Вы находитесь на странице: 1из 10

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/326528268

Conversion of Food Waste to Fermentation Products

Chapter · January 2019


DOI: 10.1016/B978-0-08-100596-5.22294-4

CITATION READS

1 3,880

4 authors, including:

Muhammad Waqas Mohammad Rehan


Kohat University of Science and Technology King Abdulaziz University
29 PUBLICATIONS   303 CITATIONS    100 PUBLICATIONS   1,399 CITATIONS   

SEE PROFILE SEE PROFILE

Dr. Abdul-Sattar Nizami


King Abdulaziz University
145 PUBLICATIONS   2,774 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

analysis pm2.5 View project

An approach to non-chemical and eco-friendly weed management in maize View project

All content following this page was uploaded by Muhammad Waqas on 11 August 2018.

The user has requested enhancement of the downloaded file.


Conversion of Food Waste to Fermentation Products
Muhammad Waqasa, Mohammad Rehana, Muhammad Daud Khanb, and Abdul-Sattar Nizamia, a Center of Excellence in
Environmental Studies (CEES), King Abdulaziz University, Jeddah, Saudi Arabia; and b Department of Environmental Sciences, Kohat
University of Science and Technology (KUST), Kohat, Pakistan
© 2018 Elsevier Inc. All rights reserved.

Abstract 1
Introduction 1
Food Waste to Fermentation Products 2
Lactic Acid 2
Ethanol 3
Biohydrogen (H2) 3
Biogas 4
Volatile Fatty Acids (VFAs) 5
Techno-Economical Approaches and Prospective 5
Technical Challenges and Solutions 5
Economics 6
Towards Commercialization of Food Waste Fermentation 6
Future Research and Conclusions 7
References 7

Abstract

The landfill disposal of the massive amount of food waste without treatment and resource recovery is resulting in several public and
environmental health concerns. Several technologies have emerged for the conversion of food waste to lactic acid, ethanol, biogas,
biohydrogen and volatile fatty acids (VFAs) as value-added products. Food waste is a rich source of essential components such as
protein, carbohydrate, oil, mineral, and fat that can be converted to many value-added products as mentioned above. The conver-
sion of food waste to fermentation products such as organic acids, gases, and alcohols requires precise control and optimization of
operational conditions, including pretreatment, pH, temperature, and microbes. Therefore, the fermentation technologies for food
waste are still developing to solve the technical challenges of pretreatment such as the process economics, reactor design and infra-
structure cost and lack of homogeneity in the results of laboratory and large-scale plants. A potential way forward is to optimize the
fermentation process conditions along with implementing the strategies to integrate different waste treatment technologies to
produce high-quality and cost-effective value-added products at commercial scale.

Introduction

Food waste, from kitchen, canteen, food-processing and restaurant waste, is an essential component of municipal solid waste
(MSW) and its production has become a global concern (Ren et al., 2017). According to FAO (2012), about 1.3 billion tons of
food in the form of fruits, bakery, bread, vegetables, dairy products, and meat are lost every year through food supply chain world-
wide. With increasing population, economic growth and living standards, the food waste is projected to further increase in next 25
years (Kiran et al., 2014). In the United States (US), about 38 million tons of food waste is produced every year that is 50%
increase as compared to 1947 (USEPA, 2016; Posmanik et al., 2017). In the European Union (EU), the food waste generation
is up to 98 million tons per year and projected to reach around 139 million tons by 2020 (European Communities, 2010). In
Asian countries, China is the largest waste producing country with the production of more than 90 million tons of food waste,
which is about 37%–62% of the total MSW of China (Zhang et al., 2014). In developing countries, most of the food waste is
disposed to landfills as an easy option due to a limited budget, infrastructure, and resources. However, such landfill disposal
of food waste, containing high organic contents, result in serious public and environmental health issues such as air, soil and
groundwater contamination, disease-causing vectors, greenhouse gas (GHG) emissions, waterborne pollutants, waste leachate
and odors (Waqas et al., 2018).
The conventional methods of waste treatment like composting, incineration, animal feed production and anaerobic digestion
(AD) are used to manage food waste (Thi et al., 2015). Food waste is a rich source of various vital components such as protein,
carbohydrate (hemicellulose, cellulose, starch, and sugar like sucrose, fructose, and glucose), oil, mineral, and fat that can be
used in a wide range of enzymatic and microbial processes (Pham et al., 2015). The total protein and sugar contents in food waste
are around 22%–60% respectively (Table 1). Hydrolysis of carbohydrates present in the food waste results in the bond breakage of
glycoside with the release of monosaccharides and oligosaccharides which are much acquiescent to the fermentation process. Food

1
2 Conversion of Food Waste to Fermentation Products

Table 1 Percentage composition of FW

Study MC OM Starch Sugars Lipid Cellulose Protein

Ohkouchi and Inoue, 2006 75.9 – 29.3 42.3 – – 3.9


Tang et al., 2008 80.3 95.4 – 59.8 15.7 1.6 21.8
Wang et al., 2008 75.5 – 46.1 50.2 18.1 – 15.6
He et al., 2012 81.7 87.5 – 35.5 24.1 3.9 14.4
Vavouraki et al., 2013 81.5 94.1 24 55 14 16.9 16.9

waste is also rich in moisture content, and organic matter with high biodegradability rates which make it a promising feedstock for
biogas production using AD process (Ren et al., 2017). Food waste has also been used as a sole microbial feedstock to produce
various bioproducts such as ethanol, methane (CH4), biohydrogen (H2), organic acids, enzymes and biopolymers (Kiran et al.,
2014).
Intensive research work has been carried out on the biochemical conversion of food waste to biofuels, chemicals, biodegradable
polymers and chemical intermediates (Ren et al., 2017; Kiran et al., 2014). Lactic acid, succinic acid, butanol, and 3-hydroxybutyrate
have been successfully produced by fermentation of food waste (Maina et al., 2017). Recently, the waste valorization for the extrac-
tion of different marketable components has gained significant interest in both public and scientific community. Therefore, this
module is especially designed to review the conversion of food waste to various fermentation products. A techno-economic analysis
and future perspective of food waste to fermentation products is also provided.

Food Waste to Fermentation Products


Lactic Acid
Lactic acid is a naturally occurring organic material in the form of hydroxycarboxylic acid. In 1780, it was first defined by
a Swedish chemist Scheele from sour milk (Wang et al., 2015). Due to its wide range of applications as a flavor enhancer, acidu-
lant, and preservative, it has occupied a critical position in the pharmaceutical, food, chemical and cosmetic industries. Lactic
acid is a chiral compound consist of two optical isomers, L-(þ)-LA and D-()-LA. The pure form of lactic acid is highly preferred
for its specific applications. For instance, L-(þ)-LA is easily assimilated by the human body. It is the preferred isomer in the drug
and food industries (Abdel-Rahman et al., 2011). Moreover, lactic acid can be commercially prepared using the chemical or
biotechnological processes through lactic acid fermentation (Wang et al., 2015). About 90% of lactic acid is prepared using
microbial fermentation and 10% by chemical synthesis. Due to low environmental concerns, energy requirements, and produc-
tion temperature with high purity, the microbial fermentation has become the primary method to produce lactic acid (Wee et al.,
2006).
The cost of raw materials used for lactic acid fermentation is more than 34% of the total production cost (Åkerberg and Zacchi,
2000). Therefore, to overcome the challenge of high production cost, different organic waste sources like organic fraction of munic-
ipal solid waste (OFMSW), potato peel, fruit and vegetable wastes and food waste as substrate have been successfully examined for
lactic acid fermentation (Tang et al., 2017). It has been observed that the microbial fermentation of organic waste results in higher
lactic acid yield (Tang et al., 2016). However, the major challenge of utilizing such renewable materials for lactic acid production is
the cost of pretreatment. Without pretreatment, such renewable materials are not readily available for fermentation due to the pres-
ence of impurities, their close association with lignin and lack of production of a hydrolytic enzyme by lactic acid producing strains
(Abdel-Rahman et al., 2011). In order to overcome the limitation of pretreatment, many studies have been conducted to produce
lactic acid from acidogenic fermentation during the AD. The process of the AD is comprised of four steps such as hydrolysis, acido-
genesis, acetogenesis, and methanogenesis. During the first two steps, lactic acid is produced (Tang et al., 2017). Optimizing the
operating conditions like inoculum pH, temperature, and C/N ratio is crucial in obtaining high yield lactic acid during the AD
process (Tang et al., 2016).
The pH has a critical influence on hydrolysis, acidogenesis and microbial communities during anaerobic lactic acid fermenta-
tion. Using food waste as a substrate, the pH increase from 4 to 5 promotes the hydrolysis rate and acidification process (Wu
et al., 2015). However, pH rise above 6 further converts the produced lactic acid to biogas and volatile fatty acids (VFAs) (Probst
et al., 2015). The reason behind the pH effect is the metabolic pathways that play a vital role in dominating specific microbial
communities for the production of various intermediates (Wu et al., 2015). Inoculum is also an essential factor that affects fermen-
tative pathways (Liang et al., 2016). For instance, inoculation of axenic microbial cultures like Lactococcus and Lactobacillus gives
a high yield, and optically pure lactic acid can be generated using food waste, starch, green biomass, lignocellulosic and refined
sugars as feedstocks (Wakai et al., 2014; Tashiro et al., 2016). Similarly, among the operational parameters temperature is another
critical factor that influences substrate conversion rate, microbial activity and the economic ratio (Tang et al., 2017). Kim et al.
(2012) have found optimum lactic acid yield (23 g COD/L) at 50–55  C. Higher temperature promotes the hydrolysis rate.
Conversely, Liang et al. (2014) have studied the effect of different temperature on lactic acid fermentation and found that the lactic
acid yield was decreased from 0.22 g/g-TS at 37  C to 0.088 g/g-TS lactic acid at 55  C. The possible reason for a lower yield of lactic
acid at 55  C could be the unsuitability of high temperature for the growth of lactic acid bacteria (Zhang et al., 2007a,b,c).
Conversion of Food Waste to Fermentation Products 3

Ethanol
Ethanol is one of the most promising product obtained from fermentation of renewable materials. In addition to its applications as
a fuel, it is widely used as a feedstock to produce different industrial materials such as ethylene which has the annual market
demand of above 140 million tons. Ethylene is further used in the manufacturing of polyethylene and other polymers. Ethanol
is mainly produced by microbial catalyzed fermentation using yeast. The most common feedstock for ethanol fermentation is
starchy raw materials such as sorghum corn, potato and wheat (Pietrzak and Kawa-Rygielska, 2014). Recently, food waste in the
form of wheat-rye bread, kitchen waste, mixed food waste, potato and banana peel has been successfully examined as feedstocks
for bioethanol production (Table 2). However, the complex lignocellulosic nature of food waste requires pretreatment including
enzymatic and thermal process, and alkali and acid treatment with an ambition to enhance the digestibility of cellulose and starch
(Pham et al., 2015).
Among the pretreatment methods, enzymatic hydrolysis is the standard technique. It is a two-step process in which the lique-
faction of starch is first carried out by a-amylase (EC 3.2.1.1) to produce short-chained dextrins. During this step, the breakage of a-
1,4-glycosidic bond occurs at amylopectin and amylose chains. In the second step, glucoamylase (EC 3.2.1.3) saccharifies the
dextrins to produce monomeric sugars (glucose). However, to produce free amino nitrogen and fermentable sugars, different
enzymes like pullulnases, cellulases and proteases are added as a nutrient source for microbes (yeast) (Sapi nska et al., 2013). After
hydrolysis, the obtained mash is subjected to ethanol fermentation through inoculation with yeast. Distillation is carried out after
fermentation to obtain pure ethanol. This process of ethanol production is known as separate hydrolysis and fermentation (SHF).
Ethanol yields of 29.1 and 32.2 g/L from food waste treated with amylases were reported by Moon et al. (2009) and Uncu and
Cekmecelioglu (2011) respectively. The recent development in this technology of making ethanol from starchy feedstock is the
direct conversion of starch by granular starch hydrolyzing enzymes (GSHE). These enzymes are obtained from genetically modified
Trichoderma reesei and show the activity of glucoamylase and a-amylase. Pietrzak and Kawa-Rygielska (2014) have compared the
waste bread fermentation with different pretreatment methods and without pretreatment. They proved that fermentation with
pretreatment techniques such as sonification, microwave irradiation, and enzymatic pre-hydrolysis increased the ethanol yields
to about 12–35 g/kg in comparison to fermentation without pretreatment. However, high yield of ethanol (354 g/kg of raw mate-
rial) was observed in the direct conversion of waste bread to ethanol using GSHE.

Biohydrogen (H2)
Food waste can be converted to biohydrogen (H2) through a number of biotechnological processes. These processes include one-
stage H2 fermentation, two-stage H2þCH4 fermentation, and photo and dark fermentation coupled with the AD (Alibardi and
Cossu, 2016). In the biorefinery concept, the central role for H2 production from food waste is being favored by dark fermentation
due to its low energy requirements (Tawfik et al., 2011; Alibardi et al., 2014). In comparison to photo-fermentation, dark fermen-
tation has a faster H2 production. Likewise, utilizing cheaply available feedstock such as food waste could further enhance the
economic benefit of the process. Therefore, food waste utilization for H2 production is simultaneously solving waste problem
and producing renewable energy (Table 3). However, food waste to H2 fermentation is limited by various factors such as the
requirement of pretreatment processes, substrate type, origin and type of inoculum, reactor configuration, temperature, and avail-
ability of micro-nutrients. Considerable variation and gaps in the scientific data have been reported because of the influence of these
factors on H2 production using organic waste (De Gioannis et al., 2013).

Table 2 Ethanol production from FW

Study Reactor type Capacity (mL) Mic: Inoculum Time (hours) Yield (g/g FW)

Tang et al., 2008 Separate 450 Saccharomyces cerevisiae 15 0.03


Continuous
Ma et al., 2008 Simultaneous batch 250 Zymomonas mobilis 14 0.07
Kim et al., 2008 Separate vessel 500 S. cerevisiae 16 0.49
Uncu and Cekmecelioglu, 2011 Separate fermenter 250 S. cerevisiae 96 0.20
Hong and Yoon, 2011 Simultaneous 250 S. cerevisiae 48 0.22

Table 3 Biohydrogen production from FW

Study Reactor type pH Temp ( C) Yield (mL/gVS)

Boni et al. (2013) Batch 5.5 36 70.34


Favaro et al. (2013) Batch 7.0 35 16–70
Redondas et al. (2012) Continuous 5.5 34 13.1–20.5
Shin et al. (2004) Batch 6.5 50 91.5
Han and Shin (2004) Continuous – 34 155
4 Conversion of Food Waste to Fermentation Products

The composition of waste is another important characteristic that affects H2 yield in the fermentation process (Alibardi and
Cossu, 2016). It has been reported that the substrates rich in carbohydrates result in high H2 yield in comparison to protein and
lipid-rich substrates (De Gioannis et al., 2013). However, the stored nutrients in the form of macromolecules need to be broken
down into accessible forms like free amino nitrogen and glucose before microbial utilization for fermentative H2 production
(Han et al., 2015). Therefore, various pretreatments techniques have been developed for the conversion of macromolecules into
essential components (De Gioannis et al., 2013). A promising technique is an enzymatic hydrolysis that in addition to nutrient
release could also speed up the hydrolysis of food waste. Other pretreatment techniques include sonification and heat treatment.
Several research studies have found that in comparison to untreated food waste, heat treatment results in high H2 production
without harming H2 consuming bacteria. Elbeshbishy et al. (2011) have applied sonification of food waste without inoculum addi-
tion for enhanced H2 production. Their results showed that pretreatment is an essential parameter to enhance H2 production from
food waste.
There are several other factors such as moisture content, organic matter, nutrient concentration, particle size, chemical oxygen
demand (COD) and biodegradability that also influence the yields of H2 (Zhang et al., 2007a,b,c). The research study of Ismail et al.
(2009) has reported the optimum H2 yield and production of 120 mL/g carbohydrate and 35.69 mL/h respectively at controlled
COD of 200 g/L food waste. Similarly, Han and Shin (2004) have obtained high H2 yield at a controlled moisture content of
food waste. The required C:N ratio for optimal yield is up to 20. The high carbon content and presence of indigenous microbial
consortium of food waste solely make it a suitable feedstock for H2 production. Numerous studies have reported higher H2 yields
using food waste as feedstock without adding the inoculum (Kim et al., 2011).
The H2-consuming bacteria is one of the critical yield-limiting factors that needs to be controlled to favor the growth of H2
producing bacteria (Yasin et al., 2013). Food waste contains a mixture of different microbial consortiums like H2 producing
bacteria, and acid, and CH4 producing bacteria. Therefore, the mixed microbial culture needs to be pretreated by chemical, heat,
or pH shock to eliminate H2 consuming bacteria and promote the germination of H2 producing bacteria (Kim and Shin, 2008).
H2 producing bacteria is resistant to high heat and pH shock. Kim et al. (2011, 2009) have treated food waste at 90  C for
20 min for promoting the germination of H2 producing bacteria. They have found optimum H2 yield and production (1.98 mol
H2/mol hexoseconsumed and 148.7 mL H2/g VSadded) by increasing the pretreatment temperature to 90  C. In another study, Kim
et al. (2010) have applied potassium hydroxide (alkali treatment with pH 12.5) to increase H2-production with the seeded culture
of heat treated sewage sludge.

Biogas
Biogas production is one of the most promising solutions for organic waste management due to a renewable energy source, less
production cost and low production of residual waste (Kiran et al., 2014). In addition, the digestate produced by the AD is
a nutrient-rich product that could be used as a soil conditioner and organic fertilizer. Various research studies conducted on
food waste to biogas production are summarized in Table 4. Among the feedstocks for biogas production, food waste is most prom-
ising due to its wide availability and heterogeneous composition with high energy content (Paritosh et al., 2017). Forster-Carneiro
et al. (2008a) have studied the process yield of the AD using food waste and a shredded OFMSW. They have found the CH4 yield of
0.18 m3/kg volatile solid added (VSadded) for food waste. During the 1950s, the successful design of pilot and commercial AD plants
received substantial attention worldwide (Karagiannidis and Perkoulidis, 2009). Several research studies have shown that 1 m3 of
biogas produced via AD is equivalent to about 21 MJ energy, which can generate around 2.04 kWh of electricity at 35% process
efficiency (Murphy et al., 2004). However, the one drawback of biogas production through the AD is its long operation period
that ranges from 20 to 40 days. Zhang et al. (2007a,b,c) have conducted a batch study on methanization of food waste for 10
and 28 days. They observed the optimum CH4 yield (0.435 m3/kg VS) after 28 days of digestion with VS removal of 81% followed
by 0.348 m3/kg VS after 10 days of digestion.
Food waste, being a rich source of organic components, is an excellent choice for the AD, but the presence of high salt concen-
trations and cations such as potassium, sodium, magnesium, and calcium could inhibit the digestion process (Chen et al., 2008). In
addition, the organic and nitrogen (N) rich feedstocks produce a high concentration of free ammonia (NH3) that could be toxic to
methanogenic bacteria (Chen et al., 2008). Co-digestion with wastes containing lower lipid and nitrogen contents are preferably
used to control such issues. Co-digestion generally neutralizes the substrates and decreases N concentration, thus reduces the

Table 4 Biogas production from FW

Study Pretreatment Reactor type Time (days) Yield (mL/g VS) Efficiency (VS %)

Heo et al., 2004 Freeze drying 2 stage vessel with 8 L capacity 120 Up to 482 90
Trzcinski and Stuckey, Enzymatic 2 stage vessel with 2.7 L capacity 75 – 61
2011 pretreatment
Latif et al., 2012 Blending Two-stage Hydrolytic (10 L) and methanogenic 19 357 81
(3 L) reactor
Zhang and Jahng, 2012 Adding trace element Single stage vessel with 150 mL capacity 368 Up to 450 –
Dai et al., 2013 No Single stage vessel with 3 L capacity 72 455 92.2
Conversion of Food Waste to Fermentation Products 5

accumulation of intermediate products such as NH3 and volatile compounds (Castillo et al., 2006). Different research studies
related to the AD have proved that co-digestion of food waste with MSW have enhanced biogas yield by 40%–50% in comparison
to food waste as a mono-feedstock. Parawira et al. (2004) have carried out co-digestion batch tests with different combinations of
sugar beet leaves and potato waste. They observed that high CH4 yield was 0.68 m3/kg VSadded for mixing at 16:24% total solid. The
observed CH4 yield from potato waste alone was 0.42 m3/kg VS.
Alvarez and Lidén (2008) have examined various combinations of food waste, animal manure and slaughterhouse waste under
mesophilic anaerobic conditions. Their results demonstrated that higher CH4 yields (0.3–1.3 m3/kg VSadded) were recorded for
anaerobic co-digestion. They have further concluded that anaerobic co-digestion process facilitates the degradation of wastes
that cannot be efficiently treated alone. To further optimize the biological process yield, quality of the substrate and pretreatments
such as thermal, chemical, physical and biological techniques play a fundamental role in the mass transformation of the substrate
during each phase of the AD process (Pham et al., 2015).
In addition to pretreatment techniques, various types of reactors have been designed and used for the AD process. There are three
reactor systems commonly used for the AD, including continuous one stage, continuous two stage and batch scale reactors. Forster-
Carneiro et al. (2008b) have optimized the AD process of food waste at two different inoculums ratios (20%–30% of mesophilic
sludge) using 6 reactors with different total solid (TS) concentrations of 20%, 25%, and 30%. They have found that the best perfor-
mance in terms of higher CH4 production (0.49 m3 kg1 VSadded) and food waste digestion was observed for 30% of inoculums and
20% TS during 20 and 60 days of operation respectively.

Volatile Fatty Acids (VFAs)


VFAs are among the essential intermediates produced when organic waste is treated in the AD process. They are produced during
acidogenesis and acetogenesis stages of the AD. Being a potential renewable carbon source, VFAs have various applications
including biodiesel production, polymers synthesis and N removal (Chen et al., 2013; Zhou et al., 2018). The storage of VFAs is
much safer and easier than biogas. Moreover, VFAs have a higher economic value of up to 130 $/ton (Fei et al., 2015) than biogas
(0.72 $/m3) (Oleskowicz-Popiel et al., 2012). Therefore, VFAs are considered as a more attractive product from food waste fermen-
tation. During the AD process, the hydrolysate monomers are first converted to propionate, acetate, alcohols, butyrate, CO2 and H2
by acidogenic bacteria. Afterward, propionate, acetate, alcohols, and butyrate are further converted to acetate through acetogenic
pathways. The main VFAs products of acidogenic fermentation are butyrate, acetate, and propionate (Jiang et al., 2013).
The strategies for increasing the production of VFAs during the AD process include 1) improving the process of acidogenesis, 2)
eliminating the inhibiting factors and 3) improving the hydrolysis rate to produce soluble substrates. Improving the hydrolysis rate
is targeted to increase the availability of carbon for its conversion to VFAs. Therefore, during acidogenic fermentation, hydrolysis is
considered as a rate-limiting step (Kim et al., 2005). However, the optimization of hydrolysis rate is directly related to operational
parameters such as pretreatment, pH, and temperature of food waste before fermentation. Pretreatment of food waste during anaer-
obic fermentation enhances the production of soluble COD that is an essential intermediate in linking the hydrolysis and acido-
genesis (Fdez-Güelfo et al., 2011). Therefore, pretreatment of food waste for the acidogenic fermentation is a promising technique
to increase the production of VFAs. There are several methods used for pretreatment of food waste such as biological (enzymes),
physical (microwave, thermal and ultrasound) and chemical (alkaline and acid) methods. Kim et al. (2005) have studied the acido-
genic fermentation of food waste under the effect of enzymatic, thermal and the combined thermal-enzymatic pretreatment. They
have found that all the tested pretreatment techniques have increased the production of sCOD generation and VFAs production.
However, the optimum production of VFAs was observed for the combined thermal-enzymatic treatment.
The food waste degradation also requires high efficient microbial communities. Therefore the critical drivers for process speed up
are specific functioning inoculums. The activities of methanogens must be inhibited in order to reduce the consumption but
improve the yield of VFAs. Several techniques have been successfully applied for inhibiting the activities of methanogens. These
include pH control, thermal pretreatment, and the addition of the inhibitor. It has also been reported that extremely high or
low pH also inhibits the activities of methanogens. The favorable pH for the activities of methanogens ranges from 7.8 to 8.2, hence
pH of the digester needs to be adjusted to inhibit methanogenesis and favor VFAs production (Chaganti et al., 2011; Wang et al.,
2014). Temperature, as like pH, also affects microbial biomass, enzymatic activities, and hydrolysis of the substrate (Kim et al.,
2003). Many research studies have been conducted to improve VFAs production by heating the inoculum to 100  C or above before
fermentation to inactivate the non-pore forming (Yan et al., 2014). Usually, mesophilic range (35  C) is considered as economical
and efficient to produce VFAs (Jiang et al., 2013). However, the optimal temperature varies by examining the composition of VFAs.
For instance, propionate and acetate are generated at 45 and 35  C during food waste fermentation whereas butyrate is produced
above 55  C, followed by propionate and acetate (Jiang et al., 2013).

Techno-Economical Approaches and Prospective


Technical Challenges and Solutions
Food waste is a potential and cost-effective feedstock to produce fermentation products, but several challenges need to be addressed
appropriately to achieve high product quality and yield. Food waste to various fermentation products is an emerging area of
research, and therefore an in-depth understanding of all the aspects of food waste could help to overcome these challenges. For
6 Conversion of Food Waste to Fermentation Products

instance, a collection of segregated food waste is a major challenge as food waste is thrown and mixed with all other waste. In order
to overcome this limitation, a proper campaign highlighting the importance of food waste segregation at sources such as housing,
urban planning departments, and food industries is highly required (Karmee, 2016). In addition, policies to transport food waste to
collection facilities with proper sorting and preparation methods of food waste from non-biological wastes should be designed to
further facilitate the processing and utilization of the collected food waste. Starting large industries for food waste recycling would
need a continuous supply of food waste. However, it should be kept under consideration that sorting and separation would vary
from region to region depending on types of food waste. In this context, a large industry of sustaining food waste couldn't be prac-
tically implemented. Therefore, small processing plants could be attached to various food waste producers such as parks and restau-
rants. Such formulation would also reduce transportation cost.
The composition of food waste depends upon the local eating habits, area, and eating periods. Thus, the chemical composition
needs to be determined before utilizing it as a resource to produce fermented products (Karmee, 2016). In comparison to other
feedstocks like corn, plant oils, and lignocellulosic materials, food waste is more complex as it contains carbohydrates, amino acids,
lipid, vitamins, phosphates, and nutrients. Thus, proper characterization should be carried out for various types of food waste. The
separation and purification cost of carbohydrates, lipids and other organic materials from food waste alone is high enough in addi-
tion to the further requirement of volatile organic solvents which may cause environmental and public health issues. Instead, single
reaction system for simultaneous production of bioethanol, bio-oil and biodiesel should be developed without any isolation and
purification of carbohydrates, lipids and organic matter. Such approaches would reduce the operational cost and make the process
simpler (Karmee, 2016).
There are technical challenges for each technology converting food waste to fermentation products. The primary technical chal-
lenge is the difficult control of process conditions that result in the production of harmful intermediate compounds causing low
products yield and reducing the system stability. The high lipids and protein contents in food waste lead to the production of
hydrogen sulfide, NH3, and long chain fatty acids during fermentation (Xu et al., 2018). The other technical challenges are scaling
up of the technology, purification of end products and estimation of biomass that encourage the researchers to brainstorm to find
the possible solutions (Singhania et al., 2009). For example, scaling up has been a major challenge for solid state fermentation (SSF)
for a long time. However, the recent advent of biochemical engineering resulted in the designing of a number of bioreactors with
large-scale waste treatment capacity along with on-line monitoring of different process parameters including heat and mass transfer.
Even though, product recovery and, purification is still much expensive (Couto and Sanromán, 2005). Therefore, a detailed
economic and technical feasibility study must be conducted before process scale up.

Economics
On the basis of biofuel-based energy demand many developed countries in Europe, Asia and US have designed different economic
policies. In the near future, waste to biofuel market would be a major driving force for the economic growth. The availability of
a wide range of biomass, its collection, and conversion to value-added products would employ more people in the near future as
compared to conventional fossil fuel-based technologies (Kartha and Larson, 2000). Moreover, growing biofuel production would
reduce the cost of conventional fuels by reducing the dependency on petroleum fuel. In addition, the bioethanol and biodiesel, from
food waste and other renewable sources, as a transportation fuel could replace gasoline and diesel. However, the cost and availability
of feedstock significantly affect the price of biofuels. For instance, food waste is discarded and mixed with other waste hence the main
cost is the collection, sorting, transportation, and pretreatment. The techno-economic analysis would provide sufficient information
on methodology development, estimation cost of biofuel plant, cost of the production facility and real market data and cost of bio-
fuel (Karmee, 2016). Currently, there are limited detailed techno-economic reports been published on small or medium scale bio-
ethanol and biodiesel production from food waste. In addition to economics, there are also environmental benefits of using
biomass-based energy as transportation fuels, electricity and heat which are carbon dioxide neutral by recycling same carbon atoms
(Demirbas, 2008). Research investigation proved lower greenhouse gas (GHG) emissions from biofuels in comparison to conven-
tional fossil fuels (Huang et al., 2013).
Several factors need to be considered for biomass to biofuels technologies by both industrialized and developing nations. In
addition, strong public support system would also benefit biofuels production through indicative national targets. Such supports
include the encouragement of the feedstocks supply, i.e., cultivation of raw materials for biofuels production. Similarly, a significant
role could also be played by designing and encouraging the demand for biofuels. Approaches to such support are tax reductions and
biofuel obligations. With a significant amount of available biomass, it has been projected that by 2050 about one-half of total
energy demand would be fulfilled by biomass energy in the developing countries. In the future energy system, various estimations
have been put forward for biomass biofuel. For instance, the primary and essential energy source from biomass is methanol, hence
up to some extent the fuel choice for each sector is either transportation. The biomass availability determines electricity and heat. In
addition, the production cost also influences each biofuel conversion technology e.g. the conversion of biomass to H2 is cheaper
and more energy efficient technology as compared to methanol (Azar et al., 2003).

Towards Commercialization of Food Waste Fermentation


Development of sustainable waste fermentation industry requires an extensive supply chain (Hoekman, 2009). The first primary
element in this supply chain is the development of large industries for sufficient feedstock production around the year for
Conversion of Food Waste to Fermentation Products 7

fermentation plants. The organic waste could contribute significantly to meet this increasing demand of the feedstock. However,
bioengineering tools are required to be developed for selecting suitable technology and understanding the composition of feedstock
for fermentation. The development of cost-effective methods for sorting food waste, and its collection, transportation, storage and
pre-processing are highly needed. In addition, the continuous and reliable feedstock supply is also required for commercial-scale
biofuel production. For biofuel production, it is necessary to improve the process for cost-effective operation. In this regard, inte-
grated biorefineries should be developed and deployed in addition to the effective utilization of coproducts (heat and electricity) for
economic operation. The distribution of biofuel is the next step after the biofuel production. The most important factor is the
compatibility with existing infrastructure components such as storage, transportation, and dispensing in addition to quality control
to ensure that the produced biofuel meet all the standards and product specifications. The final stage is the consumption of biofuel
in the form of fuel, energy and other applications. For the consumer satisfaction towards biofuel consumption, it is highly required
to implement biofuels compatible equipment along with the performance equivalent to conventionally used equipment. The final
major deliberation in the entire supply chain is the process economics. The commercial success of each point along the chain for
waste to value-added products require favorable economics (Hoekman, 2009).

Future Research and Conclusions

The major challenges in the valorization of food waste to fermentation products are high construction cost of the digester, control of
the process conditions, and low-quality end products (Xu et al., 2018). The novel approaches to integrate food processing facilities
under a biorefinery concept could be developed to produce economical value-added products from food waste along with the
generation of heat and power to support the biorefinery. Such approaches would enhance the food waste valorization with better
economics and lower environmental impacts. For instance, the hydrothermal liquefaction of food waste to oil followed by
a carbon-rich feed of hydrothermal extract to produce biogas using the AD process (Posmanik et al., 2017). The key advantage
of this integration is to increase the energy recovery from food waste and easy handling of the end products. Similarly, the available
knowledge of enzymes and microbes in the fermentation of food waste is not enough to provide an effective control system. For
instance, how micronutrients and their concentrations affect the activities of microbes and enzymes (Romero-Güiza et al., 2016).
Therefore, the industrial scale digesters are still performing a trial and error approach for the supplementing micronutrients.
Further research is required to achieve homogeneous results at both lab and large-scale applications. For example, in a laboratory
scale fermenter, homogeneity for sufficient agitation and precise pH control could be guaranteed. However, the large-scale practical
implementation is still uncertain whether the pH would be precise and uniform throughout the fermenter, as insufficient agitation
may lead to dead zone resulting in the disturbance of the accuracy of pH control (Amanullah et al., 2001).
The scale-up challenges and the technological gaps might lead to lower process control and product quality (Moon et al., 2009).
The existing infrastructure of biogas facilities for treating on-farm manure and sewage sludge is also desirable for treating food waste
to save the additional labor and infrastructure cost. Xu et al. (2018) have suggested to add food waste to the existing biogas plants to
achieve higher electricity and heat production. The food waste composition dictates the choice of each fermentation technology. For
example, food waste rich in carbohydrates is suitable to be treated in a two-stage fermentation process for co-production of CH4 and
H2. In addition, carbohydrate-rich food waste could also be converted to heat. Whereas, food waste rich in protein, COD and total
solid could be directly converted to biogas via single stage fermentation.
The food waste could also be directly hydrolyzed to form hydrolysate using proteases enzyme. The produced hydrolysate
comprised of phosphates, amino acids, carbohydrates and nutrients that could be used as growth medium for microbial growth
(Pleissner et al., 2014). For instance, microalgae that is rich in lipid biomass could be grown on hydrolysate. The extracted lipids
from microalgae could be converted to epoxides, surfactants, and biodiesel (Gude et al., 2013). The hydrolysate is also rich in carbo-
hydrate and hence could be used for bioethanol production. However, based on this concept more research is needed in addition to
cross public-private and industrial collaborations for designing such sustainable systems.

References

Abdel-Rahman, M.A., Tashiro, Y., Sonomoto, K., 2011. Lactic acid production from lignocellulose-derived sugars using lactic acid bacteria: overview and limits. J. Biotechnol. 156,
286–301.
Åkerberg, C., Zacchi, G., 2000. An economic evaluation of the fermentative production of lactic acid from wheat flour. Bioresour. Technol. 75, 119–126.
Alibardi, L., Cossu, R., 2016. Composition variability of the organic fraction of municipal solid waste and effects on hydrogen and methane production potentials. Waste Manag. 36,
147–155.
Alibardi, L., Muntoni, A., Polettini, A., 2014. Hydrogen and waste: illusions, challenges and perspectives. Waste Manag. 34, 2425–2426.
Alvarez, R., Lidén, G., 2008. Semi-continuous Co-digestion of solid slaughterhouse waste, manure, and fruit and vegetable waste. Renew. Energy 33 (4), 726–734.
Amanullah, A., McFarlane, C.M., Emery, A.N., Nienow, A.W., 2001. Scale-down model to simulate spatial pH variations in large-scale bioreactors. Biotechnol. Bioeng. 73,
390–399.
Azar, C., Lindgren, K., Andersson, B.A., 2003. Global energy scenarios meeting stringent CO2 constraintsdcost-effective fuel choices in the transportation sector. Energy Policy 31,
961–976.
Boni, M.R., Sbaffoni, S., Tuccinardi, L., 2013. The influence of slaughterhouse waste on fermentative H2 production from food waste: preliminary results. Waste Manag. 33 (6),
1362–1371.
Castillo, M.E.F., Cristancho, D.E., Arellano, A.V., 2006. Study of the operational conditions for anaerobic digestion of urban solid wastes. Waste Manag. 26 (5), 546–556.
8 Conversion of Food Waste to Fermentation Products

Chaganti, S.R., Kim, D.H., Lalman, J.A., 2011. Flux balance analysis of mixed anaerobic microbial communities: effects of linoleic acid (LA) and pH on biohydrogen production. Int. J.
Hydrogen Energy 36 (21), 14141–14152.
Chen, H., Meng, H., Nie, Z., Zhang, M., 2013. Polyhydroxyalkanoate production from fermented volatile fatty acids: effect of pH and feeding regimes. Bioresour. Technol. 128,
533–538.
Chen, Y., Cheng, J.J., Creamer, K.S., 2008. Inhibition of anaerobic digestion process: a review. Bioresour. Technol. 99 (10), 4044–4064.
Couto, S.R., Sanromán, M.A., 2005. Application of solid-state fermentation to ligninolytic enzyme production. Biochem. Eng. J. 22 (3), 211–219.
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 Manag. 33 (2), 308–316.
De Gioannis, G., Muntoni, A., Polettini, A., Pomi, R., 2013. A review of dark fermentative hydrogen production from biodegradable municipal waste fractions. Waste Manag. 33,
1345–1361.
Demirbas, A., 2008. Biofuels sources, biofuel policy, biofuel economy and global biofuel projections. Energy Convers. Manag. 49 (8), 2106–2116.
Elbeshbishy, E., Hafez, H., Nakhla, G., 2011. Ultrasonication for biohydrogen production from food waste. Int. J. Hydrogen Energy 36, 2896–2903.
European Communities, 2010. Preparatory Study on Food Waste across Eu 27, October. URL: http://ec.europa.eu/environment/eussd/pdf/bio_foodwaste_report.pdf.
FAO, 2012. Towards the Future We Want: End Hunger and Make the Transition to Sustainable Agricultural and Food Systems. Food Agriculture Organization of the United
Nations, Rome.
Favaro, L., Alibardi, L., Lavagnolo, M.C., Casella, S., Basaglia, M., 2013. Effects of inoculum and indigenous microflora on hydrogen production from the organic fraction of
municipal solid waste. Int. J. Hydrogen Energy 38 (27), 11774–11779.
Fdez-Güelfo, L.A., Álvarez-Gallego, C., Sales, D., Romero, L.I., 2011. The use of thermochemical and biological pretreatments to enhance organic matter hydrolysis and solubilization
from organic fraction of municipal solid waste (OFMSW). Chem. Eng. J. 168 (1), 249–254.
Fei, Q., Fu, R., Shang, L., Brigham, C.J., Chang, H.N., 2015. Lipid production by microalgae Chlorella protothecoides with volatile fatty acids (VFAs) as carbon sources in
heterotrophic cultivation and its economic assessment. Bioprocess Biosyst. Eng. 38 (4), 691–700.
Forster-Carneiro, T., Perez, M., Romero, L.I., 2008a. Thermophilic anaerobic digestion of source-sorted organic fraction of municipal solid waste. Bioresour. Technol. 99 (15),
6763–6770.
Forster-Carneiro, T., Perez, M., Romero, L.I., 2008b. Influence of total solid and inoculum contents on performance of anaerobic reactors treating food waste. Bioresour. Technol. 99
(15), 6994–7002.
Gude, V.G., Grant, G.E., Patil, P.D., Deng, S., 2013. Biodiesel production from low cost and renewable feedstock. Central Eur. J. Eng. 3 (4), 595–605.
Han, S.K., Shin, H.S., 2004. Biohydrogen production by anaerobic fermentation of food waste. Int. J. Hydrogen Energy 29, 569–577.
Han, W., Liu, D.N., Shi, Y.W., Tang, J.H., Li, Y.F., Ren, N.Q., 2015. Biohydrogen production from food waste hydrolysate using continuous mixed immobilized sludge reactors.
Bioresour. Technol. 180, 54–58.
Heo, N.H., Park, S.C., Kang, H., 2004. Effects of mixture ratio and hydraulic retention time on single-stage anaerobic co-digestion of food waste and waste activated sludge.
J. Environ. Sci. Health, Part a 39 (7), 1739–1756.
He, M., Sun, Y., Zou, D., Yuan, H., Zhu, B., Li, X., Pang, Y., 2012. Influence of temperature on hydrolysis acidification of food waste. Procedia Environ. Sci. 16, 85–94.
Hoekman, S.K., 2009. Biofuels in the US–challenges and opportunities. Renew. Energy 34 (1), 14–22.
Hong, Y.S., Yoon, H.H., 2011. Ethanol production from food residues. Biomass Bioenergy 35 (7), 3271–3275.
Huang, H., Khanna, M., Önal, H., Chen, X., 2013. Stacking low carbon policies on the renewable fuels standard: economic and greenhouse gas implications. Energy Policy 56,
5–15.
Ismail, F., Nor’Aini, A.R., Abd-Aziz, S., Chong, M.L., Hassan, M.A., 2009. Statistical optimization of biohydrogen production using food waste under thermophilic conditions. Open
Renew. Energy J. 2, 124–131.
Jiang, J., Zhang, Y., Li, K., Wang, Q., Gong, C., Li, M., 2013. Volatile fatty acids production from food waste: effects of pH, temperature, and organic loading rate. Bioresour.
Technol. 143, 525–530.
Karagiannidis, A., Perkoulidis, G., 2009. A multi-criteria ranking of different technologies for the anaerobic digestion for energy recovery of the organic fraction of municipal solid
wastes. Bioresour. Technol. 100 (8), 2355–2360.
Karmee, S.K., 2016. Liquid biofuels from food waste: current trends, prospect and limitation. Renew. Sustain. Energy Rev. 53, 945–953.
Kartha, S., Larson, E.D., 2000. Bioenergy Primer: Modernised Biomass Energy for Sustainable Development, Technical Report UN Sales Number E.00.III.B.6. United Nations
Development Programme, United Nations Plaza, New York, NY, USA.
Kim, D.H., Kim, S.H., Shin, H.S., 2009. Hydrogen fermentation of food waste without inoculums addition. Enzyme Microbes Technol. 45, 181–187.
Kim, D.H., Kim, S.H., Kim, K.Y., Shin, H.S., 2010. Experience of a pilot-scale hydrogen producing anaerobic sequencing batch reactor (ASBR) treating food waste. Int. J. Hydrogen
Energy 35, 1590–1594.
Kim, D.H., Kim, S.H., Jung, K.W., Kim, M.S., Shin, H.S., 2011. Effect of initial pH independent of operational pH on hydrogen fermentation of food waste. Bioresour. Technol. 102,
8646–8652.
Kim, D.H., Lim, W.T., Lee, M.K., Kim, M.S., 2012. Effect of temperature on continuous fermentative lactic acid (LA) production and bacterial community, and development of LA-
producing UASB reactor. Bioresour. Technol. 119, 355–361.
Kim, H.J., Choi, Y.G., Kim, D.Y., Kim, D.H., Chung, T.H., 2005. Effect of pretreatment on acid fermentation of organic solid waste. Water Sci. Technol. 52 (1–2), 153–160.
Kim, M., Gomec, C.Y., Ahn, Y., Speece, R.E., 2003. Hydrolysis and acidogenesis of particulate organic material in mesophilic and thermophilic anaerobic digestion. Environ. Technol.
24 (9), 1183–1190.
Kim, S.H., Shin, H.S., 2008. Effects of base-pretreatment on continuous enriched culture for hydrogen production from food waste. Int. J. Hydrogen Energy 33, 5266–5274.
Kim, J.K., Oh, B.R., Shin, H.J., Eom, C.Y., Kim, S.W., 2008. Statistical optimization of enzymatic saccharification and ethanol fermentation using food waste. Process Biochem. 43
(11), 1308–1312.
Kiran, E.U., Trzcinski, A.P., Ng, W.J., Liu, Y., 2014. Bioconversion of food waste to energy: a review. Fuel 134, 389–399.
Latif, M.A., Ahmad, A., Ghufran, R., Wahid, Z.A., 2012. Effect of temperature and organic loading rate on upflow anaerobic sludge blanket reactor and CH4 production by treating
liquidized food waste. Environ. Prog. Sustain. Energy 31 (1), 114–121.
Liang, S.B., Gliniewicz, K., Gerritsen, A.T., McDonald, A.G., 2016. Analysis of microbial community variation during the mixed culture fermentation of agricultural peel wastes to
produce lactic acid. Bioresour. Technol. 208, 7–12.
Liang, S.B., McDonald, A.G., Coats, E.R., 2014. Lactic acid production with undefined mixed culture fermentation of potato peel waste. Waste Manag. 34, 2022–2027.
Maina, S., Kachrimanidou, V., Koutinas, A., 2017. A roadmap towards a circular and sustainable bioeconomy through waste valorization. Curr. Opin. Green Sustain. Chem. 8,
18–23.
Ma, H., Wang, Q., Zhang, W., Xu, W., Zou, D., 2008. Optimization of the medium and process parameters for ethanol production from kitchen garbage by Zymomonas mobilis. Int.
J. Green Energy 5 (6), 480–490.
Moon, H.C., Song, I.S., Kim, J.C., Shirai, Y., Lee, D.H., Kim, J.K., Chung, S.O., Kim, D.H., Oh, K.K., Cho, Y.S., 2009. Enzymatic hydrolysis of food waste and ethanol fermentation.
Int. J. Energy Res. 33, 164–172.
Murphy, J.D., McKeogh, E., Kiely, G., 2004. Technical/economic/environmental analysis of biogas utilization. Appl. Energy 77 (4), 407–427.
Ohkouchi, Y., Inoue, Y., 2006. Direct production of L (þ)-lactic acid from starch and food wastes using Lactobacillus manihotivorans LMG18011. Bioresour. Technol. 97 (13),
1554–1562.
Conversion of Food Waste to Fermentation Products 9

Oleskowicz-Popiel, P., Kadar, Z., Heiske, S., Klein-Marcuschamer, D., Simmons, B.A., Blanch, H.W., Schmidt, J.E., 2012. Co-production of ethanol, biogas, protein fodder and
natural fertilizer in organic farming–evaluation of a concept for a farm-scale biorefinery. Bioresour. Technol. 104, 440–446.
Parawira, W., Murto, M., Zvauya, R., Mattiasson, B., 2004. Anaerobic batch digestion of solid potato waste alone and in combination with sugar beet leaves. Renew. Energy 29 (11),
1811–1823.
Paritosh, K., Kushwaha, S.K., Yadav, M., Pareek, N., Chawade, A., Vivekanand, V., 2017. Food waste to energy: an overview of sustainable approaches for food waste management
and nutrient recycling. Biomed Res. Int. https://doi.org/10.1155/2017/2370927.
Pham, T.P.T., Kaushik, R., Parshetti, G.K., Mahmood, R., Balasubramanian, R., 2015. Food waste-to-energy conversion technologies: current status and future directions. Waste
Manag. 38, 399–408.
Pietrzak, W., Kawa-Rygielska, J., 2014. Ethanol fermentation of waste bread using granular starch hydrolyzing enzyme: effect of raw material pretreatment. Fuel 134, 250–256.
Pleissner, D., Kwan, T.H., Lin, C.S.K., 2014. Fungal hydrolysis in submerged fermentation for food waste treatment and fermentation feedstock preparation. Bioresour. Technol.
158, 48–54.
Posmanik, R., Labatut, R.A., Kim, A.H., Usack, J.G., Tester, J.W., Angenent, L.T., 2017. Coupling hydrothermal liquefaction and anaerobic digestion for energy valorization from
model biomass feedstocks. Bioresour. Technol. 233, 134–143.
Probst, M., Walde, J., Pümpel, T., Wagner, A.O., Insam, H., 2015. A closed loop for municipal organic solid waste by lactic acid fermentation. Bioresour. Technol. 175, 142–151.
Redondas, V., Gómez, X., García, S., Pevida, C., Rubiera, F., Morán, A., Pis, J.J., 2012. Hydrogen production from food wastes and gas post-treatment by CO2 adsorption. Waste
Manag. 32 (1), 60–66.
Ren, Y., Yu, M., Wu, C., Wang, Q., Gao, M., Huang, Q., Liu, Y., 2017. A comprehensive review on food waste anaerobic digestion: research updates and tendencies. Bioresour.
Technol. 247, 1069–1076.
Romero-Güiza, M.S., Vila, J., Mata-Alvarez, J., Chimenos, J.M., Astals, S., 2016. The role of additives on anaerobic digestion: a review. Renew. Sustain. Energy Rev. 58,
1486–1499.
Sapinska, E., Balcerek, M., Pielech-Przybylska, K., 2013. Alcoholic fermentation of high-gravity corn mashes with the addition of supportive enzymes. J. Chem. Technol. Biotechnol.
88 (12), 2152–2158.
Shin, H.S., Youn, J.H., Kim, S.H., 2004. Hydrogen production from food waste in anaerobic mesophilic and thermophilic acidogenesis. Int. J. Hydrogen Energy 29 (13), 1355–1363.
Singhania, R.R., Patel, A.K., Soccol, C.R., Pandey, A., 2009. Recent advances in solid-state fermentation. Biochem. Eng. J. 44 (1), 13–18.
Tang, J., Wang, X.C., Hu, Y., Zhang, Y., Li, Y., 2017. Effect of pH on lactic acid production from acidogenic fermentation of food waste with different types of inocula. Bioresour.
Technol. 224, 544–552.
Tang, J., Wang, X., Hu, Y., Zhang, Y., Li, Y., 2016. Lactic acid fermentation from food waste with indigenous microbiota: effects of pH, temperature and high OLR. Waste Manag.
52, 278–285.
Tashiro, Y., Inokuchi, S., Poudel, P., Okugawa, Y., Miyamoto, H., Miayamoto, H., Sakai, K., 2016. Novel pH control strategy for efficient production of optically active lactic acid from
kitchen refuse using a mixed culture system. Bioresour. Technol. 216, 52–59.
Tang, Y.Q., Koike, Y., Liu, K., An, M.Z., Morimura, S., Wu, X.L., Kida, K., 2008. Ethanol production from kitchen waste using the flocculating yeast Saccharomyces cerevisiae strain
KF-7. Biomass Bioenergy 32 (11), 1037–1045.
Tawfik, A., Salem, A., El-Qelish, M., 2011. Two stage anaerobic baffled reactor for biohydrogen production from municipal food waste. Bioresour. Technol. 102, 8723–8726.
Thi, N.B.D., Kumar, G., Lin, C.Y., 2015. An overview of food waste management in developing countries: current status and future perspective. J. Environ. Manag. 157, 220–229.
Trzcinski, A.P., Stuckey, D.C., 2011. Parameters affecting the stability of the digestate from a two-stage anaerobic process treating the organic fraction of municipal solid waste.
Waste Manag. 31 (7), 1480–1487.
Uncu, O.N., Cekmecelioglu, D., 2011. Cost-effective approach to ethanol production and optimization by response surface methodology. Waste Manag. 31 (4), 636–643.
USEPA, 2016. Advancing Sustainable Materials Management: 2014 Fact Sheet. United States Environ. Prot. Agency, Off. L. Emerg. Manag, Washington, DC.
Vavouraki, A.I., Angelis, E.M., Kornaros, M., 2013. Optimization of thermo-chemical hydrolysis of kitchen wastes. Waste Manag. 33 (3), 740–745.
Wang, Q., Ma, H., Xu, W., Gong, L., Zhang, W., Zou, D., 2008. Ethanol production from kitchen garbage using response surface methodology. Biochem. Eng. J. 39 (3), 604–610.
Wakai, S., Yoshie, T., Asai-Nakashima, N., Yamada, R., Ogino, C., Tsutsumi, H., Hata, Y., Kondo, A., 2014. L-lactic acid production from starch by simultaneous saccharification and
fermentation in a genetically engineered Aspergillus oryzae pure culture. Bioresour. Technol. 173, 376–383.
Wang, K., Yin, J., Shen, D., Li, N., 2014. Anaerobic digestion of food waste for volatile fatty acids (VFAs) production with different types of inoculum: effect of pH. Bioresour. Technol.
161, 395–401.
Wang, Y., Tashiro, Y., Sonomoto, K., 2015. Fermentative production of lactic acid from renewable materials: recent achievements, prospects, and limits. J. Biosci. Bioeng. 119 (1),
10–18.
Waqas, M., Nizami, A.S., Aburiazaiza, A.S., Barakat, M.A., Ismail, I.M.I., Rashid, M.I., 2018. Optimization of food waste compost with the use of biochar. J. Environ. Manag. 216,
70–81.
Wee, Y.J., Kim, J.N., Ryu, H.W., 2006. Biotechnological production of lactic acid and its recent applications. Food Technol. Biotechnol. 44, 163–172.
Wu, Y.Y., Ma, H.L., Zheng, M.Y., Wang, K.J., 2015. Lactic acid production from acidogenic fermentation of fruit and vegetable wastes. Bioresour. Technol. 191, 53–58.
Xu, F., Li, Y., Ge, X., Yang, L., Li, Y., 2018. Anaerobic digestion of food waste–challenges and opportunities. Bioresour. Technol. 247, 1047–1058.
Yan, B.H., Selvam, A., Wong, J.W.C., 2014. Application of rumen microbes to enhance food waste hydrolysis in acidogenic leach-bed reactors. Bioresour. Technol. 168, 64–71.
Yasin, N.H.M., Mumtaz, T., Hassan, M.A., 2013. Food waste and food processing waste for biohydrogen production: a review. J. Environ. Manag. 130, 375–385.
Zhang, L., Jahng, D., 2012. Long-term anaerobic digestion of food waste stabilized by trace elements. Waste Manag. 32 (8), 1509–1515.
Zhang, C., Su, H., Baeyens, J., Tan, T., 2014. Reviewing the anaerobic digestion of food waste for biogas production. Renew. Sustain. Energy Rev. 38, 383–392.
Zhang, R., El-Mashad, H.M., Hartman, K., Wang, F., Liu, G., Choate, C., Gamble, P., 2007c. Characterization of food waste as feedstock for anaerobic digestion. Bioresour. Technol.
98, 929–935.
Zhang, R., El-Mashad, H.M., Hartman, K., Wang, F., Liu, G., Choate, C., Gamble, P., 2007b. Characterization of food waste as feedstock for anaerobic digestion. Bioresour. Technol.
98 (4), 929–935.
Zhang, Z.Y., Jin, B., Kelly, J.M., 2007a. Production of lactic acid from renewable materials by Rhizopus fungi. Biochem. Eng. J. 35, 251–263.
Zhou, M., Yan, B., Wong, J.W., Zhang, Y., 2018. Enhanced volatile fatty acids production from anaerobic fermentation of food waste: a mini-review focusing on acidogenic metabolic
pathways. Bioresour. Technol. 248, 68–78.

View publication stats

Вам также может понравиться