Você está na página 1de 20

Bioresource Technology xxx (2017) xxxxxx

Contents lists available at ScienceDirect

Bioresource Technology
journal homepage: www.elsevier.com/locate/biortech

Review

Anaerobic bioconversion of food waste into energy: A critical review


Camilla M. Braguglia , Agata Gallipoli, Andrea Gianico, Pamela Pagliaccia
Istituto di Ricerca sulle Acque (IRSA-CNR), Area della Ricerca RM1, Via Salaria km 29,300, 00015 Monterotondo, Italy

h i g h l i g h t s

 More than 400 papers in the last 5 years have been published on FW into energy.
 Food waste complexity and composition affects anaerobic conversion.
 Scaling up of the AD process is fundamental to assess the real methane potential.
 Methanogenesis is often rate limiting step leading to pH drop and instability.
 Strategies to improve AD performances and stability are here reviewed.

a r t i c l e i n f o a b s t r a c t

Article history: hhh


Received 28 April 2017 2017 Elsevier Ltd. All rights reserved.
Received in revised form 23 June 2017
Accepted 24 June 2017
Available online xxxx

Keywords:
Food waste
Anaerobic digestion
Fermentation
Pre-treatment
Co-digestion

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2. Food waste characteristics and impact on AD process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3. Anaerobic conversion of FW into methane. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3.1. Batch AD tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3.2. Continuous AD tests for methane production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
4. Anaerobic conversion of food waste into hydrogen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
5. Strategies to improve the anaerobic digestion of food waste . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
5.1. Anaerobic co-digestion with other substrates to improve stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
5.2. Food waste pre-treatments and impact on AD performances . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
6. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00

Corresponding author.
E-mail address: braguglia@irsa.cnr.it (C.M. Braguglia).

http://dx.doi.org/10.1016/j.biortech.2017.06.145
0960-8524/ 2017 Elsevier Ltd. All rights reserved.

Please cite this article in press as: Braguglia, C.M., et al. Anaerobic bioconversion of food waste into energy: A critical review. Bioresour. Technol. (2017),
http://dx.doi.org/10.1016/j.biortech.2017.06.145
2 C.M. Braguglia et al. / Bioresource Technology xxx (2017) xxxxxx

latin america

north africa, west and central asia

USA, Canada and Oceania

japan, china and south korea

europe

0 20 40 60 80 100
composition (%) of food waste at household

Cereals fruits and vegetables meat and fish milk and eggs

Fig. 1. Composition (%) of Food Waste at household in different regions of the world.

1. Introduction a comprehensive sustainable FW management system. In this con-


text, anaerobic digestion (AD) is considered as one of the best
Food waste was defined by the UN Food and Agriculture Orga- environmental-friendly alternatives for the FW management,
nization (FAO) and includes any healthy or edible substance that is because of its limited environmental footprints (Capson-Tojo
wasted, lost, degraded at every stage of the food supply chain. et al., 2016), high potential for energy recovery (Capson-Tojo
Every year, between 1.3 and 1.6 billion tons of food, such as fresh et al., 2016; Kuruti et al., 2017; Zamanzadeh et al., 2016;
vegetables, fruit, and meat, bakery and dairy products, are lost Zhang et al., 2014) producing carrier material for biofertilizers
along the food-supply chain, and this accounts for one third of (Kuruti et al., 2017; Shen et al., 2013). In 2012, 90% of the FW trea-
the food produced globally for human consumption, affecting sev- ted in Europe was processed biologically with both AD and com-
eral natural resources. Food waste, in fact, cost the global economy posting according to the Eurostat waste treatment statistics. The
around USD 990 billion annually, and consumes in fact about a use of AD as a treatment for FWs and other organic wastes has
quarter of all the water used for agriculture purposes, and is increased in Europe with a current reported capacity of almost 9
responsible for an estimated 8% of total anthropogenic global million tons per year in 2015 due to 290 full-scale plants (De
greenhouse gas emission, contributing to biodiversity loss. An Baere and Mattheeuws, 2014) producing methane. Mesophilic
ever-increasing amount of food waste (FW) is generated, owing digesters operate at a temperature between 35 C and 40 C, while
to population growth and rising living standards. Globally, around thermophilic digesters operate between 50 C and 55 C. Mesophi-
2 billion tons of municipal solid waste are formed annually, of lic digestion has always been predominant, because of the small
which 3453% is organic biodegradable waste (defined as energy needs, and greater stability. Nevertheless, thermophilic
OFMSW), mainly food waste which collected from households, digestion has always played an important role in the digestion of
and restaurants, but the composition may vary from country to food waste (Montecchio et al., 2016). In recent years, different
country. The total FW quantity produced each year in Europe has novel reactor designs, such as for example two stage or multiple-
been estimated to be around 90 million tons Mt in 2012, of which stage reactors have been investigated and developed to reach
an estimated 47 Mt is collected from household (average of 92 kg stable conditions under high organic loading rates (OLRs) of food
per capita). The United Kingdom generates most FW in Europe, waste or OFMSW (Dong et al., 2010). However, the AD of organic
namely 14 Mt, corresponding to almost 135 kg per capita yearly waste generally relies on single-stage systems, which account for
against the 62 kg per capita per year wasted in Italy. Annually, in more than 95% of Europes full-scale plants (De Baere and
the United States the amount of food wasted is nearly 61 Mt, in Mattheeuws, 2014).
South Korea around 6.2 Mt and in China, despite the low food Investigating AD processes for FW conversion has become, for
waste per capita (55 kg annually), is 195 Mt. The EU Commission all these reasons, an exciting research field. For this review, the dis-
estimates an increase in the FW amounts to 120 Mt by 2020, which tribution of the 410 peer-reviewed articles extracted from litera-
is mainly due to increased FW generation in households, that ture (Scopus database) in the last 5 years (20122017) containing
almost doubled from 2004 to 2012, while FW from food manufac- the fixed term food waste in the title combined with biogas
turing and agriculture showed decreasing trend. The composition (56), or anaerobic digestion (113), or methane (56) or hydrogen
of the food wasted at household and food service level (restau- (54), or co-digestion (89) or pre-treatment (42) was analyzed also
rants, canteen, etc.) varies from region to region in the world. In on the basis of affiliations country of the authors.
fact, as showed in Fig. 1, in Europe, FW is composed by 40% vegeta- China is unconditionally the most productive country with the
bles and fruit, 33% pasta and bread, 17% of dairy products (includ- largest number of publications in this research field (more than
ing eggs) and 9% of meat and fish residues, while in Asia, in 40% of the papers) followed by South Korea (12%), USA (10%) and
particular in Japan, China and South Korea, FW is composed by a Japan (5%). In Europe, Italy and Spain researchers are the most
56% of vegetables and fruits, 34% of rice and noodle, and only a active in this field.
small fraction (around 10%) due to fish, meat residues, and dairy This review focuses on the anaerobic conversion of FW for
products. energy recovery, in terms of hydrogen and methane production,
Although reduction is the most preferred option in the FW man- reporting the most important results, performances and
agement hierarchy, subsequent approaches such as recovery in developments taken mainly during the last decade as regards dif-
terms of waste-to-energy also require attention and technical ferent reactor configurations (one or multi-stage, dry or wet) and
development from the research community in order to promote effects of inhibition compounds and parameters on AD, in batch

Please cite this article in press as: Braguglia, C.M., et al. Anaerobic bioconversion of food waste into energy: A critical review. Bioresour. Technol. (2017),
http://dx.doi.org/10.1016/j.biortech.2017.06.145
C.M. Braguglia et al. / Bioresource Technology xxx (2017) xxxxxx 3

or continuous mode. Different strategies to improve the anaerobic the amount of methane produced. Acidification through reactor
conversion, including co-digestion and feedstock pre-treatments, overload is one of the most common reasons for process deteriora-
are also reviewed and discussed. tion in anaerobic digesters, and occurs because of the accumulation
of volatile fatty acids fatty acids (VFAs) due to the kinetic uncou-
pling between acid producers and consumers (El-Mashad and
2. Food waste characteristics and impact on AD process Zhang, 2010; Kawai et al., 2014; Nagao et al., 2012; Ventura
et al., 2014). Methanogenesis resulted therefore the rate-limiting
The FW composition clearly affects the physicochemical charac- step, in particular for carbohydrate-rich FW anaerobic digestions,
teristics of FW investigated in different regions in the world. Food opening controversial questions on the effective kinetic limiting
waste consisting of rice, pasta and and vegetables is abundant in stage of complex heterogeneous organic material as food waste.
carbohydrates while food waste consisting of meat, fish and eggs The pH is one of the most important parameters influencing AD,
contains high quantity of proteins and lipids. However, FW pre- because all the involved microorganisms are very sensitive to pH
sents general features that can be extrapolated worldwide, with variations, and each process step shows a different pH sensitivity.
a moisture content of 7490%, high volatile solids fraction around For fermentative bacteria, a comprehensive pH range from 4 to 8.5
85 5%, and a mean acidic pH of 5.1 0.7 (Fisgativa et al., 2016; is suitable, while most methanogens work optimally in a pH range
Zhang et al., 2007. Typical FW is mainly composed of degradable from 6.5 to 7.2, with a methanogenesis step failure below pH of 5.5.
carbohydrates (4162%), proteins (1525%) and lipids (1330%). To ensure microorganisms activity, a good nutritional balance is
On general, FW proved to have varying proportions of nutrients necessary, both in terms of carbon and nitrogen ratio, and of
and micronutrients and low presence of heavy metals, but the vari- micronutrients as Ca, K, Mg, Na, P and Fe, indispensable for
ability was very high (Fisgativa et al., 2016). FW typically has a rel- microorganisms. The feedstock total solids content (as conse-
atively low C/N ratio, varying between 13.2 and 24.50, lower with quently also the volatile solids) affect the AD process of food waste.
respect to the optimal range of 2535 assuring efficient digestion So far, three main types of AD technologies have been developed
conditions (Capson-Tojo et al., 2016; Chen et al., 2008; Zhang according to the total solids (TS) content of feedstocks: conven-
et al., 2007). In fact, an excessively high C/N ratio causes the tional wet (<10% TS), semi-dry (1020% TS) and dry (2040% TS)
increase in acid formation inhibiting methane production, while processes. Dry anaerobic digestion, so called high-solids technol-
at low C/N ratio nitrogen is converted to ammonium at a faster rate ogy, facilitates the dry substrate manipulation with batch and
than it can be assimilated by the methanogens. For biodegradable semi-continuous process, but higher structure cost may be present.
substrates, the optimum C/N ratio is in the range 2025. However, With specific reference to the full-scale industrial application, dur-
for materials that are resistant to microbial degradation, the C/N ing the last 5 years dry digestion accounts for about 70% of the
ratio can be also as high as 40. Due to its fundamental characteris- installed capacity, resulting in a cumulative market share of about
tics such as wide availability, high biodegradable organic fraction 62% (De Baere and Mattheeuws, 2014). The predominance of dry
and in particular high carbohydrate content, FW has been consid- digesters was due to the reduced volume of reactors and wastew-
ered an attractive economical source for energy production, and ater production. In the case of the biological production of hydro-
substrate characteristics influenced largely the AD process perfor- gen by fermentation, no full-scale applications have been realised
mances (Fisgativa et al., 2016; Zhang et al., 2013a). Organic feed- to date (De Gioannis et al., 2013).
stocks undergo different degradation steps during the AD process
(Fisgativa et al., 2016; Patinvoh et al., 2017). Step one is hydrolysis,
where the feedstock is disintegrated by the action of a diverse 3. Anaerobic conversion of FW into methane
community of hydrolytic bacteria producing simple sugars, amino
acids, and fatty acids. This step has been reported as being the rate Extensive effort has been devoted to the production of bio-
limiting step for complex, hard biodegradable, organic substrates methane from FW through anaerobic processes, but its economic
(Ariunbaatar et al., 2014; Carlsson et al., 2012; Gianico et al., viability is largely dependent on the efficiency of FW type, pre-
2013; Izumi et al., 2010; Zhang et al., 2014), and can be accelerated treatment and scale (Ma et al., 2017). Scale effect of anaerobic
by pre-treating the substrate before digestion (see paragraph 5.2). digestion tests in batch and semi-continuous mode for the techni-
Step two is acidogenesis, where monomers from the hydrolysis are cal and economic feasibility of a full-scale digester treating FW is,
fermented into short chain organic acids (acetic, propionic, butyric in our opinion, very important. Therefore, the first part of this
and others), alcohols, together with hydrogen and carbon dioxide. chapter is dedicated to batch systems and the second one is dedi-
The products formed vary with the types of bacteria as well as cated to continuous reactors. The difference is extremely impor-
environmental conditions. Hydrogen, as the most valuable fermen- tant in particular by digesting complex heterogeneous feedstocks
tation product, is widely discussed in paragraph 4. Acidogenesis is as FW, where the organic load (in particular the readily biodegrad-
the fastest step in the AD process, so if the feedstock does not have able one) affects significantly the performances of the process. In
buffering capacity and the organic loading rate is too high, the fact, in batch systems, a reactor is loaded with feed and run to com-
accumulation of volatile fatty acids can result in a pH drop, which pletion until methane production stops. Batch reactors benefit
would inhibit the methanogens that produce methane in the final from technical simplicity, low operating costs, and, last but not
step (Nagao et al., 2012; Zhang et al., 2014). Step three is the ace- least, short digestion times (in 2540 days, the test is completed).
togenesis, here the homoacetogenic microorganisms reduce hydro- On the contrary, in continuous systems, reactors are continuously
gen and carbon dioxide to acetic acid. In this step, the acetogenic fed with the feedstock, allowing a steady-state to be reached in
bacteria can only survive at a very low hydrogen concentration, the reactor with a constant methane yield, after 12 HRTs (hydrau-
so excessive production of hydrogen from the acidogenesis step lic residence time). Although continuous reactors have higher
can inhibit these bacteria. The last step is the methanogenesis, operating costs due to storage and requirements, these reactors
where methane production takes place under strict anaerobic con- are able to maintain and adapt microorganisms within the system,
ditions, by utilizing the intermediate products (as H2 and acetate) thereby avoiding lag times associated with microorganism growth
from the preceding stages. Because of these complex succession of in batch reactors, but methane yields depend strictly on the
biological steps and of the close connection of the degradation organic load (OLR) and HRT applied. In fact, food waste feedstock,
phases, AD of organics is based on a delicate balance that may generally rich of biodegradable organics, can rapidly lead to an
affect the instability of the digesters treating food waste, and thus imbalance in the production-consumption of VFAs, overwhelming

Please cite this article in press as: Braguglia, C.M., et al. Anaerobic bioconversion of food waste into energy: A critical review. Bioresour. Technol. (2017),
http://dx.doi.org/10.1016/j.biortech.2017.06.145
4 C.M. Braguglia et al. / Bioresource Technology xxx (2017) xxxxxx

Table 1
Methane yields from batch digestion of food waste.

Food Waste type Volume of Temperature Duration VS Methane yield (L CH4 Initial F/I CH4 Initial References
batch removal kg 1VSfed) (or load) (%) pH
reactors (%)
Synthetic FW 1L Meso 25 d N.R. 469 6.8 0.5 N.R. N.R. Ariunbaatar
et al. (2015)
Canteen FW 5L Meso 25 d N.R. 467 0.33 N.R. N.R. Browne and
Murphy
(2013)
Canteen FW 0.5 L Meso 25 d N.R. 529 0.33 N.R. N.R. Browne and
Murphy
(2013)
FW 35 L 30 C 60 d 42 314 0.57a 65 5.7 Dong et al.
(2010)
Waste sorted by waste company 1L Meso 30 d 82 353 10 2 gVS/L 54 N.R. El-Mashad and
Zhang (2010)
Source segregated FW N.R. Meso 40 d N.R. 434 40 (inoculum A); 0.30.4 N.R. 7.1 Facchin et al.
338 30 (inoculum B) (2013)
Source segregated FW + trace metal N.R. Meso 40 d N.R. 487 0.30.4 N.R. N.R. Facchin et al.
mix (50% of typical requirement) (2013)
a
Disposer grinded standard Japanese 2L Meso 16 d N.R. 417 N.R. N.R. 6.7 Izumi et al.
FW (2010)
Standard FW 2L Meso 45 d N.R. 435 0.33 N.R. 7.5 Kawai et al.
(2014)
Synthetic FW (+0.4 g ZVI/g VS FW) 0.25 L Meso 1221 h N.R. 495 8.5 40 gVS/L 60 N.R. Kong et al.
(2016)
Vegetable waste from supermarket Double stage Meso/Meso 55 d N.R. 445 18.9 N.R. N.R. L et al. (2012)
(1 L each)
Synthetic FW 0.5 L Meso 100 h N.R. 82 7.5 48 10 N.R. Nathao et al.
(2013)
Synthetic FW Double stage Meso/Meso 100 h N.R. 94 7.5 N.R. N.R. Nathao et al.
(0.5 L each) (2013)
Local waste management company 0.5 L Meso 30 d N.R. 492.7 9.3 0.5 N.R. N.R. Voelklein et al.
(2016)
Canteen FW 1L Meso 28 d N.R. 410 8 gVS/L 75 7.3 Zhang et al.
(2013a)
Restaurant FW 1L 50 C 25 d 93.6 510 1.6 65.6 7.2 Liu et al.
(2009)
Canteen FW 0.5 L Thermo 28 d 95.85 178a 1.5 54 8 Yang et al.
(2015)
Waste sorted by waste company 1L Thermo 28 d 81 435 2 gVS/L 73 7.6 Zhang et al.
(2007)

N.R. not reported.


a
Calculated from data provided in the manuscript.

the methanogenesis and eventually leading to the failure of the for its determination have been proposed (Holliger et al., 2016).
process due to VFA accumulation and decrease in the pH. Alterna- Under mesophilic conditions biomethane potential is variable
tively, nitrogen is released during the digestion process and, between 200 and 570 mL CH4 g 1VSadded (Ariunbaatar et al.,
depending on pH, OLR and temperature, this may lead to high con- 2014; Izumi et al., 2010; Kawai et al., 2014; Liu et al., 2012;
centrations of free ammonia (FAN) in the digester. Inhibition due to Zhang et al., 2007, 2013a) as summarized in Table 1.
high nitrogen content of the substrate has been reported (Chen Some papers reported relatively low methane yields ranging
et al., 2008). The different feedstock composition and characteris- from 100 to 250 mL CH4 g 1VSadded (Capson-Tojo et al., 2016;
tics are hence responsible of methane yields fluctuation depending Nathao et al., 2013; Yang et al., 2015;) probably due to acidification
on the specific FW processed (Tables 1 and 2). In case of AD insta- during food waste digestion (Liu et al., 2009). An important factor
bility remedial measures were introduced, such as alkalinity addi- affecting the performances of batch AD is the substrate to inocu-
tion (Liu et al. 2013; Ventura et al., 2014), feed interruption, lum (S/I) ratio (or the inverse factor, namely inoculum to substrate
mixing, or trace elements addition (Facchin et al., 2013; Qiang ratio, ISR) used. The major task in a one-stage batch reactor is to
et al., 2013; Yirong et al., 2015). prevent VFAs accumulation inside the inoculum particles beyond
their assimilative methanogenic capacity. This accumulation can
3.1. Batch AD tests be prevented by increasing the amount of inoculum, in order to
overcome irreversible acidification during start-up (Kawai et al.,
Methane production varies with many factors such as inocu- 2014). In single-stage batch tests, it is common practice to load
lum, volatile solids, ammonia, VFA, pH and temperature. A method inoculum and substrate based on VS ratios (Pagliaccia et al.,
to source and determine the feasibility of a material to serve as a 2016). Although theoretically, the S/I ratio has an effect only on
substrate in anaerobic digestion is the Biochemical Methane Poten- the kinetics, the influence of S/I on biomethane production has
tial (BMP) test. Such a test monitors the gas production following been commonly investigated during batch tests, and specific stud-
the incubation of an organic material with an anaerobic bacteria ies on the effect of this ratio for different FW have been carried out
mixture under well-controlled conditions. Because the BMP of (Liu et al., 2009; L et al., 2012; Pagliaccia et al., 2016). It is worth
the organic material is very important in the design, installation, to note that methane yields, obtained from single-stage mesophilic
and operation of an anaerobic digester, comprehensive protocols batch tests operated at S/I 0.5, range all between 417 and 529 L

Please cite this article in press as: Braguglia, C.M., et al. Anaerobic bioconversion of food waste into energy: A critical review. Bioresour. Technol. (2017),
http://dx.doi.org/10.1016/j.biortech.2017.06.145
http://dx.doi.org/10.1016/j.biortech.2017.06.145
Please cite this article in press as: Braguglia, C.M., et al. Anaerobic bioconversion of food waste into energy: A critical review. Bioresour. Technol. (2017),

Table 2
Best semi-continuous performances of single and two-stage anaerobic digestion systems in terms of methane conversion.
1 1 1
FW type/source one or two stage reactors (volume) Temperature HRT (d) OLR (g VS L d ; kg VS Duration VS removal Methane yield (L CH4 kg References
m 3d 1 (d) (%) VSfed)
Kitchen garbage 2(1 L; 5 L) Thermo/ 2; 10 N.R. 60 364 Chu et al. (2012)
thermo
Kitchen 1(5 m3) Meso N.R. 3.79 6 months 96 380 Grimberg et al.
(2015)
2(5 m3; 5 m3) Meso/meso N.R. 0.78 400 93 446
Dining hall raw food waste 2(8 L; 40 L) Thermo/ 2.87; 14.4 15 78.7 (as 470 Kobayashi et al.

C.M. Braguglia et al. / Bioresource Technology xxx (2017) xxxxxx


thermo COD) (2012)
1 1
Dining hall 2(10 L; 40 L) Thermo/ 1.9/7.7 93 g COD L d ; 100 450.6 Lee et al. (2010)
thermo 8.4 g COD L 1d 1
Incoming FW at the WWTP 2(200 L; 380 L) Thermo/ 3.3/12.6 18.4/4.8 140 476 Micolucci et al.
thermo (2014)
Local waste management comp 1(3 L working volume) Meso 16 9.2 45 91.8 455 Nagao et al. (2012)
Synthetic 1(12 L working vol) Thermo 50 6.3 kg COD m3 day 1
38 72 430 2 Qiang et al. (2013)
Synthetic + TEs 1(12 L working vol) Thermo 30 6.3 kg COD m3 day 1
120 78 475 95
FW + fruit and vegetable waste 1(8 L) Meso 30 13.5 210 328544 Shen et al. (2013)
2(5 L; 8 L) Meso/meso 10/10 1stage: 2.010 210 198546
2stage: 1.05.0
Source segregated domestic food 1(1 L) Meso 78 3 100 77.7 483 13 Tampio et al. (2014)
waste
Food waste recycling company 2(10 L; 30 L) Meso/meso 5.0/15 3.2 195 78.5 380 Ventura et al. (2014)
2(10 L; 30 L) Meso/termo 5.0/15 4.4 109 81.7 440
2(10 L; 30 L) Thermo/ 5.0/15 4 43 79.2 370
meso
Local waste management comp 2(5 L; 5 L) Meso/meso 4/12 15; 5 N.R. 389.2 31.8 Voelklein et al.
(2016)
1(5 L) Meso 16 4 N.R. 316.4 17.9
Restaurant 2(450 L; 1500 L) 40 C/40 C 160 h(SRT); 22.65; 4.61 125 546 Wang and Zhao
26.67 d (2009)
1 1 *
Synthetic 2(1.6 CSTR; 1.3 AFBR; with Thermo/ 3.7/1.5 3.4/6.1 g COD L d 30 334.7( ) Yeshanew et al.
recirculation) meso (2016)
Source segregated domestic FW 1(5 L) Thermo N.R. 2 50 400 Yirong et al. (2015)
(with TEs)
Pasteurized 1(10 L) Meso 20 3 152 480 33 Zamanzadeh et al.
(2016)
1(10 L) Thermo 20 3 152 448 44
Canteen 1(1 L) Meso 25 7 N.R. 405 Zhang et al. (2013a)
Campus restaurant (with TEs) 1(6 L) Meso 40 4.5 22 460 Zhang et al. (2015a)

CSTR: continuous stirred tank reactor; AFBR anaerobic fixed bed reactor; N.R.: not reported; (*) in L CH4 kg 1
COD d 1
.

5
Table 3

6
http://dx.doi.org/10.1016/j.biortech.2017.06.145
Please cite this article in press as: Braguglia, C.M., et al. Anaerobic bioconversion of food waste into energy: A critical review. Bioresour. Technol. (2017),

Best operating conditions and hydrogen yields from food waste fermentation, in single and dual stage.

Waste Inoculum Inoculum pre-treatment Operative conditions H2 Yield References


Reactor type Temperature HRT (h) Organic pH Duration
(C) Load
Single-Stage
1
FW Mesophilic Batch 35 NR S/I: 6 NR 44 h 39 mL g VSfed Pan et al.
1
anaerobic sludge 55 NR S/I: 7 44 h 57 mL g VSfed (2008)
Thermophilic
anaerobic sludge
FW hydrolysate Anaerobic sludge Heat pretreatment (T = 100C; t = 6 h) Batch (0.5 L) 37 NR NR 4.04.6 75 d 219.9 mL Han et al.
g 1VSfed (2015)
(1.56 mol
mol 1 glucose
added)
Heat pretreated FW None Batch (0.2 L) 35 NR NR Initial pH:8 100 h 1.92 mol mol 1 Kim et al.
(T = 90C; t = 20 min); Operational hexose added (2011)
pH:5

C.M. Braguglia et al. / Bioresource Technology xxx (2017) xxxxxx


Alkali pretreated FW None Batch (0.3L) 37 NR NR 6.0 1.2 d 162 mL Jang et al.
(pH:12) g 1VSfed (2015)
(1.71 mol
mol 1 hexose
added)
FW (ultrasound pre- Anaerobic sludge Heat pretreatment (T = 90C; t = 20 Batch (0.125 L) 37 NR 4 gCOD 5.5 NR Max 149 mL Gadhe et al.
treatment) min) g 1VSS g 1VSfed (2014)
(FW TS = 8%;
UStime:15
min)
FW (pretreatment with T80 Sewage sludge Heat pretreatment (T = 100C; t = 15 Batch (0.18 L) 55 NR 70:30% NR 60 h 80.6 mL Elsamadony
and PEG6000) min) (v/v) g 1COD et al. (2015)
22.6 (VS/ (2.8% T80)
VS)a 82.4 mL
g 1COD
(1.7 g L 1 PEG)
85.6 mL
g 1COD
(2.8% T80 + 1.7
g L 1 PEG)
Kitchen waste Digested slurry Heat pretreatment (T = 100C; t = 30 Inclined plug flow n.r. HRT:168a 8.19 gVS 5.6 NR 72 mL g 1VSfed Jayalakshmi
min) reactor (150 L) pilot L 1d 1a et al. (2009)
scale
OFMSW (ultrasonic + acid None Batch (0.2 L) 37 NR NR Initial pH:5.5 360 h 118 mL Elbeshbishy
pretreated) g 1VSfed et al. (2011a)
OFMSW Anaerobic sludge Ultrasonic pre-treatment CSTR with ultrasonic 37 24 14.5 gVSS 56 45 d 332 mL Elbeshbishy
(Espec = 79 kJ kg 1TS) probe (2 L) L 1d 1 g 1VSSfed et al. (2011b)
continuous mode
Municipal food waste UASB reactor ABR (26 L) 35 38.4 29 5.05.2 250 d 245 mL Tawfik and El-
+ kitchen waste sludge for H2 (1.6 d) gCODtot g 1CODremoved Qelish (2012)
production L 1d 1 558 mL
g 1VSremoveda
FW Acidogenic sludge CSTR (3 L) 34 48 11.4 gVS 5.5 10 HRT 20.5 mL Redondas
Semi-continuous L 1d 1 g 1VSfed et al. (2012)
mode
FW hydrolysate WWTP sludge Aeration (30 d) CMISR (3.2 L) packing NR 6 40 kgVS >4 5d Best rate: 354 Han et al.
3
ratio:15% m d 1 mL h 1L 1 (2015)
of activated carbon 16 kgVS Best yield: 85.6
3
m d 1 mL g 1FW
Table 3 (continued)
http://dx.doi.org/10.1016/j.biortech.2017.06.145
Please cite this article in press as: Braguglia, C.M., et al. Anaerobic bioconversion of food waste into energy: A critical review. Bioresour. Technol. (2017),

Waste Inoculum Inoculum pre-treatment Operative conditions H2 Yield References


Reactor type Temperature HRT (h) Organic pH Duration
(C) Load
1
FW + OFMSW (20:80) No SSRT 55 SRT:45.6 66 gVS 5.54 >3 SRT 38 mL g VSfed Angeriz-
(1.9 d) L 1d 1 Campoy et al.
(2015)
FW Anaerobic sludge CSTR (4.5 L) 55 120 23.6 NR NR 70.7 mL Algapani et al.
kgCOD g 1VSfed (2016)
m 3d 1
Two-Stage
1
FW No SCRD (200 L) 40 SRT:160 22.65 5.65.7 125 d 65 mL g VSfed Wang and
kgVS Zhao (2009)
m 3d 1
FW Anaerobic sludge Heat pretreatment (T = 90C; t = 30 Semi-continuous 55 45.6 39 (gCOD 5.45.7 100 d 114 mL Lee et al.
min) + acclimatization reactor (10 L) (1.9 d) L 1d 1) g 1VSfed (2010)
(2.5 mol mol 1
hexose)

C.M. Braguglia et al. / Bioresource Technology xxx (2017) xxxxxx


FW Anaerobic sludge Heat pretreatment (T = 90C; t = 30 Semi-continuous 55 68.9 NR 5.45.6 110 147.3 mL Kobayashi
min) + acclimatization reactor (8 L) (2.87 d) g 1VSfed et al. (2012)
Kitchen garbage Anaerobic sludge Heat pretreatment (T = 70C; t = 30 CSTR (1 L) 55 2d NR 5.5 60 d 66 mL g 1VSfed Chu et al.
min) Semi-continuous (2012)
1
FW Full scale UASB Heat pretreatment (T = 90C; t = 30 Two stage batch 37 NR S/I : 7.5 Initial pH:6 45 h 55 mL g VSfed Nathao et al.
reactor sludge min) fermentation (2013)
(0.5 L)
Source-segregated FW WWTP mesophilic CSTR (1 L) 52 72 20 gVS 5.22 NR 117 mL Chinellato
(+2nd phase recycled digestate Semi-continuous L 1d 1 g 1VSfed et al. (2013)
sludge)
1 1
FW Mesophilic CSTR (5 L) 35 10 2 gVS l NR 55d 28 mL g VSfed Shen et al.
anaerobic sludge semi continuous d 1 (2013)
mode
1
OFMSW (+2nd phase No CSTR (200 L) Semi- 55 79.2 18.4 kgVS 5.2 310 d 60 mL g VSfed Micolucci
recycled sludge) continuous (3.3 d) m 3d 1 et al. (2014)
Source-segregated FW Anaerobic sludge Heat pretreatment (T = 120C; t = 20 CSTR (1.35 L) 35 96 9 gVS 5.5 NR 11.8 mL Voelklein
min) + 1 M HCl + acclimatization (20 Semi-continuous L 1d 1 g 1VSfed et al. (2016)
d) (tested (1.711.8)
range 6
15)
FW Anaerobic sludge CSTR (1.6 L) 55 88.8 3.4 gVS 5.05.3 53 115.2 mL Yeshanew
Semi-continuous (3.7 d) L 1d 1 g 1VSfed et al. (2016)
FW (no pretreatment) UASB granular Heat pretreatment (T = 80C; t = 15 Batch (1 L) 35 NR S/I: 0.3 6.0 88 ha 55.3 mL Rafieenia et al.
FW (aeration digestate min) g 1VSfed (2017)
pretreatment; air
flow = 5L h 1; t = 24 h) 44.4 mL
g 1VSfed

ABR: anaerobic buffled reactor; CSTR: continuous stirred tank reactor; CMISR: continuous mixed immobilized sludge reactor; SCRD: semi-continuous rotating drum; SSRT: semi-continuous stirred tank reactor; UASB: upflow
anaerobic sludge blanket reactor; NR: Not reported; v/v = volume/volume.
a
Calculated.

7
8 C.M. Braguglia et al. / Bioresource Technology xxx (2017) xxxxxx

CH4 kg 1VSfed (Ariunbaatar et al., 2015; Browne and Murphy, 3.2. Continuous AD tests for methane production
2013; Facchin et al., 2013; Kawai et al., 2014; Voelklein et al.,
2016) evidencing stability of the process (Table 1). Most studies The effectiveness of methane AD production from FW process-
suggested that using S/I ratio below 1.0 was enough to prevent ing is currently limited by solubilization and acidogenesis, in par-
acidification (Elbeshbishy et al., 2012; Kawai et al., 2014). Methane ticular by semi-continuous fed reactors. The microbial
yields lower than 100 L CH4 kg 1VSfed are generally related to the community and inoculum quality used for the start-up of an anaer-
extremely high S/I ratios applied, and short duration time obic reactor are also critical factors for successful biogas produc-
(Nathao et al., 2013). Liu et al. (2009) performed thermophilic tion (Elbeshbishy et al., 2012; Capson-Tojo et al., 2016). In fact,
BMP assays on food waste at S/I ratios from 1.6 to 5.0 and reported the slow growth of methanogens compared to acidogens makes
that anaerobic biodegradability decreased with the rise of S/I long acclimatization periods and gradual operational changes nec-
ratios. However, when the inoculum was completely mixed with essary. Stable digestion can be assumed when the measured
the waste, methanogenesis was not initiated even at S/I of 0.9 parameters, such as the specific CH4 production, digestate VS con-
(L et al., 2012). In contrast, when the inoculum was initially sep- tent or pH levels are consistently maintained within 10% of their
arated from the waste in a double-reactor system (and connected average values, for a minimum period of one HRT. For all these rea-
by the controlled exchange of fermented waste), the waste could sons, the semi-continuous tests need long time duration, but
be effectively transformed to methane even at S/I 18.9 (L et al., sometimes literature data originate from too rushed digestion tests
2012). The instability of the process due to acidification results also with duration shorter than the applied HRT (see Table 2).
directly affected by the amount of labile organic fraction (LOF) Table 2 summarizes the best performances in terms of methane
immediately transformed to VFAs in the initial phase of AD yields and VS removal (if available) for single and dual stage diges-
(Kawai et al., 2014). In fact, the methane yields from the same tion systems reported in literature. Stable AD of FW was usually
FW, without LOF rich supernatant, were relatively stable in all S/I attained for HRT ranging between 16 and 40 days and OLRs lower
conditions, although the maximum methane yield was lower com- than 4.5 gVS L 1d 1 (Table 2). High buffer capacity, due to released
pared to the raw FW (Kawai et al., 2014). Moreover, excessive total ammoniacal nitrogen (TAN) allows AD to operate at higher
reduction of the particle size of the substrate resulted in VFA accu- organic loading rates (OLRs), thus resulting in higher biomethane
mulation, and decreased methane production (Izumi et al., 2010). production without experiencing a pH drop (Ariunbaatar et al.,
The impact of the solubilized matter, in particular sugars, on VFA 2015). Nevertheless, Chen et al. (2008) reported that TAN concen-
accumulation and consequent process instabilities should be trations ranging from 1700 mg L 1 to 14,000 mg L 1 all decreased
therefore taken into account when feedstock pre-treatments are methane yield by more than 50%, depending on feedstock, inocu-
selected and optimized (see paragraph 5.2), otherwise the strategy lum source and operation parameters employed in the studies.
can be detrimental rather than beneficial. Zero-valent iron signifi- Serna-Maza et al. (2014) studied on the efficient performance
cantly enhanced the conversion of butyric acid to acetic acid dur- and stability of 35 L anaerobic digesters fed on food waste coupled
ing AD of FW at very high initial loading rate (42.32 gVS L 1) to side-stream ammonia stripping columns, but high stripping
attaining a final CH4 yield of 380 mL g 1VSadded, suggesting inhibi- temperature (70 C) and a pH of 10 are needed. Selenium, which
tion of excessive acidification (Kong et al., 2016). is present only in low concentrations in FW, has been shown to
It is however also acknowledged that thermophilic process tem- be essential in recovering a digester suffering from a propionic acid
perature results in larger degree of imbalance and higher risk for accumulation due to elevated ammonia concentrations, while at
ammonia inhibition than mesophilic process temperatures, and higher organic loading rates additional cobalt is also required
in fact only few studies have focused on the thermophilic AD of (Yirong et al., 2015; Zhang et al., 2015a).
FW (Liu et al., 2009; Yang et al., 2015; Zhang et al., 2007). Contro- Trace metals (Fe, Co, Ni) requirement per COD removed was
versial results regard the impact of inoculum acclimation on obtained experimentally for mesophilic and thermophilic high-
methane conversion rate during BMP tests. While Browne and solid FW digestion evidencing higher Fe requirement per COD
Murphy (2013) highlighted that for accurate BMP assessment the removed operating in thermophilic conditions, probably because
inoculum should be sourced from a stable AD process and prefer- of the high metabolic rates of thermophilic anaerobic microorgan-
ably acclimatized to the substrate, a recent published cross- ism. In fact, ratios of required Fe to Ni and Co were about two times
comparison of methodologies used in different laboratories evi- higher in thermophilic system than those in mesophilic system
denced that it was not necessary that the inoculum was specifically (Qiang et al., 2013).
adapted to the substrates to be tested (Holliger et al., 2016). Effect A novel approach was the application of a dual solidliquid
of the inoculum source was observed also by Facchin et al. (2013) (ADSL) system, involving the use of two digesters for FSW (food
where using the inoculum originated from a reactor co-digesting solid waste) and FLW (food liquid waste without waste oil) diges-
FW and sludge generated more methane than using an inoculum tion, where optimum OLRs for FSW, FLW and raw FW were 9, 4 and
originated from an anaerobic reactor treating only FW. The pri- 7 gVS L 1d 1, and the corresponding methane yields 540, 390 and
mary parameter for the selection of inoculum is the soluble COD 405 mL g 1VS, respectively (Zhang et al., 2013a). Ghanimeh et al.
concentration, which reflects the health status of the provenance (2012) examined the effect of mixing and non-mixing strategies
digester. In the study of Elbeshbishy et al., 2012, in fact, the FW on the performance of thermophilic digesters, and the digester
digester inoculum had five times higher soluble COD concentration with mixing showed better stability minimizing VFA production.
than the municipal sludge digester, suggesting important differ- Considering the option of operating at thermophilic conditions, lit-
ences between the ratio of fermentative bacteria, i.e. hydrolyzers erature data show that by increasing the temperature, the propor-
and degraders to acidformers and methanogens between the two tion of the most toxic form of ammonia, namely FAN, also
inocula. The growth of methanogens is dependent on many ions increased, with detrimental effect on thermophilic process stability
such as sodium, nickel, cobalt, iron, zinc, magnesium, calcium and methane yields (Chen et al., 2008; Yirong et al., 2015;
and potassium cations and molybdate or tungstate and phosphate Zamanzadeh et al., 2016). The acquisition of the conductivity and
anions. Various concentrations of different trace elements (TEs) the online alkalinity measure are robust indirect parameters to
have been studied for the AD of FW. Facchin et al. (2013) achieved predict the content and the variation of ammonia concentration
a 4565% higher methane production yield from FW with supple- to avoid inhibitory issues (Micolucci et al., 2014). In the recent
mentation of TE (Co, Mo, Ni, Se, and W) cocktail, and stressed the years, considerable attention has been paid towards the develop-
importance of Se and Mo for the biomethane production. ment of high rate bioreactors with maximum treatment efficiency,

Please cite this article in press as: Braguglia, C.M., et al. Anaerobic bioconversion of food waste into energy: A critical review. Bioresour. Technol. (2017),
http://dx.doi.org/10.1016/j.biortech.2017.06.145
C.M. Braguglia et al. / Bioresource Technology xxx (2017) xxxxxx 9

basically designed to minimize HRT and increase rate of biogas two-stage thermophilic reactors without recirculation generated
production hence marking the fact that the reactor design has a low methane yields despite the low TS content of the feed (Chu
strong effect on digester performance. The performance of dry et al., 2012). Moreover, some studies have compared different
AD of FW was investigated under mesophilic conditions using an combinations of mesophilic and thermophilic conditions in each
horizontal-type cylindrical reactor, and stable dry AD was achieved of the stages. Ventura et al. (2014) reported the highest methane
by HRT control (more than 40 d) without the addition of alkali yields with a mesophilic first stage and a thermophilic second
agents, attaining average CH4 production rate of 2.51 0.17 m3 - stage, but the process was found to be less stable at higher temper-
m 3 d 1 (Cho et al., 2013). The rapid acidification phenomena atures in the second reactor.
observed at loading 4.9 kg VS m 3 d 1 to the high rate biomethana-
tion technology called Anaerobic gas lift reactor (AGR) was wor-
thy from the perspective of giving readily available soluble 4. Anaerobic conversion of food waste into hydrogen
compounds to the methanogens in the digester and decreasing
HRT for biodegradable organic matter removal (Kuruti et al., Among alternative energy carriers, hydrogen is recognized to
2017). However, this phenomenon could cause instability in the have an important role in the development of sustainable energy
digester due to pH drop unless buffering capacity of the digester systems, having the highest energy per unit weight (143 kJ kg 1)
was enhanced externally. To overcome this problem, milk of lime of any known gaseous fuel. Dark fermentation (DF) of low value
was added in the feed tank to maintain the pH of the feed slurry waste biomass is the most feasible biological technology for hydro-
around neutrality. Two-stage AD systems separate acid fermenta- gen production due to its low energy requirements and high pro-
tion and methanogenesis for the purpose of optimizing reactor duction rates meeting the circular economy requirements. H2 is
conditions for the distinctly different microbes that carry out these produced via dark fermentation (DF), mediated by bacteria capable
functions. The first stage (acid fermentation) is maintained at low of synthesizing hydrogen producing enzymes, such as hydroge-
pH and short hydraulic residence times (HRT; 23 days) resulting nase. In general, biohydrogen is generated from carbohydrate
in a washout of acid-consuming organisms. The second stage degradation through the acidogenesis and acetogenesis route, thus
(methanogenesis) is operated at HRT of 2030 days and pH of 6 FW represents a suitable feedstock. Different metabolic pathways
8, facilitating proliferation of slow-growing methanogenic archaea. can occur during DF of FW, depending on the adopted operating
Two stage AD systems may be less susceptible to system overload- conditions, resulting in either H2 production or depletion. In carbo-
ing with increased specific activity of methanogens resulting in a hydrates fermentation, hydrogen evolution is related to acetate
higher methane yield (Shen et al., 2013; Wang and Zhao, 2009). and butyrate pathways, which involve the production of, respec-
Optimizing two-stage conditions may result in the production of tively, 4 and 2 mol of molecular H2 per mol of glucose degraded
hydrogen from the primary reactor and methane from the second (i.e. 544 and 272 mL H2/g glucose at 25 C). However, in mixed
reactor, making it a very attractive bio-energy producing system. microbial cultures, different biochemical pathways leading to the
Numerous studies have been conducted on the optimization of formation of other products (propionate, ethanol and lactic acid)
such systems with reactors both at mesophilic and/or thermophilic lower the H2 production (de Gioannis et al., 2013). As a result,
temperatures, optimizing HRT and OLR in particular of the first the actual hydrogen yield is always lower than the theoretical
reactor (see Table 3). Best operative conditions and results as one. Operational parameters including substrate type, inoculum
regards on hydrogen production are widely discussed in the next type and origin, pre-treatment type, temperature, pH, reactor con-
paragraph. In terms of methane yields, the two-stage AD process figuration, OLR and HRT are the main factors affecting the H2 con-
may provide unique benefits when treating high strength waste version rate (de Gioannis et al., 2013). Recently, a wide number of
such as FW at high load. In fact, by digesting FW with fruit and veg- scientific studies on H2 production through DF of FW has been pub-
etable waste at low OLRs (<2.0 g VS L 1 d 1), single-phase AD lished; most of these experiments have been conducted at lab or
resulted better than two-phase AD in terms of CH4 production pilot scale (Jayalakshmi et al., 2009; Wang and Zhao, 2009), while
(up to +4%), while at higher level of OLR two-phase digestion no data on full-scale hydrogen fermentation plants are currently
achieved higher CH4 production (Shen et al., 2013). Single and available. A broad range of operating conditions have been applied,
two-phase operations were compared also at higher scale using a so that the reported results are difficult to compare and even con-
digester system consisting of 5 m3 reactors treating FW and higher flicting (Table 3).
methane yields were obtained by two-phase mesophilic digestion An attempt to discuss, critically, the relative importance of each
compared to the single-stage operation (380 vs 446 L CH4 kg VS 1 parameter, as well their mutual interactions, on H2 production
L 1), but it must be pointed out that the applied VS loading rate for efficiency is here reported. For each operating parameter, the
the single stage was significantly higher with respect to the two- results reported in Table 3 refer to the best H2 yield or optimal sys-
stage (Grimberg et al., 2015). Yan et al. (2016) optimized acido- tem conditions for 2-stage processes.
genic off-gas utilization in a methanogenic UASB reactor for uti- The seed microorganisms used as inoculum represent a key fac-
lization of H2 and CO2 via direct hydrogenotrophic tor for H2 fermentation. In fact, most researchers use mixed micro-
methanogenesis and methane recovery increased up to 38.6%. bial cultures, such as sludge, soil or compost, because of low costs,
The introduction of external hydrogen into the anaerobic digester easiness in control and high versatility (de Gioannis et al., 2013).
and the consequent promotion of the hydrogenotrophic methano- However, by using mixed microflora, the coexistence of H2- pro-
genic community is a challenging strategy for in situ biogas ducing and consuming bacteria should be considered. In order to
upgrading. enrich the seed inoculum in hydrogen producers and to inactivate
High rate digester, based on two-stage technology, was the consumers, namely the hydrogenotrophic methanogens, inocu-
designed for the treatment of FW at TS concentration of 1821%, lum is generally pre-treated. This process step relies on the better
operating at high HRT 100 days attaining an average biogas pro- chance of hydrogen producer bacteria to survive when exposed to
duction of 0.16 m3 kg 1 VS d 1 with 5060% of methane (Dahiya harsh environmental conditions, due to their ability to form spores.
and Joseph, 2015). Thermophilic temperatures have been effec- The reported pre-treatments are heat-shock treatment (HST), aer-
tively applied in 2-stage systems digesting FW, obtaining high ation, addition of chemical compounds, acid or alkali treatment,
methane yields and avoiding inhibition by optimization of OLRs freezing and thawing. The most common approach is HST, which
and HRTs applying methanogenic sludge recirculation (Chinellato is generally carried out at temperature of 80100 C for 10
et al., 2013; Lee et al., 2010; Micolucci et al., 2014), while operating 30 min (Chu et al., 2012; Elsamadony et al., 2015; Gadhe et al.,

Please cite this article in press as: Braguglia, C.M., et al. Anaerobic bioconversion of food waste into energy: A critical review. Bioresour. Technol. (2017),
http://dx.doi.org/10.1016/j.biortech.2017.06.145
Table 4

10
http://dx.doi.org/10.1016/j.biortech.2017.06.145
Please cite this article in press as: Braguglia, C.M., et al. Anaerobic bioconversion of food waste into energy: A critical review. Bioresour. Technol. (2017),

Operative conditions and anaerobic performances of co-digestion of food waste with other substrates.

Co-substrates Pretreatment Process type Operative conditions Performances References


(Shredding excluded)
Temp HRT OLR or S/I Duration (d) VS CH4 yield (L H2 yield (L H2
(C) (d) removal CH4 kg 1 VS) kg 1 VS)
(%)
FW + Animal Waste
FW + PMa (VS ratio) None Batch (0.5 L) 37 S/I = 3 (VS based) 35 NR Dennehy et al. (2016)
(20/80) 54% 320
(40/60) 63% 443
(60/40) 68% 489
(80/20) 71% 521
FW + DMb (VS ratio) None Batch (1 L) 35 Initial 30 NR El-Mashad and Zhang
1
loading = 3 gVS L (2010)
(32/68) 60% 282
(48/52) 68% 311
FW + DMb (VS ratio) None CSTR (2 L) 36 80 2 gVS L 1
d 1
180 NR 630 (fine- NR Agyeman and Tao
grinded FW) (2014)

C.M. Braguglia et al. / Bioresource Technology xxx (2017) xxxxxx


(50/50) 470 (coarse-
grinded FW)
FW + ChMc None CSTR (3.5 L) 35 35 2.5 gVS L 1
d 1
225 NR 508 NR Wang et al. (2014)
(Alternate feeding)
FW + CSf (VS ratio) None CSTR (75 L) 36 30 2 gVS L 1
d 1
308 NR 220 NR Zhang et al. (2012)
(20/80)
FW + CMd (w/w) None Induced Bed Reactor 55 20 5.53 gCOD L 1
d 1
40 82.5% 330 NR Castrilln et al. (2013)
(10/90) (19 L)
FW + CoMg None SBR (24 L) 20 NR 0.84.2 gVS L 1
d 1
205 (Cycle NR 477 88 NR Rajagopal et al. (2017)
(NR) length = 7d) (last 24 cycles)
FW + Sewage Sludge
FW + PSh + WASi (v/v) None Batch (0.1 L) NR S/I = 4 NR NR NR Zhu et al. (2008)
(75/12.5/12.5) (Volume based) 104
(50/25/25) 112
(25/37.5/37.5) 10
FW + SSj (TS ratio) Batch (0.4 L) 37 S/I = 5 (TS based) 35 NR Zhang et al. (2016)
(50/50) None 50% 297
(50/50) Microwaves 48.1% 311
FW + WASi (w/w) None Two-stage batch (0.1 L) 37 S/I = 2.3 55 (h) 18%50% 25106 Liu et al. (2013)
(Several tests with FW contents = 10%; S/I = 3.0 (VS based) 1000 (h) 22%32% 216354
20%; 30%; 40%; 54%; 85%)
FW + WASi (TS ratio) None CSTR (2 L) 16.7 2 gVS L 1
d 1
NR Gou et al. (2014)
d
(33.3/66.6) 35 160 d 62% 250
(33.3/66.6) 45 178 d 60% 290
(33.3/66.6) 55 188 d 58% 370
FW + DSk (VS ratio) None CSTR (6 L) 35 20 NR NR Dai et al. (2013)
(30/70) 6.3 gVS L 1 d 1 46% 258
(50/50) 7.2 gVS L 1 d 1 58% 332
(70/30) 7.6 gVS L 1 d 1 67% 380
FW + WASi (VS ratio) None Single-stage pilot CSTR 35 20 0.85 gVS L 1 d 1 350 46% 960 (biogas) NR Zhang et al. (2013b)
(16.5/83.5) (20 tonns)
FW + WASi (VS ratio) None Two-stage pilot CSTR 35 NR NR 350 48% 440 (biogas) NR
(16.5/83.5) (4.5 + 15.5 tonns)
FW + Green and Agro/Waste
FW + OHl (w/w) None Batch (0.3 L) 37 S/I = 0.6 30 49% 505 84 Pagliaccia et al. (2016)
Table 4 (continued)
http://dx.doi.org/10.1016/j.biortech.2017.06.145
Please cite this article in press as: Braguglia, C.M., et al. Anaerobic bioconversion of food waste into energy: A critical review. Bioresour. Technol. (2017),

Co-substrates Pretreatment Process type Operative conditions Performances References


(Shredding excluded)
Temp HRT OLR or S/I Duration (d) VS CH4 yield (L H2 yield (L H2
(C) (d) removal CH4 kg 1 VS) kg 1 VS)
(%)
(67/33) Thermal (VS based) 53% 91 87
FW + GWm (VS ratio) None Batch (1 L) 50 S/I = 1.6 25 90.8% 430 NR Liu et al. (2009)
(50/50) 50 S/I = 3.1 82.5% 359
50 S/I = 4.0 80.8% 390
50 S/I = 5.0 60.2% 337
35 S/I = 3.1 48.7% 185
(VS/VSS based)
FW + Sn (w/w) None Batch (1 L) 35 Initial NR NR NR Yong et al. (2015)
1
loading = 5 gVS L
(20/80) 171
(60/40) 299
(80/20) 313
FW + RHo (VS ratio) None Batch (1 L) 37 S/I = 0.25 45 55% 557 (biogas) NR Rizwan Haider et al.
(2015)

C.M. Braguglia et al. / Bioresource Technology xxx (2017) xxxxxx


(98/2) S/I = 0.5 48% 458 (biogas)
S/I = 1.0 31% 267 (biogas)
S/I = 1.5 27% 97 (biogas)
S/I = 2.0 24% 71 (biogas)
(VS based)
FW + MHp (w/w) None CSTR (5 L) 37 68 1 gVS L 1 d 1
120 80.7% 400 NR Owamah and Izinyon
(2015)
(75/25) 27 2.5 gVS L 1 d 1 120 76.5% 408
19 3.5 gVS L 1 d 1 120 74.3% 447
15 4.5 gVS L 1 d 1 120 78.3% 482
FW + FVWq (VS ratio) None One-stage CSTR (8 L) 35 30 Single stage: 1.0 210 NR 328544 Shen et al. (2013)
3.5 gVS L 1 d 1
(81/19) Two-stage CSTR (5 L 10 1stage: 2.0 140 528
+ 8 L) 10 gVS L 1 d 1
10 2stage: 1.0 210 198546
5.0 gVS L 1 d 1
FW + RTLr (w/w) None Pilot scale one stage 35 30 6.8 gVS L 1 d 1 112 80.4% 153 NR Ratanatamskul and
(2500 L) Manpetch (2016)
1 1
(95/5) Pilot scale two stage 30 9.5 gVS L d 89.2% 283
(1000 L + 2500 L)
s 1 1
FW + LW None High Solids Digester 35 175 2 gCOD L d 141 NR 229 NR Drennan and Di
(280 L) (SRT) Stefano (2014)
(FW and LW mixed to obtain constant
OLR and reactor TS = 20%)
FW + RHo None Pilot scale Plug-Flow 37 26 5 gVS L 1
d 1
27 82.4% 446 (biogas) NR Jabeen et al. (2015)
(80 L)
1 1
(mixed to obtain a C/N = 28) 25 6 gVS L d 52 73.1% 399 (biogas)
1 1
14 9 gVS L d 30 35.4% 215 (biogas)
FW + Other Waste
FW + GTWu None CSTR (6 L) 35 30 NR 180 77.4% 620 Wu et al. (2016)
(NR) TPADx (10 L) 55 6 73.3% 520
+ 35 + 24
TPADry (7.5 L) 55 3 78.8% 470 13.5
+ 35 + 12
v 1 1
FW + BW (v/v) None Acidogenic 35 2 18 gVS L d 47 29% 78.5 Paudel et al. (2017)
(70/30) CSTR (8 L) 35 1 35 gVS L 1 d 1 34 27% 73.6
35 0.5 71 gVS L 1 d 1 27 26% 80.5
35 0.33 106 gVS L 1 d 1 25 NR 99.8

(continued on next page)

11
12
Table 4 (continued)
http://dx.doi.org/10.1016/j.biortech.2017.06.145
Please cite this article in press as: Braguglia, C.M., et al. Anaerobic bioconversion of food waste into energy: A critical review. Bioresour. Technol. (2017),

Co-substrates Pretreatment Process type Operative conditions Performances References


(Shredding excluded)
Temp HRT OLR or S/I Duration (d) VS CH4 yield (L H2 yield (L H2
(C) (d) removal CH4 kg 1 VS) kg 1 VS)
(%)
Methanogenic 35 20 1.24 gVS L 1 d 1 70 52.5% 728 Zhang et al. (2015b)
CSTR (30 L) 35 15 1.76 gVS L 1 d 1 80 44% 676
FW + FILw (VS ratio) None CSTR (0.3 L) 37 20 4.1 gVS L 1 d 1 70 78% 479
(77/23) 15 6.2 gVS L 1 d 1 70 76% 452
10 8.3 gVS L 1 d 1 70 72% 506

NR = Not Reported; S/I = Substrate/Inoculum ratio; w/w = weight/weight; v/v = volume/volume.


e
SS = Sewage Sludge; tOFMSW = Organic Fraction of Municipal Solid Waste.
.
a
PM = Pig Manure.
b
DM = Dairy Manure.
c
ChM = Chicken Manure.
d
CM = Cattle Manure.
f

C.M. Braguglia et al. / Bioresource Technology xxx (2017) xxxxxx


CS = Cattle Slurry.
g
CoM = Cow Manure.
h
PS = Primary Sludge.
i
WAS = Waste Activated Sludge.
j
SS = Sewage Sludge.
k
DS = Dewatered Sludge.
l
OH = Olive Husk.
m
GW = Green Waste.
n
S = Straw.
o
RH = Rice Husk.
p
MH = Maize Husk.
q
FVW = Fruit and Vegetable Waste.
r
RTL = Rain Tree Leafs.
s
LW = Landscape Waste.
u
GTW = De-oiled Grease Trap Waste.
v
BW = Brown Water.
w
FIL = Fresh Incineration plant Leachate.
x
TPAD = Temperature Phased Anaerobic Digestion.
y
TPADr = TPAD with biomass recycling.
C.M. Braguglia et al. / Bioresource Technology xxx (2017) xxxxxx 13

2014; Han et al., 2015; Jayalakshmi et al., 2009; Kobayashi et al., in the active biomass composition. Concerning batch systems, also
2012; Lee et al., 2010; Nathao et al., 2013; Rafieenia et al., 2017; in the case of hydrogen production, some studies proved that S/I
Voelklein et al., 2016;). Only few studies have been conducted ratio influences the efficiency of the process, and, contrary to what
without inoculum pre-treatment with the aim of reducing costs was observed for methane yields, efficient H2 production are
(Algapani et al., 2016; Shen et al., 2013; Tawfik and El-Qelish, strictly related to high S/I ratios (Nathao et al., 2013; Pan et al.,
2012; Yeshanew et al., 2016). As FW itself is a source of indigenous 2008;). Under thermophilic conditions, a maximum H2 yield of
microorganisms, also some DF experiments were performed with- 57 mL g 1VS was obtained at S/I of 7, while under mesophilic con-
out inoculum addition (Elbeshbishy et al., 2011a; Jang et al., 2015; ditions the highest yield of 39 mL H2 g 1VS was attained at S/I of 6
Kim et al., 2011; Wang and Zhao, 2009), by pre-treating the sub- (Pan et al., 2008). For semi-continuous and continuous operations,
strate, aimed at selecting fermentative microorganisms, with the the HRT is one of the most crucial parameter, controlling substrate
suppression of lactic acid bacteria (LAB). Based on the results of hydrolysis, conversion efficiency, but also the type of active micro-
their previous tests, Kim et al. (2011), applied HST treatment to bial population and the metabolic pathways established in the sys-
FW without inoculum addition, as they found that high tempera- tem. It is generally assumed that short HRTs promote H2
ture pre-treatment depressed the lactate production, contem- production due to methanogens wash-out from the reactor, while
porarily increasing the H2/butyrate production. The effect of long HRTs favour the growth of methanogenic activity and non H2-
alkali-shock was investigated by Jang et al. (2015) and maximal producing acidogens inside the active biomass. Therefore, both
H2 yield in batch was obtained by pre-treating the FW at pH 11 OLR and HRT need optimization as controlling parameters to inhi-
12; otherwise, when the batch process switched to continuous bit the activity of H2 consuming bacteria in DF process. Comparing
mode, H2 production dropped significantly due to the increased the data reported in Table 3, most researchers, with very few
activity of H2-consumers. exceptions, have set HRT in the range 14 days (de Gioannis
Studies on H2 production from FW were performed both under et al., 2013; Redondas et al., 2012), operating at high OLRs from
mesophilic and thermophilic conditions (Table 3). Many authors 8 to 40 kg VS m 3d 1. Nevertheless, hydrogen yield up to 38 L H2
agree that thermophilic conditions are suitable to optimize the kg 1VSfed was reported by Angeriz-Campoy et al. (2015) operating
enzymatic activity of hydrogenase during fermentation by Clostri- a CSTR in thermophilic conditions at OLR 66 g VS L 1d 1 with a
dia, to inhibit the activity of H2 consumers and to suppress the mixture of food waste and OFMSW. The stepwise increase of the
growth of LABs, thus assuring higher H2 yields. Process tempera- OLR from 15.10 to 37.75 kg VSfed m 3d 1 may cause a decrease of
ture shows a significant effect on the metabolic pathways, dictat- H2 production and, at the same time, an increase of lactic acid con-
ing the composition of soluble microbial products (SMPs) in the centration by an integrated two-stage process (Wang and Zhao,
fermentation medium: lactate was found to be predominant at 2009). Also Voelklein et al. (2016) noted a close correlation
35 C, while butyrate was the main product at 50 C. pH is consid- between increasing OLRs of fermentation reactor (615 kg VS L 1-
ered one of the most crucial operational parameter in H2 fermenta- d 1), with fixed SRT of 4 days, and high quantities of ethanol and
tion, as it determines the hydrogenase activity, metabolic pathway, lactic acid in the fermentation broth. The continuous stirred tank
and dominant microbial population. Optimal pH values have been reactors (CSTR) is the most commonly used reactor type, as the
reported in the range of 5.06.5 when the predominant metabolic biomass is suspended and well mixed in the fermentation liquid,
pathways are those associate with acetate and butyrate produc- this configuration guarantees an efficient substrate uptake by the
tion. On the contrary, neutral or higher pH seemed to favour etha- anaerobic bacteria. However, due to CSTR vulnerability to environ-
nol and propionate production. Moreover, if pH drops at very low mental shocks, other types of reactors have been exploited, too.
values (4.5) a metabolic shift could occur which leads to lactic Continuous mixed immobilized sludge reactor (CMISR) and anaer-
acid accumulation (zero-H2 pathway) and solvent production (sol- obic baffled reactor (ABR) were used, respectively, by Han et al.
ventogenesis). To avoid reactor acidification, the pH of the system (2015) and Tawfik and El-Qelish (2012) in order to enhance bio-
is usually maintained at the optimal operating value by buffer or mass retention in the reactor by physical immobilization of sludge
alkali solutions addition (Han et al., 2015; Kim et al., 2011; Jang through various carriers. These systems allow operating the reac-
et al., 2015), particularly in continuous H2-producing reactors. tors at higher OLRs compared to CSTRs, maintaining a stable fer-
Moreover, the influence of initial pH was correlated to lag phase mentation process and representing a practical solution at
duration, synthesis of enzymes and spore germination (in heat industrial scale. As regards to one stage H2 fermentation, higher
pre-treated inoculum) and the maximal H2 production was gas yields are generally reported in batch experiments rather than
observed for initial pH around 8 (de Gioannis et al., 2013; Kim with semi-continuous or continuous operations (Table 3). This is
et al., 2011). As discussed in the precedent paragraph, in 2-stage probably related to the short period in batch fermentation, which
AD systems, internal recirculation of the effluent from the metha- permits to easier re-establish proper environmental conditions in
nogenic phase has been proposed as an efficient strategy to control the system, resulting in high H2 productions and yields. On the
pH in the optimal range for the hydrogen producing bacteria, thus contrary, for long term operations, the continuous feeding of a
improving the efficiency and economics of the DF process low buffering capacity substrate as FW could likely lead to system
(Chinellato et al., 2013; Yeshanew et al., 2016). Nevertheless, con- failure due to pH drop, low nitrogen input, and biomass metabolic
troversial results were obtained by Kobayashi et al. (2012) because shifting. Unfortunately, even under optimized process conditions, a
by the recirculation of active methanogenic sludge an inhibitive considerable portion of the substrate energy content remains in
effect on H2 production was observed, probably due to the high the effluent from the H2 fermentation, promoting research activity
hydrogen-consuming activity of microorganisms present in the cir- and technological development towards the just mentioned two-
culated sludge. Moreover, in long-term operation, special attention phase AD systems to improve the overall FW conversion yields
must be paid also to the increase of ammonia concentrations in the recovering both hydrogen, from first, and methane, from second
reactors as TAN is recycled. Most of the DF hydrogen production reactor. In two-stage systems, the first reactor is generally oper-
studies have been performed by means of small laboratory vessels ated at lower OLRs with respect to those applied for one stage reac-
or stirred reactors, under wet conditions, operated under batch, tors (Table 3), and H2 yields in two-stage systems operating at
semi-continuous or continuous mode. The organic load applied thermophilic conditions vary between 60 and 150 mL H2 g 1VS,
to the reactor greatly affects the system performances, as it deter- and are significantly higher with respect to those obtained in
mines VFAs accumulation and pH changes, but also modifications mesophilic ones.

Please cite this article in press as: Braguglia, C.M., et al. Anaerobic bioconversion of food waste into energy: A critical review. Bioresour. Technol. (2017),
http://dx.doi.org/10.1016/j.biortech.2017.06.145
14 C.M. Braguglia et al. / Bioresource Technology xxx (2017) xxxxxx

5. Strategies to improve the anaerobic digestion of food waste higher methane yields, with respect to those observed with WAS
(Dai et al., 2013). As just highlighted for animal waste, also for sew-
Complex heterogeneous organic materials such as FW range age sludge hydrogen production from co-fermentation with FW is
from highly recalcitrant material to extremely biodegradable com- still limited (Liu et al., 2013). Hydrogen yields ranging from 10 to
pounds, so the best option to improve the AD performances is a 104 L H2 kg 1VSfed were obtained in batch mode with high S/I
case-specific and properly designed strategy considering not only (up to 4) treating different mixtures of FW, primary sludge and
FW characteristics and composition but also the final aim of the WAS. As expected, the higher the FW content in the mixture, the
study. In fact, feedstock pre-treatment is suitable in the case of higher was the H2 yield (Zhu et al., 2008). Liu et al. (2013) obtained
recalcitrant, lignocellulosic FW to improve AD performances when up to 106 L H2 kg 1VSfed and 354 L CH4 kg 1VSfed with the 85% of
hydrolysis is the rate limiting step, while, in the case of biodegrad- FW in a two-stage co-digestion batch system with WAS. Co-
able FW, the long-term digestion leads often to inhibition phenom- digestion of FW with sewage sludge results advantageous where
ena due to intermediate accumulation suggesting the need of co- spare digestion capacity of existing WWTPs is available (Mata-
digestion with other substrates to stabilize the entire process Alvarez et al., 2014) offering beneficial synergies for the water
(Wang et al., 2014). In the subsequent paragraphs both strategies industry and authorities responsible for FW management (Zhang
are reviewed and critically discussed. et al., 2013b). Nevertheless, several bottlenecks as inert impurities
in FW, regulatory uncertainty, lack of suitable options for biogas
5.1. Anaerobic co-digestion with other substrates to improve stability utilization, impact on bio-solids agricultural use, have been identi-
fied. A multi-disciplinary approach is therefore necessary to face
Co-digestion of FW with other organic substrates has currently these challenges and to promote co-digestion as a key technology
attracted increasing interest, making AD more efficient, enhancing for a circular economy (Nghiem et al., 2017). Besides sewage
biogas production, promoting synergistic effects of microorgan- sludge or animal manures, agro-food waste and green waste are
isms, stabilizing digestate and increasing its amount of key nutri- other possible co-substrates with FW, mainly because of the very
ents (El-Mashad and Zhang, 2010; Gou et al., 2014; Liu et al., low cost associated with their collection (Chen et al., 2014). More-
2013; Mata-Alvarez et al., 2014; Nghiem et al., 2017). In the liter- over, the mono-digestion of lignocellulosic green/agro waste faces
ature of the last years, various organic waste streams have been different challenges, because of the poor nutrient content, long
investigated to co-digest FW for biogas production, and the most retention times and potential high levels of inhibitory compounds
frequent are animal manures, sewage sludge, green waste and (Gianico et al., 2013). At the same time, the high content of recal-
agro-waste (Mata-Alvarez et al., 2014). The mixture ratios are citrant lignin in green/agro waste might reduce the biodegradable
selected in order to favor positive interactions as for instance the rate of FW (Drennan and Di Stefano, 2014), thus reducing the risk
right nutrients and moisture balance, to avoid inhibition and to of VFAs accumulation (Chen et al., 2014; Rizwan Haider et al.,
optimize methane production (Zhang et al., 2012). However, the 2015). The relatively low biodegradability of landscape waste,
wrong combination of co-substrates can lead to negative results. makes it an unsuitable substrate for-co-digestion with the aim to
The most important literature results regarding co-digestion of increase the C/N ratio (Drennan and Di Stefano, 2014). Co-
FW with different organic waste residues are summarized in digestion of FW with a substrate with high content of bioavailable
Table 4. carbon is therefore recommended to allow AD at loading rates up
Generally, animal waste is considered an excellent co-substrate to 15 g COD L 1d 1. Jabeen et al. (2015) co-digested FW with rice
because of its alkalinity, low C/N ratio and for its wide variety of husk in a pilot scale plug-flow reactor and, biogas production, reac-
macro- and micronutrients needed by the anaerobic consortium tor stability and VS removal efficiency decreased by OLR increase
of microorganisms. Zhang et al. (2012) confirmed that the co- and HRT reduction. Conversely, Owamah and Izinyon (2015)
digestion of FW with cattle slurry, in 20:80 ratio, permitted to obtained increasing methane yields (up to 482 L CH4 kg 1VS) by
reach greater stability operating at OLR of 2 kg VS m 3 day 1. increasing the OLR, up to 4.5 g VS L 1d 1, in a co-digestion of FW
Operating at the same OLR, Agyeman and Tao (2014) investigated with maize husk, probably due to the appropriate high FW fraction
the effects of FW particle size on co-digestion with dairy manure (i.e. 75%). An increase in methane yields by increasing the FW con-
(50:50), finding that specific methane yield increased up to 630 L tent in batch mesophilic co-digestion with straw was reported by
CH4 kg 1VSfed feeding fine-grinded FW in mesophilic conditions. Yong et al. (2015). As regards batch co-digestion, also in this case
Lowering the temperature to 20 C assured good methane yields an inverse correlation between CH4 yield and load (expressed as
and stability by co-digesting FW and cow manure (Rajagopal S/I) was observed, during mesophilic and thermophilic co-
et al., 2017). It is worth to note that in most published papers, digestion of FW with green waste or rice husk (Liu et al., 2009;
the mixtures of FW and animal waste matter are typically com- Rizwan Haider et al., 2015). Interesting prototypes as single-stage
posed by low percentages of FW (ideal substrate for H2 production (2500 L) and two-stage (1000 L + 2500 L) anaerobic digesters treat-
due to the high content of sugars), consequently no data are avail- ing FW and rain tree leaf have been developed and operated at high
able in the literature as regards hydrogen generation during co- OLRs, resulting in high VS removal efficiencies, but poor CH4 yields
digestion. Another typical co-substrate of FW is sewage sludge, (153 L CH4 kg 1VSfed and 283 L CH4 kg 1VSfed for single- and two-
characterized by low C/N ratio and low organic content that guar- stage) (Ratanatamskul and Manpetch, 2016). The feasibility of the
antee the C/N balance improving microbial activity and lowering co-digestion process of FW with a further wide range of other co-
intermediate accumulation, as ammonia (Dai et al., 2013; Liu substrates has been widely investigated. Zhang et al. (2015b) stud-
et al., 2013)). In fact, during co-digestion batch tests no process ied the co-digestion with fresh leachate rich in trace elements and
failure was reported and methane production increased signifi- alkalinity promoting therefore methane production assuring
cantly (Naran et al., 2016; Zhang et al., 2016). In semi-continuous buffering ability of AD system. Paudel et al. (2017) carried out a
co-digestion of FW and WAS (33:67 ratio) the contemporary two-stage co-digestion of FW with brown wastewater, obtaining
increase of HRT from 160 d to 188 d and process temperature up up to 99.8 L H2 kg 1VSfed in the first stage and up to 728 L CH4
to thermophilic conditions affected positively the methane yield kg 1VSfed in the second stage. The use of de-oiled gras trap waste
and VS reduction rate (Gou et al., 2014). Operating a mesophilic as co-substrate provided excellent conversion yields in single-
digester at shorter HRT (20 d) and high OLRs (up to 7.6 g VS L 1- stage digestion process and in a temperature-phased process with
d 1), co-digestion with dewatered sludge assured stability and biomass recycling system (Wu et al., 2016).

Please cite this article in press as: Braguglia, C.M., et al. Anaerobic bioconversion of food waste into energy: A critical review. Bioresour. Technol. (2017),
http://dx.doi.org/10.1016/j.biortech.2017.06.145
Table 5
http://dx.doi.org/10.1016/j.biortech.2017.06.145
Please cite this article in press as: Braguglia, C.M., et al. Anaerobic bioconversion of food waste into energy: A critical review. Bioresour. Technol. (2017),

Pre-treatment impact on solubilization and AD anaerobic digestion performances.

Type Substrate origin Pretreatment Pretreatment conditions Process type and parameters Solubilisation Results References
method
THERMAL
Source Autoclave T = 160 C; p = 6.2 bars T = 37 C; F/I = 1 (VS); Working +16% sCOD, +22% NH4-N 445 mLCH4/gVSfed vs 501 mL Tampio et al.
segregated volume = 0.4 L; CH4/gVSfed (-11%) (2014)
domestic FW Duration = 35 days
Cafeteria food Autoclave T max = 134 C; p max = 3.2 bar; T = 37 C; F/I = 0.6 (VS); pHi set +35% sCOD, +75% soluble sugars**; +58% 23 mLH2/gVSfed vs 5 mLH2/ Pagliaccia
waste Retention time = 20 min to 7 Working volume = 0.3 L; soluble proteins** gVSfed (+360%); 385 mLCH4/ et al. (2016)
Duration = 30 days gVSfed vs 446 mLCH4/gVSfed
( 14%)
University Autoclave T max = 121 C; Retention T = 35 C; Load = 0.8 gVS; No +27% sCOD, +19% soluble sugars 39.2 mLH2/gVSfed vs 0.4 mLH2/ Hu et al.
Canteen time = 15 min inoculum addition; Working gVSfed ** (+9700%) (2014)
volume = 0.03 L;
Duration = 5 days;
Kitchen waste Thermal T = 120 C; Retention time = 50 min T = 35 C; 2 phase system; 899 mLCH4/gVSfed ** vs Li and Jin
Working volume 607mLCH4/gVSfed ** (+48%) (2015)
acidogenic = 5.5 L;

C.M. Braguglia et al. / Bioresource Technology xxx (2017) xxxxxx


Duration = 21 days;
Kitchen waste; Thermal T = 175 C; Retention time = 60 min T = 35 C; F/I = 0.2; pHi set to +114% soluble sugars **, +204% soluble 7.9% on methane yield Liu et al.
vegetable/fruit 7.5; Working volume = 0.25 L; proteins ** (kitchen waste); +312% soluble mLCH4/gVSfed (kitchen (2012)
residue Duration = 15 days; sugars **, +185% soluble proteins ** waste); 11.7% on methane
(vegetable/fruit residue) yield mLCH4/gVSfed
(vegetable/fruit residue)
Synthetic food Thermal T = 80 C; Retention time = 90 min T = 33 C; F/I = 0.5 (VS); 647 mLCH4/gVSfed vs 426 Ariunbaatar
waste (various) Working volume = 1 L; mLCH4/gVSfed (+52%) et al. (2014)
Duration = 220 days;
BATCH
Dining hall Microwave Microwave frequency = 2450 MHz; T = 37 C; F/I = 5 (TS); Co- 25% sCOD **, 52% soluble sugars **, +45% 316 mLCH4/gVSfed (ratio MW- Zhang et al.
Ambient pressure; Power = 600 W; digestion with sewage sludge soluble proteins** FW:SS 3:2) vs 71 mLCH4/gVSfed (2016)
Tmax = 100 C; No retention time (ratio FW:SS = 3:2 TS based) (only FW) (+347%)
Working volume = 0.4 L;
Duration = 35 days
M-OFMSW Microwave Microwave frequency = 2450 MHz; T = 33 C; Working +21% sCOD **, +12.5% soluble proteins **
+8% on total biogas Shahriari
(representative Tmax = 115 C; Ramp time = 40 min; volume = 0.155 L; production** et al. (2012)
of Canadian Retention time = 1 min Duration = 30 days*
kitchen waste)
MECHANICAL
Campus Ultrasounds Power = 1200 W; T = 37 C; F/I = 4 (COD/VSS); +7% sCOD 48 mLH2/gVSfed vs 15 mLH2/ Gadhe et al.
cafeteria Frequency = 20 kHz; Tmax = 30 C; pHi set to 5.5; Working gVSfed (220%) (2014)
Retention time = 15 min volume = 0.125 L;
Food waste Ultrasounds Energy intensity = 7300 kJ/kgTSS**; T = 35 C; F/I (VSS) = 60**; 206 mLCH4/gVSSremoved vs Naran et al.
treatment plant (various) Retention time = 30 min Working vol = 0.06 L; 137 mLCH4/gVSremoved (+50%) (2016)
Duration = 20 days
University HVPD (High Pulse voltage = 40 kV; Electrode T = 35 C; F/I = 0.5 (VS); pHi set +16% sCOD, +53% soluble sugars, +1149% 315 mLCH4/gCODremoved vs Zou et al.
Canteen Voltage Pulse distance = 5 mm; Pulse to 7; Working volume = 1 L; soluble proteins (40 kV, 5 mm, 300 Hz, 234mLCH4/gCODremoved (2016)
Discharge); frequency = 400 Hz; Retention Duration = 45 days * 30 min); +107% sCOD, +25% soluble sugars, (+35%)
(various) time = 30 min +171% soluble proteins (34 kV, 5 mm,
400 Hz, 30 min)

(continued on next page)

15
16
Table 5 (continued)
http://dx.doi.org/10.1016/j.biortech.2017.06.145
Please cite this article in press as: Braguglia, C.M., et al. Anaerobic bioconversion of food waste into energy: A critical review. Bioresour. Technol. (2017),

Type Substrate origin Pretreatment Pretreatment conditions Process type and parameters Solubilisation Results References
method

BIOLOGICAL
University Enzymatic Glucoamylase loading = 10 U/ T = 35 C; F/I = 0.5 gVS/gVSS**; +1547% sCOD*,** 468 mLCH4/gVSfed vs 198 Kiran et al.
Cafeteria hydrolysis gdryFW T = 60 C; contact Working volume = 0.3 L**; mLCH4/gVSfed (+137%) (2015)
time = 24 h Duration = 30 days;
University Enzymatic Glucoamylase concentration = 2 g/L; T = 35 C; F/I = 5 (TS/VS)**; +25% sCOD*,** 817 mLCH4/gVSfed vs 610 Yin et al.
Canteen hydrolysis T = 60 C; contact time = 24 h Working volume = 0.16 L; mLCH4/gVSfed (+34%) (2016)
Duration = 30 days*;
Synthetic food Aeration Air flow rate = 5 L/h; Contact T = 35 C; F/I = 0.3 (VS); 2 19  33% on H2 production Rafieenia
waste time = 24 h phase system; Working mLH2/gVSfed; 10  +46% on et al. (2017)
volume = 1 L; methane production mLCH4/
Duration = 73 days* gVSfed according to food waste
type
CHEMICAL
Research Alkaline Reagent = 6 N KOH; pHfin = 12; T = 37 C; Load = 30 gCODcarb/ 162 mLH2/gVSfed (no control Jang et al.
institute Contact time = 6 h L; pHi set to 8.0; Working data were reported) (2015)

C.M. Braguglia et al. / Bioresource Technology xxx (2017) xxxxxx


cafeteria volume = 0.3 L;
Duration = 28 h*
Campus Surfactants T80 concentration = 2.8%; PEG T = 55 C; F/I = 22.6 (VS)**; pHi 86 mLH2/gCODi vs 62 mLH2/ Elsamadony
restaurant concentration = 1.7 g/L T = 37 C; set to 7.0; Working gCODi (37%) et al. (2015)
Contact time = 8 h volume = 0.18 mL;
Duration = 60 h
COMBINATION
M-OFMSW Microwave H2O2 added = 0.38 g/gTS at 30% (v/ T = 33 C; Working +35% sCOD **, +101% soluble proteins** +6% on total biogas Shahriari
(representative + H2O2 v); Contact time = 1 h; Microwave volume = 0.155 L; production** et al. (2012)
of Canadian frequency = 2450 MHz; Tmax = 85 C; Duration = 30 days*
kitchen waste) Ramp time = 40 min; Retention
time = 1 min;
Organics Ultrasounds Power = 500 W; Frequency = 20 kHz; T = 37 C; No inoculum +30% sCOD *,**; +31% soluble sugars; +35% 118 mLH2/gVSfed vs 42 mLH2/ Elbeshbishy
Processing with acid Tmax = 30 C; specific energy addition; pHi set to 5.5; soluble proteins *,** gVSfed (+181%) et al.
Facility (various) input = 7900 kJ/kgTS; pHfin = 3.0 Working volume = 0.2; (2011a)
(1 N HCl); Contact time = 24 h at 4 C Duration = 15 days *

SEMICONTINUOUS
THERMAL
Source Autoclave T = 160 C; T = 37 C; +16% sCOD, +22% NH4-N 439 mLCH4/gVSfed vs 465 Tampio et al.
segregated p = 6.2 bars OLR = 4 kgVS/m3 d; mLCH4/gVSfed (-6%) (2014)
domestic FW HRT = 47 d (Untr 58 d);
duration = 162 days**;
Working volume = 11 L
Synthetic Microwave Microwave frequency = 2450 MHz; T = 35C; +19% sCOD** 1.50 LCH4/LgVSremovedd vs Shahriari
kitchen waste Tmax = 145C; OLR = 3.26 kgVS/ m3d; 1.31 LCH4/LgVSremovedd et al. (2012)
Ramp = 2.7C/min 2 phase system; (+14.5%)
HRTacidogenic = 2d;
HRTmethanogenic = 7d;
Working volume = 0.3/0.6L
Duration = 20d*
MECHANICAL
Organics Ultrasounds Power = 500 W; T = 37C; +9% sCOD, +17% soluble sugars, +20% 247 mLH2/gVSSfed vs 180 Elbeshbishy
Processing Frequency = 20 kHz; CSTR reactor (HRT = 2 days; soluble proteins mLH2/gVSSfed (+37%) et al.
Facility Tmax = 30C; OLR = 14.5 for untreated and (2011b)
Specific energy input = 5000 kJ/kgTS; 13.4 gVSS/Ld for sonicated);
Retention time = 24 min Working volume = 2 L;
Duration = 45 days;
*
Deducted from the graph.
**
Calculated.
C.M. Braguglia et al. / Bioresource Technology xxx (2017) xxxxxx 17

5.2. Food waste pre-treatments and impact on AD performances Mechanical pre-treatments are, in general, far less sensitive to
substrate specific characteristics and they exclude the risk of inhi-
Every pre-treatment, in fact, acts with a different mechanism, bitory compounds formation, but they all require high-energy con-
with several variables playing a role and it is therefore difficult to sumption (electricity). Ultrasounds pre-treatment at 20 kHz acts as
compare them and to assess the ideal pre-treatment type because mechanical pre-treatment by disrupting the biomass due to strong
of the lack of common and standardised protocols (Carlsson et al., mechanical shear forces. Gadhe et al.(2014) and Elbeshbishy et al.
2012). In this section, the pre-treatments applied to food waste (2011a) applied ultrasounds for H2 production enhancement
have been revised and evaluated taking into consideration both sol- achieving +220% and +181% gain, respectively, by sonicating food
ubilisation and H2/CH4/biogas yields. Table 5 reports the best com- waste at high specific energy. Solubilisation was directly propor-
bination of pre-treatment parameters, leading to the best tional to sonication time and energy, but the degree of solubilisa-
performance, i.e. the one that had the highest gain in H2/CH4/biogas tion was generally lower compared to that obtained with
yield if compared to the control (untreated substrate). thermal pre-treatment. FW pre-treatment for methane production
As showed in Table 5 the solubilisation of organic matter (usu- gave a +50% gain by sonicating the substrate at 7300 kJ kg 1TSS
ally assessed by means of soluble COD measure) is not directly (Naran et al., 2016). Anyway, specific methane yield of control
related to the subsequent gain in CH4 or H2 yield, but sometimes AD with untreated FW was very low, namely 137 mL CH4 g 1-
the correlation is even negative. This leads to an urgent need in VSSrem versus a typical average value of 400 30 mL CH4 g 1 VSrem.
understanding more deeply the inner mechanisms of the various This might be due to the high value of the S/I chosen (60 on VSS
pre-treatments on the substrate characteristics, so to understand based), which might have led to inhibition of the process. Zou
how the variable contents of carbohydrates, proteins, lipids and et al. (2016) tried a first attempt of HVPD (high voltage pulse dis-
lignocellulosic fractions present in organic waste are associated charge) on FW. It consists of a pulsed power supplier and a multi-
to pre-treatment. Moreover, most FW pre-treatments have been needle-to-plate reactor, which creates a rapidly pulsing (several
still studied in lab scale batch reactors, while semi-continuous tri- kHz), high-voltage electric field to disrupt and break up the cellular
als are still very limited (Table 5). In this chapter, mechanical pre- membrane, complex organic solids, and macromolecules. HVPD
treatments aimed to reduce particle size such as grinding or pre-treated FW samples generated more methane than the control,
milling, and to screen FW (screw press, disc screens, magnets, and from the point of energy usage efficiency, output methane pro-
etc.) were not considered as they are already successfully imple- duction was calculated to be equal to 5.7 times than energy con-
mented in full-scale treatment plants. sumption, which makes it more appealing than sonication.
Thermal pre-treatments have been extensively investigated and Biological pre-treatments include both enzymatic and aerobic
are already implemented in full-scale for what concerns sewage methods (to enhance hydrolytic rate). Enzymatic pre-treatment
sludge and lignocellulosic biomasses. Its mechanism of action sig- intensify the hydrolytic activity improving biogas production.
nificantly alters both physical and chemical properties, resulting Interesting application of FW by fungal mash to obtain glucose-
generally in increased solubilisation and improved digestate dewa- rich hydrolyzate for ethanol production is reported elsewhere
terability and hygienization. Higher temperatures or longer reten- (Capson-Tojo et al., 2016). Enzymes can be added to the substrate
tion times affect significantly the degree of solubilisation, but this prior to digestion as pre-treatment or directly in the digester. As
is not always connected to a final energy yield gain. At 160 C lig- pre-treatment there is the inherent risk that released sugars are
nin solubilisation starts, but at the same time lignin solubilisation rapidly consumed by the endogenous microorganisms present in
is responsible for the formation of inhibitory phenolic compounds. the FW, so a sterilisation step might be required but this option
For temperature higher than 170 C there might be also some con- is generally too expensive to be pursued in full-scale implementa-
densation reactions (e.g. Maillard reactions) which turns out to be tion. In full-scale scenario, adding the enzymes directly in the
hardly biodegradable or even refractory (Ariunbaatar et al., 2014). digester is the most usual configuration (Carrere et al., 2016). Gen-
So, for all these reasons, increased yields have been mostly erally, even though enzymatic pre-treatments require much less
observed in the lower temperature range, and main results show energy input than mechanical and thermal ones without chemicals
a decreased yield related to high T pre-treatment (Table 5), while addition, cost, enzyme selectivity (lipase or glucoamylase) and pro-
there is an increasing trend of CH4 yield at T < 120 C. Tampio cess efficiency remain major issues. Moreover the relatively long
et al. (2014) and Liu et al. (2012), reported a 7.9  11.7% decrease contact time needed (24 h at least) might not be a feasible option
in methane yields after thermal pre-treatment at T > 160 C, even for full-scale plants, despite the high methane enhancement
though a high solubilisation degree. However, thermal pre- obtained in batch tests (Kiran et al., 2015; Yin et al., 2016). Aeration
treatments for H2 production seem to be a good option: a 92-fold prior to anaerobic digestion is another option, in order to induce
upgrade in hydrogen production by pure culture was obtained faster hydrolysis rates, because of the higher production of hydro-
(Hu et al., 2014) Most of the pure cultures tested (3 out of 4) were lytic enzymes, induced by an increased specific microbial growth
efficient in utilizing FW to produce H2 and gave a substantial yield (Ariunbaatar et al., 2014). In addition, pre-aeration reduces accu-
gain even without pH control (Hu et al., 2014). Similar results were mulation of VFAs, thus improving the start-up stability of AD.
reported for a single stage AD of pre-treated cafeteria FW, where Rafieenia et al. (2017) investigate aeration efficiency on different
sugars solubilisation enhanced significantly H2 yield, affecting neg- synthetic FW types before two-stage AD highlighting that average
atively the successive CH4 yield because of the lack of bioavailable hydrogen production decreased, probably due to the low SRT (3 d)
substance just transformed (Pagliaccia et al., 2016). Microwave in the first reactor and parallel consumption of the readily
(MW) pre-treatment is an appealing alternative method to conven- biodegradable carbon for microbial cell growth during substrate
tional heating because of lower energy requirements (no heat loss aeration. However, during the subsequent stage, CH4 production
through convection or conduction), and polarization of macro- was higher for pre-aerated protein- and carbohydrate-rich samples
molecules chains leading to the hydrogen bonds breakage. than in non-pre-aerated ones; conversely, pre-aeration of the lipid
Nonetheless, results are not that satisfactory: despite an increase rich-substrate was not effective, causing a decrease of energy
of organics solubilisation with T, biogas gain at 115 C was only potential compared to untreated sample.
8%, and AD process collapsed for the 175 C pre-treated samples Strong acids, alkalis or oxidants are commonly used reagents to
(Shahriari et al., 2012), highlighting that even MW pre-treatment solubilise biopolymers favouring the availability of the associated
is prone to melanoidins formation. Also by semi-continuous tests, organic compounds for anaerobic conversion. Solubilisation levels
the results were not satisfactory (Shahriari et al., 2013). achieved with chemical pre-treatments are usually very high,

Please cite this article in press as: Braguglia, C.M., et al. Anaerobic bioconversion of food waste into energy: A critical review. Bioresour. Technol. (2017),
http://dx.doi.org/10.1016/j.biortech.2017.06.145
18 C.M. Braguglia et al. / Bioresource Technology xxx (2017) xxxxxx

much higher with respect to those obtained with other pre- Ariunbaatar, J., Di Perta, E.S., Panico, A., Frunzo, L., Esposito, G., Lens, P.N.L., Pirozzi,
F., 2015. Effect of ammoniacal nitrogen on one-stage and two-stage anaerobic
treatments, but the risk of chemical contamination and recalcitrant
digestion of food waste. Waste Manage. 38, 388398.
compounds formation should be considered. Because of the high Browne, J.D., Murphy, J.D., 2013. Assessment of the resource associated with
solubilisation capacity, chemical pre-treatment is not suitable for biomethane from food waste. Appl. Energy 104, 170177.
easily biodegradable substrates such as FW: high rate carbohy- Capson-Tojo, G., Rouez, M., Crest, M., Steyer, J.-P., Delgens, J.-P., Escudi, R., 2016.
Food waste valorization via anaerobic processes: a review. Rev. Environ. Sci.
drates degradation and subsequent VFA accumulation convey in Bio/Technol. 15 (3), 499547.
methanogenesis inhibition (Ariunbaatar et al., 2014). A different Carlsson, M., Lagerkvist, A., Morgan-Sagastume, F., 2012. The effects of substrate
approach using surfactants (Tween 80 and polyethylene glycol, pretreatment on anaerobic digestion: a review. Waste Manage. 32, 16341650.
Carrere, H., Antonopoulou, G., Affes, R., Passos, F., Battimelli, A., Lyberatos, G., Ferrer,
individually or combined) to modify substrate structure and to I., 2016. Review of feedstock pretreatment strategies for improved anaerobic
improve enzymatic hydrolysis of particulate organic matter in digestion: from lab-scale research to full-scale application. Bioresour. Technol.
order to achieve higher H2 production was successfully reported 199, 386397.
Castrilln, L., Maran, E., Fernndez-Nava, Y., Ormaechea, P., Quiroga, G., 2013.
by Elsamadony et al. (2015). Pre-treatments are usually coupled Thermophilic co-digestion of cattle manure and food waste supplemented with
to lower the severity of each pre-treatment reducing the intrinsic crude glycerin in induced bed reactor (IBR). Bioresour. Technol. 136, 7377.
disadvantages as inhibition, recalcitrant products formation, long Chen, Y., Cheng, J.J., Creamer, K.S., 2008. Inhibition of anaerobic digestion process: a
review. Bioresour. Technol. 99, 40444064.
retention times, specificity of the enzyme, energy costs. Combina- Chen, X., Yan, W., Sheng, K., Sanati, M., 2014. Comparison of high solids to liquid
tion is also applied in the case a single pre-treatment is not suffi- anaerobic co-digestion of food waste and green waste. Bioresour. Technol. 154,
cient to obtain the desired result due to its specific mechanism 215221.
Chinellato, G., Cavinato, C., Bolzonella, D., Heaven, S., Banks, C.J., 2013. Biohydrogen
of action. Elbeshbishy et al. (2011a) tried to couple ultrasonic with
production from food waste in batch and semi-continuous conditions:
acid, alkali and heat shock pre-treatment for H2 production with- evaluation of a two-phase approach with digestate recirculation for pH
out using extra seed (FW was inoculum and substrate at the same control. Int. J. Hydrogen Energy 38, 43514360.
time). The only significant increase compared to the single pre- Cho, S.K., Im, W.T., Kim, D.H., Kim, M.H., Shin, H.S., Oh, S.E., 2013. Dry anaerobic
digestion of food waste under mesophilic conditions: performance and
treated sample was due to acid addition, resulting in +181% higher methanogenic community analysis. Bioresour. Technol. 131, 210217.
hydrogen yield than the control. Even though results are promising Chu, C.F., Xu, K.Q., Li, Y.Y., Inamori, Y., 2012. Hydrogen and methane potential based
in this case, an economic analysis is needed to determine whether on the nature of food waste materials in a two-stage thermophilic fermentation
process. Int. J. Hydrogen Energy 37 (14), 1061110618.
the combination is energetically feasible and economically conve- Dahiya, S., Joseph, J., 2015. High rate biomethanation technology for solid waste
nient. As the pretreatment of FW is relatively new, no real cost esti- management and rapid biogas production: an emphasis on reactor design
mation has been reported in literature to date. parameters. Bioresour. Technol. 188, 7378.
Dai, X., Duan, N., Dong, B., Dai, L., 2013. High-solids anaerobic co-digestion of
sewage sludge and food waste in comparison with mono digestions: stability
and performance. Waste Manage. 33, 308316.
6. Conclusions De Baere, L., Mattheeuws, B., 2014. Anaerobic digestion of the organic fraction of
municipal solid waste in Europe status, experience and prospects in: Waste
Management, vol. 3 Recycling and Recovery, TK, 517526.
This review explored recent advances in the research field of De Gioannis, G., Muntoni, A., Polettini, A., Pomi, R., 2013. A review of dark
FW conversion into energy in the framework of the circular econ- fermentation hydrogen production from biodegradable municipal waste
omy. It resulted that FW is a complex and heterogeneous substrate, fractions. Waste Manage. 33 (6), 13451361.
Dennehy, C., Lawlor, P.G., Croize, T., Jiang, Y., Morrison, L., Gardiner, G.E., Zhan, X.,
with variable biodegradability, and consequently AD results can 2016. Synergism and effect of high initial volatile fatty acid concentrations
vary significantly during the scaling up of the process, from batch during food waste and pig manure anaerobic co-digestion. Waste Manage. 56,
BMP tests to semi-continuous performances. Prudence is needed 173180.
Dong, L., Zhenhong, Y., Yongming, S., 2010. Semi-dry mesophilic anaerobic digestion
by pre-treating FW to improve solubilization, since hydrolysis is of water sorted organic fraction of municipal solid waste (WS-OFMSW).
not necessarily the limiting step, and the risks related to typical Bioresour. Technol. 101, 27222728.
acids accumulation may be emphasized. In situ methane enrich- Drennan, M.F., Di Stefano, T.D., 2014. High solids co-digestion of food and landscape
waste and the potential for ammonia toxicity. Waste Manage. 34, 12891298.
ment processes, as bubble columns for selective desorption of Elbeshbishy, E., Hafez, H., Dhar, B.R., Nakhla, G., 2011a. Single and combined effect
CO2 need more applicative research. of various pretreatment methods for biohydrogen production from food waste.
Int. J. Hydrogen Energy 36, 1137911387.
Elbeshbishy, E., Hafez, H., Nakhla, G., 2011b. Ultrasonication for biohydrogen
production from food waste. Int. J. Hydrogen Energy 36, 28962903.
Acknowledgements
Elbeshbishy, E., Nakhla, G., Hafez, H., 2012. Biochemical methane potential (BMP) of
food waste and primary sludge: influence of inoculum pre-incubation and
This work was supported by the PRIN 2012 project titled inoculum source. Bioresour. Technol. 110, 1825.
Advanced Processes to convert organic wastes in innovative, sus- El-Mashad, H.M., Zhang, R., 2010. Biogas production from co-digestion of dairy
manure and food waste. Bioresour. Technol. 101, 40214028.
tainable and useful products co-financed by the Italian Minister of Elsamadony, M., Tawfik, A., Suzuki, M., 2015. Surfactant-enhanced biohydrogen
University and Scientific Research (MIUR). Authors wish to thank production from organic fraction of municipal solid waste (OFMSW) via dry
dr. Giuseppe Mininni and Prof. Fausto Gironi for their valuable anaerobic digestion. Appl. Energy 149, 272282.
Facchin, V., Cavinato, C., Fatone, F., Pavan, P., Cecchi, F., Bolzonella, D., 2013. Effect of
support. trace element supplementation on the mesophilic anaerobic digestion of food
waste in batch trials: the influence of inoculum origin. Biochem. Eng. J. 70, 71
77.
References Fisgativa, H., Tremier, A., Dabert, P., 2016. Characterizing the variability of food
waste quality: a need for efficient valorization through anaerobic digestion.
Waste Manage. 50, 264274.
Agyeman, F.O., Tao, W., 2014. Anaerobic co-digestion of food waste and dairy
Gadhe, A., Sonawane, S.S., Varma, M.N., 2014. Ultrasonic pretreatment for an
manure: effects of food waste particle size and organic loading rate. J. Environ.
enhancement of biohydrogen production from complex food waste. Int. J.
Manage. 133, 268274.
Hydrogen Energy 39, 77217729.
Algapani, D.E., Qiao, W., Su, M., di Pumpo, F., Wandera, S.M., Adani, F., Dong, R.,
Ghanimeh, S., El Fadel, M., Saikaly, P., 2012. Mixing effect on thermophilic anaerobic
2016. Bio-hydrolysis and bio-hydrogen production from food waste by
digestion of source-sorted organic fraction of municipal solid waste. Bioresour.
thermophilic and hyperthermophilic anaerobic process. Bioresour. Technol.
Technol. 117, 6371.
216, 768777.
Gianico, A., Braguglia, C.M., Mescia, D., Mininni, G., 2013. Ultrasonic and thermal
Angeriz-Campoy, R., lvarez-Gallego, C.J., Romero-Garca, L.I., 2015. Thermophilic
pretreatments to enhance the anaerobic bioconversion of olive husks.
anaerobic co-digestion of organic fraction of municipal solid waste (OFMSW)
Bioresour. Technol. 147, 623626.
with food waste (FW): enhancement of bio-hydrogen production. Bioresour.
Gou, C., Yang, Z., Huang, J., Wang, H., Xu, H., Wang, L., 2014. Effects of temperature
Technol. 194, 291296.
and organic loading rate on the performance and microbial community of
Ariunbaatar, J., Panico, A., Frunzo, L., Esposito, G., Lens, P.N.L., Pirozzi, F., 2014.
anaerobic co-digestion of waste activated sludge and food waste. Chemosphere
Enhanced anaerobic digestion of food waste by thermal and ozonation
105, 146151.
pretreatment methods. J. Environ. Manage. 146, 142149.

Please cite this article in press as: Braguglia, C.M., et al. Anaerobic bioconversion of food waste into energy: A critical review. Bioresour. Technol. (2017),
http://dx.doi.org/10.1016/j.biortech.2017.06.145
C.M. Braguglia et al. / Bioresource Technology xxx (2017) xxxxxx 19

Grimberg, S.J., Hilderbrandt, D., Kinnunen, M., Rogers, S., 2015. Anaerobic digestion Nghiem, L.D., Koch, K., Bolzonella, D., Drewes, J.E., 2017. Full scale co-digestion of
of food waste through the operation of a mesophilic two-phase pilot scale wastewater sludge and food waste: bottlenecks and possibilities. Renewable
digester assessment of variable loadings on system performance. Bioresour. Sustainable Energy Rev. 72, 354362.
Technol. 178, 226229. Owamah, H.I., Izinyon, 2015. The effect of organic loading rates (OLRs) on the
Han, W., Liu, D.N., Shi, Y.W., Tang, J.H., Li, Y.F., Ren, N.Q., 2015. Biohydrogen performances of food wastes and maize husks anaerobic co-digestion in
production from food waste hydrolysate using continuous mixed immobilized continuous mode. Sustainable Energy Technol. Assess. 11, 7176.
sludge reactors. Bioresour. Technol. 180, 5458. Pagliaccia, P., Gallipoli, A., Gianico, A., Montecchio, D., Braguglia, C.M., 2016. Single
Holliger, C., Alves, M., Andrade, D., Angelidaki, I., Astals, S., Baier, U., Bougrier, C., stage anaerobic bioconversion of food waste in mono and co-digestion with
et al., 2016. Towards a standardization of biomethane potential tests. Water Sci. olive husks: impact of thermal pretreatment on hydrogen and methane
Technol. 74 (11), 25152522. production. Int. J. Hydrogen Energy 41 (2), 905915.
Hu, C.C., Giannis, A., Chen, C.-L., Wang, J.-Y., 2014. Evaluation of hydrogen producing Pan, J., Zhang, R., El-Mashad, H.M., Sun, H., Ying, Y., 2008. Effect of food to
cultures using pretreated food waste. Int. J. Hydrogen Energy 39, 1933719342. microorganism ratio on biohydrogen production from food waste via anaerobic
Izumi, K., Okishio, Y., Nagao, N., et al., 2010. Effects of particle size on anaerobic fermentation. Int. J. Hydrogen Energy 33, 69686975.
digestion of food waste. Int. Biodeterior. Biodegrad. 64, 601608. Patinvoh, R.J., Osadolor, O.A., Chandolias, K., Srvri Horvth, I., Taherzadeh, M.J.,
Jabeen, M., Zeshan, Yousaf, S., Rizwan Haider, M., Malik, R.N., 2015. High-solids 2017. Innovative pretreatment strategies for biogas production. Bioresour.
anaerobic co-digestion of food waste and rice husk at different organic loading Technol. 224, 1324.
rates. Int. Biodeterior. Biodegrad. 102, 149153. Paudel, S., Kang, Y., Yoo, Y.-S., Seo, G.T., 2017. Effect of volumetric organic loading
Jang, S., Kim, D.H., Yun, Y.M., Lee, M.K., Moon, C., Kang, W.S., Kwak, S.S., Kim, M.S., rate (OLR) on H2 and CH4 production by two-stage anaerobic co-digestion of
2015. Hydrogen fermentation of food waste by alkali shock pretreatment: food waste and brown water. Waste Manage. 61, 484493.
microbial community analysis and limitation of continuous operation. Qiang, H., Niu, Q., Chi, Y., Li, Y., 2013. Trace metals requirements for continuous
Bioresour. Technol. 186, 215222. thermophilic methane fermentation of high solid food waste. Chem. Eng. J. 222,
Jayalakshmi, S., Joseph, K., Sukumaran, V., 2009. Bio hydrogen generation from 330336.
kitchen waste in an inclined plug flow reactor. Int. J. Hydrogen Energy 34, Rafieenia, R., Girotto, F., Peng, W., Cossu, R., Pivato, A., Raga, R., Lavagnuolo, M.C.,
88548858. 2017. Effect of aerobic pre-treatment on hydrogen and methane production in a
Kawai, M., Nagao, N., Tajima, N., Niwa, C., Matsuyama, T., Toda, T., 2014. The effect two-stage anaerobic digestion process using food waste with different
of the labile organic fraction in food waste and the substrate/inoculum ratio on compositions. Waste Manage. 59, 194199.
anaerobic digestion for a reliable methane yield. Bioresour. Technol. 157, 174 Rajagopal, R., Bellavance, D., Rahaman, M.S., 2017. Psychrophilic anaerobic
180. digestion of semi-dry mixed municipal food waste: for North American
Kim, D.H., Kim, S.H., Jung, K.W., Kim, M.S., Shin, H.S., 2011. Effect of initial pH context. Process Saf. Environ. Prot. 105, 101108.
independent of operational pH on hydrogen fermentation of food waste. Ratanatamskul, C., Manpetch, P., 2016. Comparative assessment of prototype
Bioresour. Technol. 102, 86468652. digester configuration for biogas recovery from anaerobic co-digestion of food
Kiran, E.U., Trzcinski, A.P., Liu, Y., 2015. Enhancing the hydrolysis and methane waste and rain tree leaf as feedstock. Int. Biodeterior. Biodegrad. 113, 367374.
production potential of mixed food waste by an effective enzymatic Redondas, V., Gmez, X., Garca, S., Pevida, C., Rubiera, F., Morn, A., Pis, J.J., 2012.
pretreatment. Bioresour. Technol. 183, 4752. Hydrogen production from food wastes and gas post-treatment by CO2
Kobayashi, T., Xu, K.Q., Li, Y.Y., Inamori, Y., 2012. Effect of sludge recirculation on adsorption. Waste Manage. 32 (1), 6066.
characteristics of hydrogen production in a two-stage hydrogenmethane Rizwan Haider, M., Zeshan, Yousaf, S., Malik, R.N., Visvanathan, C., 2015. Effect of
fermentation process treating food wastes. Int. J. Hydrogen Energy 37, 5602 mixing ratio of food waste and rice husk co-digestion and substrate to inoculum
5611. ratio on biogas production. Bioresour. Technol. 190, 451457.
Kong, X., Wei, Y., Xu, S., Liu, J., Li, H., Liu, Y., Yu, S., 2016. Inhibiting excessive Serna-Maza, A., Heaven, S., Banks, C.J., 2014. Ammonia removal in food waste
acidification using zero-valent iron in anaerobic digestion of food waste at high anaerobic digestion using a side-stream stripping process. Bioresour. Technol.
organic load rates. Bioresour. Technol. 211, 6571. 152, 307315.
Kuruti, K., Begum, S., Ahuja, S., Rao Anupoju, G., Juntupally, S., Gandu, B., Kumar Shahriari, H., Warith, M., Kennedy, K.J., 2012. Anaerobic digestion of organic fraction
Ahuja, D., 2017. Exploitation of rapid acidification phenomena of food waste in of municipal solid waste combining two pretreatment modalities e high
reducing the hydraulic retention time (HRT) of high rate anaerobic digester temperature microwave and hydrogen peroxide. Waste Manage. 32 (1), 4152.
without conceding on biogas yield. Bioresour. Technol. 226, 6572. Shahriari, H., Warith, M., Hamoda, M., Kennedy, K., 2013. Evaluation of single vs.
Lee, D.-Y., Ebie, Y., Xu, K.-Q., Li, Y.-Y., Inamori, Y., 2010. Continuous H2 and CH4 staged mesophilic anaerobic digestion of kitchen waste with and without
production from high-solid food waste in the two-stage thermophilic microwave pretreatment. J. Environ. Manage. 125, 7484.
fermentation process with the recirculation of digester sludge. Bioresour. Shen, F., Yuan, H., Pang, Y., Chen, S., Zhu, B., Zou, D., Liu, Y., Ma, J., Yu, L., Li, X., 2013.
Technol. 101 (1), S42S47. Performances of anaerobic co-digestion of fruit & vegetable waste (FVW) and
Li, Y.Y., Jin, Y.Y., 2015. Effects of thermal pretreatment on acidification phase during food waste (FW): single-phase vs. two-phase. Bioresour. Technol. 144, 8085.
two-phase batch anaerobic digestion of kitchen waste. Renewable Energy 77, Tampio, E., Ervasti, S., Paavola, T., et al., 2014. Anaerobic digestion of autoclaved and
550557. untreated food waste. Waste Manage. 34, 370377. http://dx.doi.org/10.1016/j.
Liu, G., Zhang, R., El-Mashad, H.M., Dong, R., 2009. Effect of feed to inoculum ratios wasman.2013.10.024.
on biogas yields of food and green wastes. Bioresour. Technol. 100, 51035108. Tawfik, A., El-Qelish, M., 2012. Continuous hydrogen production from co-digestion
Liu, X., Wang, W., Gao, X., Zhiu, Y., Shen, R., 2012. Effect of thermal pretreatment on of municipal food waste and kitchen wastewater in mesophilic anaerobic
the physical and chemical properties of municipal biomass waste. Waste baffled reactor. Bioresour. Technol. 114, 270274.
Manage. 32, 249255. Ventura, J.-R.S., Lee, J., Jahng, D., 2014. A comparative study on the alternating
Liu, X., Li, R., Ji, M., Han, L., 2013. Hydrogen and methane production by co-digestion mesophilic and thermophilic two-stage anaerobic digestion of food waste. J.
of waste activated sludge and food waste in the two-stage fermentation Environ. Sci. 26, 12741283.
process: substrate conversion and energy yield. Bioresour. Technol. 146, 317 Voelklein, M.A., Jacob, A., OShea, R., Murphy, J.D., 2016. Assessment of increasing
323. loading rate on two-stage digestion of food waste. Bioresour Technol. 202, 172
L, F., Hao, L., Zhu, M., Shao, L., He, P., 2012. Initiating methanogenesis of vegetable 180.
waste at low inoculums-to-substrate ratio: importance of spatial separation. Wang, X., Zhao, Y., 2009. A bench scale study of fermentative hydrogen and
Bioresour. Technol. 105, 169173. methane production from food waste in integrated two-stage process. Int. J.
Ma, Y., Yin, Y., Liu, Y., 2017. A holistic approach for food waste management towards Hydrogen Energy 34, 245254.
zero-solid disposal and energy/resource recovery. Bioresour. Technol. 228, 56 Wang, M., Sun, X., Li, P., Yin, L., Liu, D., Zhang, Y., Li, W., Zheng, G., 2014. A novel
61. alternate feeding mode for semi-continuous anaerobic co-digestion of food
Mata-Alvarez, J., Dosta, J., Romero-Giza, M.S., Fonoll, X., Peces, M., Astals, S., 2014. waste with chicken manure. Bioresour. Technol. 164, 309314.
A critical review on anaerobic co-digestion achievements between 2010 and Wu, L.-J., Kobayashi, T., Kuramochi, H., Li, Y.-Y., Xu, K.-Q., 2016. Improved biogas
2013. Renewable Sustainable Energy Rev. 36, 412427. production from food waste by co-digestion with de-oiled grease trap waste.
Micolucci, F., Gottardo, M., Bolzonella, D., Pavan, P., 2014. Automatic process control Bioresour. Technol. 201, 237244.
for stable bio-hythane production in two-phase thermophilic anaerobic Yan, B.H., Selvam, A., Wong, J.W.C., 2016. Innovative method for increased methane
digestion of food waste. Int. J. Hydrogen Energy 39, 1756317572. recovery from two-phase anaerobic digestion of food waste through
Montecchio, D., Gallipoli, A., Gianico, A., Pagliaccia, P., Mininni, G., Braguglia, C.M., reutilization of acidogenic off-gas in methanogenic reactor. Bioresour.
2016. Biomethane potential of food waste: modeling the effects of mild thermal Technol. 217, 39.
pretreatment and digestion temperature. Environ. Technol. http://dx.doi.org/ Yang, L., Huang, Y., Zhao, M., et al., 2015. Enhancing biogas generation performance
10.1080/09593330.2016.1233293. from food wastes by high-solids thermophilic anaerobic digestion: effect of pH
Nagao, N., Tajima, N., Kawai, M., Niwa, C., Kurosawa, N., Matsuyama, T., Yusoff, F.M., adjustment. Int. Biodeterior. Biodegrad. 105, 153159.
Toda, T., 2012. Maximum organic loading rate for the single -stage wet Yeshanew, M.M., Frunzo, L., Pirozzi, F., Lens, P.N.L., Esposito, G., 2016. Production of
anaerobic digestion of food waste. Bioresour. Technol. 118, 210218. biohythane from food waste via an integrated system of continuously stirred
Naran, E., Toor, U.A., Kim, D.J., 2016. Effect of pretreatment and anaerobic tank and anaerobic fixed bed reactors. Bioresour. Technol. 220, 312322.
codigestion of food waste and waste activated sludge on stabilization and Yin, Y., Liu, Y., Meng, S., Kiran, E.U., Liu, Y., 2016. Enzymatic pretreatment of
methane production. Int. Biodeterior. Biodegrad. 113, 1721. activated sludge, food waste and their mixture for enhanced bioenergy recovery
Nathao, C., Sirisukpoka, U., Pisutpaisal, N., 2013. Production of hydrogen and and waste volume reduction via anaerobic digestion. Appl. Energy 179, 1131
methane by one and two stage fermentation of food waste. Int. J. Hydrogen 1137.
Energy 38, 1576415769.

Please cite this article in press as: Braguglia, C.M., et al. Anaerobic bioconversion of food waste into energy: A critical review. Bioresour. Technol. (2017),
http://dx.doi.org/10.1016/j.biortech.2017.06.145
20 C.M. Braguglia et al. / Bioresource Technology xxx (2017) xxxxxx

Yirong, C., Heaven, S., Banks, C.J., 2015. Effect of a trace element addition strategy on Zhang, C., Su, H., Baeyens, J., Tan, T., 2014. Reviewing the anaerobic digestion of food
volatile fatty acid accumulation in thermophilic anaerobic digestion of food waste for biogas production. Renewable Sustainable Energy Rev. 38, 383392.
waste. Waste Biomass Valorization 6, 112. Zhang, W., Wu, S., Guo, J., Zhou, J., Dong, R., 2015a. Performance and kinetic
Yong, Z., Dong, Y., Zhang, X., Tan, T., 2015. Anaerobic co-digestion of food waste and evaluation of semi-continuously fed anaerobic digesters treating food waste:
straw for biogas production. Renewable Energy 78, 527530. role of trace elements. Bioresour. Technol. 178, 297305.
Zamanzadeh, M., Hagen, L.H., Svensson, K., Linjordet, R., Horn, S.J., 2016. Anaerobic Zhang, W., Zhang, L., Li, A., 2015b. Anaerobic co-digestion of food waste with MSW
digestion of food waste effect of recirculation and temperature on incineration plant fresh leachate: process performance and synergistic effects.
performance and microbiology. Water Res. 96, 246254. Chem. Eng. J. 259, 795805.
Zhang, R., El-Mashad, H.M., Hartman, K., Wang, F., Liu, G., Choate, C., Gamble, P., Zhang, J., Lv, C., Tong, J., Liu, J., Liu, J., Yu, D., Wang, Y., Chen, M., Wei, Y., 2016.
2007. Characterization of food waste as feedstock for anaerobic digestion. Optimization and microbial community analysis of anaerobic co-digestion of
Bioresour. Technol. 98, 929935. food waste and sewage sludge based on microwave pretreatment. Bioresour.
Zhang, Y., Banks, C.J., Heaven, S., 2012. Co-digestion of source segregated domestic Technol. 200, 253261.
food waste to improve process stability. Bioresour. Technol. 114, 168178. Zhu, H., Parker, W., Basnar, R., Proracki, A., Falletta, P., Bland, M., Seto, P., 2008.
Zhang, C., Su, H., Tan, T., 2013a. Batch and semi-continuous anaerobic digestion of Biohydrogen production by anaerobic co-digestion of municipal food waste and
food waste in a dual solidliquid system. Bioresour. Technol. 145, 1016. sewage sludges. Int. J. Hydrogen Energy 33, 36513659.
Zhang, Z.-L., Zhang, L., Zhou, Y.-L., Chen, J.-C., Liang, Y.-M., Wei, L., 2013b. Pilot-scale Zou, L., Ma, C., Liu, J., Li, M., Ye, M., Qian, G., 2016. Pretreatment of food waste with
operation of enhanced anaerobic digestion of nutrient-deficient municipal high voltage pulse discharge towards methane production enhancement.
sludge by ultrasonic pretreatment and co-digestion of kitchen garbage. J. Bioresour. Technol. 222, 8288.
Environ. Chem. Eng. 1, 7378.

Please cite this article in press as: Braguglia, C.M., et al. Anaerobic bioconversion of food waste into energy: A critical review. Bioresour. Technol. (2017),
http://dx.doi.org/10.1016/j.biortech.2017.06.145

Você também pode gostar