacteria cannot survive in an acidic condition in the AD system. System classified into mesophilic and thermophilic temperatures. As mentioned
failure can also occur due to overfeeding. This is turn affects CH4 in Section 4.1, mesophilic and thermophilic digesters usually operate at
production rate, which is highly dependent on OLR (Kothari et al., 20–40 °C (usually 35 °C) and 50–65 °C (usually 45 °C), respectively.
2014). Hence, it is important to control the OLR of the digester. Leung Thermophilic microorganisms are more sensitive to temperature
and Wang (2016) demonstrated that lower OLR and longer HRT might changes compared to mesophilic bacteria. Banks et al. (2008) compared
prompt organic overload, and thus, CH4 production could be reduced. the thermophilic condition with the condition for source-segregated
This is likely due to the insufficient buffering capability in the digester. domestic FW. They found that the AD under thermophilic conditions

6
S.K. Pramanik, et al. Bioresource Technology Reports 8 (2019) 100310

displayed better volatile solid removal efficiency and biogas yield in operating conditions of low HRT (2–3 days) and acidic pH (5.5–6.5),
compared to mesophilic conditions. They also found that the VFA in which the acidification stage must be maintained, whereas high HRT
concentration of the thermophilic reactor (45,000 mg/L) was higher (20–30 days) and optimum pH (6–8) are required for the development
than that of the mesophilic reactor (28,000 mg/L). Q. Li et al. (2017) of gradually growing methanogenic bacteria (Xu et al., 2018). Hagos
also compared the thermophilic digestion and mesophilic digestion of et al. (2017) reported the several benefits of the two-stage process such
FW and waste activated sludge mixture. They found that the thermo- as increased CH4 production, high OLR, improved process stability,
philic system (407 mL/g VSadded) yielded better CH4 production com- higher possibility of managing pathogens, and enhanced rate of volatile
pared to the mesophilic system (350 mL/g VSadded). This was likely due solid removal efficiency. They also noted that complex maintenance,
to the high transformation ratio from protein to ammonium in the high capital, and operational cost were the major drawbacks of the two-
thermophilic reactor. stage process. Xiao et al. (2018) compared the performance and energy
recovery of single-stage and two-stage thermophilic anaerobic digestion
5.3. Batch and continuous system (TAD) of FW. They found that single-stage and two-stage TAD both
exhibited better performance with an HRT of 30 days. However, single-
A reactor that is fed at one time with fresh FW and the addition of stage TAD achieved slightly higher CH4 yield and volatile solid removal
inoculum (such as anaerobic sludge) from other reactors followed by its rate compared to the two-stage TAD. They also noted that single-stage
closing down for a certain period to degrade substrate anaerobically is TAD recovered more energy than the two-stage TAD. Another study by
called a batch system. Batch digesters have several benefits such as Micolucci et al. (2018) found that the two-stage TAD of FW displayed
being technically simple and having low investment cost and main- clearly enhanced biogas production and organic matter destruction
tenance requirements and minimum parasitic energy loss (Kothari compared to the single-stage system. The volatile solid and augmen-
et al., 2014). On the other hand, for the continuous process, fresh FW is tation removal rate of the two-stage process was increased by 17% and
continuously loaded in the reactor and the same quantity of digested 16%, respectively. They also indicated that the digestate dewatering
FW is removed. Studies have reported that biogas recirculation or post-treatment cost in the two-stage TAD was lesser than in the single-
mechanical agitators could be used in the continuous process to agitate stage, whereas the two-stage TAD produced more net energy than the
the FW and inoculum continuously (Kothari et al., 2014; Deepanraj one-stage TAD. Zhang et al. (2017) established a compact three-stage
et al., 2014). Biogas production can be kept almost constant and/or anaerobic digester for FW in which a three-stage anaerobic digestion
continuous due to the constant input of FW. Park et al. (2018) studied process was used, consisting of three independent chambers including
the effect of feeding mode for the AD of FW, and reported that the the hydrolysis stage, acidification stage, and wet methane-production
continuous feeding of diluted FW yielded constant performance com- stage. They found that the three-stage AD obtained 24%–54% more CH4
pared to the batch feeding of undiluted FW. They also noted that the yield with an OLR of 10 g VS/L compared to the one-stage and two-
potential CH4 yield was 2.78 L CH4/L/day at an OLR of 8.6 kg COD/m3/ stage AD. The three-stage AD also achieved a high volatile solid re-
day during continuous feeding of diluted FW, whereas the potential moval efficiency of 83.5%.
CH4 yield was 1.51 L CH4/L/day at OLR of 6.1 kg COD/m3/day during
batch feeding of undiluted FW.
6. Factors that improve the anaerobic digestion process
5.4. Single-stage and two/multi-stage process
6.1. Co-digestion
The single-stage process occurs when four metabolic phases—hy-
drolysis, acidogenesis, acetogenesis, and methanogenesis—take place in Co-digestion is defined as two or more types of organic waste, which
one reactor. Low OLR, long retention time, a pH range between 6 and 7, are mixed and treated simultaneously. This process is extensively
low CH4 production, less investment, and less maintenance cost are the practiced nowadays to avoid the problems associated with the mono-
major features of the single-stage process (Xu et al., 2018). The main digestion of FW (Shi et al., 2018). Anaerobic co-digestion has numerous
limitation in the single-stage reactor is the presence of acidogenic mi- advantages which are displayed in Fig. 2 (Mehariya et al., 2018; Tyagi
croorganisms (i.e., decreases pH) that result due to the quick acid- et al., 2018). However, there are a few disadvantages of co-digestion
ification of FW during the acid formation phase, which disturbs the including transport costs (Mata-Alvarez et al., 2000). The effect of co-
methanogenic bacterial groups. This is likely due to the lack of an op- digestion of FW with different substrates is summarised in Table 3. The
timum environmental condition inside the single-stage reactor, which is mono-digestion of FW often results in low buffer capacity and a high C/
required for the growth of all participating bacteria (Deepanraj et al., N ratio. This is likely due to the high biodegradability of the substrate
2014; Xu et al., 2018). However, adjusting the feeding rate via mixing (Leung and Wang, 2016). Therefore, animal manure can be considered
and adding a buffer in the single-stage reactor could be a solution to as a suitable co-substrate due to its wide variety of nutrients and high
manage the methanogenic microbial community in the reactor (Monson buffer capacity, which could enhance the maximum allowable OLR (up
et al., 2007). The addition of buffer (NaHCO3) reduced volatile solid to 10 kg VS/m3/day) and offer a more stable environment for anaerobic
and biodegradation time, which resulted in higher biogas productions bacteria (Xu et al., 2018). Sewage sludge is considered to be an ex-
compared with the performances of the other non-chemically buffered cellent co-substrate because of its low organic load and high trace
cultures studied (Abdulkarim and Evuti, 2010). Chen et al. (2015) element content (Mehariya et al., 2018). Studies have reported that
studied the effects of four alkalinity sources (i.e. lime mud from pa- large amounts of active bacteria present in sewage sludge are favorable
permaking, waste eggshell, CaCO3 and NaHCO3) on the stability of for the development of various groups of microorganisms engaged in
anaerobic digestion from food waste. They reported that adding a the AD process (Xu et al., 2018). The same study reported that CH4
buffer in the feeding stage promoted process stability and methane yield was enhanced by 1.4 times after the addition of sewage sludge
production compared to the non-chemically buffered sample. It is worth into FW. The main substrate for co-digestion was primary sludge and
noting that approximately 95% of Europe's full-scale plants usually waste activated sludge, where most of the plants operated at 35 °C
operate in a single-stage anaerobic process when treating organic waste temperature (except for the Camposampiero and Kurobe plants). The
(Nagao et al., 2012). electricity generation of the East Bay MUD plant was 11 MW, which
The two-process or multi-stage process occurs in separate anaerobic was higher than the Rovereto (0.4 MW), Camposampiero (0.4 MW),
reactors. The first reactor is used for hydrolysis, acidogenesis, and Treviso (0.125 MW), Moosburg (0.38 MW), and Kurobe (0.095 MW)
acetogenesis, whereas the second reactor is mainly employed for me- plants (Nghiem et al., 2017).
thanogenesis. Studies have shown that the acidification stage is evident

7
S.K. Pramanik, et al. Bioresource Technology Reports 8 (2019) 100310

Fig. 2. Primary benefits associated with co-digestion of food waste. Information collected from Mehariya et al. (2018) and Tyagi et al. (2018).

6.2. Pre-treatment (Liu et al., 2019). Karthikeyan et al. (2018) showed that the aim of
using pre-treatment is to increase the biogas yield along with en-
The performance of the AD process can be enhanced with the help hancement of the hydrolysis rate kinetics for proteins and lipids. They
of a pre-treatment method. As mentioned in Section 3, among the also reported that the pre-treatment process could decrease harmful
biological stages, hydrolysis is the main rate-limiting step in the AD of composites, which could disturb the AD process. In principle, pre-
FW. This is because every stage directly influences both the food treatment can minimise the quantity of sludge and enhance the re-
availability and mass transfer of the AD process. Zhang et al. (2014) duction of organic substances. Numerous pre-treatment methods have
reported that the rate of hydrolysis is affected by particle size, com- been used for the AD of FW (Table 4). Among the pre-treatment
position, and substrate component during the biodegradation of gran- methods include chemical, thermal, ultrasonic, mechanical, and biolo-
ular substrates and high molecular compounds. According to gical methods, which are further discussed in the following paragraph.
Ariunbaatar et al. (2014), FW needs to be pasteurised or sterilised be- Mechanical pre-treatment is aimed at enhancing the specific surface
fore the AD process, and pre-treatment might be the best option for area and minimising the particle size of organic polymers (Ren et al.,
achieving more energy recovery along with eradicating the additional 2018). An increased surface area offers excellent communication be-
cost for pasteurisation/sterilisation. Liu et al. (2019) reported that most tween the substrate and anaerobic microorganisms, which can in turn
biogas plants apply thermal pasteurisation according to the appropriate increase the performance of the entire AD process (Ariunbaatar et al.,
regulations. Although the design and the operational approach of the 2014; Ren et al., 2018). The principle of ultrasonic pre-treatment de-
thermal pasteurisation differ from one biogas plant to another, when it pends on the cavitation process, which is produced from massive hydro-
comes to commercially viable technology, CambiTHP™ and Biothelys® mechanical shear force through high-intensity ultrasonic waves and
are the global leading provider of thermal hydrolysis, advanced anae- sludge disruption (Deepanraj et al., 2017). Increased ultrasonic pre-
robic digestion and biogas solutions for sewage sludge and organic treatment time has many physical and chemical impacts on the sub-
waste management (Panigrahi and Dubey, 2019). Regarding techno- strate such as free radical (e.g., HO2%, OH% and H%) production, re-
economic feasibility, 6–25% of the total primary energy production can duction in partial pressure, micro-bubble production, disintegration of
be achieved in European biogas plants through thermal pasteurisation the cell walls, moderate spreading, and solubilization of solid organic

8
S.K. Pramanik, et al.

Table 3
Comparison of operational parameters for anaerobic co-digestion of food waste with other substrates and their respective methane production.
Feedstock + Co-substrate Mixing ratio Reactor type Inoculum Operating conditions Methane yield Reference
(mL/g VS)

Cucumber residues + pig manure + corn stover 3:5:2 (wet basis) 1 L glass reactor (batch) Anaerobic sludge Temperature = 35 °C, C/N = 14.5, VFA/alkalinity 305.4 T. Wang et al.
ratio = 0.76 g/L (2018)
FW + pig manure 50:50 (VS basis) 1 L glass digester (batch) Solid digestate Temperature = 37 °C, pH = 7.6–8.7 252 Y. Jiang et al.
(2018)
Municipal food waste + dairy cow slurry 67.8:32.2 (wet 15 L CSTR (continuous) Anaerobic digestate OLR = 5.04 g VS/L, HRT = 17.5 days, 444.7 Morken et al.
weight) temperature = 37 ± 2 °C, pH = 7.64 (2018)
FW + sludge 1:3 (VS basis) Glass bottle (batch) N/A Temperature = 35 ± 2 °C, pH = 5.8–7.5, C/N = 6.29 435.5 Y. Wang et al.
(2018)
Potato waste + cabbage waste 1:1 (VS basis) 5 L CSTR (semi-continuous) Anaerobic granular OLR = 3 kg VS/(m3 day), temperature = 37 ± 1 °C, 360 Mu et al. (2017)
sludge C/N = 20.1
Kitchen waste + cow manure 1:1 (wet weight 1 L lab-scale anaerobic Anaerobic sludge OLR = 8% TS, HRT = 45 days, temperature = 35 °C, 179.8 Zhai et al. (2015)
basis) digester (batch) pH = 7.5.
FW + garden waste + mixed sludge 67.5:22.5:10 (VS 7.5 L CSTR (semi- Anaerobic sludge OLR = 2.55–7.79 g VS/L, HRT = 30–10 days, 424–356 Fitamo et al.

9
basis) continuous) temperature = 55 °C. pH = 7.84–7.79 (2016)
FW + straw 5:1 1 L bottle reactor (batch) Anaerobic granular OLR = 5 g VS/L, HRT = 8 days, temperature = 35 °C, 392 Yong et al. (2015)
sludge C/N = 30.9
FW + green waste 40:60 (VS basis) 500 mL glass bottle (batch) Anaerobic sludge OLR = 5 g VS feedstock, HRT = 24.5 days, 272.1 Chen et al. (2014)
temperature = 37 ± 1 °C, C/N = 15.8
Organic fraction of municipal solid waste + fruit and 1:3 (VS basis) 2 L glass reactor (batch) Anaerobic sludge Temperature = 35 °C, pH = 7.4–8.2, C/N = 34.7, 396.6 Pavi et al. (2017)
vegetable waste (FVW) alkalinity = 860–986 mg/L HCO3−
FW + cattle manure N/A 1 L glass digesters (batch) Anaerobic activated OLR = 8 g VS/L, temperature = 35 ± 1 °C, C/N = 15.8, 388 Zhang et al. (2013)
sludge
FVW + slaughterhouse waste + manure 11:8:7 (% wet 2 L stainless steel digesters Anaerobic slurry OLR = 1.3 kg VS/(m3 day), HRT = 30 days, 320 Alvarez and Lidén
weight) (semi-continuous) temperature = 35 ± 1 °C, pH = 7.4 (2008)
Food waste + wheat straw 9:1 (VS basis) 6 L CSTR (continuous) Mesophilic anaerobic OLR = 3 kg VS/(m3 day), temperature = 35 ± 1 °C, 344 Shi et al. (2018)
sludge pH = 7.1–7.5
Food waste + wheat straw 5:5 (VS basis) 6 L CSTR (continuous) Thermophilic anaerobic OLR = 3 kg VS/(m3 day), temperature = 55 ± 1 °C, 370 Shi et al. (2018)
sludge pH = 7.1–7.5
FW + waste activated sludge 3:1 3 L CSTR (continuous) Anaerobic seed sludge OLR = 7.75 g VS/L/day, HRT = 10 days, 350 Q. Li et al. (2017)
temperature = 35 °C, pH = 7.64
FW + waste activated sludge 3:1 3 L CSTR (continuous) Anaerobic seed sludge OLR = 5.19 g VS/L/day, HRT = 15 days, 407 Q. Li et al. (2017)
temperature = 55 °C, pH = 7.86.
Bioresource Technology Reports 8 (2019) 100310
Table 4
Comparison of pretreatment methods to enhance AD of food waste.
Pretreatment conditions Inoculum AD reactor type and condition Methane/biogas yield Effects of pretreatment References

Acetic acid (0.2 M), 62.5 °C, and 30 min Anaerobic sludge 250 mL serum bottle (batch), 53.58 mL CH4/g VSinitial Dilute acetic acid pretreatment improves the surface Saha et al. (2018)
S.K. Pramanik, et al.

pH = 7.0 ± 0.2, 37 °C, 86 days roughness and porosity of pretreated FW for better
bacterial accessibility compared to untreated FW.
Crystallinity index was increased by 56% and
maximum recovery of sugar was 95%. Methane
production enhanced by 10%.
Thermal conditions: 70 min at 55–90 °C and 50 min at Anaerobic seed sludge 250 mL glass bottles, 35 °C, The cumulative biogas production Thermal pretreatment displayed no major impact on Y. Li et al.
120–160 °C 21 days increased linearly. the final content of protein, but it reduced the fat, oil, (2017b)
and grease (FOG) potential by 7–36% and improved
the lag phase for degradation of protein (35–65%) and
FOG (11–82%). The removal efficiency of VS, FOG, and
crude protein enhanced exponentially.
Ultrasound: specific energy applied for sonication was – 18.75 L induced bed reactor 520 L CH4/kg VS Ultrasound pretreatment improves the biodegradability Ormaechea et al.
1040 kJ/kg TS (continuous), 55 ± 1 °C, and the feasibility of biogas plants. COD removal was (2017)
40 days > 90%. Methane production was enhanced by about
70%. They also concluded that lower specific energy
was more efficiencies that high specific energy.
Enzyme conditions: Lipase-I, Lipase-II, and Lipase-III Anaerobic seed sludge Scrum bottle (batch), Maximum biomethane production was Enzyme pretreatment could improve hydrolysis rate Meng et al.
were hydrolyzed in the conditions of 0–36 h, 35 ± 1 °C, 60 days 189 mL for AF + Lipase II and biomethane production rate. Methane production (2017)
50–2000 μL and 30–50 °C. rate was increased by 80.8–157.7%, 26.9–53.8%, and
37.0–40.7% for AF, VO, and FG, respectively. Lipase-I
and Lipase-II could display the optimal hydrolysis
performance at 24 h, 1000–1500 μL and 40–50 °C.
Ultrasonic (20 kHz and amplitude of 80 μm). Sonication Anaerobic digestate 1000 mL glass reactor (batch), Maximum methane yield = 237 mL/g VSin Methane yield improved by > 80%. TS and VS Zeynali et al.
times was 9, 18, 27 min and specific energy was 35 °C, 25 days (sonication time = 18 min) reduction was increased compared to untreated FVW. (2017)

10
1175, 2380, 3560 kJ/kg TS Longer sonication time consumes more energy than
shorter sonication time.
Alkali concentration was 40–190 mEq/L, and time was Sewage sludge 2 L glass AD vessel (batch), 864.2 mL CH4/g VSdestructed Methane production increased up to 20% compared to Junoh et al.
1–6 h mesophilic temperature, without pretreatment. COD solubilization was (2016)
pH = 7.0, 35 days optimised at 166.98 mEq/L Ca(OH)2.
Commercial enzymes (carbohydrate, protein, and lipid Granular sludge 2.7 L UASB reactor, 35 ± 2 °C, 0.35 L-CH4/g-SCOD Combination of enzymatic hydrolysis of FW and Moon and Song
were used at a rate of 1:2:1, pH 4.5, 50 °C and 75 days methane fermentation increases the hydrolysis rate by (2011)
150 rpm for 24 h) reducing HRT, resulting in high methane production.
SCOD removal efficiency was > 95% and high
methane content (67–75%).
Ultrasound (specific energy applied for sonication was Mesophilic and 5 L CSTR, 36 °C and 55 °C Maximum methane production was Ultrasound pretreatment not only improves energy Quiroga et al.
7500 kJ/kg TS) thermophilic 0.85 L CH4/L day at 36 °C and 0.82 CH4/ production but also allows functioning at lower (2014)
digestates L day at 55 °C retention times. Methane production increase of up to
31% and 67% for mesophilic and thermophilic
temperature compared to without pretreatment.
Physical (2.5 mm–8.0 mm) Anaerobic sludge 2 L complete-mix anaerobic 570 mL CH4/g VS (for 2.5 mm), Lower food waste particle size improves digestate Agyeman and
digesters (semi-continuous), 515 mL CH4/g VS (for 4.0 mm) and dewaterability and hydrolysis. Methane production Tao (2014)
36 °C, 140 rpm 465 mL CH4/g VS (for 8 mm) enhanced by 10–29%.
Physical (Bead mill) 300 rpm Mesophilic anaerobic 2 L glass reactor (batch), 439 mLbiogas/g-TCOD Bead mill pretreatment successfully increased Izumi et al.
1000 rpm sewage sludge 37 ± 1 °C, 80 rpm, 16 days 503 mLbiogas/g-TCOD solubilization by almost 40% and methane yield by (2010)
1000 rpm 455 mLbiogas/g-TCOD 28%. Excessive size reduction of the substrate caused
4000 rpm 470 mLbiogas/g-TCOD VFAs accumulation, resulting in reduced methane
20,000 rpm 455 mLbiogas/g-TCOD yield.
40,000 rpm 404 mLbiogas/g-TCOD
Thermal (120 °C) 10 min Anaerobic sludge 250 mL glass bottles (batch), 112 mL CH4/mL Anaerobic biodegradability for liquid phase and Li et al. (2016b)
30 min 35 °C 152 mL CH4/mL methane production rate can be improved after thermal
40 min 168 mL CH4/mL pretreatment. This pretreatment also helps to recycle
50 min 161 mL CH4/mL the floating oil.
60 min 129 L CH4/mL
(continued on next page)
Bioresource Technology Reports 8 (2019) 100310
Table 4 (continued)

Pretreatment conditions Inoculum AD reactor type and condition Methane/biogas yield Effects of pretreatment References

Thermal (70 min) 55 °C Anaerobic sludge 5.5-L airtight Plexiglas reactors 939 mLbiogas/g Biogas production from kitchen waste cannot be Li and Jin (2015)
70 °C (batch), 35 °C, 1135 mLbiogas/g increased significantly by the low and high thermal
S.K. Pramanik, et al.

90 °C 1173 mLbiogas/g pretreatment temperature. Large amount of methane


Thermal (50 min) 120 °C 1200 mLbiogas/g was generated and large amount of VS, crude fats and
140 °C 885 mLbiogas/g proteins were degraded at 90 and 120 °C, with
160 °C 909 mLbiogas/g treatment durations of 70 min and 50 min.
Thermal conditions: 70 °C, time, 2 h Anaerobic microbial Hybrid anaerobic solid-liquid 0.29 dm3 CH4/g VSremoved Thermal pretreatment of food waste increased the Wang et al.
sludge system (batch), 35 °C, 14 days 0.30 dm3 CH4/g VSremoved concentration of methanogens, which helps to increase (2006)
Thermal = 150 °C, 1 h methane production. It also decreases the VFA
production and enhances the VS & COD removal.
Alkali = 0.4 N NaOH, pH = 12.7, 1 h Anaerobic seed sludge 120 mL serum bottle (batch), 339.2 mL CH4/g VSSremoved Methane production after alkali, thermal, and Naran et al.
Thermal = 120 °C for 30 min 35 ± 1 °C, 20 days at 80 rpm 480.8 mL CH4/g VSSremoved ultrasonic pretreatment increased by 25%, 77%, and (2016)
Ultrasonic (energy intensity of 360 kJ/L, 30 min) 423.6 mL CH4/g VSSremoved 56%, respectively. The solubilization of VS and COD
can be increased by alkali, thermal, and ultrasonic
pretreatment that consequently enhanced the methane
production.
pH 2 (10 N HCl), 24 h Thermophilic 1.2 L Erlenmeyer reactor 0.16 Lbiogas/g CODremoval Biogas production reduced by 54% because of the Ma et al. (2011)
Thermal condition: 120 °C (1 bar) anaerobic sludge (batch), 55 ± 2 °C, 40 days 0.36 Lbiogas/g CODremoval formation of inhibitors in acid pretreatment, whereas
Pressure: pressurized until 10 bar and depressurized 0.52 Lbiogas/g CODremoval biogas production in thermal and pressure-depressure
with CO2, few minutes pretreatment increased by 3% and 49%, respectively.
COD solubilization in acid, thermal, and pressure-
depressure method increased by 13 ± 7, 19 ± 3, and
12 ± 7%, respectively.
Commercial enzymes: 10 U/g dry FW for glucoamylase, Anaerobic sludge Batch, 35 °C and 150 rpm. 457.3 mL CH4/g VS Enzymatic pretreatment with fungal mash was Kiran et al.
at 60 °C, 100 rpm, 24 h displayed to be more effective compared to commercial (2015)
Fungal mash: 10 U/g dry FW for glucoamylase, at 60 °C, 468.2 mL CH4/g VS enzymes. Methane yield and production rate with

11
100 rpm for 24 h fungal mash were almost 2.3 and 3.5-times higher than
control, VSS removal was 80.4 ± 5%.
Bioresource Technology Reports 8 (2019) 100310
S.K. Pramanik, et al. Bioresource Technology Reports 8 (2019) 100310

substances (Ma et al., 2018). Ultrasound pretreatment has been widely enzymes is favorable in terms of energy economy compared to the use
implemented in full-scale anaerobic digesters. Sonix and Sonolyzer, of a commercial enzyme.
Biosonator, MaXonics, and Hielscher are the five most commonly ap-
plied commercially available systems (Panigrahi and Dubey, 2019; 7. Challenges and future perspectives
Tyagi et al., 2014). The full-scale studies of ultrasonic pre-treatment
proposed that net energy benefit up to seven times greater than the The AD of FW still faces numerous challenges, although this process
energy applied (Tyagi et al., 2014). Rasapoor et al. (2019) studied the is already a comparatively advanced and broadly used technology for
energy performance of an ultrasonic pre-treatment of food waste in a treating animal manure, sewage sludge, and wastewater. It has been
pilot-scale digester. They found that biogas production improved up to found that low biogas production could occur due to the deficiency of
59% and 80% for 500 and 1500 g VS/m3 day organic loading rate, re- short process control and optimisation. Toxic intermediate composites
spectively compared to the untreated food waste. Energy analysis dis- can be formed at the time of digestion, which is one of the main
played a better energy gain and energy benefit was achieved due to the technical challenges. Other challenges include VFA accumulation, high
ultrasonic pre-treatment. Thermal pre-treatment aims to improve financial cost (operation and transportation), low buffer capacity, and
anaerobic dewatering and digestibility features. Ariunbaatar et al. process instability (Xu et al., 2018). These are due to the pH drop at an
(2014) reported that thermal pre-treatment is a process of producing initial phase of the digestion and inhibition of ammonia, long-chain
heat, which disrupts the cell structure. This process offers better hy- fatty acids, and hydrogen sulphide. The deficiency of appropriate re-
drolysis of sugars and increases the pH of the slurry, as well as reduces actor design is another important challenge, where the design of the
the toxicity effects, pathogens loads, and viscosity of the FW slurry reactor is still commonly carried out through rule of thumb (Appels
(Karthikeyan et al., 2018). et al., 2011). Overall, the FW characterisation, biodegradability, de-
Chemical pre-treatment primarily depends on strong to mild che- velopment of bacterial activities, accessibility, improvement in CH4
mical agents to change the physico-chemical and biological properties production, and nutritional balance are considered the key challenges
of FW (Karthikeyan et al., 2018). Alkaline pre-treatment was found for the efficient production of biogas (Hagos et al., 2017).
more suitable for the elimination of lignin, whereas acid pre-treatment
was reported to be efficient in solubilizing hemicellulose (Rodriguez 7.1. Future perspectives
et al., 2017). Sodium hydroxide (NaOH), potassium hydroxide (KOH),
aqueous ammonia (NH3·H2O), calcium hydroxide (Ca(OH)2), and am- i. Municipal solid waste is one of the major sources of food waste and
monia hydroxide (NH4OH) are the most common alkalis, which have municipal solid waste is a heterogeneous mixture of different types
been used to increase the performance of the AD process (Ma et al., of organic and inorganic waste. Several problems can be occurred
2018; Rodriguez et al., 2017; Zheng et al., 2014). Inhibitory effects of due to the presence of undesired materials in the reactor, which
accompanying cations present in some reagents (e.g., potassium, so- could lead to the shutdown of the reactor. This problem can be
dium, magnesium, and calcium) cannot be neglected. These accom- solved through the separation of food waste from municipal solid
panying cations inhibit the activity of some methanogenic micro- waste. The separation of food waste from municipal solid waste is
organisms, since the chemicals are added mostly as salts or hydroxides not practiced in many countries. In the future, proper steps should
of these cations (Ariunbaatar et al., 2014). They reported that 0.2 g/L be taken for the separation of food waste from municipal solid
and 0.72 g/L are the optimum concentrations of calcium and magne- waste, as it helps to improve renewable energy production along
sium, respectively. Whereas, the maximum tolerable limit of potassium with the municipal solid waste management.
and sodium is up to 8 g/L and < 5 g/L, respectively for anaerobic mi- ii. Anaerobic digestion is a biological process and therefore, food waste
croorganisms. is continuously broken down by anaerobic microorganisms and
Acid pre-treatment can be performed either by using a concentrated transformed into biogas. Every phase of anaerobic digestion is
acid or a dilute acid (Zheng et al., 2014). Generally, high-concentration carried out by dissimilar groups of microorganisms and each group
acid pre-treatment is carried out at low temperatures, whereas low- of microorganisms cannot survive without their individual re-
concentration acid pre-treatment is conducted at high temperatures. spective optimum environment. The correlation between the dy-
Both organic and inorganic acid such as hydrochloric acid, sulphuric namic behaviour of the bacterial communities and food waste
acid, phosphoric acid, acetic acid, nitric acid, and hydrogen peroxide composition has not been completely studied. Further investigation
are generally used for acid pre-treatment to enhance the AD perfor- is necessary to determine the impacts of numerous aspects on the
mance of FW (Ma et al., 2018; Rodriguez et al., 2017; Zheng et al., micro-ecology, bacterial development features, and structure and
2014). Strong acid pretreatment is commonly avoided, since it may dynamic behaviour of the numerous bacterial groups. This needs to
result in the formation of inhibitory by-products, like hydroxyl methyl be investigated cautiously at different steps of the anaerobic di-
furfural (HMF) and furfural (Ariunbaatar et al., 2014; Mao et al., 2015). gestion process during biogas production from food waste under
Dilute acid is more favored, since it has some advantage including less various circumstances (e.g., high salt, high oil and fat).
corrosive, less expensive, and less toxic (Panigrahi and Dubey, 2019). iii. The anaerobic co-digestion process is extensively practiced nowa-
They pointed out that dilute acid needs less quantity of neutralizing days to avoid the problems associated with the mono-digestion of
reagents and it doesn't need expensive material for digester construc- food waste and co-digestion process helps to increase biogas pro-
tion. duction compared to mono-digestion. Most wastewater treatment
Biological pre-treatment is a generally slow process that requires plants using anaerobic digestion for the treatment of sewage sludge,
longer retention time and the microbes utilise the free and readily and food waste can be used as co-substrate with sludge to these
available sugars as main carbon source throughout the pre-treatment existing plants in the future. This can help not only to recover high
period (Karthikeyan et al., 2018). Commercial enzymes are expensive, energy but also to eradicate the high capital cost for constructing
and have been used to increase the hydrolysis of starch in FW new biogas plants. Therefore, further studies should be performed
(Karthikeyan et al., 2018; Kiran et al., 2015). However, the pre-treat- in order to determine the optimal operating factors for the devel-
ment of FW with a single commercial enzyme seems to be less efficient opment of the large-scale co-digestion system.
than when multiple commercial enzymes are used (Kiran et al., 2015). iv. The performance of anaerobic digestion and the degradation rate of
To increase the economical aspect of the enzymatic hydrolysis of FW, food waste could be enhanced using different pre-treatment
different types of FWs have been used to create enzymes such as cel- methods. More research is therefore needed not only for process
lulases, proteases, lipases, amylases, and pectinases mainly via solid- optimisation but also for the suitable characterisation of the food
state fermentation. Karthikeyan et al. (2018) noted that the use of crude waste structure. This includes reducing the production of inhibitory

12
S.K. Pramanik, et al. Bioresource Technology Reports 8 (2019) 100310

or less biodegradable compounds for chemical pre-treatment, in- Appels, L., Lauwers, J., Degrve, J., Helsen, L., Lievens, B., Willems, K., Van Impe, J.,
creasing the hydrolysis rate for biological pre-treatment, and Dewil, R., 2011. Anaerobic digestion in global bio-energy production: potential and
research challenges. Renew. Sust. Energ. Rev. 15, 4295–4301.
minimising the higher energy consumption for thermal and ultra- Ariunbaatar, J., 2014a. Methods to enhance anaerobic digestion of food waste. Agric. Sci.
sonic pre-treatments. It is noticed that most of the studies on pre- Univ. Paris-Est. 123, 143–156.
treatment are confined to laboratory scale and results obtained from Ariunbaatar, J., Panico, A., Esposito, G., Pirozzi, F., Lens, P.N.L., 2014. Pretreatment
methods to enhance anaerobic digestion of organic solid waste. Appl. Energy 123,
those studies may not be suitable for pilot scale systems. In the long- 143–156.
term development perspective, further researches should be con- Banks, C.J., Chesshire, M., Stringfellow, A., 2008. A pilot-scale comparison of mesophilic
ducted from laboratory to pilot scale systems in order to maximize and thermophilic digestion of source segregated domestic food waste. Water Sci.
Technol. 58, 1475–1481.
the accuracy of results in energy and mass balance calculations. Baroutian, S., Munir, M.T., Sun, J., Eshtiaghi, N., Young, B.R., 2018. Rheological char-
v. Food waste-based bio-refinery concept needs to be developed soon, acterisation of biologically treated and non-treated putrescible food waste. Waste
as the anaerobic digestion of food waste faces several problems Manag. 71, 494–501.
Bong, C.P.C., Lim, L.Y., Lee, C.T., Klemeš, J.J., Ho, C.S., Shin, H.W., 2018. The char-
associated with the minimum cost of the final product and max-
acterisation and treatment of food waste for improvement of biogas production
imum expense of the feedstock transportation, reactor structure, during anaerobic digestion – a review. J. Clean. Prod. 172, 1545–1558.
and performance. Food waste-based bio-refinery may produce en- Browne, J.D., Murphy, J.D., 2013. Assessment of the resource associated with biomethane
zymatic bioremediation, a variety of high-value products (e.g., from food waste. Appl. Energy 104, 170–177.
Chen, X., Yan, W., Sheng, K., Sanati, M., 2014. Comparison of high-solids to liquid
bioplastics, bio-oil, lactic acid, wax, biosurfactants or bio-lu- anaerobic co-digestion of food waste and green waste. Bioresour. Technol. 154,
bricants), nanoparticles and bioenergy (e.g., biodiesel, bioethanol). 215–221.
Currently, the food waste-based bio-refinery is still in the initial Chen, S., Zhang, J., Wang, X., 2015. Effects of alkalinity sources on the stability of
anaerobic digestion from food waste.
phase of theoretical research. In the near future, the food waste- Chen, Y., Xiao, K., Jiang, X., Shen, N., Zeng, R.J., Zhou, Y., 2018. Long solid retention
based bio-refinery concept is expected to play an important part in time (SRT) has minor role in promoting methane production in a 65 °C single-stage
material, chemical, and energy production. anaerobic sludge digester. Bioresour. Technol. 247, 724–729.
Clercq, D.D., Wen, Z., Gottfried, O., Schmidt, F., Fei, F., 2017. A review of global stra-
vi. Biogas comprises of CH4, CO2, various gases (e.g., H2S), impurities tegies promoting the conversion of food waste to bioenergy via anaerobic digestion.
(e.g., siloxanes), and portions of water vapour. Biogas is a pure and Renew. Sust. Energ. Rev. 79, 204–221.
eco-friendly renewable energy source, which can replace fossil fuels Deepanraj, B., Sivasubramanian, V., Jayaraj, S., 2014. Biogas generation through anae-
robic digestion process-an overview. Res. J. Chem.Environ 18, 80–93.
in electricity production through combined heat and power. Biogas Deepanraj, B., Sivasubramanian, V., Jayaraj, S., 2017. Effect of substrate pretreatment on
technology needs to be upgraded not only for usage as a vehicle fuel biogas production through anaerobic digestion of food waste. Int. J. Hydrog. Energy
and as a fuel cell but also for injection into natural gas grids. 42, 26522–26528.
Dobre, P., Nicolae, F., Matei, F., 2014. Main factors affecting biogas production - an
Pressure swing adsorption, water scrubbing, amine scrubbing,
overview. Rom. Biotechnol. Lett. 19, 9283–9296.
cryogenic and membrane separation are the biogas upgrading Edwiges, T., Frare, L., Mayer, B., Lins, L., Mi Triolo, J., Flotats, X., de Mendonça Costa,
technologies which are available in the commercial level. Biological M.S.S., 2018. Influence of chemical composition on biochemical methane potential of
and hybrid methods-based technologies are still under develop- fruit and vegetable waste. Waste Manag. 71, 618–625.
Fernández-Rodríguez, J., Pérez, M., Romero, L.I., 2014. Dry thermophilic anaerobic di-
ment. Further research should be focused on the economic aspect gestion of the organic fraction of municipal solid wastes: solid retention time opti-
and technological advancement in biogas upgrading processes. mization. Chem. Eng. J. 251, 435–440.
Fisgativa, H., Tremier, A., Dabert, P., 2016. Characterizing the variability of food waste
quality: a need for efficient valorisation through anaerobic digestion. Waste Manag.
8. Conclusions 50, 264–274.
Fitamo, T., Boldrin, A., Boe, K., Angelidaki, I., Scheutz, C., 2016. Co-digestion of food and
Anaerobic digestion is one of the most significant technologies in garden waste with mixed sludge from wastewater treatment in continuously stirred
tank reactors. Bioresour. Technol. 206, 245–254.
treating food waste via its transformation into biogas. The two-stage Gerardi, M.H., 2003. The Microbiology of Anaerobic Digesters.
anaerobic digestion process could increase methane production. Gokarn, S., Kuthambalayan, T.S., 2017. Analysis of challenges inhibiting the reduction of
However, the co-digestion of food waste could help to enhance the waste in food supply chain. J. Clean. Prod. 168, 595–604.
Gonzalez-Fernandez, C., Sialve, B., Molinuevo-Salces, B., 2015. Anaerobic digestion of
stability of the anaerobic digestion process along with achieving a
microalgal biomass: challenges, opportunities and research needs. Bioresour.
higher energy production. Food waste hydrolysis can be increased via Technol. 198, 896–906.
numerous pre-treatment technologies, which can in turn improve me- Granado, R.L., Maria, A., Antune, D.S., Valéria, F., 2017. Technology overview of biogas
production in anaerobic digestion plants: a European evaluation of research and
thane production. However, further research needs to be done on the
development. 80, 44–53.
anaerobic digestion of food waste for not only to advance the process Grimberg, S.J., Hilderbrandt, D., Kinnunen, M., Rogers, S., 2015. Anaerobic digestion of
design and optimisation but also to achieve economic cost. food waste through the operation of a mesophilic two-phase pilot scale digester -
assessment of variable loadings on system performance. Bioresour. Technol. 178,
226–229.
Declaration of competing interest Gustavsson, J., Cederberg, C., Sonesson, U., Otterdijk, R. van, Meybeck, A., 2011. Global
food losses and food waste - extent, causes and prevention. Glob. Food Losses Food
No conflict of interest. Waste “Food Agric. Organ. United Nations (2011)” Rome, Italy ((Accessed 23 August
2017)).
Hagos, K., Zong, J., Li, D., Liu, C., Lu, X., 2017. Anaerobic co-digestion process for biogas
Acknowledgments production: progress, challenges and perspectives. Renew. Sust. Energ. Rev. 76,
1485–1496.
Hu, Y., Kobayashi, T., Qi, W., Oshibe, H., Xu, K.Q., 2018. Effect of temperature and or-
This research was funded by the Lembaga Lebuhraya Malaysia (KK- ganic loading rate on siphon-driven self-agitated anaerobic digestion performance for
2018-016). food waste treatment. Waste Manag. 74, 150–157.
Huttunen, S., Rintala, J.A., 2007. Laboratory investigations on co-digestion of energy
crops and crop residues with cow manure for methane production: effect of crop to
References
manure ratio. 51, 591–609.
Izumi, K., Okishio, Y. ki, Nagao, N., Niwa, C., Yamamoto, S., Toda, T., 2010. Effects of
Abdulkarim, B.I., Evuti, A.M., 2010. Effect of buffer (NaHCO3) and waste type in high particle size on anaerobic digestion of food waste. Int. Biodeterior. Biodegrad. 64,
solid thermophilic anaerobic digestion 2. pp. 980–984. 601–608.
Agyeman, F.O., Tao, W., 2014. Anaerobic co-digestion of food waste and dairy manure: Jiang, J., Li, L., Cui, M., Zhang, F., Liu, Y., Liu, Y., Long, J., Guo, Y., 2018a. Anaerobic
effects offood waste particle size and organic loading rate. J. Environ. Manag. 133, digestion of kitchen waste: the effects of source, concentration, and temperature.
268–274. Biochem. Eng. J. 135, 91–97.
Alvarez, R., Lidén, G., 2008. Semi-continuous co-digestion of solid slaughterhouse waste, Jiang, Y., Dennehy, C., Lawlor, P.G., Hu, Z., Zhan, X., Gardiner, G.E., 2018b. Inactivation
manure, and fruit and vegetable waste. Renew. Energy 33, 726–734. of enteric indicator bacteria and system stability during dry co-digestion of food
Appels, L., Baeyens, J., Degrève, J., Dewil, R., 2008. Principles and potential of the waste and pig manure. Sci. Total Environ. 612, 293–302.
anaerobic digestion of waste-activated sludge. Prog. Energy Combust. Sci. 34, Jo, Y., Kim, J., Hwang, K., Lee, C., 2018. A comparative study of single- and two-phase
755–781. anaerobic digestion of food waste under uncontrolled pH conditions. Waste Manag.

13
S.K. Pramanik, et al. Bioresource Technology Reports 8 (2019) 100310

78, 509–520. digestion of food waste. Bioresour. Technol. 118, 210–218.


Junoh, H., Yip, C., Kumaran, P., 2016. Effect on Ca(OH)2 pretreatment to enhance biogas Naran, E., Toor, U.A., Kim, D.J., 2016. Effect of pretreatment and anaerobic co-digestion
production of organic food waste. IOP Conf. Ser. Earth Environ. Sci. 32, 012013. of food waste and waste activated sludge on stabilization and methane production.
Karthikeyan, O.P., Trably, E., Mehariya, S., Bernet, N., Wong, J.W.C., Carrere, H., 2018. Int. Biodeterior. Biodegrad. 113, 17–21.
Pretreatment of food waste for methane and hydrogen recovery: a review. Bioresour. Naroznova, I., Møller, J., Scheutz, C., 2016. Characterisation of the biochemical methane
Technol. 249, 1025–1039. potential (BMP) of individual material fractions in Danish source-separated organic
Khalid, A., Arshad, M., Anjum, M., Mahmood, T., Dawson, L., 2011. The anaerobic di- household waste. Waste Manag. 50, 39–48.
gestion of solid organic waste. Waste Manag. 31, 1737–1744. Nghiem, L.D., Koch, K., Bolzonella, D., Drewes, J.E., 2017. Full scale co-digestion of
Kothari, R., Pandey, A.K., Kumar, S., Tyagi, V.V., Tyagi, S.K., 2014. Different aspects of wastewater sludge and food waste: bottlenecks and possibilities. Renew. Sust. Energ.
dry anaerobic digestion for bio-energy: an overview. Renew. Sust. Energ. Rev. 39, Rev. 72, 354–362.
174–195. Ormaechea, P., Castrillón, L., Marañón, E., Fernández-Nava, Y., Negral, L., Megido, L.,
Krishna, D., Kalamdhad, A.S., 2014. Pre-treatment and anaerobic digestion of food waste 2017. Influence of the ultrasound pretreatment on anaerobic digestion of cattle
for high rate methane production – a review. J. Environ. Chem. Eng. 2, 1821–1830. manure, food waste and crude glycerine. Environ. Technol. (United Kingdom) 38,
Leung, D.Y.C., Wang, J., 2016. An overview on biogas generation from anaerobic di- 682–686.
gestion of food waste. Int. J. Green Energy 13, 119–131. Ostrem, K., 2004. Greening waste: anaerobic digestion for treating the organic fraction of
Li, Y., Jin, Y., 2015. Effects of thermal pretreatment on acidification phase during two- municipal solid waste. Dep. Earth Environ. Eng. Fu Found. Sch. Eng. Appl. Sci.
phase batch anaerobic digestion of kitchen waste. Renew. Energy 77, 550–557. Columbia Univ. 1–59.
Li, L., He, Q., Ma, Y., Wang, X., Peng, X., 2015. Dynamics of microbial community in a Panigrahi, S., Dubey, B.K., 2019. A critical review on operating parameters and strategies
mesophilic anaerobic digester treating food waste: relationship between community to improve the biogas yield from anaerobic digestion of organic fraction of municipal
structure and process stability. Bioresour. Technol. 189, 113–120. solid waste. Renew. Energy 143, 779–797.
Li, Y., Jin, Y., Li, J., Li, H., Yu, Z., 2016a. Effects of thermal pretreatment on the bio- Papargyropoulou, E., Lozano, R., Steinberger, K., Wright, J., Ujang, N., Bin, Z., 2014. The
methane yield and hydrolysis rate of kitchen waste. Appl. Energy 172, 47–58. food waste hierarchy as a framework for the management of food surplus and food
Li, Y., Jin, Y., Li, J., 2016b. Enhanced split-phase resource utilization of kitchen waste by waste. J. Clean. Prod. 76, 106–115.
thermal pre-treatment. Energy 98, 155–167. Park, J.H., Kumar, G., Yun, Y.M., Kwon, J.C., Kim, S.H., 2018. Effect of feeding mode and
Li, Y., Jin, Y., Borrion, A., Li, H., Li, J., 2017a. Effects of organic composition on meso- dilution on the performance and microbial community population in anaerobic di-
philic anaerobic digestion of food waste. Bioresour. Technol. 244, 213–224. gestion of food waste. Bioresour. Technol. 248, 134–140.
Li, Q., Li, H., Wang, G., Wang, X., 2017b. Effects of loading rate and temperature on Pavi, S., Kramer, L.E., Gomes, L.P., Miranda, L.A.S., 2017. Biogas production from co-
anaerobic co-digestion of food waste and waste activated sludge in a high frequency digestion of organic fraction of municipal solid waste and fruit and vegetable waste.
feeding system, looking in particular at stability and efficiency. Bioresour. Technol. Bioresour. Technol. 228, 362–367.
237, 231–239. Quiroga, G., Castrillón, L., Fernández-Nava, Y., Marañón, E., Negral, L., Rodríguez-
Li, Y., Jin, Y., Li, J., Li, H., Yu, Z., Nie, Y., 2017c. Effects of thermal pretreatment on Iglesias, J., Ormaechea, P., 2014. Effect of ultrasound pre-treatment in the anaerobic
degradation kinetics of organics during kitchen waste anaerobic digestion. Energy co-digestion of cattle manure with food waste and sludge. Bioresour. Technol. 154,
118, 377–386. 74–79.
Li, Y., Jin, Y., Li, H., Borrion, A., Yu, Z., Li, J., 2018. Kinetic studies on organic de- Rasapoor, M., Adl, M., Baroutian, S., Iranshahi, Z., Pazouki, M., 2019. Energy perfor-
gradation and its impacts on improving methane production during anaerobic di- mance evaluation of ultrasonic pretreatment of organic solid waste in a pilot-scale
gestion of food waste. Appl. Energy 213, 136–147. digester. Ultrason. - Sonochemistry 51, 517–525.
Liu, C., Wang, W., Anwar, N., Ma, Z., Liu, G., Zhang, R., 2017. Effect of organic loading Ren, Y., Yu, M., Wu, C., Wang, Q., Gao, M., Huang, Q., Liu, Y., 2018. A comprehensive
rate on anaerobic digestion of food waste under mesophilic and thermophilic con- review on food waste anaerobic digestion: research updates and tendencies.
ditions. Energy and Fuels 31, 2976–2984. Bioresour. Technol. 247, 1069–1076.
Liu, X., Lendormi, T., Lanoisell, J., 2019. Overview of hygienization pretreatment for Rodriguez, C., Alaswad, A., Benyounis, K.Y., Olabi, A.G., 2017. Pretreatment techniques
pasteurization and methane potential enhancement of biowaste: challenges, state of used in biogas production from grass. Renew. Sust. Energ. Rev. 68, 1193–1204.
the art and alternative technologies. 236. Saha, S., Jeon, B.H., Kurade, M.B., Jadhav, S.B., Chatterjee, P.K., Chang, S.W.,
Ma, J., Duong, T.H., Smits, M., Verstraete, W., Carballa, M., 2011. Enhanced biometha- Govindwar, S.P., Kim, S.J., 2018. Optimization of dilute acetic acid pretreatment of
nation of kitchen waste by different pre-treatments. Bioresour. Technol. 102, for mixed fruit waste increased methane production. J. Clean. Prod. 190, 411–421.
592–599. Shi, X., Guo, X., Zuo, J., Wang, Y., Zhang, M., 2018. A comparative study of thermophilic
Ma, Y., Liu, Y., 2019. Turning food waste to energy and resources towards a great en- and mesophilic anaerobic co-digestion of food waste and wheat straw: process sta-
vironmental and economic sustainability: an innovative integrated biological ap- bility and microbial community structure shifts. Waste Manag. 75, 261–269.
proach. Biotechnol. Adv (Article in press). Slorach, P.C., Jeswani, H.K., Cuéllar-franca, R., Azapagic, A., 2019. Environmental sus-
Ma, C., Liu, J., Ye, M., Zou, L., Qian, G., Li, Y.Y., 2018. Towards utmost bioenergy con- tainability of anaerobic digestion of household food waste. J. Environ. Manag. 236,
version efficiency of food waste: pretreatment, co-digestion, and reactor type. Renew. 798–814.
Sust. Energ. Rev. 90, 700–709. Tyagi, V.K., Lo, S., Appels, L., Dewil, R.A.F., 2014. Ultrasonic treatment of waste sludge: a
Magnet, 2018. Food waste around the world. https://www.magnet.co.uk/advice- review on mechanisms and applications 1220–1288.
inspiration/blog/2018/February/food-waste-around-the-world/. Tyagi, V.K., Fdez-Güelfo, L.A., Zhou, Y., Álvarez-Gallego, C.J., Garcia, L.I.R., Ng, W.J.,
Mao, C., Feng, Y., Wang, X., Ren, G., 2015. Review on research achievements of biogas 2018. Anaerobic co-digestion of organic fraction of municipal solid waste (OFMSW):
from anaerobic digestion. Renew. Sust. Energ. Rev. 45, 540–555. progress and challenges. Renew. Sust. Energ. Rev. 93, 380–399.
Mata-Alvarez, J., Macé, S., Llabrés, P., 2000. Anaerobic digestion of organic solid wastes. Kiran, E.U., Trzcinski, A.P., Liu, Y., 2015. Enhancing the hydrolysis and methane pro-
An overview of research achievements and perspectives. Bioresour. Technol. 74, duction potential of mixed food waste by an effective enzymatic pretreatment.
3–16. Bioresour. Technol. 183, 47–52.
Mehariya, S., Patel, A.K., Obulisamy, P.K., Punniyakotti, E., Wong, J.W.C., 2018. Co- Vasco-correa, J., Khanal, S., Manandhar, A., Shah, A., 2018. Anaerobic digestion for
digestion of food waste and sewage sludge for methane production: current status and bioenergy production: global status, environmental and techno-economic implica-
perspective. Bioresour. Technol. 265, 519–531. tions, and government policies. Bioresour. Technol. 247, 1015–1026.
Meng, Y., Li, S., Yuan, H., Zou, D., Liu, Y., Zhu, B., Chufo, A., Jaffar, M., Li, X., 2015. Wang, J.Y., Liu, X.Y., Kao, J.C.M., Stabnikova, O., 2006. Digestion of pre-treated food
Evaluating biomethane production from anaerobic mono- and co-digestion of food waste in a hybrid anaerobic solid-liquid (HASL) system. J. Chem. Technol.
waste and floatable oil (FO) skimmed from food waste. Bioresour. Technol. 185, Biotechnol. 81, 345–351.
7–13. Wang, L., Shen, F., Yuan, H., Zou, D., Liu, Y., Zhu, B., Li, X., 2014. Anaerobic co-digestion
Meng, Y., Luan, F., Yuan, H., Chen, X., Li, X., 2017. Enhancing anaerobic digestion per- of kitchen waste and fruit/vegetable waste: lab-scale and pilot-scale studies. Waste
formance of crude lipid in food waste by enzymatic pretreatment. Bioresour. Technol. Manag. 34, 2627–2633.
224, 48–55. Wang, P., Wang, H., Qiu, Y., Ren, L., Jiang, B., 2018a. Microbial characteristics in
Micolucci, F., Gottardo, M., Pavan, P., Cavinato, C., Bolzonella, D., 2018. Pilot scale anaerobic digestion process of food waste for methane production–a review.
comparison of single and double-stage thermophilic anaerobic digestion of food Bioresour. Technol. 248, 29–36.
waste. J. Clean. Prod. 171, 1376–1385. Wang, T., Yang, Pupu, Zhang, X., Zhou, Q., Yang, Q., Xu, B., Yang, Pinghua, Zhou, T.,
Monnet, F., 2003. An introduction to anaerobic digestion of organic wastes. Remade 2018b. Effects of mixing ratio on dewaterability of digestate of mesophilic anaerobic
Scotl. Rep. 1–48. co-digestion of food waste and sludge. Waste and Biomass Valorization 9, 87–93.
Monson, K.D., Esteves, S.R., Guwy, A.J., Dinsdale, R.M., 2007. Anaerobic Digestion of Wang, Y., Li, G., Chi, M., Sun, Y., Zhang, J., Jiang, S., Cui, Z., 2018c. Effects of co-di-
Biodegradable Municipal Wastes: A Review. gestion of cucumber residues to corn stover and pig manure ratio on methane pro-
Moon, H.C., Song, I.S., 2011. Enzymatic hydrolysis of foodwaste and methane production duction in solid state anaerobic digestion. Bioresour. Technol. 250, 328–336.
using UASB bioreactor. Int. J. Green Energy 8, 361–371. Wu, L.J., Kobayashi, T., Li, Y.Y., Xu, K.Q., 2015. Comparison of single-stage and tem-
Morken, J., Gjetmundsen, M., Fjørtoft, K., 2018. Determination of kinetic constants from perature-phased two-stage anaerobic digestion of oily food waste. Energy Convers.
the co-digestion of dairy cow slurry and municipal food waste at increasing organic Manag. 106, 1174–1182.
loading rates. Renew. Energy 117, 46–51. Xiao, B., Qin, Y., Wu, J., Chen, H., Yu, P., Liu, J., Li, Y.Y., 2018. Comparison of single-
Mu, H., Zhao, C., Zhao, Y., Li, Y., Hua, D., Zhang, X., Xu, H., 2017. Enhanced methane stage and two-stage thermophilic anaerobic digestion of food waste: performance,
production by semi-continuous mesophilic co-digestion of potato waste and cabbage energy balance and reaction process. Energy Convers. Manag. 156, 215–223.
waste: performance and microbial characteristics analysis. Bioresour. Technol. 236, Xiao, B., Zhang, W., Yi, H., Qin, Y., Wu, J., Liu, J., Li, Y., 2019. Biogas production by two-
68–76. stage thermophilic anaerobic co-digestion of food waste and paper waste: effect of
Nagao, N., Tajima, N., Kawai, M., Niwa, C., Kurosawa, N., Matsuyama, T., Yusoff, F.M., paper waste ratio. Renew. Energy 132, 1301–1309.
Toda, T., 2012. Maximum organic loading rate for the single-stage wet anaerobic Xu, Z., Zhao, M., Miao, H., Huang, Z., Gao, S., Ruan, W., 2014. In situ volatile fatty acids

14
S.K. Pramanik, et al. Bioresource Technology Reports 8 (2019) 100310

influence biogas generation from kitchen wastes by anaerobic digestion. Bioresour. Zhang, L., Lee, Y.W., Jahng, D., 2011. Anaerobic co-digestion of food waste and piggery
Technol. 163, 186–192. wastewater: focusing on the role of trace elements. Bioresour. Technol. 102,
Xu, F., Li, Y., Ge, X., Yang, L., Li, Y., 2018. Anaerobic digestion of food waste – challenges 5048–5059.
and opportunities. Bioresour. Technol. 247, 1047–1058. Zhang, C., Xiao, G., Peng, L., Su, H., Tan, T., 2013. The anaerobic co-digestion of food
Yadvika, Santosh, Sreekrishnan, T.R., Kohli, S., Rana, V., 2004. Enhancement of biogas waste and cattle manure. Bioresour. Technol. 129, 170–176.
production from solid substrates using different techniques - a review. Bioresour. Zhang, C., Su, H., Baeyens, J., Tan, T., 2014. Reviewing the anaerobic digestion of food
Technol. 95, 1–10. waste for biogas production. Renew. Sust. Energ. Rev. 38, 383–392.
Yi, J., Dong, B., Jin, J., Dai, X., 2014. Effect of increasing total solids contents on anae- Zhang, W., Zhang, L., Li, A., 2015a. Anaerobic co-digestion of food waste with MSW
robic digestion of food waste under mesophilic conditions: performance and micro- incineration plant fresh leachate: process performance and synergistic effects. Chem.
bial characteristics analysis. PLoS One 9. Eng. J. 259, 795–805.
Yong, Z., Dong, Y., Zhang, X., Tan, T., 2015. Anaerobic co-digestion of food waste and Zhang, A., Shen, J., Ni, Y., 2015b. Anaerobic digestion for use in the pulp and paper
straw for biogas production. Renew. Energy 78, 527–530. industry and other sectors: an introductory mini-review. Bioresources 10,
Zeynali, R., Khojastehpour, M., Ebrahimi-Nik, M., 2017. Effect of ultrasonic pre-treatment 8750–8769.
on biogas yield and specific energy in anaerobic digestion of fruit and vegetable Zhang, J., Loh, K.C., Li, W., Lim, J.W., Dai, Y., Tong, Y.W., 2017. Three-stage anaerobic
wholesale market wastes. Sustain. Environ. Res. 27, 259–264. digester for food waste. Appl. Energy 194, 287–295.
Zhai, N., Zhang, T., Yin, D., Yang, G., Wang, X., Ren, G., Feng, Y., 2015. Effect of initial Zheng, Y., Zhao, J., Xu, F., Li, Y., 2014. Pretreatment of lignocellulosic biomass for en-
pH on anaerobic co-digestion of kitchen waste and cow manure. Waste Manag. 38, hanced biogas production. Prog. Energy Combust. Sci. 42, 35–53.
126–131.

15

You might also like