Sei sulla pagina 1di 10

Applied Energy 140 (2015) 65–74

Contents lists available at ScienceDirect

Applied Energy
journal homepage: www.elsevier.com/locate/apenergy

A low-energy, cost-effective approach to fruit and citrus peel waste


processing for bioethanol production
In Seong Choi a, Yoon Gyo Lee a, Sarmir Kumar Khanal b, Bok Jae Park c, Hyeun-Jong Bae a,d,⇑
a
Department of Wood Science and Landscape Architecture, Chonnam National University, Gwangju 500-757, Republic of Korea
b
Department of Molecular Biosciences and Bioengineering, University of Hawai’i at Mānoa, Honolulu, HI 96822, United States
c
Division of Business and Commerce, Chonnam National University, Yeosu 550-749, Republic of Korea
d
Department of Bioenergy Science and Technology, Chonnam National University, Gwangju 500-757, Republic of Korea

h i g h l i g h t s

 Simple bioprocess of bioethanol production from fruit wastes containing D-limonene.


 Two in-house enzymatic bioconversion rates were approximately 90%.
 Limonene recovery column (LRC) was designed for absorption of D-limonene.
 Ethanol production by immobilized yeast fermentation and LRC was 12-fold greater.

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

Article history: Large quantities of fruit waste are generated from agricultural processes worldwide. This waste is often
Received 22 March 2014 simply dumped into landfills or the ocean. Fruit waste has high levels of sugars, including sucrose, glu-
Received in revised form 17 November 2014 cose, and fructose, that can be fermented for bioethanol production. However, some fruit wastes, such as
Accepted 29 November 2014
citrus peel waste (CPW), contain compounds that can inhibit fermentation and should be removed for
efficient bioethanol production. We developed a novel approach for converting single-source CPW (i.e.,
orange, mandarin, grapefruit, lemon, or lime) or CPW in combination with other fruit waste (i.e., banana
Keywords:
peel, apple pomace, and pear waste) to produce bioethanol. Two in-house enzymes were produced from
Citrus peel waste
Bio ethanol
Avicel and CPW and were tested with fruit waste at 12–15% (w/v) solid loading. The rates of enzymatic
Enzymatic hydrolysis conversion of fruit waste to fermentable sugars were approximately 90% for all feedstocks after 48 h. We
D-Limonene extract also designed a D-limonene removal column (LRC) that successfully removed this inhibitor from the fruit
Continuous immobilized yeast fermentation waste. When the LRC was coupled with an immobilized cell reactor (ICR), yeast fermentation resulted in
ethanol concentrations (14.4–29.5 g/L) and yields (90.2–93.1%) that were 12-fold greater than products
from ICR fermentation alone.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction (GHG) such as nitrogen oxide (NOx) and sulfur oxide (SOx). Biofuels
are and alternative energy source that reduce the production of
The world consumed approximately 89 million barrels of crude pollution gases [2]. The production of nonpetroleum liquid fuels,
oil per day in 2013. Consumption of liquid fuels (mainly petro- such as biofuels, from food crops is not sustainable due to compe-
leum) is expected to increase to 115 million barrels per day by tition for materials and high production costs. Therefore, cheap and
2040, which is a 63% overall increase in total liquid fuel consumed. abundant nonfood materials are required as alternative biomass
The consumption of liquid fuel by the transportation sector will feedstocks (e.g., agricultural byproducts, woody biomass, or energy
increase by 57% by 2040 [1]. The transportation sector is a source crops) and processes must be developed that can efficiently and
of emissions of carbon dioxide (CO2) and other greenhouse gases economically convert these types of lignocellulosic and cellulosic
biomass into biofuels, such as bioethanol [3].
Fruit waste is generated in large quantities from the processing
⇑ Corresponding author at: Department of Bioenergy Science and Technology, of agricultural products. Examples of such waste include citrus,
Chonnam National University, Gwangju 500-757, Republic of Korea. Tel.: +82 62 banana, apple, and pear residues remaining after industrial pro-
530 2097; fax: +82 62 530 0029. cessing. Citrus, which includes oranges, grapefruits, lemons, limes,
E-mail address: baehj@chonnam.ac.kr (H.-J. Bae).

http://dx.doi.org/10.1016/j.apenergy.2014.11.070
0306-2619/Ó 2014 Elsevier Ltd. All rights reserved.
66 I.S. Choi et al. / Applied Energy 140 (2015) 65–74

mandarins, are the most abundant crops in the world. Over 115 is because although CPW is rich in various soluble and insoluble
million tons of citrus fruits are produced annually, and about 30 sugars, making it an ideal feedstock, it also contains a strong
million tons are processed industrially for juice production. After microbial inhibitor referred to as D-limonene. The production of
industrial processing, citrus peel waste (CPW) accounts for almost D-limonene from citrus peel is economically viable, as this byprod-
50% of the wet fruit mass. The annual production of bananas, uct has high added value as a flavoring agent and for various
apples, and pears are approximately 107.1, 75.5, and 24.0 million applications in the chemical industry. Thus, removing and recover-
tons, respectively, and 25–40% of this mass remains as waste after ing D-limonene prior to the yeast fermentation process serves two
processing (Fig. 1A) [4]. Fruit waste serves as cattle feed, but purposes: high-value utilization and enhanced fermentation of
because of its low protein content, it is not a high-value feedstock, CPW-derived sugars. The efficient removal of D-limonene from
and much of it is dumped into landfills or disposed of in the ocean. CPW requires a pretreatment step. Most pretreatment methods
Because fruit waste is rich in sugars and other nutrients, these are based on thermochemical or thermophysical processes such
forms of disposal may cause environmental problems. Disposal of as milling or steam explosion as shown in Fig. 1B and Table 1
waste is also becoming increasingly expensive. For example, Euro- [7–14]. A major disadvantage of these methods is the elevated
pean Union (EU) landfill directives have caused landfill gate fees to temperature and prolonged extraction time, which can cause
increase in some cases because of land limitations and transport chemical modification of the volatile molecules, including D-limonene,
and labor costs [5]. In America, the annual cost of apple pomace as well as loss of sugars for ethanol production. We developed a
disposal alone is $10 million USD [6]. Fruit waste is rich in ferment- new technique that uses raw cotton and activated carbon to
able soluble sugars such as glucose, fructose, and sucrose along remove and recover D-limonene, and requireds with less energy.
with structural cellulose and hemicellulose. These chemical con- Sorbents should have high oleophilic and hydrophobic properties,
stituents, along with the fact that fruit waste is in abundant supply, and can be classified into three groups based on the material
suggest that fruit waste may be an excellent source of waste bio- source (natural materials, treated cellulose, or petrochemical poly-
mass for ethanol production. mers). The most commonly used polymers are petrochemical poly-
However, among the variety of fruit wastes available, CPW mers such as polypropylene, polyethylene, and polyurethane.
requires additional processing before bioethanol production. This However, these polymers are non-biodegradable materials and

Fig. 1. Citrus and fruit production and schematic representation of bioethanol production processes. (A) Annual production of citrus and major fruit worldwide, (B)
traditional processes for citrus peel bioethanol production, and (C) schema of the study process. In most cases, steam explosion pretreatment is convention process to remove
the fermentation inhibitor D-limonene. Citrus peel was hydrolyzed by commercial enzymes, including pectinase, cellulase, and b-glucosidase.
I.S. Choi et al. / Applied Energy 140 (2015) 65–74 67

Table 1
Citrus waste as substrate for bioethanol production.

Substrate Pretreatment Enzymesa Fermentation process Microorganism Ethanol production References


c
Orange peel Milling Pectinase, cellulase, glucosidase HF S. cerevisiae 4.7 [7]
Orange peel Milling Pectinase, cellulase, glucosidase HF Escherichia coli KO11 2.76c [8]
Orange peel Steam explosion Pectinase, cellulase, glucosidase SSF S. cerevisiae 3.96c [9]
Orange peel Steam explosion Pectinase, cellulase, glucosidase SSF Kluyveromyces marxianus 3.45c [10]
Orange peel Acidic steam explosion Pectinase, cellulase, glucosidase SSF S. cerevisiae 2.7c [11]
Mandarin peel Steam explosion Pectinase, cellulase, glucosidase SSF S. cerevisiae 59.3d [12]
Mandarin peel Popping Pectinase, xylanase, glucosidase SHEFb S. cerevisiae 46.2d [13]
Lemon peel Steam explosion Pectinase, cellulase, glucosidase SSF S. cerevisiae 67.8d [14]
a
Commercial enzymes used for hydrolysis.
b
SHEF, separate hydrolysis and fermentation with vacuum evaporation.
c
Ethanol yields presented in%, w/v.
d
Ethanol concentration expressed in g ethanol per g of 1000 kg of fresh substrate.

can become environmental pollutants. Raw cotton is a natural Gwangju, Korea). Citrus, apple, and pear waste was collected after
material that hydrophobic, with a high sorption capacity, and is juice extraction (Hurom, Seoul, Korea). Banana waste was
easily biodegradable [15]. Activated carbon possesses a high removed, lyophilized ( 50 °C), and stored at 20 °C. CPW and fruit
degree of micro-porosity for absorption and is commonly used wastes, individually or mixed in equal ratios, waste were used for
for water treatment, detoxification, and separation of components hydrolysis and fermentation.
in flow systems.
In the enzymatic hydrolysis phase, cellulolytic, xylanolytic, and 2.2. Chemical composition
pectinolytic enzymes are often used to degrade plant cell walls and
catalyze the breakdown of complex carbohydrates into their The content of soluble sugar was analyzed by high performance
monosaccharide components (i.e., saccharification) [12,13]. Etha- liquid chromatography (HPLC) with a refractive index detector
nol production from CPW has largely been conducted using com- (2414, Waters, USA), REZEX RPM (Phenomenex, USA) column
mercial enzymes; thus, the cost of cellulosic ethanol is very high. (300  7.8 mm) at 85 °C and eluted with deionized water at a flow
The cost can be drastically reduced if in-house-produced enzymes rate of 0.6 mL per min. Insoluble solids were analyzed for neutral
are used for saccharification [16]. Trichoderma and Aspergillus are sugar content using gas chromatography (GC) [20,21]. Samples
among the most common microorganisms that produce abundant were hydrolyzed with 72% sulfuric acid for 45 min at room temper-
cellulolytic, xylanolytic, and pectinolytic enzymes. Trichoderma ature and diluted with distilled water to 4% sulfuric acid, followed
species have been studied for their cellulolytic enzymes content by autoclaving for 1 h at 121 °C. The neutral sugar composition was
[17]. In addition, Aspergillus species often have xylanolytic and pec- measured with alditol acetates containing myo-inositol as an inter-
tinolytic enzymes [18]. Both fungal species are considered extra- nal standard. Gas chromatography (GC-2010, Shimadzu, Japan)
cellular producers of cell wall-degrading enzymes that have was used, and the analysis conditions, using a DB-225 capillary col-
potential for important industrial applications. umn (30 m  0.25 mm i.d., 0.25 lm film thickness, J&W) operated
Following completion of enzymatic hydrolysis, fermentation is with He, injector temperature of 220 °C, flame ionization detector
necessary for bioethanol production. Generally, ethanol production (FID) at 250 °C, and oven temperature programming, were 100 °C
may occur through separate hydrolysis and fermentation (SHF) for 1.5 min and 5 °C/min to 220 °C.
processes, or simultaneous saccharification and fermentation The D-limonene content was determined according to a previ-
(SSF). Continuous fermentation is also considered an efficient fer- ous study [13]. Briefly, CPW was homogenized in 10 mL hexane,
mentation process, because it has many advantages, including which had a known amount of camphor as an internal standard.
the ability to separate immobilized yeast from the ethanol product, After treatment for 3 h, 5 mL of supernatant was transferred to a
thus allowing the immobilized yeast to be reused for further fer- test tube. A quantity of 0.2 mL potassium hydroxide (2 N) was
mentation. In addition, immobilizing the yeast cell wall confers added to methanol and mixed for 1 min. After the addition of
higher ethanol tolerance and cell concentrations, shorter fermenta- 1 mL distilled water, the samples were shaken and centrifuged
tion time, enhanced fermentation productivity, and lower costs of for 5 min at 3000 rpm. The hexane phase was measured by GC
recovery and recycling [19]. (CP-9100, Chrompack), using a CP-Sil 5 CB fused silica capillary col-
Considering the above-mentioned limitations of conventional umn (25 m  0.32 mm i.d., 1.2 lm film thickness, Chrompack)
processes, we explored the possibility of directly converting fruit operated with He, injector temperature 280 °C, FID at 280 °C, and
waste to ethanol without pretreatment (Fig. 1B and C). This oven temperature programming at 110 °C for 5 min and 20 °C/
involved developing an efficient enzymatic hydrolytic process, as min to 220 °C, which was then held constant for 10 min.
well as an effective, low-cost strategy to remove the fermentation
inhibitor D-limonene. The utility of this approach was examined by 2.3. In-house enzyme production
evaluating ethanol production efficiency during continuous fer-
mentation with immobilized yeast cells. Furthermore, because Aspergillus citrisporus (KCCM 11449) was obtained from the Kor-
feedstock flexibility is important for successful commercial ethanol ean Culture Center of Microorganisms (KCCM), and Trichoderma
production, the feasibility of using CPW alone or in combination longibrachiatum (KCTC 6507) was purchased from the Korean Col-
with other fruit waste was also examined. lection for Type Cultures (KCTC). The lyophilized fungi were revi-
talized on potato dextrose broth with 1.2% (w/v) agar (PDA) and
2. Materials and methods incubated for spore production for 7 days at 25oC. One hundred
of potato dextrose broth (PDB) was sterilized in 500 mL Erlen-
2.1. Raw materials meyer flasks.
Two types of carbon sources were used to produce extracellular
Citrus (orange, mandarin, grapefruit, lemon, and lime), apple, enzymes for fruit waste hydrolysis. The medium contained either
banana, and pear were obtained from a local market (Homeplus, 20 g/L MP or Avicel as the carbon source. The other components
68 I.S. Choi et al. / Applied Energy 140 (2015) 65–74

were similar for both media (in g/L): 40, peptone; 24, KH2PO4; 5, 2.5. D-Limonene recovery column design
(NH4)2SO4; 4.7, K2C4H4O64H2O; 2, urea; 1.2, MgSO47H2O and (in
mg/L) 10; ZnSO47H2O, 9.3; MnSO47H2O, 8.7; CuSO47H2O with The D-limonene removal column (LRC) is a tubular apparatus,
1 mL Tween 80. The pH was adjusted to 5.0, using hydrochloric consisting of an internal diameter (ID) of 1.5 cm and 7.0 cm length.
acid. The medium was sterilized at 121 °C for 15 min. Cultures LRC was packed with raw cotton (100–300 mg) and activated-car-
were conducted in a 10 L fermenter (Fermentec, Korea) equipped bon (0–2 g) to optimize limonene removal. LRC was connected to
with a 7 L working volume for 7 d. The culture broth was centri- the fermentation reactor for D-limonene removal and recovery
fuged and the supernatant was stored at 4 °C. from the hydrolysate prior to fermentation. The fermentation pro-
cess was conducted with- or without LRC on a fermentation reac-
tor. After fermentation, D-limonene was recovered from LRC using
2.4. In-house enzyme activity and hydrolysis hexane, and the recovery rate was determined using GC as
described in 2.2.
Two enzymes, produced in-house from A. citrisporus (In-house
enzyme A [HEA]) and T. longibrachiatum (In-house enzyme B
[HEB]), were evaluated. The protein concentration was measured 2.6. Continuous immobilized yeast fermentation
using the Lowry method, with bovine serum albumin (BSA) as a
protein standard [22]. Enzyme activities were assayed with a spe- Saccharomyces cerevisiae KCTC 7906 was obtained from the
cific substrate solution consisting of 50 mM citrate phosphate buf- KCTC and activated in 4 mL yeast peptone dextrose media (YPD).
fer, pH 4.8 (at 45 °C for 30 min), and appropriately diluted enzyme The yeast inoculum was placed in a 500 mL Erlenmeyer flask con-
concentrations. Endoglucanase (CMCase) and exoglucanase (Avice- taining 100 mL autoclaved YPD media for 24 h at 30 °C.
lase) activities were measured with 1% carboxymethylcellulose To prepare for immobilization, 100 mL S. cerevisiae cells were
(CMC, Sigma) and microcrystalline cellulose (Avicel, Sigma) as harvested at the exponential growth phase and mixed with 2%
the substrates, respectively. Xylanase activity was measured by sodium alginate solution prepared by dissolving 8 g sodium algi-
the same procedure described for endo- and exo-glucanase, but nate in 300 mL deionized water [24]. Using a syringe, the alginate
with beechwood xylan (Sigma) as the substrate. Pectinase activity drops were deposited in a 0.1 M CaCl2 solution to produce beads.
was measured with a 0.5% polygalacturonic acid (Sigma) in 50 mM The beads were stored after washing with deionized water to
citrate phosphate buffer (pH 4.8) at 45 °C for 5 min. Reducing sug- remove any remnant CaCl2. The 3.8 mm beads were uniformly
ars were quantified with dinitrosalicylic acid (DNS) at an absor- packed and stored in deionized water at 4 °C.
bance of 540 nm [23]. One unit of activity was defined as the The immobilized cell reactor (ICR) was used in continuous fer-
amount of enzyme required to release one lmol of glucose, xylose, mentation of the CPW hydrolysate. ICR consists of a tubular col-
or galacturonic acid per min. Specific activities were expressed as umn, constructed with a 2.1 cm ID and 25 cm length. About 70%
enzyme units per milligram of protein. of the column was packed with immobilized yeast cells. The med-
HEA or HEB enzymes were added to fruit waste at concentra- ium was fed into the reactor from the feed stock, and a peristaltic
tions of 12–16 and 10–25 mg protein/g fruit waste, respectively. pump (EP-1 Econopump, Biorad) was used to transfer the feed
Enzymatic hydrolysis was performed on 1% matter (w/v) with cit- medium. The volumes of the reactor before and after immobilized
rate phosphate buffer (pH 4.8) at 180 rpm for 48 h at 45 °C. Optimi- yeast cell packing were 80 and 42 mL, respectively. The fresh feed
zation of enzyme loading volume and the influence of biomass was pumped in an up-flow manner and the total sugar and ethanol
hydrolysis time during enzymatic hydrolysis were measured using concentrations were monitored during fermentation. Prior to being
HPLC as described in 2.2. fed into reactor, the pH of the CPW hydrolysates was adjusted to a

Table 2
Chemical compositions of fruit waste.

(% Dry matter) Rhamnose Arabinose Xylose Mannose Galactose Sucrose Glucose Fructose FSa Total
OP 2.1 ± 0.0 5.6 ± 0.2 2.2 ± 0.0 2.4 ± 0.1 2.7 ± 0.1 5.6 ± 0.2 35.5 ± 0.5 12.1 ± 0.4 53.2 ± 0.4 68.2 ± 0.5
MP 2.9 ± 0.1 3.3 ± 0.1 2.4 ± 0.1 2.3 ± 0.1 3.9 ± 0.1 7.4 ± 0.2 39.4 ± 1.1 10.3 ± 0.8 57.1 ± 0.6 71.9 ± 0.9
GP 3.4 ± 0.0 4.8 ± 0.2 2.3 ± 0.1 2.2 ± 0.0 3.5 ± 0.2 1.4 ± 0.1 30.6 ± 0.8 8.2 ± 0.3 43.2 ± 0.7 59.4 ± 0.8
LeP 2.1 ± 0.1 5.2 ± 0.3 2.6 ± 0.2 2.1 ± 0.1 4.6 ± 0.1 ND 27.9 ± 0.4 3.3 ± 0.1 31.2 ± 0.4 47.8 ± 0.3
LiP 2.5 ± 0.2 8.5 ± 0.4 2.5 ± 0.1 2.0 ± 0.1 4.3 ± 0.1 ND 22.5 ± 1.2 0.7 ± 0.0 23.2 ± 1.0 43.0 ± 0.6
AP 1.7 ± 0.1 5.5 ± 0.1 6.2 ± 0.3 2.8 ± 0.1 4.2 ± 0.3 9.2 ± 0.1 25.2 ± 2.8 24.7 ± 0.3 59.1 ± 1.8 79.5 ± 1.5
BP 0.6 ± 0.1 4.4 ± 0.3 5.6 ± 0.4 3.6 ± 0.1 2.8 ± 0.1 ND 30.1 ± 0.8 15.2 ± 0.7 45.3 ± 0.3 71.5 ± 0.6
PP 1.3 ± 0.1 6.0 ± 0.3 20.2 ± 0.9 2.4 ± 0.0 4.5 ± 0.3 1.9 ± 0.1 21.1 ± 0.6 14.1 ± 0.5 37.1 ± 0.8 62.3 ± 0.7
MixP 2.7 ± 0.1 5.6 ± 0.2 2.4 ± 0.2 2.2 ± 0.0 3.8 ± 0.0 2.9 ± 0.1 32.0 ± 0.8 6.8 ± 0.3 41.7 ± 1.1 58.4 ± 0.8
TFW 2.0 ± 0.1 5.4 ± 0.4 5.5 ± 0.6 2.5 ± 0.1 3.8 ± 0.2 3.2 ± 0.3 29.0 ± 1.7 11.1 ± 0.9 43.4 ± 1.9 62.5 ± 1.7

Abbreviations used: OP, orange peel; MP, mandarin peel; GP, grapefruit peel; LeP, lemon peel; LiP, lime peel; AP, apple pomace; BP, banana peel; PP, pear peel; MixP, mixed
citrus peel; TFW, mixed total fruit wastes; ND, not detected.
Values represent the average of three replicates.
a
FS: Fermentable sugars are the sum of sucrose, glucose, and fructose, which are fermented by S. cerevisiae.

Table 3
Comparison of specific activities for the in-house enzymes used in the study.

Endoglucanase (U/mg protein) Exoglucanase (U/mg protein) Xylanase (U/mg protein) Pectinase (U/mg protein)
In-house enzyme A (HEA) 8.41 ± 0.11 0.18 ± 0.01 170.95 ± 1.81 17.90 ± 0.43
In-house enzyme B (HEB) 13.22 ± 1.21 1.26 ± 0.17 4.34 ± 0.52 1.11 ± 0.21
I.S. Choi et al. / Applied Energy 140 (2015) 65–74 69

pH of 4.8 by the addition of CaCO3. The flow rate of feed in the

88.7 ± 1.6
88.4 ± 1.6
88.4 ± 2.1
70.9 ± 1.0
84.1 ± 2.0
packed-bed reactor column was 0.08 mL/min. The ICR was main-
tained in an incubator at 30 °C, and samples were withdrawn

Abbreviations used: OP, orange peel; MP, mandarin peel; GP, grapefruit peel; LeP, lemon peel; LiP, lime peel; MixP, mixed citrus peel; TFW, mixed total fruit wastes; HEA, in-house enzyme A; HEB, in-house enzyme B.
LiP
aseptically from the bioreactor periodically during a 10-day per-
iod to analyze sugar and ethanol concentrations.

71.6 ± 2.2
85.7 ± 2.1
89.0 ± 1.2

89.4 ± 1.7
89.1 ± 0.9
Group B
3. Results and discussion

LeP
3.1. Fruit waste composition

79.6 ± 1.1
84.7 ± 1.9

91.2 ± 2.1
73.0 ± 1.0

91.1 ± 2.0
Conversion rates for various types of citrus peel waste (CPW), alone or in combination with other fruit wastes after treatment with in-house enzymes (HEA and HEB) at different loading volumes.
The carbohydrate composition of the various fruit wastes dif-

TFW
fered as shown in Table 2. The total carbohydrate contents of the
fruit wastes were separated into soluble sugars, which dissolve

73.1 ± 1.7
84.1 ± 1.5
87.8 ± 2.2

90.4 ± 1.7
90.0 ± 2.0
easily in water, and insoluble sugars (cellulose and hemicellu-

MixP
lose) in the cell walls. Although arabinose and xylose were pres-
ent, they appeared in low concentrations in the fruit waste. We
mainly focused on fermentable sugars (FS), namely, glucose, fruc-

72.5 ± 1.6
80.1 ± 1.5
82.7 ± 1.7

90.3 ± 1.9
90.1 ± 1.0
tose, and sucrose. All fruit wastes presented were high in FS con-
tent (Table 2). Sucrose and fructose were present as soluble free

GP
sugars, whereas, glucose was part of the fruit waste structural
components and present as a free sugar. FS contents in the vari-

16 mg HEA /g fruit waste

73.1 ± 1.2
82.4 ± 2.9
86.1 ± 1.6
90.8 ± 1.2
90.7 ± 1.5
ous fruit wastes ranged from 23.2% to 59.1%. Orange peel (OP),
mandarin peel (MP), grapefruit peel (GP), apple pomace (AP),

MP
and banana peel (BP) waste contained 53.2%, 57.1%, 43.2%,
59.1%, and 45.3% FS, respectively. Lemon peel (LeP), lime peel
(LiP), and pear pomace (PP) showed moderate FS levels of

70.4 ± 2.2
83.4 ± 3.5
85.7 ± 2.1
90.2 ± 1.4
90.5 ± 2.0
Group A
31.2%, 23.2%, and 37.1%, respectively. FS contents in the CPW
mixture (MixP) and CPW, in combination with other fruit waste

OP
(TFW), were 41.7% and 43.4%, respectively.

70.6 ± 1.2

88.7 ± 1.2
88.9 ± 1.7
81.1 ± 0.8

88.9 ± 2.0
3.2. In-house enzyme production and fruit waste hydrolysis

LiP
The current cost of pretreatment and enzymes for biomass

70.1 ± 1.9

89.0 ± 2.1
89.1 ± 1.8
89.1 ± 2.1
82.1 ± 2.0
hydrolysis are major obstacles to large-scale ethanol production
Group B

[25]. The cost of cellulase is estimated at a minimum of $10/kg


LeP

protein [16]. Accordingly, it is necessary to reduce the cost and


amount of enzymes required for biomass hydrolysis to industrial-
ize the process. Here, we produced a suitable enzyme complex for
69.1 ± 1.7
72.8 ± 1.6
79.8 ± 2.1
84.1 ± 1.8
85.8 ± 3.5
fruit waste hydrolysis using CPW or Avicel as the carbon source.
TFW

In addition, we also report on the efficacy of the in-house enzyme


activities and fruit waste hydrolysis.
66.8 ± 1.5
75.4 ± 1.2
77.1 ± 3.3
76.2 ± 2.2
83.4 ± 1.1

3.2.1. In-house enzyme activity and effective loading volumes for


MixP

hydrolysis
Between the two in-house enzymes evaluated (Table 3), HEA
exhibited the highest level of xylanase activity. Its pectinase
71.6 ± 1.5
73.6 ± 3.1
78.7 ± 1.6
80.1 ± 1.2
64.0 ± 0.9

activity was moderate, and both exoglucanase and endoglucan-


ase activities were observed. The activity of xylanase and pectin-
GP

ase were lower for HEB, compared to that of HEA, but


12 mg HEA/g fruit waste

endoglucanase and exoglucanase activities were higher or HEB.


65.8 ± 1.2

76.1 ± 1.8
81.2 ± 2.2
82.4 ± 3.3
73.1 ± 0.9

Interestingly, HEA was produced using CPW as the carbon source,


and it showed high xylanase activity. This may have occurred
MP

because hemicellulose forms a large component of the polysac-


charides in CPW [13,26], and xylanase produced monosaccha-
63.2 ± 1.8
72.1 ± 1.2
75.4 ± 3.1
80.4 ± 2.2
81.2 ± 2.6
Group A

rides by CPW hydrolysis for fungal survival. Microorganisms


produce the appropriate complex enzymes for hydrolysis during
OP

growth on a given substrate. The presence of hemicellulose-


derived saccharides in CPW is thought to be important for HEA
Mg HEB/g fruit waste

induction. Based on the above mentioned HEA and HEB activities,


Conversion rate (%)

we designed a synergistic cooperation between cellulolytic,


xylanolytic and pectinolytic enzyme mixtures to hydrolysis. To
determine the amount of enzyme necessary for fruit waste
hydrolysis, different enzyme volumes were loaded onto OP, MP,
Table 4

15

25
10

20
0

GP, LeP, LiP, MixP, and TFW substrates. Data for the hydrolysis
of various fruit wastes by HEA and HEB are shown in Table 4.
70 I.S. Choi et al. / Applied Energy 140 (2015) 65–74

Fig. 2. Waste-to-FS conversion rates as influenced by substrate loadings (%, w/w).

Table 5
Influence of enzymatic hydrolysis time on various kinds of citrus and mixed fruit waste.

Time (h) 3 6 9 12 15 18 21 24 48
Group A (12%) OP 53.0 ± 2.3 64.0 ± 2.0 76.2 ± 2.7 84.8 ± 3.1 86.2 ± 1.3 87.3 ± 1.8 88.8 ± 2.0 89.5 ± 2.2 90.2 ± 1.7
MP 51.8 ± 1.9 66.3 ± 3.0 78.8 ± 2.1 85.7 ± 2.7 87.4 ± 3.1 88.5 ± 2.0 89.3 ± 3.3 89.7 ± 2.6 90.8 ± 2.1
GP 51.8 ± 2.5 65.3 ± 3.1 75.6 ± 2.9 85.7 ± 2.1 87.4 ± 3.5 88.5 ± 2.5 89.3 ± 2.5 89.7 ± 3.0 90.1 ± 2.3
MixP 51.8 ± 1.5 65.0 ± 2.4 76.3 ± 2.3 85.7 ± 2.4 87.4 ± 2.7 88.5 ± 3.0 89.3 ± 3.2 89.7 ± 2.8 90.0 ± 1.8
TFW 52.1 ± 2.1 67.2 ± 2.0 79.1 ± 2.6 87.8 ± 2.1 89.2 ± 2.3 90.1 ± 1.8 90.3 ± 2.7 90.8 ± 2.1 91.4 ± 2.1
Group B (15%) LeP 45.8 ± 2.1 61.2 ± 2.6 72.4 ± 1.9 80.1 ± 2.6 85.3 ± 1.9 87.0 ± 2.2 87.8 ± 1.7 89.3 ± 2.6 89.0 ± 1.7
LiP 47.2 ± 2.9 64.3 ± 1.9 73.8 ± 1.9 79.4 ± 3.0 86.7 ± 2.7 87.8 ± 1.9 88.0 ± 1.9 88.0 ± 2.4 88.7 ± 2.0

Abbreviations used: OP, orange peel; MP, mandarin peel; GP, grapefruit peel; LeP, lemon peel; LiP, lime peel; MixP, mixed citrus peel; TFW, mixed total fruit wastes.
Group A and B concentrations were 12% and 15% (w/w) solid loading, respectively.

Table 6
Summary of ethanol production from LRC–ICR system.

Fermentable sugar content (g/L) Enzymatic hydrolysate (g/L, %a) Ethanol concentration (g/L) Ethanol yield (%) Productivity (g/L/h)
OP 63.8 57.5/90.2% 27.1 92.4 3.01
MP 68.5 62.2/90.8% 29.5 93.1 3.28
GP 51.8 46.7/90.1% 21.6 90.7 2.40
MixP 50.0 44.4/90.0% 20.4 90.2 2.27
TFW 47.4 43.3/91.4% 20.3 91.8 2.26
LeP 46.8 42.1/89.0% 19.6 91.1 2.18
LiP 34.8 31.0/88.7% 14.4 90.8 1.60

Ethanol yield was calculated based on the fermentable sugars obtained from the hydrolysis of fruit waste.
Theoretical ethanol yield was assumed to be 0.51 g/g sugar.
a
Enzymatic hydrolysis efficiency.

Although effects on fruit waste hydrolysis rates were species- therefore, increasing accessibility and probability that the cellulose
dependent, the relatively low loading of HEA supplemented with will become hydrolyzed [27,28]. A combination of 16 mg HEA and
HEB achieved a high overall hydrolysis rate. An increase in FS con- 20 mg HEB, or 12 mg HEA and 15 mg HEB per g fruit waste, was
centrations from OP, MP, GP, MixP, and TFW (group A) was used in all further experiments for group A or B, respectively. In
observed when HEA levels were increased from 12 to 16 mg HEA this study, treatment with HEA invertase resulted in a decrease
with 20 mg HEB per g fruit waste. The FS from LeP and LiP (group in sucrose levels (through hydrolysis) and corresponding increases
B) increased at lower enzyme loadings (HEA 12 and HEB 15 mg/g in monomers fructose and glucose (Supplementary Fig. S1). The
fruit waste) compared to loadings required in group A. The differ- hydrolysis of sucrose can be an issue in continuous bioethanol pro-
ent chemical components of fruit waste may lead to differences in duction. This is because S. cerevisiae shows preferential consump-
enzymatic hydrolysis processes. However, FS concentration was tion of glucose and fructose over sucrose during fermentation,
not increased significantly, even added more enzymes to group A and, as a result, sucrose is only consumed when the former two
and B. This may have occurred because the hydrolysis of hemicel- substrates are exhausted. These differences in the kinetics of sugar
lulose and pectin increases the surface area of the fruit waste and, consumption may limit bioethanol production from fruit waste.
I.S. Choi et al. / Applied Energy 140 (2015) 65–74 71

Fig. 3. Limonene removal and recovery. (A) The sorbent column was filled with raw cotton and activated carbon. (B) Citrus peel and mixed fruit waste contained different
D-limonene concentrations. (C) D-limonene from orange peel (black arrow) was detected by gas chromatography, before and after recovery, and (D) D-limonene was recovered
after fermentation.

According to Ghorbani et al. [29], to increase fermentation effi- CPW hydrolysis within the first hours of the reaction, followed
ciency and avoid limitations, sucrose must be hydrolyzed to glu- by a significant decrease. An FS conversion rate of approximately
cose and fructose via sucrose hydrolysis enzymes. 85% was achieved within the first 12 h for group A. However, group
B required 15 h to reach a similar level of conversion. Conversion
3.2.2. Influences of fruit waste concentration and time on enzymatic rates did not increase significantly after 12 and 15 h in groups A
hydrolysis and B, respectively. From an economic perspective, these results
Based on the enzyme loading results in Table 4, we evaluated are favorable given that they permit a high degree of conversion,
the effects of varying fruit waste concentrations on the enzymatic which is necessary to maximize yield of ethanol during fermenta-
conversion of waste to FS. The conversion rates were calculated tion. Moreover, the overall hydrolysis time required was relatively
based on the total FS of fruit waste. Fruit waste solid loadings var- short compared to previous studies examining ethanol production
ied from 3% to 18% (Fig. 2). For group A (OP, MP, GP, MixP, and from lignocellulosic biomass [19,27,30].
TFW) and group B (LeP and LiP), conversion rates decreased only
slightly as substrate loadings increased from 3% to 12% and 3% to 3.3. D-limonene recovery and continuous immobilized yeast
15%, respectively. After this point, further increases in substrate fermentation
loading significantly decreased conversion rates in group A
(>12%) and group B (>15%). Based on these results, all further
experiments used solid waste loadings of 12% for group A wastes, 3.3.1. Development of a D-limonene adsorbent column
and 15% for group B wastes. In addition to the influence of sub- Citrus contains D-limonene, a terpenoid essential oil that inhib-
strate loading, we examined the influence of hydrolysis time on its yeast fermentation. In conventional processing, D-limonene is
waste-to-FS conversion (Table 5). FS conversion rates were high removed and recovered using energy-intensive methods, such as
within the first 3 h, and considerable conversion of CPW to FS steam explosion. In contrast, conventional pretreatment method
was achieved within 9 h of hydrolysis. This kinetic behavior is in has a major disadvantage, in that carbohydrate content of the feed-
agreement with our previous work [13], which showed rapid stock may decrease to as low as 10% after pretreatment due to
72 I.S. Choi et al. / Applied Energy 140 (2015) 65–74

Fig. 4. Comparisons of fermentable sugar conversion and ethanol concentrations in ICR vs. LRC–ICR fermentation. (A) Initial FS concentrations in OP, (B) MixP, and (C) TFW
were 57.5, 44.4, and 47.4 g/L, respectively. (D) The glucose-to-ethanol conversion rates obtained after 10 days of fermentation. Black solid lines indicate the amount of FS from
ICR ( ) or LRC–ICR ( ) fermentation, and gray solid lines indicate the amount of ethanol produced from ICR ( ) or LRC–ICR ( ) fermentation.

losses resulting from the Maillard reaction, caramelization, and 3.3.2. Immobilized yeast fermentation
oxidation [13,31–33]. With the aim of developing a more cost- When immobilizing cells onto a solid matrix such as calcium
effective, low-energy solution to this issue, we devised an LRC alginate beads, a number of factors can affect the penetration of
made of raw cotton and activated carbon, and we evaluated its cells into the bead and ultimately the conversion of FS to ethanol.
removal efficiency through gas chromatography (GC). Factors affecting bead penetration include alginate content, the
To evaluate the effects of column packing on D-limonene ratio of yeast cells to alginate, and pore size. In a previous study,
removal rate, various weights of raw cotton and activated carbon these factors were optimized and we identified a suitable alginate
were used to construct the LRCs. Decreasing D-limonene concen- microlattice matrix for our bioethanol reactor, known as an egg-
trations in the filtrate were observed when the raw cotton weight box structure [24].
was increased from 100 to 300 mg per column, as well as when the Calcium alginate beads and cultured S. cerevisiae were used to
quantity of activated carbon was increased from 0 to 1.5 g. How- construct an ICR, with which we evaluated fermentation efficiency
ever, further improvements in D-limonene adsorption were not using a number of different fruit waste hydrolysates. Total FS, eth-
observed when using >2.0 g activated carbon. Thus LRCs containing anol concentrations, and FS-to-ethanol conversion rates were
300 mg raw cotton and 1.5 g activated carbon (as shown in Fig. 3A) obtained using two types of fermentation processes: ICR alone
were used in all further experiments. Analysis of the fresh CPW and LRC followed by ICR (LRC–ICR). The volume metric ethanol
showed contents of 0.321–1.858% (w/w) D-limonene (Fig. 3B); productivity (g/L/h) was calculated by dividing final ethanol con-
however, after passing through the LRC, D-limonene was undetect- centration with respect to fermentation time (Table 5). The initial
able. This result represented an improvement in D-limonene FS concentrations in OP, MP, GP, LeP, LiP, MixP, and TFW were
removal compared to the conventional method used in previous 57.5, 62.2, 46.7, 42.1, 31.0, 44.4, and 43.3 g/L, respectively. The rel-
studies. In previous studies, inhibition of fermentation processes ative FS concentrations decreased with increasing time in the LRC–
was observed at concentrations greater than or equal to 0.1% (v/ ICR, especially over the first 9 h, whereas ethanol concentrations
w) [9]. Grohmann et al. [7] have shown inhibitory minimum con- increased. After 9 h, FS concentrations in OP, MixP, and TFW had
centrations, between 0.05% and 0.15% (v/w), that can affect the fer- fallen to 2.4, 2.3, and 2.6 g/L, respectively, and ethanol concentra-
mentation process. Moreover, about 90% of the D-limonene was tions had increased to 27.1, 29.5, and 20.3 g/L, respectively
recovered after a 10-day fermentation period with a 0.08 mL/min (Fig. 4A–C). When using ICR alone, without prior removal of D-lim-
flow rate (Fig. 3C and D). These removal rates could be due to onene, FS concentrations were subsequently lower. After 9 h, FS
the fact that D-limonene concentration and viscosity are low in concentrations for OP, MixP, and TFW were 51.9, 17.9, and
the hydrolysate. Because oil penetration rate into the internal sur- 20.8 g/L, respectively, whereas ethanol concentrations were 2.7,
face of sorbents is inversely proportional to oil viscosity and con- 15.9, and 12.9 g/L, respectively. In the LRC–ICR system, the FS-to-
centration [34], adsorption should be high in the pores of the ethanol conversion rates for OP, MixP, and TFW feedstocks were
raw cotton and activated carbon in the LRC. Importantly, FS con- 92.4%, 90.2% and 91.8%, respectively, after 10 d, whereas no further
centrations remained unchanged in the LRC filtrate. ethanol production was observed in the ICR system after the first
I.S. Choi et al. / Applied Energy 140 (2015) 65–74 73

9 h (Fig. 4D). These results are likely due to high D-limonene con- Korea (NRF) funded by the Ministry of Education, Science and
centrations and its inhibitory effect on fermentation in the ICR sys- Technology, and by a grant (S211314L010120) from Forest Science
tem. Regarding the remaining feedstocks, MP and GP showed high & Technology Projects, Forest Service, Republic of Korea.
FS contents and low ethanol concentrations after ICR fermentation,
similar to the results obtained for OP, MixP, and TFW following ICR
Appendix A. Supplementary material
fermentation. After ICR fermentation, LeP and LiP showed FS con-
tents of 5.5 and 3.8 g/L, respectively, and ethanol concentrations
Supplementary data associated with this article can be found, in
of 20.2 and 15.1 g/L, respectively (Supplementary Fig. S2A–D).
the online version, at http://dx.doi.org/10.1016/j.apenergy.201
The FS-to-ethanol conversion rates for LeP and LiP in the ICR sys-
4.11.070.
tem were only slightly lower compared to the LRC–ICR system
(Supplementary Fig. S2a–d). This may have occurred because the
initial D-limonene concentrations in the LeP and LiP hydrolysates References
were insufficient to inhibit fermentation. However, these results
[1] International Energy Outlook. The International Energy Agency; 2013. <http://
indicate that the LRC–ICR fermentation system improved the FS- www.eia.gov/forecasts/ieo/pdf/0484(2013).pdf>.
to-ethanol conversion rates and ethanol concentrations even at [2] Liu T, McConkey B, Huffman T, Smith S, MacGregor B, Yemshanov D, et al.
low D-limonene concentrations. Interestedly, the ethanol produc- Potential and impacts of renewable energy production from agricultural
biomass in Canada. Appl Energy 2014;130:222–9.
tivity of OP and MP, which are major citrus biomass sources, were [3] Demirbas A. Competitive liquid biofuels from biomass. Appl Energy
3.01 and 3.28 g/L/h, respectively, through the LRC–ICR system. In 2011;88:17–28.
other words, 1000 kg fresh OP and MP (19.8% and 20.1% of mois- [4] Food and Agriculture Organization of the United Nations. FAOSTAT; 2013.
<http://faostat3.fao.org/faostat-gateway/go/to/browse/Q/QC/E>.
ture contents) would be converted into 44.8 and 49.5 L of bioetha- [5] Lin CSK, Pfaltzgraff LA, Herrero-Davila L, Mubofu EB, Abderrahim S, Clark JH,
nol, respectively (Table 6). et al. Food waste as a valuable resource for the production of chemicals,
Several previous studies have examined the effects of pretreat- materials and fuels. Current situation and global perspective. Energy Environ
Sci 2013;6:426–64.
ment on CPW composition and subsequent bioethanol production; [6] Shalini R, Gupta DK. Utilization of pomace from apple processing industries: a
however, ethanol concentrations and productivities obtained in review. J Food Sci Technol-Mysore 2010;47:365–71.
this study were similar to or greater than those observed in previ- [7] Grohmann K, Baldwin EA, Buslig BS. Production of ethanol from enzymatically
hydrolyzed orange peel by the yeast Saccharomyces cerevisiae. Appl Biochem
ous studies. For example, ethanol production from OP, using steam Biotechnol 1994;45:315–27.
explosion combined with acid pretreatment, produced 25–27 g/L [8] Grohmann K, Baldwin EA, Buslig BS, Ingram LO. Fermentation of galacturonic
ethanol concentration with around 0.5 g/L/h productivity [11]. acid and other sugars in orange peel hydrolysates by the ethanologenic strain
of Escherichia coli. Biotechnol Lett 1994;16:281–6.
Another study reported that the fermentation of MP and LeP after
[9] Wilkins MR, Widmer WW, Grohmann K. Simultaneous saccharification and
steam explosion produced approximately 60 L/1000 kg (fresh mat- fermentation of citrus peel waste by Saccharomyces cerevisiae to produce
ter) of ethanol concentration with 0.5–0.94 g/L/h productivity, ethanol. Process Biochem 2007;42:1614–9.
respectively [12,14]. [10] Widmer WW, Narciso JA, Grohmann K, Wilkins MR. Simultaneous
saccharification and fermentation of orange processing waste to ethanol
Fruit waste and other solid residues, such as coffee waste and using Kluyveromyces marxianus. J Biol Eng 2009;2:17–29.
rice, from agricultural by-products were considered bioethanol [11] Widmer WW, Zhou W, Grohmann K. Pretreatment effects on orange
production materials [13,21,33]. One main obstacle to achieving processing waste for making ethanol by simultaneous saccharification and
fermentation. Bioresour Technol 2010;101:5242–9.
efficient bioethanol production is the cost of production. The com- [12] Boluda-Aguilar M, García-Vidal L, González-Castañeda FP, López-Gómez A.
mercial success of ethanol production depends on productivity, in Mandarin peel wastes pretreatment with steam explosion for bioethanol
terms of volume and concentration. Notably, our new process production. Bioresour Technol 2010;101:3506–13.
[13] Choi IS, Kim J-H, Wi SG, Kim K-H, Bae H-J. Bioethanol production from
achieved high ethanol production without costly pretreatment, mandarin (Citrus unshiu) peel waste using popping pretreatment. Appl Energy
suggesting utility in industrial ethanol production applications. 2013;102:204–10.
The high ethanol production during the validation experiment [14] Boluda-Aguilar M, López-Gómez A. Production of bioethanol by fermentation
of lemon (Citrus limon L.) peel wastes pretreated with steam explosion. Ind
could be due to several factors, including suitable inhibitor Crop Prod 2013;41:188–97.
removal conditions, enzyme production, loading volume, and con- [15] Radetic M, Ilic V, Radojevic D, Miladinovic R, Jocic D, Jovancic P. Efficiency of
tinuous yeast fermentation. recycled wool-based nonwoven material for the removal of oils from water.
Chemosphere 2008;70:525–30.
[16] Klein-Marcuschamer D, Oleskowicz-Popiel P, Simmons BA, Blanch HW. The
challenge of enzyme cost in the production of lignocellulosic biofuels.
4. Conclusion Biotechno Bioeng 2012;109:1083–7.
[17] Martinez D, Berka RM, Henrissat B, Saloheimo M, Arvas M, Baker SE, et al.
Fruit waste is an attractive biomass alternative for bioethanol Genome sequencing and analysis of the biomass-degrading fungus
Trichoderma reesei (syn. Hypocrea jecorina). Nat Biotechnol 2008;26:553–60.
production because it has high levels of FS such as sucrose, glucose,
[18] de Vries RP, Visser J. Aspergillus enzymes involved in degradation of plant cell
and fructose. In this study, these sugars were hydrolyzed and fer- wall polysaccharides. Microbiol Mol Biol Rev 2001;65:497–522.
mented without an energy-intensive conventional pretreatment. [19] Lin Y, Tanaka S. Ethanol fermentation from biomass resources: current state
After enzymatic hydrolysis with two in-house enzymes, D-limo- and prospects. Appl Microbiol Biotechnol 2006;69:627–42.
[20] Wi SG, Chung BY, Lee YG, Yang DJ, Bae H-J. Enhanced enzymatic hydrolysis of
nene was removed using an adsorbent column containing raw cot- rapeseed straw by popping pretreatment for bioethanol production. Bioresour
ton and activated carbon and directly conducted to an immobilized Technol 2011;102:5788–93.
reactor (LRC–ICR) for fermentation. Ethanol production in this [21] Wi SG, Choi IS, Kim KH, Kim HM, Bae H-J. Bioethnol production from rice straw
by popping pretreatment. Biotechnol Biofuels 2013;6:166.
LRC–ICR system was 12-fold greater than that observed without [22] Lowry OH, Rosenbrough NJ, Fair AL, Randall RJ. Protein measurement with the
prior use of the sorbent column (LRC) to remove the fermentation Folin-phenol reagents. J Biol Chem 1951;193:265–75.
inhibiting D-limonene. This new approach to removing D-limonene [23] Miller GL. Use of dinitrosalicylic acid reagent for determination of reducing
sugar. Anal Chem 1959;31:426–8.
and enhancing immobilized yeast fermentation could potentially [24] Lee KH, Choi IS, Kim YG, Yang DJ, Bae H-J. Enhanced production of bioethanol
be useful in more cost-effective bioethanol production. and ultrastructural characteristics of reused Saccharomyces cerevisiae
immobilized calcium alginate beads. Bioresour Technol 2011;102:8191–8.
[25] Berlin A, Gilkes N, Kilburn D, Bura R, Markov A, Skomarovsky A, et al.
Acknowledgements Evaluation of novel fungal cellulase preparations for ability to hydrolyze
softwood substrates – evidence for the role of accessory enzymes. Enzyme
Microb Technol 2005;37:175–84.
This work was supported by Priority Research Centers Program [26] Van Dyk JS, Gama R, Morrison D, Swart S, Pletschke BI. Food processing waste:
(2010-0020141) through the National Research Foundation of problems, current management and prospects for utilization of the
74 I.S. Choi et al. / Applied Energy 140 (2015) 65–74

lignocellulose component through enzyme synergistic degradation. Renew [31] Cantarella M, Cantarella L, Gallifuoco A, Spera A, Alfani F. Effect of inhibitors
Sust Energ Rev 2013;26:521–31. released during steam-explosion treatment of poplar wood on subsequent
[27] Sun Y, Cheng J. Hydrolysis of lignocellulosic materials for ethanol production: enzymatic hydrolysis and SSF. Biotechnol Prog 2004;20:200–6.
a review. Bioresour Technol 2002;83:1–11. [32] Jørgensen H, Kristensen JB, Felby C. Enzymatic conversion of lignocellulose
[28] Alvira P, Tomás-Pejó E, Ballesteros M, Negro MJ. Pretreatment technologies for into fermentable sugars: challenges and opportunities. Biofuels Bioprod
an efficient bioethanol production process based on enzymatic hydrolysis: a Biorefining 2007;1:119–34.
review. Bioresour Technol 2010;101:4851–61. [33] Choi IS, Wi SG, Kim S-B, Bae H-J. Conversion of coffee residue waste into
[29] Ghorbani F, Younesi H, Esmaeili Sari A, Najafpour G. Cane molasses bioethanol with popping pretreatment. Bioresour Technol 2012;125:132–7.
fermentation for continuous ethanol production in an immobilized cells [34] Srinivasan A, Viraraghavan T. Removal of oil by walnut shell media. Bioresour
reactor by Saccharomyces cerevisiae. Renew Energy 2011;36:503–9. Technol 2008;99:8217–20.
[30] Wyman CE. What is (and is not) vital to advancing cellulosic ethanol. Trends
Biotechnol 2007;25:153–7.

Potrebbero piacerti anche