Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Environment International 127 (2019) 625–644

Contents lists available at ScienceDirect

Environment International
journal homepage: www.elsevier.com/locate/envint

Review article

Production of bioplastic through food waste valorization T


a,1 b,1 c,1 a d,1 e
Yiu Fai Tsang , Vanish Kumar , Pallabi Samadar , Yi Yang , Jechan Lee , Yong Sik Ok ,
Hocheol Songf, Ki-Hyun Kime, , Eilhann E. Kwonf, , Young Jae Jeong
⁎ ⁎

a
Department of Science and Environmental Studies, The Education University of Hong Kong, Tai Po, New Territories, Hong Kong
b
National Agri-Food Biotechnology Institute (NABI), S.A.S. Nagar, Punjab 140306, India
c
Department of Civil & Environmental Engineering, Hanyang University, 222 Wangsimni-Ro, Seoul 04763, Republic of Korea
d
Department of Environmental and Safety Engineering, Ajou University, Suwon 16499, Republic of Korea
e
Korea Biochar Research Center, O-Jeong Eco-Resilience Institute (OJERI), Division of Environmental Science and Ecological Engineering, Korea University, Seoul,
Republic of Korea
f
Department of Environment and Energy, Sejong University, Seoul 05006, Republic of Korea
g
Department of Microbiology, Pukyong National University, Pusan 48513, Republic of Korea

ARTICLE INFO ABSTRACT

Keywords: The tremendous amount of food waste from diverse sources is an environmental burden if disposed of in-
Food waste appropriately. Thus, implementation of a biorefinery platform for food waste is an ideal option to pursue (e.g.,
Valorization production of value-added products while reducing the volume of waste). The adoption of such a process is
Biorefinery expected to reduce the production cost of biodegradable plastics (e.g., compared to conventional routes of
Polyhydroxyalkanoates (PHA)
production using overpriced pure substrates (e.g., glucose)). This review focuses on current technologies for the
Biopolymer
Bioplastic
production of polyhydroxyalkanoates (PHA) from food waste. Technical details were also described to offer clear
insights into diverse pretreatments for preparation of raw materials for the actual production of bioplastic (from
food wastes). In this respect, particular attention was paid to fermentation technologies based on pure and mixed
cultures. A clear description on the chemical modification of starch, cellulose, chitin, and caprolactone is also
provided with a number of case studies (covering PHA-based products) along with a discussion on the prospects
of food waste valorization approaches and their economic/technical viability.

1. Introduction keting strategy. In general, the most FW is being disposed of via land-
filling, composting, and fermentation. Although European Union
The Food and Agriculture Organization (FAO) of the United Nations guidelines stated that food waste should preferentially be used as an-
reported that around 1.3 billion tons of food is lost or wasted every year imal feed, it became illegal because of disease control concerns
globally. It is found that this amount corresponds to one-third of all (Salemdeeb et al., 2017; Cerda et al., 2018). Thus, the valorization of
food resources produced for human consumption. Note that sources of food waste through production of value-added products can be an ideal
food waste include household, commercial, industrial, and agricultural and practical end use.
residues, while the compositional matrix of food wastes varies broadly Depending on geographically-specific circumstances, the generation
based on source and type (Xue et al., 2017). Food waste means a sub- patterns of FW may differ greatly across the world. In broad terms,
stantial loss of other resources such as land, water, energy, and labor. waste generation is affected by a list of variables, including crop pro-
FAO defines food waste as “food losses of quality and quantity through the duction options/models, internal infrastructure/capabilities, distribu-
process of the supply chain taking place at production, post-harvest, and tion chains/channels, and purchase/usage habits of consumers. Table 1
processing stages.” In more specific terms, food losses occurring at the summarizes the types of FW and their origins in the food industry. In
end of the food chain correspond to “food waste (FW),” which is con- the UK, 15 million tons of food were wasted annually (Salemdeeb et al.,
tingent on consumer behavior, purchase intention, and retailer mar- 2017). In Malaysia, it is estimated that 6.7 million tons of FW are


Corresponding authors.
E-mail addresses: kkim61@hanyang.ac.kr (K.-H. Kim), ekwon74@sejong.ac.kr (E.E. Kwon).
1
These authors contributed equally to this study.

https://doi.org/10.1016/j.envint.2019.03.076
Received 11 January 2019; Received in revised form 10 March 2019; Accepted 30 March 2019
0160-4120/ © 2019 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/BY/4.0/).
Y.F. Tsang, et al. Environment International 127 (2019) 625–644

Table 1
Types of wastes and their origins in the food industry.
Order Types of FW Origins Reference

1 Waste from preparation, processing, and rendering of Slaughter house, butcher shops, fish processing plants, egg processing plants, (Jayakumar et al., 2008)
meat, fish, and other food stuffs originating from animals tallow processing plants
2 Waste from preparation and processing of fruit, Fruit and vegetable processing plant; starch manufacturer; malt house; grist and (Preethi et al., 2012)
vegetables, grain, edible oil, cocoa, coffee, and production husting mill; oil mill; manufacturer of coffee, tea, cocoa, and canned foods;
of canned foods. tobacco processing plants.
3 Waste from sugar factory Sugar manufacturers (Albuquerque et al.,
2010)
4 Waste from dairy factory Dairies (Bosco and Chiampo,
2010)
5 Waste from production of baked foods and sweets Bakeries, confectioners, candy producers (Leung et al., 2012)
6 Waste from production of both non-alcoholic and Breweries, wineries, liqueur producers, distilleries, and non-alcoholic beverage (Preethi et al., 2012)
alcoholic beverages and fruit juice producers

generated annually in 2020 (Bong et al., 2017). Approximately 567 to 2. Current status of bioplastic production and food waste
726 million tons of FW (equivalent to up to 40 wt% of the total food feedstock collection and sorting
production) are generated annually in the USA (US EPA, n.d.). This
quantity of FW is equivalent to USD 218 billion (Gunders and Bloom, 2.1. Desirable bioplastic production from food waste
2012). Likewise, it is also comparable to the annual energy loss of over
456,250 cal per capita, which is enough to support a male adult (daily FW is being generated from all stages of the food supply chain in-
calorie intake = 2500) for 180 days (Buzby et al., 2014). In Spain (an cluding post-production, handling/storage, manufacturing, wholesale/
agriculturally-focused country), about 8 million tons of FW were pro- retail, and consumption stages (Ravindran and Jaiswal, 2016). In gen-
duced in 2014 (González-Torre and Coque, 2016). If such waste was eral, around 30 wt% food becomes FW. In 2015, 39.6 million tons of FW
properly converted into value-added products, huge economic and en- (15.1 wt% of MSW) were generated in the US, but only 5.3 wt% of them
ergy losses can be saved. In France, the government has already im- is used for anaerobic digestion and composting (Gunders and Bloom,
plemented a policy for stimulating valorization of FW through recovery 2012). The European Union generates 90 million tons of FW annually.
of energy (e.g., biogas) and value-added materials (e.g., bioplastic) with Among them, 38 wt% is originated from the food manufacturing sectors
a punitive law amid a FW epidemic (De Clercq et al., 2017); im- (Pfaltzgraff et al., 2013). Thus, the conversion of FW into value-added
plementation of this policy was estimated to save 88 million tons of FW chemicals can be the desired end use of food waste for increasing global
per year with a corresponding cost of USD 167 billion (Gore-Langton, sustainability. The water content of FW generally ranges from 75 to
2017). 85 wt%. The mass portion of organic matters is 60 to 70 wt% (dry basis)
Although several papers reported the simultaneous conversion of (Rhu et al., 2003). Table 2 summarizes the properties of FW that are
FW into energy and bioplastics, most of them emphasized bioplastic suitable for production of bioplastics and bioenergy. Fig. 2 provides a
production processes, operating conditions, and new bacterial/archaeal full-screen of not only process of bioplastics production but also routes
species used in the fermentation process (Khanna and Srivastava, 2005; of waste biomass transfer and energy consumption. Starch, cellulosic,
Chee et al., 2010; Chen and Patel, 2011; Bugnicourt et al., 2014; Kiran oily plant, human being, and livestock interacted with each other for
et al., 2014; Koutinas et al., 2014; Salgaonkar and Bragança, 2017). energy transfer through physical/chemical/biological way during the
Among them, only a limited number of studies reported the potential conversion process. The figure provides an ideal and economic platform
production of polyhydroxyalkanoates (PHA) from a single species of to optimize expense and energy (fuels) recovery.
FW, such as waste cooking oil (Desroches et al., 2012) and cheese whey Production of bioplastics such as PHA is an ideal strategy for FW
(Valentino et al., 2015). disposal. One reason for this is that FW is landfilled and yields un-
The production of synthetic plastic from the irreversible process desirable results, such as greenhouse gas (GHG) emissions and
(e.g., petroleum extraction) is an environmental burden (e.g., compared groundwater contamination. Production of bioplastics from FW is a
to bioplastic). Also, microorganisms in nature have not evolved to ef- renewable sustainable process, in which materials are synthesized from
ficiently degrade petroleum-derived polymers (Harding et al., 2007). the carbon neutral resources. Some bioplastics are biodegradable and
The average energy requirement of bioplastics production is obviously compostable under industrial conditions (Dietrich et al., 2017). Com-
less than traditional petro polymer (57 MJ kg−1 compared to postable bioplastics break down by 90% in several months but can also
77 MJ kg−1) to be beneficial toward global warming problem (Gironi undergo industrial composting processes (Siracusa et al., 2008). Such
and Piemonte, 2011). Therefore, due to similar functions of bioplastic bioplastics should meet the specifications and evaluation criteria of
and conventional polymer, bioplastic is an ideal alternative in the international standards (e.g., EN 13432, EN 14995, and ASTM D6400)
context of environmental sustainability. Over the past decade, en- for biodegradable plastics/products, such as composability.
ormous efforts have been put into converting FW into PHA as a prac- The environmental benefits from utilization of bioplastics are one of
tical option for FW valorization (Fig. 1). Indeed, with increased atten- the driving forces to expand their further use. Ceresana in Constance,
tion on sustainable development and renewable materials worldwide, Germany, predicted that the world market for bioplastics in 2021 will
issues of efficient conversion of FW into PHA have been gaining more be three times larger than that in 2014, generating a total of USD 5.8
attention. Yet, the technical completeness to convert FW into PHA is an billion in revenue. For a specific case, NatureWorks in Minnetonka, MN,
initial stage of development. However, it shows great potential for USA announced that polylactic acid (PLA) was produced at a capacity
commercialization with economic viability. Therefore, this review of 1.5 × 105 tons in 2009. This capacity is expected to increase to
highlights the potential of FW processing techniques for production of 8.0 × 105 tons in 2020. Note that Ceresana and NatureWorks are
bioplastics (e.g., PHA) and their economic viabilities. Specifically, we leading international market research and consultancy companies.
reviewed 193 studies, including 145 articles published after 2010, to Approximately 80% of the polymer market may be replaced by bio-
summarize the current knowledge about valorization techniques for FW plastics in Western Europe (Shen et al., 2009). Groot and Borén (2010)
with the main emphasis on PHA production. Special case studies are performed an LCA of industrial production of PLA in Thailand, in-
also provided at the end of this work. dicating that bioplastics can offset environmental problems compared

626
Y.F. Tsang, et al. Environment International 127 (2019) 625–644

Fig. 1. Science citation index publications on PHA and food waste from web of science.

with using petro-derived polymers (e.g., GHG emissions and climate PHA. The procedure mainly goes through as: substrate preparation,
change). PHA-accumulating fermentation, and PHA extraction (Serafim et al.,
2008). Their physico-chemical properties are determined by the oper-
2.2. Current status of bioplastic production ating parameters and the bacterial species chosen for the fermentation
process (Weber et al., 2002).
PHA is one of the key elements to drive the market for biodegrad- PHA has been widely used for various purposes, including packa-
able polymers. PHA is an important polymer family that has been in ging, medical applications, energy, and fine chemicals (Chen and Wu,
development stage for a while but to finally enter the commercial 2005; Chen, 2009; Chen and Patel, 2011; Liang and Qi, 2014; Koch and
market at of which production capacities are estimated to quadruple in Mihalyi, 2018). Commercialized PHA production has already been
the next five years (Chen et al., 2016; Briassoulis and Giannoulis, 2018). implemented by many manufacturers such as Metabolix® (Woburn, MA,
It has been proven to have the great potential as a substitute for tra- USA), Tianjin Green Bioscience Co., Ltd. (Tianjin, China), and Biocycle
ditional plastics due to its biodegradability and rubbery-like properties. PHB Industrial SA (Serrano, SP, Brazil). Global generation of PHA from
The unique properties of PHA are recognized as better oxygen commercial manufacturers reached 2.05 million tons in 2017 (Eur-
barrier (than both non-biodegradable polypropylene (PP) and poly- opean Bioplastics). As summarized in Table 3, most raw materials used
ethylene terephthalate), better water vapor barrier (than PP), and the in the PHA industry are food crops, sugarcane, and vegetable oil. The
fat/odor barrier. Such superior physico-chemical properties of PHA current industrial expense for PHA production is 5–10 times higher
(e.g., in reference to PP) have promoted its usage in various fields in- than that of petroleum-derived polymers. Therefore, the major hurdles
cluding food packaging (Reddy et al., 2003; Chen, 2009; Ahmed et al., for commercial production of PHA and its practical use are high pro-
2018). The poly‑4‑hydroxybutyrate (P4HB), polyhydroxyvalerate duction costs, low-efficiency downstream processing, and high energy
(PHV), polyhydroxyhexanoate (PHH), polyhydroxyoctanoate (PHO), consumption for sterilization of the fermenters (e.g., pure glucose)
and their copolymers (e.g., poly‑3‑hydroxybutyrate (P3HB)) are the (Koller et al., 2017). In particular, the feedstock cost for PHA produc-
most commonly used bioplastics (Figure 3) (Ren et al., 2011). PHA can tion represents half of the overall production cost (Salgaonkar and
be stored as granules in the cell cytoplasm by microorganisms under Bragança, 2017). In an effort to seek low-cost feedstocks at large-scale
stress conditions caused by the limitation of a nutrient, electron donor, (relative to traditional raw materials), enormous studies have been
or acceptor. Fig. 4 presents a scheme for the biosynthesis process of conducted. Among various alternative feedstocks for PHA production,

Table 2
Properties of potential food waste applied in PHA production.
Order Type of food waste Potential materials Characteristics Reference

1 Used cooking oil Palm oil, rapeseed oil, soybean, High content of lipid (fat) can be converted into biodiesel (fatty (Mozejko and Ciesielski, 2013)
sunflower seed acid methyl esters: FAMEs)
2 Animal by-products Blood, fats, residues from High nitrogen content or high levels of BOD and COD (Lin et al., 2013)
intestines
3 Organic crop Straw, stover, peels, fruit pomace These fractions consist of important sources of sugars, lipids, (Bussemaker and Zhang, 2013; Cesário
residues carbohydrates, and mineral acids et al., 2014b)
Provided water, soluble sugar, and cellulose
Succinic acid production
4 Mixed domestic Cheese whey, waste bread, nuts High content of protein, starch, fat and fatty acids (Pais et al., 2014)
waste and nut shell

627
Y.F. Tsang, et al. Environment International 127 (2019) 625–644

Fig. 2. Routes for transforming energy from biomass.

strategic means for lowering the overall production cost (Chee et al.,
CH3 O 2010).
In Europe, lactic acid, PLA, PHA, and bioethanol are produced from
----

a surplus whey containing lactose. Considering that the sugar and food
oil industries generate a tremendous amount of lignocellulosic waste
H (e.g., empty fruit bunch (EFB), palm oil mill effluent (POME), and palm
OH
O kernel shell (PKS)), FW containing sufficient amounts of carbon sources
and nutrients can be a great potential for PHA production through
n bacterial fermentation. Thus, various types of FW also offers a strategic
means for diversifying microbial strains in the fermentation process.
Fig. 3. Structure of poly‑3‑hydroxybutyrate (P3HB), the most common type of
Such technical advancements provide an innovative production
PHA.
pathway for bioplastics via mixed cultures or excess activated sludge
from wastewater treatment plants (Lam et al., 2017). Comparing with
agro-industrial waste streams (e.g., glycerol from biodiesel production, pure culture technology, mixed culture cost less on sterilization. In
lignocellulosic waste from the food industry and forestry, and petro- reference to the pure culture technology, the cost for mixed culture in
chemical plastic waste) have gained great attention since FW provides a line with sterilization during the treatment process is lower which is an

Fig. 4. PHA biosynthesis process scheme.


628
Y.F. Tsang, et al. Environment International 127 (2019) 625–644

Table 3
Global PHA producers and their substrates.
Order Company PHA type Substrates Production capacity (tons per year) Reference

1 Metabolix, (USA) P3HB Corn 50,000 www.metabolix.com


2 MHG Bio, (USA) Mcl-PHA Canola oil 20 www.mhgbio.com
3 Biocycles, (Brazil) P3HB Cane sugar 100 www.biocycle.com.b
4 Bio-On, (Italy) PHA Beet sugar 10,000 www.bio-on.it
5 TianAn Biopolymer, (China) P3HB, P3HBV Corn 10,000 www.tianan-enmat.com
6 Tianijin GreenBio (China) Etyl 3-HB Sucrose 10,000 www.tjgreenbio.com
P(3,4HB)

important technical advantage (Nielsen et al., 2017). featured strong economics and low environmental impacts, although
the limited daily processing capacity and waste utilization rate were
2.3. Food waste feedstock collection and sorting technologies still problematic.

Most parts of FW can be recycled (if separated), and recycling will 3. Technologies for converting food waste to fermentable
help reduce the overall expenditures on waste management. Also, it is substrates
important to improve the recycling rate of food waste, especially for FW
from complex MWF. Collecting and sorting from the source of waste Although FW is a good initial feedstock for production of bioplas-
generation can effectively reduce the cost of the subsequent steps to tics, it must be pretreated to improve or modify the physico-chemical
offer a strategic means for maximizing yield and profit, to reduce en- and biological properties. This section discusses the commonly used
vironmental burden, and to improve the reuse efficiency of material. pretreatment technologies (i.e., physical, chemical, and biological
Based on generation source, FW can be classified as industrial, agri- processes) and enzymatic hydrolysis and their effects on bioplastic
cultural, and household FW. The total amount of FW from industrial production yield. Successful conversion processes refer to partial or
processes and agriculture is large, but the composition of FW is rela- total liberation of monomers from the FW (e.g., lignocellulosic com-
tively simple. On the contrary, the composition of household FW is very ponents) with increasing accessibility of proteins, lipids, and poly-
heterogeneous. saccharides (e.g., starch and cellulose) for subsequent enzymatic hy-
In large-scale food production facilities, such as farms and food drolysis and fermentation (Barisik et al., 2016; Kim, 2018). Moreover,
industrial plants, recycle systems are generally reserved for unpacka- several methods can be integrated into a single treatment system to
ging and peeling functions. For example, a turbo separator can deal achieve better performance. For example, bioplastic production with
with 8–10 tons of FW per hour (Scott®, USA). In commercial applica- physical/acid treatment (i.e., 60 min heating at 121 °C followed by 2%
tions, one system can serve two functions of collection and classifica- sulfuric acid digestion) of industrial FW mixture was able to achieve the
tion. STREAM® Corporation (Malaysia) effectively collects wet FW from highest 3‑hydroxyvaleric acid (3 HV) mole fraction of 22.9% in ex-
kitchens and transfers waste in containers or delivers them directly to periments (Elbeshbishy et al., 2011; Ahn et al., 2016b). Table 4 sum-
FW treatment plants. A specialized organic waste transport system in- marizes the sub-products generated/separated from FW transition
tegrated with a dehydrator, full vacuum, and gravity vacuum can be processes and their applications.
applied in kitchens of airport catering centers, restaurants, food courts,
and food processing plants to facilitate more efficient collection and 3.1. Mechanical and thermal conversion of food waste to fermentable
sorting of FW. organic compounds
Different colored waste bags are being used for different waste
streams. In Hong Kong, green waste bags are used to separate FW from Physical pretreatment is conducted by the mechanical and thermal
municipal solid waste (MSW), while residual MSW can be packed in a conversion processes. Ultrasound, microwaves, milling, and heating
common plastic bag (Lo and Woon, 2016). However, due to the com- methods are also being used for pulverization to increase surface area,
plex composition and relatively high cost of MSW, large-scale proces- separation rate, biological conversion, or fermentable substrates (in-
sing for bioplastic production requires further research in many areas. cluding glucose, proteins, fats, fatty acids, and starch) (Sasmal et al.,
The results of life cycle assessment (LCA) of two different waste col- 2012; Bussemaker and Zhang, 2013; Pagliaccia et al., 2016). In la-
lection systems indicated that automated separation of FW from MSW is boratory-scale PHA production, a study synthesized PHB using Jambul
more environmentally friendly than conventional methods, such as seeds dried in an oven at 60 °C to reduce the moisture content and then
manual collection and truck transportation (Aranda Usón et al., 2013). milled the seeds into fine particles. Lignocellulosic waste can be pre-
In an actual case study, the performance of a pilot-scale FW collection treated by a steam-explosion method (i.e., 160 to 260 °C and 0.7 to
system in Suzhou, China, was analyzed and evaluated based on local 4.8 MPa), which is a commonly used process to separate lignin and
economic conditions, municipal facility conditions, and operation effi- hemicellulose from cellulose (Agbor et al., 2011; Pielhop et al., 2016).
ciency (Wen et al., 2015). The results indicated that integrated systems Physical pretreatment is usually applied at the beginning to change the

Table 4
Organic residues and their applications.
Order Organic residues Applications in bioplastic synthesis Reference

1 Fruits and vegetables Antioxidants, flavonoids, phenols, carotenoids, lipids, (Mirabella et al., 2014)
phytochemicals, carbon source
2 Animal wastes, meat, and Proteins, enzymes, collagen, gelatin, nitrogen source, trace materials (Jayathilakan et al., 2012; Mirabella et al., 2014; Pleissner
derivatives et al., 2016)
3 Dairy products Carbon and nitrogen sources (Mirabella et al., 2014)
4 Waste oil Fatty acid and methyl esters, glycerol, erucic acid (Bardhan et al., 2015)
5 Cellulosic biomass Phytosterols, polypropylene, acrylic acid, isobutanol, thioester, esters (Bardhan et al., 2015)
6 Lignin PHA, adipic acid (Linger et al., 2014; Vardon et al., 2015)

629
Y.F. Tsang, et al. Environment International 127 (2019) 625–644

particle size or to separate the materials for the processing in the next
steps. Thus, the physical pretreatment would not be applied alone but is

Kourmentza et al. (2018a,


(Saratale and Oh, 2015)
Kourmentza et al. (2018a,

(Vega-Castro et al., 2016)


always combined with other treatment methods.

(Valentino et al., 2018)

(Sindhu et al., 2013)

(Ahn et al., 2016b)


(Tyagi et al., 2018)
3.2. Chemical conversion of food waste to fermentable sugars

Reference
Acid pretreatment is common for FW, especially for lignocellulosic

Reference

2018b)
materials (Mussoline et al., 2013). Acid treatment increases the acces-

2018b)
sibility of such low-cost and abundant feedstocks to hydrolytic enzymes
while producing small amounts of cell growth inhibitors (Monavari

−1
et al., 2011). Acid-treated lignocellulose to lignocellulose-derived by-

Yield (g bioplastic g
products can inhibit or deactivate enzymes as well as influence the

−1
Yield (g bioplastic g
performance of bacteria used in the fermentation step. The inhibitors
include furan derivatives (e.g., furfural and 5‑hydroxymethylfurfural

substrate)
(HMF)), lignin-derived phenolics (e.g., phenols are inhibitors to cellu-

substrate)

0.77

0.89

0.87
lolytic enzyme), and carboxylic acids (e.g., weak organic acids such as

0.7
acetic, formic, and levulinic acids) (Monavari et al., 2011; Barisik et al.,

0.77

0.08
0.4

0.9

Rice paddy straw


2016; Kim, 2018). The performance of microbial bioplastic synthesis is
contingent on the operating conditions of acid pretreatment (Ahn et al.,

Domestic FW mixture

Spent coffee
Spent coffee grounds

Rice straw

Rice straw
2016a, 2016b). Alkali pretreatment involves the use of alkaline solu-

Example

grounds
tions such as NaOH, Ca(OH)2, and ammonia. Solvation is the first re-

Pineapple peel
action that occurs during alkali pretreatment, leading to solid expan-

Sugarcane
Example
sion (Carlsson et al., 2012). Alkaline pretreatment effectively

2% H2SO4, 60 min autoclaving


hydrolyzes ester bonds between plant polysaccharides and lignin
(Alvira et al., 2010) for lignin solubilization. Elbeshbishy et al. (2011)

2% NaOH, 121 °C, 30 min

2% H2SO4, 121 °C, 15 lb.


studied three combined physical/chemical pretreatments, including

Operating conditions
ultrasonication with heat, ultrasonication with acid (i.e., 1 N HCl,

Mechanoacoustic and nonchemical effects; high recoveries of cellulose and

Improve surface area and better accessibility of the substrate to enzymatic


pH = 3.0, and 24 h at 4 °C), and ultrasonication with base (i.e., 1 N

pressure for 1 h

121 °C, 15 psi


NaOH, pH = 11.0, and 24 h at 4 °C). The results showed that alkaline
ultrasonication can achieve the highest increase rate of 30% of soluble
chemical oxygen demand (COD) and 40% of soluble protein. The cor-


responding hydrolysis yield was 0.84 g g−1 at a loading rate of 20

Enhances enzyme hydrolysis for fermentable substrates


FPU g−1 of pretreated paddy straws. Table 5 summarizes the raw ma-
terials from FW through various pretreatment methods. In PHA pro-

Make products more accessible to hydrolytic enzymes (produce


duction, a combination of physical/chemical treatments is useful to
achieve better performance or to enhance the final yield. In general, the
Reductions in particle size to extract sugars
hemicellulose and lignin removal product

combined pretreatment processes aim to increase in accessible surface

Make products more accessible to hydrolytic enzymes


smaller amounts of inhibitors and lower sugar losses)
area while lowering in degree of polymerization of raw materials (e.g.,
Provided characteristics and effects
Raw materials extracted from food waste using physical and chemical conversion methods.

cellulose).

3.3. Biological conversion of food waste to fermentable substrates


Provided characteristics and effects

White rot fungus aids in delignification, which in turn improves


enzymatic saccharification rate and productivity (Kalyani et al., 2013).
Provide higher digestibility
Provide higher digestibility

Several studies have adopted fungi as a FW pretreatment method (Isroi


hydrolysis

et al., 2011; Vasmara et al., 2015). Cianchetta et al. (2014) evaluated


selected white-rot fungi to improve carbohydrate yield from wheat
straw and investigated the effects of five different fungi on enzymatic
Thermal pretreatment

hydrolysis of wheat straw. The results of an optimized fungal strain-


biomass combination showed that Ceriporiopsis subvermispora provided
the highest yield of net carbohydrate and minimized weight and cel-
Technology

Ultrasound

lulose losses (Saha et al., 2016). Thus, appropriate fungal strain selec-
Grinding

Alkali pretreatment
Alkali pretreatment
Milling

Acid pretreatment

Acid pretreatment

tion influences the specific biomass pretreatment. In reference to the


chemical treatment, biological pretreatment of lignocellulosic biomass
Technology

offers a way to save energy consumption while reducing intractability


toward cellulolytic enzymes. Despite such benefits, the slow reaction
kinetics and loss of polysaccharides during the biological process have
Mechanical and thermal
(A) Results for physical methods

(B) Results of chemical methods

been pointed out as the main technical demerits (Lemée et al., 2012).
Bule et al. (2016) performed biodegradation of wheat straw using a
white rot fungus that secretes three major classes of ligninolytic en-
conversion

zymes, namely lignin peroxidase, manganese peroxidase, and laccase. It


conversion
Chemical

was suggested that these unique lignin modification patterns are asso-
ciated with the white rot (P. radiate) extracellular proteome which was
expressed during the solid-state fermentation (pretreatment) of wheat
Table 5

Order

Order

straw as an efficient biodegradation method. Biological pretreatments


1

3
4

are promising eco-friendly processes based on addition of specific


2

630
Y.F. Tsang, et al. Environment International 127 (2019) 625–644

enzymes (e.g., peptidase, carbohydrolases, and lipases) from micro-


organisms.

Amount of food waste required for


3.4. Enzymatic hydrolysis of food waste

Hydrolysis is the main mechanism to breakdown polymers into their


corresponding monomers and/or intermediates. Enzymatic hydrolysis

production (tons)
promotes the hydrolytic ability of FW and reduces volatile suspended
solids. Converting polymeric structures into fermentable products (e.g.,

2.7 × 104

3.2 × 105

2.1 × 104

1.5 × 106

1.9 × 106
lignocellulose into carbohydrate and animal fat into fatty acids) is a
critical step in this process. FWs containing lignocellulosic are a com-
plex matrix of cellulose, hemicellulose, and lignin. Although cellulose

Waste cooking oil


and hemicellulose yield fermentable sugars via an enzyme hydrolysis

Bakery product

Bakery product

Bakery product

Bakery product
step, lignin is one of the most recalcitrant structures as it consists of

Type of food
phenylpropanoid units (Saritha et al., 2012). Kiran et al. (2015) studied
the effect of pretreatment with commercial enzymes on hydrolytic so-

waste

waste

waste

waste

waste
lubilization of raw FW collected from a cafeteria. The authors found
that protease exhibited the highest reduction rate of volatile suspended
solids among three types of enzymes, namely carbohydrases, proteases,

Amount of glucose required for


and lipases, and the mixed enzyme treatment exhibited better reduction
efficiency than that of single-enzyme treatment. Heng et al. (2017)
optimized a three-step PHA production process of alkaline pretreat-
ment, enzymatic hydrolysis, and biosynthesis conversion from rice

production (tons)
husks with Burkholderia cepacian USM. Thus, multiple-step treatment
methods can increase the biodegradability of most carbon sources, re-

1.3 × 104

1.6 × 105

1.0 × 104

1.5 × 105

9.3 × 105
sulting in enhanced PHA production performance and efficiency.
Certain types of FW do not contain sufficient amounts of nutrients to
maintain biological activity during fermentation, which can be resolved
by utilizing a mixture of different FWs. Mixed enzymes produced
Production (kg kg−1 of

through solid-state fermentation can hydrolyze proteins and sugary


compounds (e.g. polysaccharides) in FW through the different path-
ways (Koller et al., 2005; Kwon et al., 2005; Matsakas et al., 2017;
Paritosh et al., 2017). Hydrolysates and/or protein-rich FW can also
glucose)

substitute for commercial nutrient supplements for bioplastic produc-


1.16

0.95

1.16

0.54

0.43
tion (Franek et al., 2000; Fitzpatrick and O'keeffe, 2001; Vázquez and
Murado, 2008). Platform chemicals for bioplastics, namely succinic
acid, lactic acid, fumaric acid, and PHB, can be obtained from fer-
Worldwide production

mentation of sugars in household FW, such as dried foods and bakery


products (Lin et al., 2013). Table 6 lists a production potential of
platform chemicals for bioplastic production (Thompson and
1.5 × 104

1.5 × 105

1.2 × 104

8.0 × 104

4.0 × 105

Thompson, 2004; FitzPatrick et al., 2010; Leung et al., 2012; Lin et al.,
(tons)

2013; Deng et al., 2016).

4. Bioconversion of fermentable substrates from food waste to


Polytrimethylene terephthalate (PTT), polyurethane,

bioplastics
Production potential of platform chemicals for bioplastic production.

4.1. Biological synthesis of bioplastics


Food packaging materials, biocomposites

Although 250 types of natural PHA producers have been identified,


only a few bacteria have been adopted for commercial production of
PHA. Such bacteria, including Alcaligenes latus, Bacillus megaterium,
Polybutyrate, polyamides

Cupriavidus necator, and Pseudomonas oleovorans, are found to convert


different kinds of carbon sources into PHA. In particular, C. necator is
copolyester, ethers

one of the most widely utilized microbial strains to produce PHA


Polyester resin
Applications

(Reddy et al., 2003). Marine bacteria manifested a huge potential for


bioplastic production, but their utilization for PHA production has been
poorly reported (Takahashi et al., 2017). The latest research showed
PLA

that the bacterium Halomonas hydrothermalis, H. campaniensis LS21 can


grow in fabricated sea water as well as in FW-like mixed substrates
Platform chemical

1,3‑Propanediol

consisting of cellulose, proteins, fats, fatty acids, and starch. The pro-
Succinic acid

Fumaric acid

cess produced approximately 70% PHB with a pH of 10 at 37 °C (Yue


Lactic acid

et al., 2014). In addition, Pandian et al. (2010) investigated B. mega-


PHB

terium SRKP-3-produced PHB from dairy waste and sea water, demon-
strating a PHB yield of 0.1 g g−1.
Table 6

Order

Natural bioplastic producers are classified according to an accu-


1

mulation mechanism. One group of microorganisms requires limited


2

631
Y.F. Tsang, et al. Environment International 127 (2019) 625–644

nutrients (for instance, nitrogen, phosphorus, and oxygen) but excess of bioplastic. Batch and fed-batch reactors have been widely adopted in
the carbon sources for the biosynthesis process. The other group of industrial fermentation processes. Fed-batch technology demonstrates
microorganisms accumulates PHA through the exponential growth higher PHA yield than batch cultivation methods. Since the con-
phase without the carbon source-starvation conditions (Khanna and centration of N/P is limited in this process, the cell concentration can be
Srivastava, 2005). These PHA-producing microorganisms integrated the easily controlled by adjusting the feeding rate of the carbon source.
carbon catabolic mechanisms with the PHA anabolic pathways of dif- Therefore, a high initial concentration of carbon sources can be avoided
ferent kinds of carbon. Researchers found that the culture methodolo- to provide high osmotic pressure to the PHA producers.
gies can be segregated into unadulterated cultures and mixed cultures A two-stage cultivation method has most commonly been adopted
on the basis of cultivation strategy. for industrial-scale production of copolymer (i.e., P(3HB-co-3HV))
(Hafuka et al., 2011). In the first step, bacterial cells are grown until a
4.1.1. Unadulterated culture pre-determined cell mass concentration is achieved without being
PHA production from unadulterated culture can be segregated into limited by the nutrient. Subsequently, cells are moved to the second-
two key phases (Chen and Patel, 2011). The first is associated with cell stage medium with restricted nutrients and consume only carbon source
growth due to available nutrients. Moreover, the biosynthesis of PHA is for production of PHA. Cells cannot multiply during the nutrient-defi-
prevalent in the second phase under starvation conditions for nitrogen cient stage. Nevertheless, cells increase in size and weight due to in-
(N) and/or phosphorus (P). Nevertheless, growth-associated PHA-pro- tracellular accumulation of PHA as a storage product. PHB is produced
ducing bacteria, for instance, A. latus and recombinant Escherichia coli, by microorganisms like Ralstonia eutrophus and B. megaterium in re-
do not require limited nutrients. Nonetheless, the nutrient-feeding sponse to conditions of physiological stress; they can be created in ei-
strategy can be applied for PHA production in a fed-batch mode. In ther pure or mixed cultures (Laycock et al., 2014). Solid-state fermen-
regard to the fed-batch cultures of growth-associated PHA-producing tation is performed in the absence or near-absence of free water
bacteria, a nutrient-feeding strategy is essential for high-yield PHA (Pandey, 2003). This process has the characteristics of low energy
production. This reflects that both cell growth and PHA synthesis can be consumption, high volumetric productivity, and high titer of value-
enhanced as they occur simultaneously. Furthermore, the two actions added products, low waste generation, and low catabolic repression
require balance to avoid low PHA level. (Hölker et al., 2004). As reported, various types of wastes were effi-
ciently applied as substrates for microbial solid-state fermentation
4.1.2. Mixed culture through which various products were economically produced. Easy
Mixed microbial cultures are cohorts of different microorganisms pretreatment of solid waste can facilitate microbial colonization; this
that can grow together on the same culture medium. When the nutrient process requires grinding and material classification using various
growth is limited, three main steps are used to produce PHA from mixed granulometries to obtain proper material homogenization and to ensure
cultures: anaerobic-aerobic process, aerobic dynamic feeding system that this parameter has less of an effect on the succeeding steps. Some
(feast and famine), and fed-batch process (Samorì et al., 2015; Hilliou typical FW used for fermentation processes, namely sugarcane mo-
et al., 2016a, 2016b). As revealed in key research works, activated lasses, pressed juice, and wheat straw, are summarized in Table 8.
sludge obtained from the wastewater treatment process has been as-
sessed for potential mixture with FW for production of PHA. Different 4.2. Physico/chemical modification of bioplastics
mixed culture modes have been investigated, including aerobic se-
quencing batch reactor (SBR) fed with brewery wastewater (Ben et al., The conventional synthesis approach for bioplastics involves com-
2016), dairy wastewater activated sludge fed with cheese whey (Bosco plete utilization of biomass or its single component (e.g., fiber, starch,
and Chiampo, 2010), and activated sludge from a waste stabilization cellulose, sugar, and lipid) by physical mixing and/or chemical cross-
pond fed with POME (Mohd et al., 2012). Venkateswar Reddy and linking. Recent studies placed great emphasis on the conversion of
Venkata Mohan (2012) carried out research on the consortia attained biomass through separation of monomers or oligomers to create new
by operating an activated sludge system (applied for wastewater polymers using industrial chemical biotechnology. Polysaccharides,
treatment) for PHA production through feeding with fermented FW. such as starches, celluloses, and chitin, are sources of hemiacetal or
According to 16S rRNA gene analysis, the major groups of bacteria were hemi-ketal linkages (Abdul Khalil et al., 2012). These compounds lead
γ‑proteobacteria (39%) and uncultured bacteria (16%). In comparison to short oligosaccharide sequences or polymeric repeating units con-
with unadulterated culture, using activated sludge as a mixed culture nected to other bioplastics. Altogether, polysaccharides comprise ap-
eliminates the need for aseptic conditions, resulting in the lower op- proximately 22 to 37% of FW resources (Tommonaro et al., 2016).
erating cost. Polysaccharides are recognized as unmodified polymers. Hence, for
The ideal industrial production of PHA is highly contingent on the conversion of polysaccharides into bioplastics, bulk or surface chemical
development of strains that are capable of achieving high final cell modification is required on the hydroxyl groups present in their back-
concentrations within 60 h, including 48 h fermentation and 12-h bone structures (Cunha and Gandini, 2010). Such modification ne-
turnaround processes. In addition, obtaining a higher PHA content from cessitates the formation of derivatives such as chitosan. Surface mod-
a direct and economical medium is also a key factor (Khanna and ification then allows for compatibility and minimization of
Srivastava, 2005). Genetic engineering is a critical means of developing hydrophilicity of natural fibers via covalent bonds between the surface
strains that are capable of efficiently producing PHA from affordable and matrix of the fibers. Modification of natural polysaccharides is
renewable resources. Due to the development of genetically engineered generally based on decreasing hydrophilicity by reducing surface en-
bacteria, microorganisms like recombinant E. coli produce PHA that ergy or by generating an adequate surface morphology (Cunha and
contains 3HB, in addition to 3HHx and 3HO monomers from soybean Gandini, 2010; Sanchez-Vazquez et al., 2013). Fig. 5 demonstrates
oil. Recombinant C. necator was found to contain the Aeromonads caviae various bioplastics and their natural resources. However, compared
PHA synthase gene, which successfully produced copolymer P(3HB-co- with the biological synthesis of bioplastics, higher energy input is re-
3HHx) from PKS (Loo et al., 2005; Fonseca and Antonio, 2006). Table 7 quired for chemical synthesis, especially purification of polysaccharides
presents a comparison of unadulterated and mixed cultures. from FW (Tommonaro et al., 2016).

4.1.3. Fermentation technology 4.2.1. Starch-based bioplastics via physico/chemical modification


With the progress in the cultivation techniques applied in large- Chemical modification of starch is well compiled in the literature.
scale production of PHA, it is identified that the type of fermentation For example, starch nanocrystals can be prepared from partial acid
used by PHA-producing bacteria plays a pivotal role in the production hydrolysis on amorphous regions of starch granules (Angellier-Coussy

632
Y.F. Tsang, et al. Environment International 127 (2019) 625–644

Table 7
Comparison of PHA production using unadulterated and mixed cultures.
Order Biosynthesis process Unadulterated culture Mixed culture Reference

1 Conditions External nutrient limitation and excess carbon Internal nutrient limitation, anaerobic/aerobic or Albuquerque et al. (2010)
source aerobic dynamic feeding
2 Substrate requirements Single substrate Waste materials Colombo et al. (2016)
3 Growth type Separately Simultaneously Kourmentza et al. (2009)
4 Media Synthetic media Complex media Khanna and Srivastava (2005)
5 Fed-batch Sequential batch reactor Reactor configuration Salehizadeh and Van Loosdrecht
(2004)
6 Advantages Higher volumetric productivity Cheap substrate Gurieff and Lant (2007)
7 Disadvantages Expensive substrate, expensive equipment for Low volumetric productivity Kourmentza et al. (2009)
maintaining aseptic operation

et al., 2009). Global generation of bread waste is as high as 27 mil- linkages of hydrophilic functional chains should effectively decrease
lion kg per year. Given that the main constituent of bread waste is a both hydration and flow properties (HPS et al., 2016). Conversion of
starch-based material, bioplastics derived from a succinic acid cellulose into polyols can also be achieved through liquefaction during
monomer are a viable option for producing bioplastics from bread manufacture of polyurethane polyesters and foams (Yu et al., 2006;
waste (on the basis of the general alternation yield of 0.55 g suc- Wang et al., 2008).
cinic acid g−1 bread). The properties and structure of starch can be
altered through hydrothermal treatment without thermal degradation
of the original morphology (Hoover, 2010). PLA can be produced by 4.2.3. Chitin-based biopolymers via physico/chemical modification
chemical conversion of corn or other carbohydrate sources into dex- Chitin is another type of biopolymer available in crustaceans and
trose. Dextrose is fermented to lactic acid followed by polycondensation insect exoskeletons. Chitin is also found in other organisms, such as
of lactic acid monomers or lactide. The starch can be customized under mushrooms and yeasts. Approximately 18 Tg of shell waste is produced
starch-filled polymer systems to improve the chemical compatibility annually and is a major source of chitin. Considering the low solubility
between synthetic and starch polymers (Rahmat et al., 2009). Another of chitin, it cannot be used directly as an initial feedstock for bioplas-
type of starch polymer, known as thermoplastic starch, can be synthe- tics. However, chitin can be chemically modified to chitosan by alkaline
sized through the addition of plasticizers (e.g., polyols, glycerol, fruc- N-deacetylation (Kumar, 2000; Rinaudo, 2006). Deacetylation results in
tose, and urea). Both the mechanical and thermal energy transferred to generation of the corresponding primary amino functional group. The
starch dough during extrusion can help break hydrogen bonds in starch degree of deacetylation can be described as an actual percentage con-
granule crystallites (Liu et al., 2009). However, the produced thermo- version of acetyl glucosamine to glucosamine. Deacetylation can sig-
plastics may lack mechanical strength and show high sensitivity to nificantly modify various properties (physical, chemical, and biolo-
moisture. To improve the mechanical strength (or to chemically modify gical) of chitin (Hirano et al., 1989). Both α-/β-forms of chitin are
the surfaces) of these thermoplastics, they can be mixed with suitable insoluble in a majority of solvents, whatever the natural variation in
nanofillers or blended with diverse forms of hydrophobic polymers crystallinity. This insolubility serves as the main obstacle in the utili-
(e.g., polyvinyl alcohol (PVA), polycaprolactone (PCL), polybutylenes zation of chitin. Chitin can be completely solubilized in acidic condi-
succinate (PBS), PHB, and P(3HB-co-3HV) (Wang et al., 2008; Tang and tions when it is converted into its deacetylated product (chitosan)
Alavi, 2011). (Franca et al., 2011). Similarly, the literature provides a score of re-
views on chemical modification of chitin and chitosan including den-
drimer-linked hyperbranched polymers, poly (dimethylsiloxane)s, poly
4.2.2. Cellulose-based bioplastics via physico/chemical modification (2‑hydroxyalkanoate)s, poly(ethylene‑imine)s, block polyethers, poly
Chemically modified cellulose is commonly employed. Cellulose (2‑alkyl‑oxazolines), and polyurethanes (Zohuriaan-mehr, 2005).
serves as a linear polymer with repeating units of Phosphorylation, acylation, alkylation, sulfonation, thiolation, and
anhydro‑D‑glucopyranose, the monomer of which contains three hy- other chemical modifications of chitin have been explored for various
droxyl groups. High-molecular-weight cellulose is highly crystalline. purposes (Kurita, 2006; Jayakumar et al., 2008). For example, qua-
Conversely, cellulose exhibits poor solubility in aqueous media because ternized chitosan derivatives gained importance due to their high an-
of its strong inter- and intra-molecular hydrogen bonds among and timicrobial activity (Liu et al., 2012). Moreover, N‑carboxy‑methylated
within the individual chains (Sandhya et al., 2013). FW with high chitosan is being used in a variety of biomedical applications including
cellulose content includes peanut husks, citrus peels, straw, and corn. wound dressings, antifungals, antioxidant agents, anionic poly-
Many studies have focused on surface chemical modification of cellu- saccharide across the intestinal epithelia, artificial bone and skin,
lose to enhance adhesion among polar OH cellulose fiber groups and apoptosis inhibitor, and blood anticoagulants (Chen and Tan, 2006).
also non-polar polymer backbones through creation of covalent active Fig. 6 presents a schematic diagram of chitin to chitosan conversion
sites. Ethers, esters, and acetals are the most commonly used derivatives along with representative images of their natural resources.
for modification of hydroxyl groups in cellulose structures. Chitosan derivatives are different from pristine chitin. The rigid
Over the past years, various technical advancements in line the crystalline structure of chitosan shows high hydrogen bonding. The free
chemical modification of cellulosic bioplastic were achieved. In the protonate-able amino groups of the chitosan are soluble in mildly acidic
papermaking industry, the following methods are commonly adopted: aqueous solutions. However, they are insoluble in alkaline media and
cellulose sialylation, esterification of the cellulose nanofibers, and cel- water (Pillai et al., 2009). Amino and hydroxyl moieties present in
lulose ester elaboration (Cunha and Gandini, 2010). Pristine cellulose chitosan may be subject to chemical modification, allowing use in
does not exhibit the physico-chemical properties of thermoplastics. further diversified applications. In contrast, chitin only features two
However, plastic properties can be introduced into cellulose fibers by available hydroxyl groups for modification. Modifications of chitin and
mechanical treatments. Furthermore, esterification of the hydroxyl chitosan do not alter their original physicochemical and biochemical
group with the acid on the cellulose structure determines the properties properties (Pillai et al., 2009).
of bioplastics (e.g., fluidity, resistance, and durability), which may be
comparable to those of synthetic polymers. Formation of covalent

633
Y.F. Tsang, et al. Environment International 127 (2019) 625–644

4.2.4. Caprolactone-based bioplastics via physico/chemical modification

Kachrimanidou et al. (2016)


The ring-opening polymerization of ε‑caprolactone using stannous

Bengtsson et al. (2010)


octoate as a catalyst is an important method for preparing PCL. Due to

Colombo et al. (2016)


Cesário et al. (2014b)
Cesário et al. (2014a)
Zahari et al. (2012)s
its low glass transition temperature, the polymer chain was reported to

Mohd et al. (2012)


Ben et al. (2016)
exhibit segmental motion which should help transfer ions from one
complex site to another (Ravi et al., 2016). The main procedure is
Reference

chemical treatment of saccharides which requires a two-step conver-


sion: (1) fermentation of saccharides to ethanol (and acetic acid) and
(2) conversion of ethanol to cyclohexanone using chromic acid. Oxi-
dation to cyclohexanone is achieved via a Bayer-Villiger reaction
(Figure 7) of caprolactone. In general, these kinds of products are added
Bioplastic content (%)

with other natural materials to afford quality biodegradable polymers


at low costs. Also, copolymerisation of lactic acid and ε‑caprolactone
has been investigated in an effort to increase degradation rates and to
improve processability (Yu et al., 2014). Ishiaku et al. (2002) fabricated
37.3
30.0
70.0
55.6
78.9

starch-PCL, which was later mixed with material exhibiting desired


39
70
58

mechanical parameters. To this end, they employed sago starch which


was extruded with a Rheomax at 80–95 °C under 350 bars. A desiccator
Pressed juice from oil palm frond

having 95% H2SO4 was later placed with the extrudates for moisture
absorption. PCL pellets (Tm = 60 °C) were placed in an internal mixer at
Rapeseed meal, wheat bran

90 °C, followed by dried sago starch powder for 10 min. Hot starch–PCL
Fermented cheese whey

blends were rolled with a two-roll mill to obtain thin sheets, which were
Palm oil mill effluent
Brewery wastewater
Sugarcane molasses

stored in the desiccator immediately (Ishiaku et al., 2002).


Crude glycerol
Wheat straw

4.3. PHA production using food waste as a substrate


Medium

The feasibility of PHA production using food waste as a substrate


has been intensively evaluated, including information related to max-
imum PHA accumulation capacity, storage yield, and production rate.
Operation mode

Saccharides (e.g., fructose, glucose, maltose, lactose, and xylose arabi-


nose), together with n‑alkanes (e.g., hexane, octane, and dodecane),
Fed batch
Fed batch
Fed batch

Fed batch

n‑alkanoic acids (e.g., acetic acid, propionic acid, butyric acids, valeric
acid, lauric acid, and oleic acid), n-alcohols (e.g., methanol, ethanol,
SBR

SBR
SBR

octanol, and glycerol), gases (e.g., methane and carbon dioxide), and
acid (e.g., fatty acid and succinic acid) are considered key carbon
sources for biosynthesis of PHA (Santhanam and Sasidharan, 2010; Poli
1.6 L continuous stirred tank reactor

et al., 2011). Most of the carbon applied in commercial PHA production


is relatively expensive, such as unadulterated carbohydrates (e.g., glu-
cose and sucrose), alkanes, and fatty acids. Unadulterated nutrients
250 mL Erlenmeyer flask
2 L stirred tank reactor

(e.g., amino acids and phosphate) are unaffordable, resulting in im-


Fermentation scale

practicality of the majority of developed biotechnological mechanisms.


600 mL reactor

Cost-effective realization of the biotechnological processes is dependent


1 L bioreactor
100 mL flask

on the existence of affordable nutrients that are put to use as sources of


6 L reactor
1 L reactor

carbon, nitrogen, and/or phosphate for microbes (Liang and Qi, 2014).

4.3.1. Food waste as a carbon source


Fruits and vegetables can be used as the carbon substrates in the
Examples of PHA production from food waste by fermentation.

P(3HB-co-3HV)
P(3HB-co-3HV)

fermentation process to produce PHA. Some fruit products (e.g., citrus


juice) are made through 50% extraction of fresh fruit, with the residues
Bioplastic

subsequently remaining as wastes. These wastes contain high amounts


P(3HB)
P(3HB)
PHA

PHA
PHA
PHA

of sugars but low quantities of proteins. Citrus wastes are used to


produce enzymes (in particular pectinase) of citric acid, succinic acid,
dietary fiber, prebiotic oligosaccharides, and natural antioxidants,
Burkholderia sacchari DSM 17165

which can improve profitability (Matsumoto and Taguchi, 2013;


Cupriavidus necator CCUG 52238

Cupriavidus necator DSM 7237

Mamma and Christakopoulos, 2014). Similarly, direct fermentation of


Activated sludge consortia
Activated sludge consortia
Activated sludge consortia

potato residues, in combination with potato starch, into lactic acid was
carried out in septic systems (Smerilli et al., 2015). Preethi et al. (2012)
Defluviicoccus vanus

Cupriavidus necator

employed hydrolyzed Jambul seeds as a sole carbon source in the


production medium, and PHA accumulation in R. eutropha and SPY-1
reached 41.7 and 42.2%, respectively. Follonier et al. (2014) in-
Unadulterated culture

vestigated nine kinds of pomace fruits, which included apricots, cher-


Strain

ries, and grapes, as key carbon substrates. In addition, recovered po-


Mixed culture

mace of apricots, cherries, and Solaris grapes were found to contain


glucose concentrations of 47, 49, and 106 g L−1, respectively. The
Table 8

Order

eventual biomass concentration was 10.2 g L−1, which included 12.4 wt


1
2
3
4
5

6
7
8

% mcl-PHA, corresponding to volumetric productivity of mcl-PHA of

634
Y.F. Tsang, et al. Environment International 127 (2019) 625–644

Pretreatment of food waste


(physical/chemical/biological)

Caprolactone Cellulose

Chitin Starch

Biosynthesis

PHA

Fig. 5. List of bioplastics and their natural resources.

0.03 g L−1 h−1. Solaris grapes featured the highest amount of fructose; resinovorans.
on the other hand, apricots had the lowest number of growth inhibitors. Baei et al. (2010) reported production of poly-3(HB-co-27%-HV)
Assessment of both of these pomaces was carried out to determine their from whey hydrolysate with the use of the Azohydromonas lata DSM
appropriateness for production of medium chain-length (mcl) PHA (i.e., 1123. Furthermore, B. megaterium CCM 2032 accumulated more than
6–14 carbon atoms) in bioreactor fermentation with Pseudomonas 50% of its biomass (w/w) in the supplemented whey media (Obruca

Crab Prawn Fungi Insects

Natural resources

Deacetylation

Fig. 6. Schematic of conversion from chitin to chitosan and their natural resources.

635
Y.F. Tsang, et al. Environment International 127 (2019) 625–644

Fig. 7. Graph of Bayer–Villiger reaction.

et al., 2011). Rhizobium etli and Pseudomonas stutzeri were grown in a recovered PHB granules manifested a high level of purity above 90%; in
medium containing whey as the source of carbon, and their PHB con- contrast, the yield of product on the substrate was 0.77 ± 0.04 g g−1
tents were 25 and 28%, respectively (Baei et al., 2010). B. thuringenesis (Cruz et al., 2015). Also, palm oil-based waste cooking oil was utilized
IAM12077 was selected for accumulation of PHB from the FW mixture. to enhance the manufacture of P(3HB) with C. necator H16 strain. It was
The PHB content was 25.28% which was comparable to pure glucose generated 80% P(3HB) content (Kamilah et al., 2018). PHA was pro-
(37.17%) (Shivakumar, 2012). Cheese whey can be used without any duced with the help of C. necator H16 in residual frying oil that had a
pretreatment. Obruca et al. (2011) carried out a direct transformation 0.21 g L−1 h−1 volumetric productivity (Obruca et al., 2013). In fed-
of the affordable residue of the cheese whey into PHB with the help of batch mode, wasted oil was used as the substrate of C. necator H16 for
the bacterial strain B. megaterium CCM 2037. Using this mixture, the PHA production, wherein the corresponding biomass and PHA pro-
PHB content was enhanced by approximately 40% through introduc- duction were 138 g L−1 and 105 g L−1, respectively. The yield coeffi-
tion of 1% ethanol into the medium during the stationary stage of de- cient and volumetric productivity were 0.83 g PHA g−1 oil and
velopment (biomass of 2.87 g L−1; PHB of 1.48 g L−1). Laboratory-scale 1.46 g L−1 h−1, respectively (Obruca et al., 2010). PHB produced from
assays and semi-productive fermenters are needed for testing pro- used cooking oil was exploited as a key carbon source with help of B.
ductivity when the medium is subjected to elevated cell density culti- thailandensis (Kourmentza et al., 2018a, 2018b).
vation. There are diverse carbon sources of bioplastic production such Butyrate can be a precursor, which has a potential to increase
as waste cooking oil, which contains plant, animal, or synthetic fat from (R)‑3‑hydroxyhexanoate content in poly(3‑hydroxybutyrate
frying, baking, or other types of cooking. C. necator has been used to ‑co‑3‑hydroxyhexanoate) (PHBH) production via genetically modified
produce the homopolymer PHB from rapeseed oil (Verlinden et al., C. necator H16 strains (Sato et al., 2015). Oleochemical industries
2011). Martino et al. used frying oil as the sole carbon source for PHB generate large amounts of fatty acids and glycerol as by-products, both
production in a batch cultivation system with C. necator DSM 428. The of which can be immediately put to use as feedstock for microbial PHA
yield of PHB reached 0.29 g g−1 with volumetric productivity of production (Fadzil and Tsuge, 2017). In total, 172 million kg of glycerol
0.14 g L−1 h−1 (Martino et al., 2014). are generated annually through biodiesel production across the globe
C. necator was utilized in other experiments as well; its cultivation (Tan et al., 2013). Glycerol is well-known for its significant potential as
was performed with used cooking oil as the only means of carbon for a carbon substrate (e.g., relative to other affordable substrates like
production of P(3HB). The biomass approached a concentration limit of whey, sugarcane bagasse, and corn steep liquor). Validation of the
11.6 ± 1.7 g L−1, which had a polymer content of 63.0 ± 10.7% (w/ optimization model with glycerol waste was carried out, wherein a
w) and volumetric productivity of 0.15 g L−1 h−1. Moreover, the maximum of 1.36 g L−1 of PHA was achieved after 96 h. PHA

636
Y.F. Tsang, et al. Environment International 127 (2019) 625–644

generation of 1.87 g L−1 was observed in 5 L lab-scale bioreactors with at C/N ratios of 65 and 80, respectively.
help of Pannonibacter phragmitetus ERC8 (Ray et al., 2016). As revealed in the former studies, the C/N/P ratio that governs the
Spent coffee grounds (SCG) are substantial fraction of residues ob- PHA synthesis is very case sensitive. Wang et al. (2007) reported that
tained from the coffee sector. Every year, substantial amounts of SCG the highest PHA production yield was achieved at the C/N ratio of 100.
are generated across the globe, followed by disposal as a solid residue. Albuquerque et al. (2010) obtained C/N/P ratio data of 100:3:1, and
Cruz et al. (2014) cultivated C. necator DSM 428 in a 2 L bioreactor with PHA content reached 74.6%. Ahn et al. (2015) investigated the effects
extracted SCG oil as an individual source of carbon in a bid to produce of C/N ratio in simulated rice straw hydrolysates using glucose and
PHA. The culture reached a yield of 0.77 g g−1, 78.4% polymer content, ammonium chloride on PHA collected by C. necator. Overall, the PHA
and productivity of 0.20 g L−1 h−1. Obruca et al. (2014b) examined collection rate was greater when subjected to more severe N-deficient
several detoxification methods for conversion of hydrolysates of SCG conditions (for instance, C/N ratio of 360:1) in comparison with the
into PHA by Burkholderia cepacian. Addition of ethanol prior to hy- weaker N-deficient conditions (for instance, C/N ratio of 3.6:1 or 36:1).
drolysis of SCG increased PHA yield by 25%. Extraction of oil from SCG Bioconversion yield relays controlled the conditions of microbial sub-
also yielded high biomass and PHB yields (0.82 g g−1 of oil) compared strates. FW served as a promising source of P. For example, P recycling
with the other waste oils. The results may reflect the positive correla- efficiency from waste is 51% in France (Senthilkumar et al., 2014).
tion between coffee oil and PHB yield in the presence of free fatty acids Thus, options for P recycling and their application in PHA production
(Obruca et al., 2014a). require improvement for customization of P use.
Lignocellulosic substances (e.g., wheat straw and rice straw) are
renewable and are food process waste. They primarily consist of cel- 5. Case studies of PHA products from food/agriculture waste
lulose, in addition to hemicellulose and lignin. Both cellulose and
hemicellulose are good sources of carbon in various biological me- The bioplastic production efficiency from food/agricultural waste/
chanisms subsequent to hydrolysis of monomeric compounds (e.g., biomass and its economic viability are contingent on several factors
glucose, xylose, and arabinose) (Kamm and Kamm, 2007; Sandhya (e.g., type of waste generated in the selected area, processing cost,
et al., 2013). The performance of lignocellulosic materials varies from availability of the biomass, transport expenses, raw materials expenses,
case to case. Hydrolysate of alkali-acid pretreatment paddy straw as a and other common industrial expenses). Apart from high productivity,
source of carbon with R. eutropha MTCC 1472 achieved and 37.55% industrial bioplastic production should comply with government stan-
PHA accumulation at CDW of 19.2 g L−1 (Sandhya et al., 2013); wheat dards for environmental pollution. As the type of waste produced in
straw hydrolysates as the primary source of carbon for production of P each season can vary, the productivity of bioplastics can be affected by
(3HB-co-4HB) with Burkholderia saccharin in fed-batch cultures showed changes in season, which hampers the supply of certain raw materials
a productivity of 0.7 g L−1 h−1 (Cesário et al., 2014a). Sindhu et al. in a certain period. In general, only the feasibility of the synthesis cost
(2013) achieved the highest performance values ever reported for Ba- has been tested at laboratory scale (Elissen and Kootstra, 2016). How-
cillus species, PHB yield of 0.26 g g−1 and 89% polymer content. Also, ever, for industrial-scale production, one needs to consider or evaluate
pentose-sugar-rich hydrolysates can be produced from acid-pretreated techno-economic/profitability, raw material supply, and production
rice straw as an individual source of carbon throughout production. product markets.
Crude sugars were prepared from hydrolyzed corn stover with an on- Several research groups focused on using the developed methods for
site-prepared cellulose cocktail using a co-culture of Trichoderma reesei bioplastics through a cost analysis of the methods. However, limited
and Aspergillus niger. Moreover, PHA represented up yield of case studies of biorefineries have been published. Here, considering the
0.243 g g−1 and 72.4% content from corn stover (Sawant et al., 2015). type of raw material and location for installing biorefineries, we re-
A comparison between wheat straw and the mix of commercial carbon viewed some case studies that directly provide information on the
sources (C6 as well as C5 sugars) was performed. Bioplastic content production of biomass-based products, techno-economical possibilities,
reached 60 and 70%, yield of 0.19 and 0.18 g g−1, and volumetric and cost analysis.
productivity of 0.12 and 0.11 g L−1 h−1, respectively (Cesário et al.,
2014b). This previous work provided a description of the fed-batch 5.1. Case studies on palm tree biomass-based refineries
cultivations of PHA generation using actual lignocellulosic hydrolysates
and demonstrated that wheat straw is valuable for biopolymer pro- The efficient, cost-effective, and environmentally safe production of
duction. bioplastics depends upon several variables. The major factor considered
Animal products and by-products (e.g., milk, cheese fats, and re- for initialization of bioplastic production is availability of raw mate-
sidues from intestines) can also be applied as substrates for PHA pro- rials. The location of a biorefinery should allow a good supply of raw
duction. The experimental results showed a maximum specific growth materials from neighboring areas. For instance, Malaysia has millions of
rate of 0.10 g L−1 h−1 in two different fermentation temperature set- hectares of palm trees. Thus, the availability of raw material in
ups (Muhr et al., 2013). The volumetric productivity of bioplastic Malaysia made it a good destination to develop bioplastic industries or
reached 0.036 and 0.050 g L−1 h−1, and the PHA contents were 20.1 biorefineries (Shuit et al., 2009). In 2006, Malaysia was the largest
and 26.6%, respectively. Table 9 presents the yield comparison of exporter and largest producer (produced 15.88 million tons, equivalent
studies wherein FW was used as a carbon source. to 43% of total world supply) of palm oil. The milling activity for palm
oil generation and plantations left a huge amount of waste biomass,
4.3.2. Food waste as N and P sources such as EFB. It was estimated that 50–70 tons of biomass waste can be
N, P, and minerals are the essential nutrients for PHA production. generated from 1 ha of palm oil plantations. Thus, the waste biomass
FW, particularly animal waste, meat derivatives, dairy products, and can be further utilized for generation of biofuels (e.g., bioethanol;
some vegetables, are capable of providing N and P sources for bioplastic 55.73 million tons of palm oil biomass is an oil equivalent of
synthesis (Jayathilakan et al., 2012; Mirabella et al., 2014). Accumu- 15.81 million tons).
lating lipids and unsaturated fatty acids are dependent on the carbon- Moreover, some plants like palm costs can be further reduced by
to-nitrogen (C/N) ratio and in other cases, carbon-to-phosphorus (C/P) optimizing the location of the production plant and the type of biomass
ratio. Constraints in N favor the collection of lipids and fatty acids in the feedstock (Shuit et al., 2009). Several initiatives have been im-
oleaginous microbes. However, the adjustment of N content makes it plemented in Malaysia including the Small Renewable Energy Power
difficult to use complex organic substrates. Ryu et al. (2013) enriched (SREP) program, Biogen program, EC-ASEAN COGEN program, Chubu
the consumed yeast lysate with glycerol to increase C/N ratio from 20 Electric Power, and construction of a bioethanol plant to utilize biomass
to 35. Smaller amounts of lipid content (by 50 and 61%) were achieved in energy production and to find cost-effective renewable energy

637
Table 9
Comparison of biosynthesis yield on different types of food waste as carbon sources.
Ord- Subst- Biopl- Micro- Biopl- Maxi- Y Ystorage Ytot Biopl- Biopl- Chara- Cultiv- Scale Refer- Re-
er rate astic bial astic mum ferment- (g bio- (g astic astic cteri- ation ence mar-
Y.F. Tsang, et al.

from strain end test con- ation plasti- bio- con- pro- zation time ks
food (g bio- centra- (g org- c g−1 plastic tent duc- techni- (h)
waste plasti- tion anic a- or- g (%) tivity ques
−1
c kg−1 (g - cid - ganic sub- (g -
VSS) L−1) g−1 acid) strate) L−1 -
sub- h−1)
strate)

1 Brew- PHA Activ- 0.4 0.43 39 GC–M- 1 L Ben Mix-


ery ated S SBR et al. ed
waste- sludge (2016) cul-
water con- ture
sortia
2 Ferm- PHA Activ- 814 0.24 0.70 58 0.45 GC–M- 1L Colo- Mix-
ented ated S SBR mbo ed
cheese sludge et al. cul-
whey con- (2016) ture
sortia
3 Chees- PHA Sporo- 60–70 0.4–0- 0.6–0- 0.24–- 0.45–- GC/ 1L Colo- Un-
e genous .6 .7 0.42 1.17 MS SBR mbo adu-
whey acido- et al. lter-
genic (2016) ated
cul-
ture
4 Chees- PHA Reco- 18.88 0.11 0.04 0.71 38.65 0.33 GC fed- Pais Un-

638
e mbi- batch et al. adu-
whey nant bior- (2014) lter-
E. coli eactor ated
cul-
ture
5 Dairy PHB Bacill- 11.32 FITR 36 3L Pandi- Un-
waste us fer- an adu-
mega- ment- et al. lter-
terium er (2010) ated
SRKP- cul-
3 ture
6 Paddy PHA R. 7.21 37.55 FITR, 72–96 2L Sand- Un-
straw eu- Ther- stirre- hya adu-
tropha mo- d-tank et al. lter-
MTCC gravi- re- (2013) ated
1472 metric actor cul-
ana- ture
lysis,
DSC
7 Pinea- PHA Bacill- 1.86 0.31 53.66 0.08 NMR, 72 5L Suwa- Un-
pple us sp. DSC fer- nnasi- adu-
waste SV13 ment- ng lter-
er et al. ated
(2015) cul-
ture
13
8 Spent P Cupri- 0.75–- 81 C 1L Rao Un-
palm (3HB- avidus 0.80 NMR, Erlen- et al. adu-
oil co- ne- DSC, meyer (2010) lter-
4HB) cator flask ated
(continued on next page)
Environment International 127 (2019) 625–644
Table 9 (continued)

Ord- Subst- Biopl- Micro- Biopl- Maxi- Y Ystorage Ytot Biopl- Biopl- Chara- Cultiv- Scale Refer- Re-
er rate astic bial astic mum ferment- (g bio- (g astic astic cteri- ation ence mar-
Y.F. Tsang, et al.

from strain end test con- ation plasti- bio- con- pro- zation time ks
food (g bio- centra- (g org- c g−1 plastic tent duc- techni- (h)
waste plasti- tion anic a- or- g (%) tivity ques
−1
c kg−1 (g - cid - ganic sub- (g -
VSS) L−1) g−1 acid) strate) L−1 -
sub- h−1)
strate)

FTIR, cul-
GC ture
9 Used PHB Cupri- 0.29 37 0.14 Wide- 27 10 L Marti- Un-
cooki- avidus angle batch no adu-
ng oil ne- X-ray bior- et al. lter-
cator dif- eactor (2014) ated
DSM frac- cul-
428 tion ture
mea-
sure-
ments
10 Wheat P Burkh- 105 0.22; 0.60; NMR, fed- Cesári- Un-
straw, (3HB- olderia 1.60 0.19 SEC, batch o et al. adu-
ligno- co- sac- DSC, in 2 L (2014- lter-
cellu- 4HB) chari TGA stirred b) ated
losic DSM tank cul-
17165 re- ture
actor

639
Bioplastic end test (bioplastic at the end of the test).
Yfementation (conversion yield from soluble COD to organic acids).
Ystorage (conversion yield from organic acids to bioplastic).
Ytot (conversion yield from soluble COD of the initial substrate to bioplastic).
Environment International 127 (2019) 625–644
Y.F. Tsang, et al. Environment International 127 (2019) 625–644

sources (Ludin et al., 2004; Shuit et al., 2009; Ong et al., 2011). to be USD 3.44 kg−1 P(3HB) which is 41% lower than using commercial
In 2013, Indonesia produced 30 thousand tons of oil palm EFB, glucose (Zahari et al., 2015). The price fluctuation corresponding to the
which consists of around 35 wt% of glucan, 20.3 wt% of xylan, and used methods can be stabilized by optimizing the methods relative to
3.1 wt% of arabinan (Kresnowati et al., 2015). In light of the avail- yield. In the study, Sugarcane bagasse was 0.0035 USD kg−1, the total
ability of the biomass, Indonesia may also be a good choice for pro- production cost of PHB was 3.32 USD kg−1, and the total cost of re-
duction of bioethanol and xylitol. On the basis of total cost for pro- finery by-product ethanol was 0.57 USD L−1. According to the market
duction of xylitol, a gross profit of 4.3 USD kg−1 can be realized if price, PHB was 3.12 USD kg−1 while fuel ethanol was 1.24 USD L−1. It
biomass is used as a raw material. Likewise, non-food sugars produced was estimated that use of a biorefinery can reduce the selling price of
from oil palm frond (i.e., around 42.8 wt% of glucan, 12.5 wt% of PHB by 50% (i.e., from 3.12 USD kg−1), making it competitive with
xylan, and 2.3 wt% of arabinan) can be used for economical production conventional plastic (polypropylene: 1.58 USD kg−1) (Moncada et al.,
of P(3HB) (Zahari et al., 2015). It was estimated that the production 2013).
cost of P(3HB) can be reduced to 3.44 USD kg−1 by using renewable
sugars generated from a palm frond, which was 41% lower than the cost 6. Conclusions
of P(3HB) produced from commercial glucose. Apart from P(3HB)
production, EFB was also explored for diverse applications such as Bioplastic is a natural polymeric material that has been developed
biofuel for combined heat and power plants, ethanol production, me- extensively over the last two decades due to its good biocompatibility,
thane recovery, composting (Chiew and Shimada, 2013). The use of biodegradability, and material properties. As such, bioplastic produc-
EFB can lead to production of ethanol, methane gas, briquette, elec- tion has become one of the most active research areas in recent years.
tricity, and paper (Sompong et al., 2012; Chiew and Shimada, 2013). Bioplastic can be applied in packaging industries, spray materials, ap-
pliance materials, electronic products, agricultural products, automa-
5.2. Case studies on banana biomass-based refineries tion products, chemical media, and solvents. In the production of bio-
plastics, the interlinkage of biotechnology processes is a key strategy
In 2012, the rejected and wasted proportions of total banana crops aimed at maximizing the use of food waste and increasing the potential
were 26.46 and 6.67%, respectively (Quinaya and Alzate, 2014). These revenue of the entire bioprocessing chain. Given that mass generation
rejected agro and food products are equally effective to produce bio- of FW is inevitable, the environmental burdens arising from waste
fuels, sugars, and PHB. It was theoretically demonstrated that the disposal (e.g., water contamination and GHG emissions) should be
quantities of glucose, ethanol, and PHB produced per ton of banana mitigated. Therefore, this review demonstrated the potential of FW as a
were 316, 238, and 31.5 kg, respectively (Naranjo et al., 2014; Quinaya raw material for bioplastic production to address important environ-
and Alzate, 2014). Moreover, banana peels are also important feed- mental problems. Hence, physical, thermo-chemical, and biological
stocks for production of diverse products. Generally, banana peels methodologies required for preparation of bioplastic raw materials
contain 40% starch that can be transformed into sugars after ripening. from FW are reported. Moreover, production of PHA based on un-
One ton of banana peels can be used to generate 57, 2, 25, and 5 kg of adulterated/mixed culture and fermentation technologies was empha-
glucose, acetic acid, and methane, respectively (Quinaya and Alzate, sized. Modifications of different FW compositions (e.g., cellulose,
2014). The overall cost and margin analysis were estimated for pro- starch, chitin, and caprolactone) for PHA-derived products were also
duction of PHB, glucose, and ethanol from bananas in three different considered. Most importantly, governments, regulations, companies,
scenarios (Naranjo et al., 2014). In the first scenario, PHB was a unique public opinions, and consumers should work together to mitigate cur-
product, while banana peels were treated as residues scenario. The rent environmental burdens arising from FW disposal.
hydrolyzed starch prepared from banana pulp was used for PHB pre-
paration. In a second scenario, glucose, ethanol, and PHB were pro- Parameters
duced in a biorefinery. In the last scenario, mass and energy integration
of the processes was proposed. The cost estimated for PHB, glucose and Bioplastic yield (g g−1): the ratio of bioplastic concentration (g L−1)
ethanol in Scenarios 1/2/3 were 2.7/2.3/1.6, -/0.9/0.7, and -/1.3/ and the substrate consumed (g L−1) during the cultivation time.
0.6 USD kg−1, respectively. In the cost analysis, Colombian labor costs Cell dry weight (CDW) (g L−1).
of USD 2.14 h−1 and USD 4.29 h−1 were used, respectively, for op- Final polymer content (%, w/w).
erators and supervisors (Posada et al., 2012). In addition, the elec- Volumetric productivity (g L−1 h−1): the value obtained by dividing
tricity, water, and vapor pressure prices used were USD 0.03044/kWh, the final bioplastic concentration over the total cultivation time (Δt, h).
USD 1.252/m3, and USD 8.18/ton, respectively. Moreover, corre-
sponding economical margins of 22/43/106%, -/0/22.2%, and -/18.2/ Acknowledgments
45.5% were demonstrated by comparing the market prices of the stu-
died products. It can be concluded that scenario 3 showed mass and KHK acknowledges support made in part by grants from the
energy integration in the biorefinery. The production of PHB from re- National Research Foundation of Korea (NRF) funded by the Ministry of
sidual banana can reduce the energy requirement by 30.6% and the Science, ICT and Future Planning (Grant No: 2016R1E1A1A01940995).
water requirement by 35% (Naranjo et al., 2014). YFT acknowledges support made in part by grants from the Research
Cluster Fund (RG50/2017-2018R) and the Dean's Research Fund (DRF/
5.3. Case studies on sugarcane biomass-based refineries SFRS-8 and DSRAF-6 SP1) of The Education University of Hong Kong.

Techno-economic and environmental analysis were carried out for References


production of ethanol, PHB, and electricity from sugarcane bagasse in
Colombia using Aspen Plus software (Moncada et al., 2013). A simu- Abdul Khalil, H.P.S., Bhat, A.H., Ireana Yusra, A.F., 2012. Green composites from sus-
lation for the techno-economical assessment was performed based on tainable cellulose nanofibrils: a review. Carbohydr. Polym. 87, 963–979.
Agbor, V.B., Cicek, N., Sparling, R., Berlin, A., Levin, D.B., 2011. Biomass pretreatment:
three scenarios according to the Cambodian conditions: (1) energy fundamentals toward application. Biotechnol. Adv. 26, 675–685.
cogeneration, (2) arbitrary distribution, and (3) preselected pathways Ahmed, T., Shahid, M., Azeem, F., Rasul, I., Shah, A.A., Noman, M., Hameed, A.,
using optimization subroutines. The price of sugarcane that can be used Manzoor, N., Manzoor, I., Muhammad, S., 2018. Biodegradation of plastics: current
scenario and future prospects for environmental safety. Environ. Sci. Pollut. R. 1–12.
for PHA production is 10 times lower than that of pure sugar (i.e., su- Ahn, J., Jho, E.H., Nam, K., 2015. Effect of C/N ratio on polyhydroxyalkanoates (PHA)
garcane 0.04 USD kg−1 vs. glucose 0.44 USD kg−1) (Suwannasing et al., accumulation by Cupriavidus necator and its implication on the use of rice straw
2015). In another case study, the P(3HB) production cost was estimated hydrolysates. Environ. Eng. Res. 20, 246–253.

640
Y.F. Tsang, et al. Environment International 127 (2019) 625–644

Ahn, J., Jho, E.H., Kim, M., Nam, K., 2016a. Increased 3HV concentration in the bacterial Bioresour. Technol. 218, 692–699.
production of 3‑hydroxybutyrate (3HB) and 3‑hydroxyvalerate (3HV) copolymer Cruz, M.V., Paiva, A., Lisboa, P., Freitas, F., Alves, V.D., Barreiros, S., Barreiros, S., Reis,
with acid-digested rice straw waste. J. Polym. Environ. 24, 98–103. M.A.M., 2014. Production of polyhydroxyalkanoates from spent coffee grounds oil
Ahn, J., Jho, E.H., Nam, K., 2016b. Effect of acid-digested rice straw waste feeding obtained by supercritical fluid extraction technology. Bioresour. Technol. 157,
methods on the 3HV fraction of bacterial poly(3‑hydro- 360–363.
xybutyrate‑co‑3‑hydroxyvalerate) production. Process Biochem. 51, 2119–2126. Cruz, M.V., Sarraguça, M.C., Freitas, F., Lopes, J.A., Reis, M.A.M., 2015. Online mon-
Albuquerque, M.G.E., Torres, C.A.V., Reis, M.A.M., 2010. Polyhydroxyalkanoate (PHA) itoring of P(3HB) produced from used cooking oil with near-infrared spectroscopy. J.
production by a mixed microbial culture using sugar molasses: effect of the influent Biotechnol. 194, 1–9.
substrate concentration on culture selection. Water Res. 44, 3419–3433. Cunha, A.G., Gandini, A., 2010. Turning polysaccharides into hydrophobic materials: a
Alvira, P., Tomás-Pejó, E., Ballesteros, M., Negro, M.J., 2010. Pretreatment technologies critical review. Part 2. Hemicelluloses, chitin/chitosan, starch, pectin and alginates.
for an efficient bioethanol production process based on enzymatic hydrolysis: a re- Cellulose 17, 875–889.
view. Bioresour. Technol. 101, 4851–4861. De Clercq, D., Wen, Z., Gottfried, O., Schmidt, F., Fei, F., 2017. A review of global stra-
Angellier-Coussy, H., Putaux, J.L., Molina-Boisseau, S., Dufresne, A., Bertoft, E., Perez, S., tegies promoting the conversion of food waste to bioenergy via anaerobic digestion.
2009. The molecular structure of waxy maize starch nanocrystals. Carbohydr. Res. Renew. Sust. Energ. Rev. 79, 204–221.
344, 1558–1566. Deng, Y., Li, B., Yu, K., Zhang, T., 2016. Biotransformation and adsorption of pharma-
Aranda Usón, A., Ferreira, G., Zambrana Vásquez, D., Zabalza Bribián, I., Llera Sastresa, ceutical and personal care products by activated sludge after correcting matrix ef-
E., 2013. Environmental-benefit analysis of two urban waste collection systems. Sci. fects. Sci. Total Environ. 544, 980–986.
Total Environ. 463, 72–77. Desroches, M., Escouvois, M., Auvergne, R., Caillol, S., Boutevin, B., 2012. From vege-
Baei, M.S., Najafpour, G.D., Lasemi, Z., Tabandeh, F., Younesi, H., Issazadeh, H., table oils to polyurethanes: synthetic routes to polyols and main industrial products.
Khodabandeh, M., 2010. Optimization PHAs production from dairy industry waste- Polym. Rev. 52, 38–79.
water (cheese whey) by Azohydromonas lata DSMZ 1123. Iran. J. Energy Environ. 1, Dietrich, K., Dumont, M.J., Del Rio, L.F., Orsat, V., 2017. Producing PHAs in the bioec-
132–136. onomy — towards a sustainable bioplastic. Sustain. Prod. Consum. 9, 58–70.
Bardhan, S.K., Gupta, S., Gorman, M.E., Haider, M.A., 2015. Biorenewable chemicals: Elbeshbishy, E., Hafez, H., Dhar, B., Nakhla, G., 2011. Single and combined effect of
Feedstocks, technologies and the conflict with food production. Renew. Sust. Energ. various pretreatment methods for biohydrogen production from food waste. Int. J.
Rev. 51, 506–520. Hydrogen Energy 36, 11379–11387.
Barisik, G., Isci, A., Kutlu, N., Bagder Elmaci, S., Akay, B., 2016. Optimization of organic Elissen, H.J.H., Kootstra, A.M.J., 2016. Production of Short-Chain Volatile Fatty Acids
acid pretreatment of wheat straw. Biotechnol. Prog. 32, 1487–1493. and Lactic Acid during Small-Scale Ensiling of Meadow Grass. PPO/Acrres,
Ben, M., Kennes, C., Veiga, M.C., 2016. Optimization of polyhydroxyalkanoate storage Wageningen UR.
using mixed cultures and brewery wastewater. J. Chem. Technol. Biotechol. 91, Fadzil, F.I.B.M., Tsuge, T., 2017. Bioproduction of polyhydroxyalkanoate from plant oils.
2817–2826. In: Microbial Applications, 2. Springer, Cham, pp. 231–260.
Bengtsson, S., Pisco, A.R., Johansson, P., Lemos, P.C., Reis, M.A.M., 2010. Molecular Fitzpatrick, J., O'keeffe, U., 2001. Influence of whey protein hydrolysate addition to whey
weight and thermal properties of polyhydroxyalkanoates produced from fermented permeate batch fermentations for producing lactic acid. Process Biochem. 37,
sugar molasses by open mixed cultures. J. Biotechnol. 147, 172–179. 183–186.
Bong, C.P.C., Ho, W.S., Hashim, H., Lim, J.S., Ho, C.S., Tan, W.S.P., Lee, C.T., 2017. FitzPatrick, M., Champagne, P., Cunningham, M.F., Whitney, R.A., 2010. A biorefinery
Review on the renewable energy and solid waste management policies towards processing perspective: treatment of lignocellulosic materials for the production of
biogas development in Malaysia. Renew. Sust. Energ. Rev. 70, 988–998. value-added products. Bioresour. Technol. 101, 8915–8922.
Bosco, F., Chiampo, F., 2010. Production of polyhydroxyalcanoates (PHAs) using milk Follonier, S., Goyder, M.S., Silvestri, A.C., Crelier, S., Kalman, F., Riesen, R., Zinn, M.,
whey and dairy wastewater activated sludge: production of bioplastics using dairy 2014. Fruit pomace and waste frying oil as sustainable resources for the bioproduc-
residues. J. Biosci. Bioeng. 109, 418–421. tion of medium-chain-length polyhydroxyalkanoates. Int. J. Biol. Macromol. 71,
Briassoulis, D., Giannoulis, A., 2018. Evaluation of the functionality of bio-based food 42–52.
packaging films. Polym. Test. 69, 39–51. Fonseca, G.G., Antonio, R.V., 2006. Polyhydroxyalkanoates production by recombinant
Bugnicourt, E., Cinelli, P., Lazzeri, A., Alvarez, V., 2014. Polyhydroxyalkanoate (PHA): Escherichia coli harboring the structural genes of the polyhydroxyalkanoate synthases
review of synthesis, characteristics, processing and potential applications in packa- of Ralstonia eutropha and Pseudomonas aeruginosa using low cost substrate. J. Appl.
ging. Express Polym Lett 8, 791–808. Sci. 6, 1745–1750.
Bule, M.V., Chaudhary, I., Gao, A.H., Chen, S., 2016. Effects of extracellular proteome on Franca, E.F., Freitas, L.C.G., Lins, R.D., 2011. Chitosan molecular structure as a function
wheat straw pretreatment during solid-state fermentation of Phlebia radiata ATCC of N‑acetylation. Biopolymers 95, 448–460.
64658. Int. Biodeterior. Biodegrad. 109, 36–44. Franek, F., Hohenwarter, O., Katinger, H., 2000. Plant protein hydrolysates: preparation
Bussemaker, M.J., Zhang, D., 2013. Effect of ultrasound on lignocellulosic biomass as a of defined peptide fractions promoting growth and production in animal cells cul-
pretreatment for biorefinery and biofuel applications. Ind. Eng. Chem. 52, tures. Biotechnol. Prog. 16, 688–692.
3563–3580. Gironi, F., Piemonte, V., 2011. Bioplastics and petroleum-based plastics: strengths and
Buzby, J.C., Farah-Wells, H., Hyman, J., 2014. The estimated amount, value, and calories weaknesses. Energy Sources, Part A 33, 1949–1959.
of postharvest food losses at the retail and consumer levels in the United States, González-Torre, P.L., Coque, J., 2016. From food waste to donations: the case of mar-
USDA-ERS Economic Information Bulletin 121. ketplaces in northern Spain. Sustain. 8, 575.
Carlsson, M., Lagerkvist, A., Morgan-Sagastume, F., 2012. The effects of substrate pre- Gore-Langton, L., 2017. France's Food Waste Ban: One Year on [WWW Document]. Food
treatment on anaerobic digestion: a review. Waste Manag. 32, 1634–1650. Navig. URL https://www.foodnavigator.com/Article/2017/03/24/France-s-food-
Cerda, A., Artola, A., Font, X., Barrena, R., Gea, T., Sánchez, A., 2018. Composting of food waste-ban-One-year-on?utm_source=copyright&utm_medium=OnSite&utm_
wastes: status and challenges. Bioresour. Technol. 248, 57–67. campaign=copyright). (accessed 3.24.17).
Cesário, M.T., Raposo, R.S., de Almeida, M.C.M.D., van Keulen, F., Ferreira, B.S., Telo, Groot, W.J., Borén, T., 2010. Life cycle assessment of the manufacture of lactide and PLA
J.P., da Fonseca, M.M.R., 2014a. Production of poly(3‑hydro- biopolymers from sugarcane in Thailand. Int. J. Life Cycle Assess. 15, 970–984.
xybutyrate‑co‑4‑hydroxybutyrate) by Burkholderia sacchari using wheat straw hy- Gunders, D., Bloom, J., 2012. Wasted: How America Is Losing up to 40 Percent of its Food
drolysates and gamma-butyrolactone. Int. J. Biol. Macromol. 71, 59–67. from Farm to Fork to Landfill. Natural Resources Defense Council, New York. http://
Cesário, M.T., Raposo, R.S., de Almeida, M.C.M.D., van Keulen, F., Ferreira, B.S., da www.indianasna.org/content/indianasna/documents/NRDC_Wasted_Food_Report.
Fonseca, M.M.R., 2014b. Enhanced bioproduction of poly‑3‑hydroxybutyrate from pdf.
wheat straw lignocellulosic hydrolysates. New Biotechnol. 31, 104–113. Gurieff, N., Lant, P., 2007. Comparative life cycle assessment and financial analysis of
Chee, J.Y., Yoga, S.S., Lau, N.S., Ling, S.C., Abed, R.M., Sudesh, K., 2010. Bacterially mixed culture polyhydroxyalkanoate production. Bioresour. Technol. 98, 3393–3403.
produced polyhydroxyalkanoate (PHA): converting renewable resources into bio- Hafuka, A., Sakaida, K., Satoh, H., Takahashi, M., Watanabe, Y., Okabe, S., 2011. Effect of
plastics. Curr. Res. Technol. Educ. Top. Appl. Microbiol. Microb. Biotechnol. 2, feeding regimens on polyhydroxybutyrate production from food wastes by
1395–1404. Cupriavidus necator. Bioresour. Technol. 102, 3551–3553.
Chen, G.Q., 2009. A microbial polyhydroxyalkanoates (PHA) based bio- and materials Harding, K.G., Dennis, J.S., von Blottnitz, H., Harrison, S.T.L., 2007. Environmental
industry. Chem. Soc. Rev. 38, 2434–2446. analysis of plastic production processes: comparing petroleum-based polypropylene
Chen, G.Q., Patel, M.K., 2011. Plastics derived from biological sources: present and fu- and polyethylene with biologically-based poly-β-hydroxybutyric acid using life cycle
ture: a technical and environmental review. Chem. Rev. 112, 2082–2099. analysis. J. Biotechnol. 130, 57–66.
Chen, Y., Tan, H.M., 2006. Crosslinked carboxymethylchitosan‑g‑poly(acrylic acid) co- Heng, K.S., Hatti-Kaul, R., Adam, F., Fukui, T., Sudesh, K., 2017. Conversion of rice husks
polymer as a novel superabsorbent polymer. Carbohydr. Res. 341, 887–896. to polyhydroxyalkanoates (PHA) via a three-step process: optimized alkaline pre-
Chen, G.Q., Wu, Q., 2005. The application of polyhydroxyalkanoates as tissue engineering treatment, enzymatic hydrolysis, and biosynthesis by Burkholderia cepacia USM (JCM
materials. Biomaterials 26, 6565–6578. 15050). J. Chem. Technol. Biotechnol. 92, 100–108.
Chen, G.Q., Jiang, X.R., Guo, Y., 2016. Synthetic biology of microbes synthesizing Hilliou, L., Machado, D., Oliveira, C.S.S., Gouveia, A.R., Reis, M.A.M., Campanari, S.,
polyhydroxyalkanoates (PHA). Synth Syst Biotechnol. 1 (4), 236–242. Villano, M., Majone, M., 2016a. Impact of fermentation residues on the thermal,
Chiew, Y.L., Shimada, S., 2013. Current state and environmental impact assessment for structural, and rheological properties of polyhydroxy(butyrate‑co‑valerate) produced
utilizing oil palm empty fruit bunches for fuel, fiber and fertilizer - a case study of from cheese whey and olive oil mill wastewater. J. Appl. Polym. Sci. 133 (2), 42818.
Malaysia. Biomass Bioeng. 51, 109–124. Hilliou, L., Teixeira, P.F., Machado, D., Covas, J.A., Oliveira, C.S.S., Duque, A.F., Reis,
Cianchetta, S., Di Maggio, B., Burzi, P.L., Galletti, S., 2014. Evaluation of selected white- M.A.M., 2016b. Effects of fermentation residues on the melt processability and
rot fungal isolates for improving the sugar yield from wheat straw. Appl. Biochem. thermomechanical degradation of PHBV produced from cheese whey using mixed
Biotechnol. 173, 609–623. microbial cultures. Polym. Degrad. Stab. 128, 269–277.
Colombo, B., Sciarria, T.P., Reis, M., Scaglia, B., Adani, F., 2016. Polyhydroxyalkanoates Hirano, S., Tsuchida, H., Nagao, N., 1989. N‑acetylation in chitosan and the rate of its
(PHAs) production from fermented cheese whey by using a mixed microbial culture. enzymic hydrolysis. Biomaterials 10, 574–576.

641
Y.F. Tsang, et al. Environment International 127 (2019) 625–644

Hölker, U., Höfer, M., Lenz, J., 2004. Biotechnological advantages of laboratory-scale Leung, C.C.J., Cheung, A.S.Y., Zhang, A.Y.Z., Lam, K.F., Lin, C.S.K., 2012. Utilisation of
solid-state fermentation with fungi. Appl. Microbiol. Biotechnol. 64, 175–186. waste bread for fermentative succinic acid production. Biochem. Eng. J. 65, 10–15.
Hoover, R., 2010. The impact of heat-moisture treatment on molecular structures and Liang, Q., Qi, Q., 2014. From a co-production design to an integrated single-cell bior-
properties of starches isolated from different botanical sources. Crit. Rev. Food Sci. efinery. Biotechnol. Adv. 32, 1328–1335.
Nutr. 50, 835–847. Lin, C.S.K., Pfaltzgraff, L.A., Herrero-Davila, L., Mubofu, E.B., Abderrahim, S., Clark, J.H.,
HPS, A.K., Syakir, M.I., Dungani, R., Saurabh, C.K., Rafatullah, M., Haafiz, M., M.K., AS, Koutinas, A.A., Kopsahelis, N., Stamatelatou, K., Dickson, F., Thankappan, S.,
A., Nurul Fazita, M.R., Abdullah, C.K., Davoudpour, Y., 2016. A review on chitosan- Mohamed, Z., Brocklesby, R., Luque, R., 2013. Food waste as a valuable resource for
cellulose blends and nanocellulose reinforced chitosan biocomposites: properties and the production of chemicals, materials and fuels. Current situation and global per-
their applications. Carbohydr. Polym. 150, 216–226. spective. Energy Environ. Sci. 6, 426–464.
Ishiaku, U.S., Pang, K.W., Lee, W.S., Ishak, Z.M., 2002. Mechanical properties and en- Linger, J.G., Vardon, D.R., Guarnieri, M.T., Karp, E.M., Hunsinger, G.B., Franden, M.A.,
zymic degradation of thermoplastic and granular sago starch filled poly(ε‑capro- Johnson, C.W., Chupka, G., Strathmann, T.J., Pienkos, P.T., Beckham, G.T., 2014.
lactone). Eur. Polym. J. 38, 393–401. Lignin valorization through integrated biological funneling and chemical catalysis.
Isroi, I., Millati, R., Niklasson, C., Cayanto, C., Taherzadeh, M.J., Lundquist, K., 2011. Proc. Natl. Acad. Sci. 111, 12013–12018.
Biological pretreatment of lignocelluloses with white-rot fungi and its applications: Liu, H., Xie, F., Yu, L., Chen, L., Li, L., 2009. Thermal processing of starch-based polymers.
review. Bioresour. 6, 5224–5259. Prog. Polym. Sci. 34, 1348–1368.
Jayakumar, R., Selvamurugan, N., Nair, S.V., Tokura, S., Tamura, H., 2008. Preparative Liu, M., Zhang, Y., Wu, C., Xiong, S., Zhou, C., 2012. Chitosan/halloysite nanotubes
methods of phosphorylated chitin and chitosan - an overview. Int. J. Biol. Macromol. bionanocomposites: structure, mechanical properties and biocompatibility. Int J
43, 221–225. Biological Macromol. 51, 566–575.
Jayathilakan, K., Sultana, K., Radhakrishna, K., Bawa, A.S., 2012. Utilization of by- Lo, I.M., Woon, K.S., 2016. Food waste collection and recycling for value-added products:
products and waste materials from meat, poultry and fish processing industries: a potential applications and challenges in Hong Kong. Environ. Sci. Pollut. Res. 23,
review. J. Food Sci. Technol. 49, 278–293. 7081–7091.
Kachrimanidou, V., Kopsahelis, N., Vlysidis, A., Papanikolaou, S., Kookos, I.K., Monje Loo, C.Y., Lee, W.H., Tsuge, T., Doi, Y., Sudesh, K., 2005. Biosynthesis and character-
Martínez, B., Escrig Rondán, M.C., Koutinas, A.A., 2016. Downstream separation of ization of poly(3‑hydroxybutyrate‑co‑3‑hydroxyhexanoate) from palm oil products in
poly(hydroxyalkanoates) using crude enzyme consortia produced via solid state a Wautersia eutropha mutant. Biotechnol. Lett. 27, 1405–1410.
fermentation integrated in a biorefinery concept. Food Bioprod. Process. 100, Ludin, N.A., Bakri, M.A., Hashim, M., Sawilla, B., Menon, N.R., Mokhtar, H., 2004. Palm
323–334. oil biomass for electricity generation in Malaysia. Pusat Tenaga Malaysia, Malaysia
Kalyani, D., Lee, K.M., Kim, T.S., Li, J., Dhiman, S.S., Kang, Y.C., Lee, J.K., 2013. Palm Oil Board, SIRIM Berhad. See also: http://www.biogen.org.my/bris/BioGen/
Microbial consortia for saccharification of woody biomass and ethanol fermentation. Tech/(d) Documents/technology (d), 7.
Fuel 107, 815–822. Mamma, D., Christakopoulos, P., 2014. Biotransformation of citrus by-products into value
Kamilah, H., Al-Gheethi, A., Yang, T.A., Sudesh, K., 2018. The use of palm oil-based waste added products. Waste and Biomass Valori. 5, 529–549.
cooking oil to enhance the production of polyhydroxybutyrate [P(3HB)] by Martino, L., Cruz, M.V., Scoma, A., Freitas, F., Bertin, L., Scandola, M., Reis, M.A.M.,
Cupriavidus necator H16 strain. Arab. J. Sci. Eng. 43, 3453–3463. 2014. Recovery of amorphous polyhydroxybutyrate granules from Cupriavidus ne-
Kamm, B., Kamm, M., 2007. Biorefineries - multi product processes. In: Adv. Biochem. cator cells grown on used cooking oil. Int. J. Biol. Macromol. 71, 117–123.
Eng. Biotechnol. Springer, Berlin, Heidelberg, pp. 175–204. Matsakas, L., Gao, Q., Jansson, S., Rova, U., Christakopoulos, P., 2017. Green conversion
Khanna, S., Srivastava, A.K., 2005. Recent advances in microbial polyhydroxyalkanoatos. of municipal solid wastes into fuels and chemicals. Electron. J. Biotechnol. 26, 69–83.
Process Biochem. 40, 607–619. Matsumoto, K., Taguchi, S., 2013. Enzyme and metabolic engineering for the production
Kim, D., 2018. Physico-chemical conversion of lignocellulose: inhibitor effects and de- of novel biopolymers: crossover of biological and chemical processes. Curr. Opin.
toxification strategies: a mini review. Molecules 23, 309. Biotechnol. 24, 1054–1060.
Kiran, E.U., Trzcinski, A.P., Ng, W.J., Liu, Y., 2014. Bioconversion of food waste to en- Mirabella, N., Castellani, V., Sala, S., 2014. Current options for the valorization of food
ergy: a review. Fuel 134, 389–399. manufacturing waste a review. J. Clean. Prod. 65, 28–41.
Kiran, et al., 2015. Enhancing the hydrolysis and methane production potential of mixed Mohd, M.F., Mohanadoss, P., Ujang, Z., van Loosdrecht, M., Yunus, S.M., Chelliapan, S.,
food waste by an effective enzymatic pretreatment. Bioresour. Technol. 183, 47–52. Zambare, V., Olsson, G., 2012. Development of bio-PORec® system for poly-
https://doi.org/10.1016/j.biortech.2015.02.033. hydroxyalkanoates (PHA) production and its storage in mixed cultu res of palm oil
Koch, D., Mihalyi, B., 2018. Assessing the change in environmental impact categories mill effluent (POME). Bioresour. Technol. 124, 208–216.
when replacing conventional plastic with bioplastic in chosen application fields. Monavari, S., Galbe, M., Zacchi, G., 2011. The influence of ferrous sulfate utilization on
Chem. Eng. Trans. 70, 853–858. the sugar yields from dilute-acid pretreatment of softwood for bioethanol production.
Koller, M., Bona, R., Braunegg, G., Hermann, C., Horvat, P., Kroutil, M., Martinz, J., Neto, Bioresour. Technol. 102, 1103–1108.
J., Pereira, L., Varila, P., 2005. Production of PHAs from agricultural waste and Moncada, J., Matallana, L.G., Cardona, C.A., 2013. Selection of process pathways for
surplus materials. Biomacromolecules 6, 561–565. biorefinery design using optimization tools: a Colombian case for conversion of su-
Koller, M., Maršálek, L., de Sousa Dias, M.M., Braunegg, G., 2017. Producing microbial garcane bagasse to ethanol, poly‑3‑hydroxybutyrate (PHB), and energy. Ind. Eng.
polyhydroxyalkanoate (PHA) biopolyesters in a sustainable manner. New Biotechnol. Chem. Res. 52, 4132–4145.
37, 24–38. Mozejko, J., Ciesielski, S., 2013. Saponified waste palm oil as an attractive renewable
Kourmentza, C., Ntaikou, I., Kornaros, M., Lyberatos, G., 2009. Production of PHAs from resource for mcl-polyhydroxyalkanoate synthesis. J. Biosci. Bioeng. 116, 485–492.
mixed and pure cultures of Pseudomonas sp. using short-chain fatty acids as carbon Muhr, A., Rechberger, E.M., Salerno, A., Reiterer, A., Schiller, M., Kwiecień, M., Adamus,
source under nitrogen limitation. Desalination 248, 723–732. G., Kowalczuk, M., Strohmeier, K., Schober, S., Mittelbach, M., Koller, M., 2013.
Kourmentza, C., Costa, J., Azevedo, Z., Servin, C., Grandfils, C., De Freitas, V., Reis, Biodegradable latexes from animal-derived waste: biosynthesis and characterization
M.A.M., 2018a. Burkholderia thailandensis as a microbial cell factory for the bio- of mcl-PHA accumulated by Ps. citronellolis. React. Funct. Polym. 73, 1391–1398.
conversion of used cooking oil to polyhydroxyalkanoates and rhamnolipids. Mussoline, W., Esposito, G., Giordano, A., Lens, P., 2013. The anaerobic digestion of rice
Bioresour. Technol. 247, 829–837. straw: a review. Crit. Rev. Environ. Sci. Technol. 43, 895–915.
Kourmentza, C., Economou, C.N., Tsafrakidou, P., Kornaros, M., 2018b. Spent coffee Naranjo, J.M., Cardona, C.A., Higuita, J.C., 2014. Use of residual banana for poly-
grounds make much more than waste: exploring recent advances and future ex- hydroxybutyrate (PHB) production: case of study in an integrated biorefinery. Waste
ploitation strategies for the valorization of an emerging food waste stream. J. Clean. Manag. 34, 2634–2640.
Prod. 172, 980–992. Nielsen, C., Rahman, A., Rehman, A.U., Walsh, M.K., Miller, C.D., 2017. Food waste
Koutinas, A.A., Vlysidis, A., Pleissner, D., Kopsahelis, N., Lopez Garcia, I., Kookos, I.K., conversion to microbial polyhydroxyalkanoates. Microb. Biotechnol. 10, 1338–1352.
Papanikolaou, S., Kwan, T.H., Lin, C.S.K., 2014. Valorization of industrial waste and Obruca, S., Marova, I., Snajdar, O., Mravcova, L., Svoboda, Z., 2010. Production of poly
by-product streams via fermentation for the production of chemicals and biopoly- (3‑hydroxybutyrate‑co‑3‑hydroxyvalerate) by Cupriavidus necator from waste rape-
mers. Chem. Soc. Rev. 43 (8), 2587–2627. seed oil using propanol as a precursor of 3‑hydroxyvalerate. Biotechnol. Lett. 32,
Kresnowati, M., Mardawati, E., Setiadi, T., 2015. Production of xylitol from oil palm 1925–1932.
empty fruits bunch: a case study on biorefinery concept. Mod. Appl. Sci. 9 (7), Obruca, S., Marova, I., Melusova, S., Mravcova, L., 2011. Production of poly-
206–213. hydroxyalkanoates from cheese whey employing Bacillus megaterium CCM 2037. Ann.
Kumar, M.N.R., 2000. A review of chitin and chitosan applications. React. Funct. Polym. Microbiol. 61, 947–953.
46, 1–27. Obruca, S., Snajdar, O., Svoboda, Z., Marova, I., 2013. Application of random mutagen-
Kurita, K., 2006. Chitin and chitosan: functional biopolymers from marine crustaceans. esis to enhance the production of polyhydroxyalkanoates by Cupriavidus necator H16
Mar. Biotechnol. 8 (3), 203–226. on waste frying oil. World J. Microbiol. Biotechnol. 29, 2417–2428.
Kwon, M.S., Dojima, T., Park, E.Y., 2005. Use of plant-derived protein hydrolysates for Obruca, S., Benesova, P., Petrik, S., Oborna, J., Prikryl, R., Marova, I., 2014a. Production
enhancing growth of Bombyx mori (silkworm) insect cells in suspension culture. of polyhydroxyalkanoates using hydrolysate of spent coffee grounds. Process
Biotechnol. Appl. Biochem. 42, 1–7. Biochem. 49, 1409–1414.
Lam, W., Wang, Y., Chan, P.L., Chan, S.W., Tsang, Y.F., Chua, H., Yu, P.H.F., 2017. Obruca, S., Petrik, S., Benesova, P., Svoboda, Z., Eremka, L., Marova, I., 2014b. Utilization
Production of polyhydroxyalkanoates (PHA) using sludge from different wastewater of oil extracted from spent coffee grounds for sustainable production of poly-
treatment processes and the potential for medical and pharmaceutical applications. hydroxyalkanoates. Appl. Microbiol. Biotechnol. 98, 5883–5890.
Environ. Technol. 38, 1779–1791. Ong, H.C., Mahlia, T.M.I., Masjuki, H.H., 2011. A review on energy scenario and sus-
Laycock, B., Halley, P., Pratt, S., Werker, A., Lant, P., 2014. The chemomechanical tainable energy in Malaysia. Renew. Sust. Energ. Rev. 15 (1), 639–647.
properties of microbial polyhydroxyalkanoates. Prog. Polym. Sci. 38, 536–583. Pagliaccia, P., Gallipoli, A., Gianico, A., Montecchio, D., Braguglia, C.M., 2016. Single
Lemée, L., Kpogbemabou, D., Pinard, L., Beauchet, R., Laduranty, J., 2012. Biological stage anaerobic bioconversion of food waste in mono and co-digestion with olive
pretreatment for production of lignocellulosic biofuel. Bioresour. Technol. 117, husks: impact of thermal pretreatment on hydrogen and methane production. Int. J.
234–241. Hydrogen Energy 41, 905–915.

642
Y.F. Tsang, et al. Environment International 127 (2019) 625–644

Pais, J., Farinha, I., Freitas, F., Serafim, L.S., Martínez, V., Martínez, J.C., Arévalo- Sasmal, S., Goud, V.V., Mohanty, K., 2012. Ultrasound assisted lime pretreatment of
Rodríguez, M., Auxiliadora Prieto, M., Reis, M.A.M., 2014. Improvement on the yield lignocellulosic biomass toward bioethanol production. Energy and Fuels 26,
of polyhydroxyalkanotes production from cheese whey by a recombinant Escherichia 3777–3784.
coli strain using the proton suicide methodology. Enzym. Microb. Technol. 55, Sato, S., Maruyama, H., Fujiki, T., Matsumoto, K., 2015. Regulation of 3‑hydro-
151–158. xyhexanoate composition in PHBH synthesized by recombinant Cupriavidus necator
Pandey, A., 2003. Solid-state fermentation. Biochem. Eng. J. 13, 81–84. H16 from plant oil by using butyrate as a co-substrate. J. Biosci. Bioeng. 120,
Pandian, S.R., Deepak, V., Kalishwaralal, K., Rameshkumar, N., Jeyaraj, M., Gurunathan, 246–251.
S., 2010. Optimization and fed-batch production of PHB utilizing dairy waste and sea Sawant, S.S., Salunke, B.K., Kim, B.S., 2015. Degradation of corn stover by fungal cel-
water as nutrient sources by Bacillus megaterium SRKP-3. Bioresour. Technol. 101, lulase cocktail for production of polyhydroxyalkanoates by moderate halophile
705–711. Paracoccus sp. LL1. Bioresour. Technol. 194, 247–255.
Paritosh, K., Kushwaha, S.K., Yadav, M., Pareek, N., Chawade, A., Vivekanand, V., 2017. Senthilkumar, K., Mollier, A., Delmas, M., Pellerin, S., Nesme, T., 2014. Phosphorus re-
Food waste to energy: an overview of sustainable approaches for food waste man- covery and recycling from waste: an appraisal based on a French case study. Resour.
agement and nutrient recycling. Biomed. Res. Int. 2017, 2370927. Conserv. Recycl. 87, 97–108.
Pfaltzgraff, L.A., Cooper, E.C., Budarin, V., Clark, J.H., 2013. Food waste biomass: a re- Serafim, L.S., Lemos, P.C., Albuquerque, M.G., Reis, M.A., 2008. Strategies for PHA
source for high-value chemicals. Green Chem. 15, 307–314. production by mixed cultures and renewable waste materials. Appl. Microbiol.
Pielhop, T., Amgarten, J., von Rohr, P.R., Studer, M.H., 2016. Steam explosion pre- Biotechnol. 81 (4), 615–628.
treatment of softwood: The effect of the explosive decompression on enzymatic di- Shen, L., Haufe, J., Patel, M.K., 2009. Product overview and market projection of merging
gestibility. Biotechnol. Biofuels 9(1), 152. bio-based plastics PRO-BIP 2009. Report for European polysaccharide network of
Pillai, C.K.S., Paul, W., Sharma, C.P., 2009. Chitin and chitosan polymers: chemistry, excellence (EPNOE) and European bioplastics, 243.
solubility and fiber formation. Prog. Polym. Sci. 34, 641–678. Shivakumar, S., 2012. Polyhydroxybutyrate (PHB) production using agro-industrial re-
Pleissner, D., Qi, Q., Gao, C., Rivero, C.P., Webb, C., Lin, C.S.K., Venus, J., 2016. sidue as substrate by Bacillus thuringiensis IAM 12077. Int. J. Chem Tech. Res. 4,
Valorization of organic residues for the production of added value chemicals: a 1158–1162.
contribution to the bio-based economy. Biochem. Eng. J. 116, 3–16. Shuit, S.H., Tan, K.T., Lee, K.T., Kamaruddin, A.H., 2009. Oil palm biomass as a sus-
Poli, A., Di Donato, P., Abbamondi, G.R., Nicolaus, B., 2011. Synthesis, production, and tainable energy source: a Malaysian case study. Energy 34, 1225–1235.
biotechnological applications of exopolysaccharides and polyhydroxyalkanoates by Sindhu, R., Silviya, N., Binod, P., Pandey, A., 2013. Pentose-rich hydrolysate from acid
Archaea. Archaea. 2011, 693253. pretreated rice straw as a carbon source for the production of poly‑3‑hydrox-
Posada, J.A., Rincón, L.E., Cardona, C.A., 2012. Design and analysis of biorefineries based ybutyrate. Biochem. Eng. J. 78, 67–72.
on raw glycerol: addressing the glycerol problem. Bioresour. Technol. 111, 282–293. Siracusa, V., Rocculi, P., Romani, S., Rosa, M.D., 2008. R18 biodegradable polymers for
Preethi, R., Sasikala, P., Aravind, J., 2012. Microbial production of polyhydroxyalkanoate food packaging: a biodegradable polymers for food packaging: a review. Trends Food
(PHA) utilizing fruit waste as a substrate. Res. Biotechnol. 3, 61–69. Sci. Tech. 19, 634–643.
Quinaya, S.H.D., Alzate, C.A.C., 2014. Plantain and banana fruit as raw material for Smerilli, M., Neureiter, M., Wurz, S., Haas, C., Frühauf, S., Fuchs, W., 2015. Direct fer-
glucose production. J. Biotechnol. 185, S34. mentation of potato starch and potato residues to lactic acid by Geobacillus stear-
Rahmat, A.R., Rahman, W.A.W.A., Sin, L.T., Yussuf, A.A., 2009. Approaches to improve othermophilus under non-sterile conditions. J. Chem. Technol. Biotechnol. 90,
compatibility of starch filled polymer system: a review. Mater. Sci. Eng. C. 29, 648–657.
2370–2377. Sompong, O.-T., Boe, K., Angelidaki, I., 2012. Thermophilic anaerobic co-digestion of oil
Rao, U., Sridhar, R., Sehgal, P.K., 2010. Biosynthesis and biocompatibility of poly(3‑hy- palm empty fruit bunches with palm oil mill effluent for efficient biogas production.
droxybutyrate‑co‑4‑hydroxybutyrate) produced by Cupriavidus necator from spent Appl. Energy 93, 648–654.
palm oil. Biochem. Eng. J. 49, 13–20. Suwannasing, W., Imai, T., Kaewkannetra, P., 2015. Cost-effective defined medium for
Ravi, M., Song, S., Wang, J., Nadimicherla, R., Zhang, Z., 2016. Preparation and char- the production of polyhydroxyalkanoates using agricultural raw materials. Bioresour.
acterization of biodegradable poly (ε‑caprolactone)‑based gel polymer electrolyte Technol. 194, 67–74.
films. Ionics 22 (5), 661–670. Takahashi, R., Castilho, N., Silva, M., Miotto, M., Lima, A., 2017. Prospecting for marine
Ravindran, R., Jaiswal, A., 2016. Exploitation of food industry waste for high-value bacteria for polyhydroxyalkanoate production on low-cost substrates. Bioeng. 4, 60.
products. Trends Biotechnol. 34, 58–69. Tan, H.W., Aziz, A.A., Aroua, M.K., 2013. Glycerol production and its applications as a
Ray, S., Prajapati, V., Patel, K., Trivedi, U., 2016. Optimization and characterization of raw material: a review. Renew. Sust. Energ. Rev. 27, 118–127.
PHA from isolate Pannonibacter phragmitetus ERC8 using glycerol waste. Int. J. Biol. Tang, X., Alavi, S., 2011. Recent advances in starch, PVOH based polymer blends, na-
Macromol. 86, 741–749. nocomposites and tehir biodegradability. Carbohydr. Polym. 85, 7–16.
Reddy, C.S.K., Ghai, R., Kalia, V.C.C., 2003. Polyhydroxyalkanoates: an overview. Thompson, S.A., Thompson, G.G., 2004. Adequacy of rehabilitation monitoring practices
Bioresour. Technol. 87, 137–146. in the Western Australian mining industry. Ecol. Manag. Restor. 5, 30–33.
Ren, X., Chen, C., Nagatsu, M., Wang, X., 2011. Carbon nanotubes as adsorbents in en- Tommonaro, G., Pejin, B., Iodice, C., Tafuto, A., De Rosa, S., 2016. Further in vitro bio-
vironmental pollution management: a review. Chem. Eng. J. 170, 395–410. logical activity evaluation of amino-, thio- and ester-derivatives of avarol. J. Enzyme
Rhu, D.H., Lee, W.H., Kim, J.Y., Choi, E., 2003. Polyhydroxyalkanoate (PHA) production Inhib. Med. Chem. 30, 333–335.
from waste. Water Sci. Technol. 48, 221–228. Tyagi, P., Saxena, N.K., Sharma, A., 2018. Production of polyhydroxyalkanoates (PHA)
Rinaudo, M., 2006. Chitin and chitosan: properties and applications. Prog. Polym. Sci. 31, from a non-lignocellulosic component of sugarcane bagasse: fueling a biobased
603–632. economy. Biofuels Bioprod. Biorefin. 12, 536–541.
Ryu, B.G., Kim, J., Kim, K., Choi, Y.E., Han, J.I., Yang, J.W., Yang, J.W., 2013. High-cell- Valentino, F., Riccardi, C., Campanari, S., Pomata, D., Majone, M., 2015. Fate of
density cultivation of oleaginous yeast Cryptococcus curvatus for biodiesel production β‑hexachlorocyclohexane in the mixed microbial cultures (MMCs) three-stage poly-
using organic waste from the brewery industry. Bioresour. Technol. 135, 357–364. hydroxyalkanoates (PHA) production process from cheese whey. Bioresour. Technol.
Saha, B.C., Qureshi, N., Kennedy, G.J., Cotta, M.A., 2016. Biological pretreatment of corn 192, 304–311.
stover with white-rot fungus for improved enzymatic hydrolysis. Int. Biodeterior. Valentino, F., Gottardo, M., Micolucci, F., Pavan, P., Bolzonella, D., Rossetti, S., Majone,
Biodegrad. 109, 29–35. M., 2018. Organic fraction of municipal solid waste recovery by conversion into
Salehizadeh, H., Van Loosdrecht, M.C.M., 2004. Production of polyhydroxyalkanoates by added-value polyhydroxyalkanoates and biogas. ACS Sustain. Chem. Eng. 6,
mixed culture: recent trends and biotechnological importance. Biotechnol. Adv. 22, 16375–16385.
261–279. Vardon, D.R., Franden, M.A., Johnson, C.W., Karp, E.M., Guarnieri, M.T., Linger, J.G.,
Salemdeeb, R., ZU Ermgassen, E.K.H.J., Kim, M.H., Balmford, A., Al-Tabbaa, A., 2017. Salm, M.J., Strathmann, T.J., Beckham, G.T., 2015. Adipic acid production from
Environmental and health impacts of using food waste as animal feed: a comparative lignin. Energy Environ. Sci. 8, 617–628.
analysis of food waste management options. J. Clean. Prod. 140, 871–880. Vasmara, C., Cianchetta, S., Marchetti, R., Galletti, S., 2015. Biogas production from
Salgaonkar, B., Bragança, J., 2017. Utilization of sugarcane bagasse by Halogeometricum wheat straw pre-treated with ligninolytic fungi and co-digestion with pig slurry.
borinquense strain E3 for biosynthesis of poly(3‑hydro- Environ. Eng. Manag. J. 14, 1751–1760.
xybutyrate‑co‑3‑hydroxyvalerate). Bioeng. 4 (2), 50. Vázquez, J.A., Murado, M.A., 2008. Enzymatic hydrolysates from food wastewater as a
Samorì, C., Abbondanzi, F., Galletti, P., Giorgini, L., Mazzocchetti, L., Torri, C., Tagliavini, source of peptones for lactic acid bacteria productions. Enzym. Microb. Technol. 43,
E., 2015. Extraction of polyhydroxyalkanoates from mixed microbial cultures: impact 66–72.
on polymer quality and recovery. Bioresour. Technol. 189, 195–202. Vega-Castro, O., Contreras-Calderon, J., León, E., Segura, A., Arias, M., Pérez, L., Sobral,
Sanchez-Vazquez, S.A., Hailes, H.C., Evans, R.G., 2013. Hydrophobic polymers from food P.J., 2016. Characterization of a polyhydroxyalkanoate obtained from pineapple peel
waste: resources and synthesis. Polym. Rev. 53, 627–694. waste using Ralsthonia eutropha. J. Biotechnol. 231, 232–238.
Sandhya, M., Aravind, J., Kanmani, P., 2013. Production of polyhydroxyalkanoates from Venkateswar Reddy, M., Venkata Mohan, S., 2012. Influence of aerobic and anoxic mi-
Ralstonia eutropha using paddy straw as cheap substrate. Int. J. Environ. Sci. Technol. croenvironments on polyhydroxyalkanoates (PHA) production from food waste and
10, 47–54. acidogenic effluents using aerobic consortia. Bioresour. Technol. 103, 313–321.
Santhanam, A., Sasidharan, S., 2010. Microbial production of polyhydroxy alkanotes Verlinden, R.A., Hill, D.J., Kenward, M.A., Williams, C.D., Piotrowska-Seget, Z., Radecka,
(PHA) from Alcaligens spp. and Pseudomonas oleovorans using different carbon I.K., 2011. Production of polyhydroxyalkanoates from waste frying oil by Cupriavidus
sources. African J. Biotechnol. 9, 3144–3150. necator. AMB Express 1, 1–8.
Saratale, G.D., Oh, M.K., 2015. Characterization of poly‑3‑hydroxybutyrate (PHB) pro- Wang, Y.J., Hua, F.L., Tsang, Y.F., Chan, S.Y., Sin, S.N., Chua, H., Yu, P.H.F., Ren, N.Q.,
duced from Ralstonia eutropha using an alkali-pretreated biomass feedstock. Int. J. 2007. Synthesis of PHAs from waste under various C:N ratios. Bioresour. Technol. 98,
Biol. Macromol. 80, 627–635. 1690–1693.
Saritha, M., Arora, A., Lata, 2012. Biological pretreatment of lignocellulosic substrates for Wang, T., Zhang, L., Li, D., Yin, J., Wu, S., Mao, Z., 2008. Mechanical properties of
enhanced delignification and enzymatic digestibility. Indian J. Microbiol. 52, polyurethane foams prepared from liquefied corn stover with PAPI. Bioresour.
122–130. Technol. 99, 2265–2268.

643
Y.F. Tsang, et al. Environment International 127 (2019) 625–644

Weber, C.J., Haugaard, V., Festersen, R., Bertelsen, G., 2002. Production and applications Yue, H., Ling, C., Yang, T., Chen, X., Chen, Y., Deng, H., Wu, Q., Chen, J., Chen, G.Q.,
of biobased packaging materials for the food industry. Food Addit. Contam. 19, 2014. A seawater-based open and continuous process for polyhydroxyalkanoates
172–177. production by recombinant Halomonas campaniensis LS21 grown in mixed substrates.
Wen, Z., Wang, Y., De Clercq, D., 2015. Performance evaluation model of a pilot food Biotechnol. Biofuels 7, 108.
waste collection system in Suzhou City. China. J. Environ. Manage. 154, 201–207. Zahari, M.A.K.M., Zakaria, M.R., Ariffin, H., Mokhtar, M.N., Salihon, J., Shirai, Y.,
Xue, L., Liu, G., Parfitt, J., Liu, X., Van Herpen, E., Stenmarck, Å., O'Connor, C., Östergren, Hassan, M.A., 2012. Renewable sugars from oil palm frond juice as an alternative
K., Cheng, S., 2017. Missing food, missing data? A critical review of global food losses novel fermentation feedstock for value-added products. Bioresour. Technol. 110,
and food waste data. Environ. Sci. Technol. 51, 6618–6633. 566–571.
Yu, F., Liu, Y., Pan, X., Lin, X., Liu, C., Chen, P., Ruan, R., 2006. Liquefaction of corn Zahari, M.A.K.M., Ariffin, H., Mokhtar, M.N., Salihon, J., Shirai, Y., Hassan, M.A., 2015.
stover and preparation of polyester from the liquefied polyol. Appl. Biochem. Case study for a palm biomass biorefinery utilizing renewable non-food sugars from
Biotechnol. 129–132, 574–585. oil palm frond for the production of poly(3‑hydroxybutyrate) bioplastic. J. Clean.
Yu, J., Lin, F., Lin, P., Gao, Y., Becker, M.L., 2013. Phenylalanine-based poly (ester urea): Prod. 87, 284–290.
synthesis, characterization, and in vitro degradation. Macromolecules 47 (1), Zohuriaan-mehr, M.J., 2005. Advances in chitin and chitosan modification through graft
121–129. copolymerization: a comprehensive review. Polym. J. 14, 235–265.

644

You might also like