You are on page 1of 29

Progress in Energy and Combustion Science 47 (2015) 60e88

Contents lists available at ScienceDirect

Progress in Energy and Combustion Science


journal homepage: www.elsevier.com/locate/pecs

Review

Challenges and opportunities in improving the production of


bio-ethanol
Jan Baeyens a, c, Qian Kang b, c, *, Lise Appels b, Raf Dewil b, Yongqin Lv c, Tianwei Tan c
a

University of Warwick, School of Engineering, Coventry, UK


KU Leuven, Department of Chemical Engineering, Process and Environmental Technology Lab, Sint-Katelijne-Waver, Belgium
c
Beijing University of Chemical Technology, College of Life Science and Technology, Key Lab of Bioprocess, Beijing, China
b

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 25 June 2014
Accepted 22 October 2014
Available online 15 December 2014

Bio-ethanol, as a clean and renewable fuel, is gaining increasing attention, mostly through its major
environmental benets. It can be produced from different kinds of renewable feedstock such as e.g. sugar
cane, corn, wheat, cassava (rst generation), cellulose biomass (second generation) and algal biomass
(third generation). The conversion pathways for the production of bio-ethanol from disaccharides, from
starches, and from lignocellulosic biomass are examined. The common processing routes are described,
with their mass and energy balances, and assessed by comparing eld data and simulations. Improvements through 5 possible interventions are discussed, being (i) an integrated energy-pinch of condensers
and reboilers in the bio-ethanol distillation train; (ii) the use of Very High Gravity (VHG) fermentation;
(iii) the current development of hybrid processes using pervaporation membranes; (iv) the substitution
of current ethanol dewatering processes to >99.5 wt% pure ethanol by membrane technology; and (v)
additional developments to improve the plant operation such as the use of microltration of the
fermenter broth to protect heat exchangers and distillation columns against fouling, or novel distillation
concepts.
Whereas the benets of introducing these techniques are recognized, extensive research is still needed
to scientically and economically justify their application. The paper nally presents a tentative economic assessment, with production costs not only depending on the extent of applying process improvements, but also on the raw material used in the process.
2014 Elsevier Ltd. All rights reserved.

Keywords:
Bio-ethanol
Characteristics
Environmental aspects
Raw materials
Energy saving
Hybrid process
Anhydrous ethanol
Aspen Plus V8.2

Contents
1.

2.

3.

Bio-ethanol: characteristics and worldwide potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61


1.1.
Ethanol and its characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
1.2.
Worldwide production and research importance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
1.2.1.
The recognized potential of bio-ethanol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
1.2.2.
The different generations of bio-ethanol production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
The uses of bio-ethanol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
2.1.
Generalities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
2.2.
The uses of bio-ethanol as fuel and feedstock in chemicals' synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
Bio-ethanol production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
3.1.
Major raw materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
3.2.
Main steps in biomass-to-ethanol processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
3.3.
Bioethanol from disaccharides- and starch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
3.4.
Lignocellulosic biomass-to-ethanol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
3.5.
Fermentation of hexoses and pentoses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

* Corresponding author. Beijing University of Chemical Technology, College of Life Science and Technology, Key Lab of Bioprocess, Beijing, China. Tel.: 86 1015210645421;
fax: 86 1064794689.
E-mail address: kangqian1985@gmail.com (Q. Kang).
http://dx.doi.org/10.1016/j.pecs.2014.10.003
0360-1285/ 2014 Elsevier Ltd. All rights reserved.

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

4.

5.

6.
7.

61

3.6.
The problem of ethanol-inhibition in the first generation feedstock fermentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
Traditional processing routes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
4.1.
Integrated saccharification/fermentation processes versus two-step processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
4.2.
Basic data and energy requirement of the process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.2.1.
Fermentation broth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.2.2.
Simulation in Aspen Plus of the basic concept (no internal energy recycle) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
Process improvements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
5.1.
Energy integration within the current production processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
5.2.
The use of VHG fermentation: principles and application results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
5.2.1.
Introduction of very high gravity (VHG) fermentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
5.2.2.
Effect of implementing VHG on the distillation thermal requirements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
5.3.
The development of hybrid (pervaporation) systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
5.3.1.
Introduction of hybrid operations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
5.3.2.
Effect of implementing pervaporation on the distillation thermal requirements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
5.4.
Overall assessment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
5.5.
The process of final ethanol dewatering to fuel grade . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
5.5.1.
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
5.5.2.
Major processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
5.5.3.
Production of anhydrous ethanol using hydrophilic membranes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
5.6.
Additional developments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
5.6.1.
Cross-flow microfiltration of bio-ethanol fermentation broth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
5.6.2.
Novel distillation concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
5.6.3.
The improved bio-ethanol production plant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
Preliminary economic assessment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

1. Bio-ethanol: characteristics and worldwide potential


1.1. Ethanol and its characteristics
Ethanol or ethyl alcohol (CH3CH2OH) is a colorless, volatile and
ammable liquid, with molecular weight of 46.07 g and density of
789 kg/m3 at 294 K. Thermal properties are given in Table 1.
It burns with a smokeless blue ame, generally invisible in
normal light.
The auto-ignition temperature is the lowest temperature at
which ethanol will spontaneously ignite in a normal atmosphere
without an external source of ignition, such as a ame or spark.
Mixtures of water and ethanol are important throughout the
bio-ethanol process. The knowledge of mixture ash points, as
presented in Fig. 1, is needed for their safe handling, storage and
transportation: the ash point is one of the most important physical properties used to determine the potential for re and explosion hazards of liquids, used for the classication and labeling of
dangerous substances and preparations. The ash point of a given
liquid is the experimentally determined temperature adjusted to
standard temperature and pressure at which a substance emits
sufcient vapor to form a combustible mixture with air. A lower
ash point value indicates that a given liquid is more hazardous
relative to a different liquid with a higher value.
The physical properties of ethanol result from the presence of
both the hydroxyl group and the shortness of the carbon chain. The
hydroxyl group is prone to hydrogen bonding, making ethanol

more viscous and less polar than organic compounds of similar


molecular weight. Ethanol is moreover miscible with water
(unlike > C3 alcohol) and with many organic solvents, e.g. acetic
acid, acetone, ether, ethylene glycol, glycerol, and toluene [1,2]. It is
also miscible with light aliphatic liquids, such as C5H12 and C6H14,
and with chlorinated aliphatics such as CH3CCl3 and Cl2CHeCHCl2
[2]. Mixing ethanol and water is slightly exothermic, releasing
~0.78 kJ/mol [3] at 298 K. Mixtures of ethanol and water at atmospheric pressure form an azeotrope of ~89 mol% ethanol and
~11 mol% water [4] at a temperature of 351 K. This azeotropic
behavior is a pronounced function of temperature and pressure and
vanishes at temperatures below 303 K or pressures below about
10 mbar [5].

Table 1
Primary properties of ethanol.
Boiling point
Flash point
Auto-ignition temperature
Heat of combustion

351.37 K
289.6 K
698 K
26,800 kJ/kg

Fig. 1. Flash points of ethanolewater mixtures.

62

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

Fig. 2. Schematics of the processes for bio-ethanol production (a) As currently applied; (b) With potential membrane applications (DDGS: Distillers Dried Grains with Solubles;
123: cellulosic materials; 23: starch or carbohydrate rich materials).

The ash point of pure ethanol is 289.6 K only, lower than the
ambient temperature [6]. The ash points of ethanolewater solutions are shown in Fig. 1 [7].
Ethanol for industrial synthesis reactions or as a solvent is
mainly produced petrochemically through the acid-catalyzed hydration of ethylene [8]:

C2 H4 H2 O/CH3 CH2 OH
The catalyst is most commonly phosphoric acid [9,10], absorbed
onto a porous carrier such as silica gel or diatomaceous earth.
Ethanol for use in alcoholic beverages, and the vast majority of
ethanol for use as bio-fuel [9,10], is produced by fermentation.
Certain species of yeast (e.g., Saccharomyces cerevisiae) or bacteria
(e.g. Zymomonas mobilis) metabolize sugars in oxygen-lean conditions and produce ethanol and carbon dioxide. The chemical reactions for glucose/fructose and for sucrose as respective raw
materials are given below:

C6 H12 O6 /2CH3 CH2 OH 2CO2 ; for glucose and fructose

(1.1)

C12 H22 O11 H2 O/4CH3 CH2 OH 4CO2

(1.2)

; for sucrose

Fig. 3. Literature (2000e2013) concerning bio-ethanol SCOPUS with keywords (


Separation;
Membrane Technology; Very High Gravity (VHG)).

The general ow sheet of the fermentative bio-ethanol production, including the starting fermentation and successive process
steps, is illustrated in Fig. 2.
The function of the different process steps, and the difference
between 1st, 2nd and 3rd generation raw materials will be discussed in Section 1.2. The possible introduction of membrane
separations is also indicated in the ow sheet, and will be detailed
in Section 5 of this paper.
The liquid waste of the fermenter (stream 8) and the stillage
recovered from the distillation (stream 5) of Fig. 2 are valuable
sources for anaerobic treatment. These streams contain moderate
concentrations of glucose and starch, respectively 1e20 g/L and up
to 200 g/L [11]. The waste water production exceeds 8 m3/ton bioethanol, providing sufcient substrate for an efcient anaerobic
digestion. The biogas produced (~670 Nm3/ton bio-ethanol) has a
sufcient energy content to make a CHP (Combined Heat and Power) application viable, generating sufcient electricity to partly
power the whole bio-ethanol plant, whilst also providing hot water
(~353 K) to be used in the boiler.
The CO2 production of fermentation nearly equals the bioethanol yield. The CO2 produced is recovered, cleaned, and

Bio-ethanol General;

Fuel-Application;

Environment and Economics;

Simulation and

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

compressed/liqueed. A review paper [12] assessed the equipment


and operating costs needed to capture and liquefy CO2 for truck
delivery from an ethanol plant. Estimated costs are provided for
food/beverage grade CO2 and also for less puried CO2 suitable for
enhanced oil recovery or sequestration. The paper includes preliminary plant and equipment designs and estimates major capital
and operating costs for each of the recovery options.
1.2. Worldwide production and research importance
1.2.1. The recognized potential of bio-ethanol
With the globalization of the increasing demand for energy,
energy shortage is a common worldwide problem. It is predicted
that the growth in the production of easily accessible oil and gas
will not match the projected rate of demand by 2040e2050, since
the level of oil demand will increase from 85 Mb$day1 in 2008 to
105 Mb$day1 in 2030 and to higher values thereafter [13], whilst
the number of vehicles will increase to 1.3 billion by 2030 and to 2
billion by 2050 [14]. In addition, concerns towards CO2 emissions

63

and associated climate change [15] have instigated an accelerated


research and production of renewable energy resources. Bioethanol is considered as the alternative renewable fuel with the
largest potential to replace fossil-derived fuels, with a world production of 50 million m3 in 2007 and in excess of 100 million m3 in
2012 [16], and with a potential for a signicant reduction of
greenhouse gas emissions [17].
Brazil and the United States represent approximately 80% of the
world supply [16]. More than 20% of Brazilian cars are able to use
100% ethanol as fuel, including ethanol-only engines and exefuel
engines. Flexefuel engines in Brazil are able to use pure ethanol, all
gasoline, or any mixtures of both. In the USA, exefuel vehicles can
run on 0%e85% ethanol (15% gasoline) with higher ethanol blends
not yet allowed or deemed less efcient. Brazil produces ethanol
from domestically grown sugar cane. Sugar cane not only has a
higher concentration of sucrose than corn (by about 30%), but is
also much easier to be extracted. The bagasse generated by the
process is not wasted, but is used to produce process steam and
electricity. The US feed stock is mostly corn and wheat [16].

Table 2
Literature (>2010) about rst generation bio-ethanol in general, and cassava in particular.
Reference

Objectives

Main results

[19]

Production of ethanol from cassava pulp via fermentation entation

[20]

Two soil derived yeasts for bio-ethanol production of Cassava starch

[21]
[22]
[18,23e25]

Process optimization for bio-ethanol production from cassava


starch
Non-thermal enzymatic saccharication of cassava pulp
Bio-ethanol in China

- Produced ethanol at 91% and 80% of theoretical yield from 5% to 10% cassava pulp
(K7G strain fermented starch and glucose).
- S. cerevisiae CHY1011 and CHFY0901 have potential use in industrial bio-ethanol
fermentation processes.
- The reaction was completed within 48.5 h at 303 1 K.

[26]

Ethanol fermentation of energy beets

[27]

Evaluation of whole Jerusalem artichoke for consolidated


bioprocessing ethanol production (CBP)

[28]

Ultrasound and technical enzymes during bio-ethanol production


from fresh cassava root

[29]

Bio-ethanol production from raw juice as intermediate of sugar beet


processing
Improvement of bio-ethanol production from corn by ultrasound
and microwave pretreatments

[30]

[31]
[32]
[33]
[34]

Production of bioethanol by immobilized Saccharomyces cerevisiae


onto modied sodium alginate gel
Real-time monitoring of fermentation process applied to sugarcane
bio-ethanol production
Cassava waste for bio-ethanol for application with dual-fuel
absorption refrigeration in Africa
Saccharication and fermentation of waste sweet potato for bioethanol production

[35]

Utilization of microwave and ultrasound pretreatments in the


production of bio-ethanol from corn

[36]

Optimal strain design of Saccharomyces cerevisiae for bio-ethanol


production

[37]

Bio-ethanol production from intermediate products of sugar beet


processing
Duckweed starch accumulation for bio-ethanol production

[38]

- More energy efcient, and more fermentable sugar yields.


- Development of non-grain ethanol;
- Cassava fermentation achieves ethanol concentration of 120 g L1;
- Cassava will be China's largest non-grain ethanol base over the next ve years.
- Ethanol yields: 0.47 g/g (sugar) and volumetric productivity: 7.81 g L-1h1
(simultaneous saccharication and fermentation (SSF));
- A suitable advanced bioenergy crop for producing industrial fermentation sugars
in the Mid-South of the U.S.
- Suggested a cost-effective CBP strategy;
- Produced 45.3 g L1 ethanol (30 h); -Productivity of 1.51 gL1h1, ethanol yield
0.32 g ethanol/g fermentable sugars (60% fermentable sugar conversion efciency).
- Increased the free sugar from 0.5% in raw material to 8%;
- Up to 100% of starch could be converted in glucose, and up to 11 vol% ethanol could
be produced.
- Optimal fermentation time: 30 h; sugar mass fraction in batch fermentation:
12.30 wt%.
- Increased the maximum ethanol concentration produced in the SSF process by
11.15% and 13.40% (ultrasound and microwave);
- Maximum ethanol concentration of 9.87 wt% and ethanol yield of 90.80% were
achieved (microwave).
- The maximum concentration, productivity and yield of ethanol were 69.68 g L1,
8.71 g L1 h1 and 0.697 g g1.
- Presented an automated, low-cost, real-time response and minimally invasive
solution for the optimization technology of bio-ethanol production.
- Proposed a simple and scalable two-step process that biologically converts the
waste streams into valuable products.
- 91.5% of the starch and sucrose were converted to glucose and fructose;
- Converted more than 90% of the total sugars into ethanol with 87.2% of the
theoretical yield and 49.07 g/L nal concentration of ethanol (Z. mobilis 8b).
- Increased the glucose concentration obtained after liquefaction by 6.82 and 8.48%
(ultrasonic and microwave);
- Increased the maximum ethanol concentration by 11.15 and 13.40% (ultrasonic
and microwave);
- A maximum ethanol concentration of 9.91% (w/w) and percentage of theoretical
ethanol yield of 92.27% were achieved.
- Conrmed the effectiveness and of a 20e30% improvement in ethanol production
during glucose fermentation;
- Three sequential gene knockout targets have been identied for eliminating over
99% of the inefcient pathways for ethanol production through the network.
- Attained efciency of sugar utilization was at least from 98 to 99 wt%;
- Maximum productivity was achieved at ~1.8 g L1 h1 for all applied yeast strains.
- The change of temperature difference between day and night affects the duckweed
starch content.

64

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

Table 3
Literature (>2010) about second generation bio-ethanol production.
Reference

Objectives

Main results

[40]

Optimal industrial symbiosis system to improve bioethanol production

[41]

Bio-ethanol production from dilute acid pretreated Indian bamboo


variety by separate hydrolysis and fermentation
Fuel ethanol production from sweet sorghum bagasse using microwave
irradiation
Ultrasonic-assisted simultaneous SSF of pretreated oil palm fronds for
bio-ethanol production
Convert sucrose and homocelluloses in sweet sorghum stalks into
ethanol
An integrated optimization model for switchgrass-based bio-ethanol
supply chain
Low-intensity pulsed ultrasound to increase bio-ethanol production

[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]

Different process congurations for bio-ethanol production from


pretreated olive pruning biomass
Bio-ethanol production from water hyacinth Eichhornia crassipes
Enhanced saccharication of biologically pretreated wheat straw for
ethanol production
Ethanol production, purication, and analysis techniques

[51]

Fermentation of biologically pretreated wheat straw for ethanol


production

[52]

Integration of pulp and paper technology with bio-ethanol production

[53]

Lignin and silica from the black liquor generated during the production
of bio-ethanol from rice straw

[54]

Pequi cake hydrolysis and fermentation to bio-ethanol

[55]

Pretreatment technologies for an efcient bio-ethanol production


process based on enzymatic hydrolysis

[56]

Production of bio-ethanol by fermentation of lemon peel wastes


pretreated with steam explosion

[57]

Sono-assisted enzymatic saccharication of sugarcane bagasse for bioethanol production

[58]

Status and barriers of advanced biofuel technologies

[59]

Sugarcane bagasse hydrolysis using yeast cellulolytic enzymes

[60]

Switchgrass for bioethanol

[61]

Tween 40 pretreatment of unwashed water-insoluble solids of reed


straw and corn stover pretreated with liquid hot water to obtain high
concentrations of bioethanol

[62]

Enzymatic saccharication of algal biomass for bio-ethanol production


(Chlorella variabilis)
Waste paper sludge as a potential biomass for bio-ethanol production

[63]

[64]

Assessment of combinations between pretreatment and conversion


congurations for bio-ethanol production

Reduce bioethanol production and logistic costs;


2nd generation biomass should be used for bioethanol production.
Bio-ethanol yield of 1.76 vol% with an efciency of 41.69%;
Bamboo can be used as feed stock for the production of bio-ethanol.
An ethanol yield based on total sugar of 480 g kg1was obtained;
Ethanol produced on marginal land at 0.252 m3 ton1 biomass.
Maximal bioethanol concentration (18.2 g/L) and yield (57.0%).

- All sugars in sweet sorghum stalk lignocellulose were hydrolysed into


fermentable sugars.
- Optimal to harvest switchgrass as loose chop; intended bioreneries with total
capacity of 2270 ML yr1 of bio-ethanol in North Dakota.
- Increase the production of bio-ethanol from lignocellulosic biomass to
52 16%.
- Ethanol concentration of 3.7 vol% was obtained.
- Yeast Saccharomyces cerevisiae TY2 produced ethanol at 9.6 1.1 g/L.
- Increase of the sugar yield from 33 to 54%, and reduction of the quantity of
enzymatic mixture by 40%.
- Utilization of lignocellulosic biomass for ethanol production is being studied
more intentionally;
- Ozonation degrades impurities, activated carbon removes impurities, and gas
stripping removes high volatile compounds without any heating.
- The highest overall ethanol yield was obtained with the yeast Pachysolen
tannophilus: yielded 163 mg ethanol per gram of raw wheat straw (23 and 35%
greater).
- Re-use existing assets to the maximum extent;
- Keep the process as simple as possible;
- Match the recalcitrance of the biomass with the severity of the pretreatment.
- Sulphuric acid is the best reagent;
- Economically more suitable; the two-step treatment is more efcient in
providing a product superior in qualities and an efuent safer to dispose off.
- 61.6% conversion of its starch content to reducing sugars;
- Produces 53 L of ethanol per ton of hydrolyzed pequi meal.
- Chemical and thermochemical are currently the most effective technologies
for industrial applications;
- Efcient conversion and utilization of hemicellulosic sugars has become an
important task to increase sugar yields.
- Reduces the residual content of essential oils below 0.025% and decreases the
hydrolytic enzyme requirements;
- Obtained ethanol production in excess of 60 L/1000 kg fresh lemon peel
biomass.
- The maximum glucose yield obtained was 91.28% of the theoretical yield and
the maximum amount of glucose obtained was 38.4 g/L (MTCC 7450);
- The hydrolyzate obtained was 91.22% of the theoretical ethanol yield (MTCC
89);
- Decreases the reaction time;
- The application of low intensity ultrasound enhanced the enzyme release and
intensied the enzyme-catalyzed reaction.
- The major barriers for the commercialization of 2nd generation ethanol
production are the high costs of pretreatment, enzymes used in hydrolysis, and
conversion of C5 sugars to ethanol;
- The residues need to be processed for byproducts through biorenery to
improve the economics of the whole process.
- This enzyme extract promoted the conversion of approximately 32% of the
cellulose;
- C. laurentii is a good b-glucosidase producer.
- Glucose yields: 70%e90%, and xylose yields: 70%e100% after hydrolysis;
- Ethanol yields range from 72% to 92% of the theoretical maximum.
- Obtain a high ethanol concentration of 56.28 g/L (reed straw) and 52.26 g/L
(corn stover);
- Ethanol yield reached a maximum of 69.1% (reed straw) and 71.1% (corn
stover).
- The enzyme addition resulted in the highest ethanol production by
fermentation.
- SSF using cellulase produced by A. cellulolyticus gave ethanol yield 0.208 (g
ethanol/g PS organic material);
- Consolidated biomass processing (CBP) technology gave ethanol yield 0.19 (g
ethanol/g Solka oc).
- The process based on dilute acid pretreatment and enzymatic hydrolysis and
co-fermentation combination shows the best economic potential;
- The cellulose hydrolysis based on an enzymatic process showed the best
energy efciency.

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

65

Table 3 (continued )
Reference

Objectives

Main results

[65]

Combined use of gamma ray and dilute acid for bio-ethanol production

[66]

Bio-ethanol production from Lantana camara

[47]

Different process congurations for bio-ethanol production from


pretreated olive pruning biomass
Ethanol production from lignocellulosic biomass (exergy analysis)

- Increasing enzymatic hydrolysis after combined pretreatment is resulting from


or decrease in crystallinity of cellulose, loss of hemicelluloses, and removal or
modication of lignin.
- 87.2% lignin removal, and 80.0% saccharication;
- 17.7 g/L of ethanol with corresponding yields of 0.48 g/g (Saccharomyces
cerevisiae).
- High ethanol concentration of 3.7 vol% was obtained.

[67]

In developing economies, food-related feed stock is preferably


replaced by non-food raw materials, such as sweet sorghum or
cassava. Sweet sorghum can be used in ethanol fermentation, and is
suitable for growing in dry conditions: the water requirement of
sweet sorghum is one-third that of sugarcane on a comparable time
scale, and it requires about 22% less water than corn does. The
world's rst sweet sorghum-based ethanol production distillery
began commercial ethanol production in 2007 in Andhra Pradesh,
India [14].
Sorghum is now also an important feedstock for bio-ethanol
production in the USA Plains states. In some parts of Europe,
particularly France and Italy, grapes have become a feedstock for
fuel ethanol from wine surplus. In Japan, it has been proposed to
use rice, normally converted into sake. Ethanol can be efciently
produced from starches, with cassava starchy crop having the
highest ethanol yield. Thailand already had a large cassava industry
in the 1990s, for use as cattle feed and as a cheap admixture to
wheat our. Nigeria and Ghana are establishing cassava-to-ethanol
plants. The world's rst large-scale cassava ethanol plant was built
in Guangxi (China) by COFCO in 2007, with an annual production
capacity of 200,000 tons [18]. China operates several bio-ethanol
production plants for a total annual production capacity of
~2,085,000 ton/year, using mostly corn and wheat (~65%) and
cassava (~35%) as feedstock.
The main reasons for the enhanced development of bio-ethanol
are related to its specic and advantageous properties, as discussed
in detail in Section 2 of the paper. Its use as a favorable and near
carbon-neutral renewable fuel, thus reducing CO2 emissions and
associated climate change; its use as octane enhancer in unleaded
gasoline; and its use as oxygenated fuel-mix for a cleaner combustion of petrol, hence reducing tailpipe pollutant emissions and
improving the ambient air quality, are the major driving forces
behind the development. The importance and potential of bioethanol are highlighted in the literature, with recent exponential
research developments illustrated in Fig. 3.
1.2.2. The different generations of bio-ethanol production
Fermentation of sugar-based raw materials is referred to as rst
generation bio-ethanol, whereas the use of lignocellulosic raw
materials is commonly called second generation bio-ethanol. The
third generation of algal bio-ethanol is at its early stages of
investigation. Recent references about rst and second generation
bio-ethanol are summarized in Tables 2 and 3, respectively.
In view of the marked research development since about 2008
(in general) and since 2010 (specic novel technologies), we have
limited the literature survey to papers of the post-2010 period,
which mostly already assess previous work, whilst discussing stateof-science development.
Fermentation of sugars is referred to as the rst generation bioethanol process, where yeast is cultured under favorable thermal
conditions to convert sugars into ethanol at around 308e313 K. The

- Lowest environmental impact for second-generation bio-ethanol production;


- Highest exergy efciency (Steam Explosion Pretreatment SSF Dehydration) reaching 79.58%.

toxicity of ethanol to yeast inhibits the conversion and limits the


ethanol concentration obtainable, as will be discussed in Section 3.
Although some expensive ethanol-tolerant strains of yeast are
known, they are not applied in producing fuel-grade ethanol due to
their cost, but are applied for wine and spirit production, e.g. Red Star
Pasteur Champagne and Lalvin EC-1118 wine yeasts, with inhibition
above 18 vol% ethanol only, or Turbo Yeast for concentrations of
20 vol% or higher. To produce ethanol from starchy materials such as
cereal grains, the starch must rst be converted into sugars. In
brewing beer, this is generally accomplished by allowing the grain to
germinate, or malt, which produces the enzyme amylase. When the
malted grain is mashed, the amylase converts the starches into
sugars. For fuel ethanol, the hydrolysis of starch into glucose can be
accomplished more rapidly by treatment with dilute sulfuric acid
[39], fungally produced amylase, or a combination of both. In bioprocesses, amylase is the preferred additive [24,25].
As illustrated in Table 3, the second generation of bio-ethanol processes uses cellulose-released sugars [68], although the cost of the
enzymes capable of hydrolyzing cellulose has been a limiting factor.
The Canadian company Iogen started the rst cellulose-based
ethanol plant in 2004 [69]. Development of this technology could
deal with a number of cellulose-containing agricultural byproducts,
such as e.g. straw, wood trimmings, sawdust, bamboo and others.
Lignocellulosic materials typically contain lignin, cellulose and
hemi-cellulose. Hydrolysis of hemicellulose yields mostly vecarbon sugars such as xylose. S. cerevisiae, the yeast commonly
used for ethanol production, cannot metabolize xylose. Other
yeasts and bacteria are under investigation to ferment xylose and
other pentoses into ethanol, and genetically engineered fungi that
produce large volumes of cellulase, xylanase, and hemicellulase
enzymes, are under investigation. These could convert agricultural
residues (e.g. corn stover, straw, sugar cane bagasse) and energy
crops (e.g. switchgrass) into fermentable sugars [67].
The third generation of bio-ethanol processes is at early stages of
investigation.
Recent references are summarized in Table 4.
Algae contain lipids, proteins, carbohydrates/polysaccharides and
have thin cellulosic walls. Whereas algal lipids are mostly extracted
and transformed into biodiesel, the left-over cake of starch (the
storage component) and cellulose (the thin wall component) can be
converted into bio-ethanol. Especially algae strains Glacilaria- and
Euglena gracilis appear promising candidates. The resulting combined biodiesel/bio-ethanol process offers moreover an added
advantage of producing fermentation CO2 that can be captured and
used. Despite these advantages, the alternative production of
numerous valuable products from the algal cake (e.g. carrageenan,
agar) offers signicant economic benets in comparison with lowpriced bio-ethanol. Additional research is certainly required, and
the process cannot be considered for industrial application yet.
Whether rst, second or third generation feed stock is used,
fermentation produces an alcohol-lean broth only, as such unusable

66

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

in most industrial and fuel applications. The ethanol must hence be


puried. Fractional distillation can concentrate ethanol to 95.6 vol%
(89.5 mol%), the azeotrope with a boiling point of 351.2 K. Further
ethanol enrichment by common distillation is impossible, but
different alternatives are possible, as discussed in detail in Section
5.5 of the present paper.
The most commonly applied methods to produce anhydrous,
fuel-grade ethanol, will be assessed in Section 5.4 of this paper.
2. The uses of bio-ethanol
2.1. Generalities
The largest single use of ethanol is as an engine fuel and fuel
additive. More than any other major country, Brazil relies on
ethanol as an engine fuel. Gasoline sold in Brazil contains at least
25% anhydrous ethanol. Hydrous ethanol (about 95% ethanol and
5% water) can be used as fuel in more than 90% of new cars sold in
the country. The US uses Gasohol (max 10% ethanol) and E85 (85%
ethanol) ethanol/gasoline mixtures. Ethanol may also be utilized as
a rocket fuel and in lightweight rocket-powered racing aircrafts
[85]. The Australian law limits the use of pure ethanol sourced from
sugarcane waste to 10% in automobiles. It has been recommended
that older cars (and vintage cars designed to use a slower burning
fuel), should have their valves upgraded or replaced [86].

The common types of available fuel-mixes are listed in Table 5.


The energy content and RON (Research Octane Number), of
ethanol, other fuels and mixes are also summarized in Table 5. The
octane rating of petrol is measured in a test engine and dened by
comparing the mixture of iso-octane and heptane that would have
the same anti-knocking capacity as the tested fuel itself. Fuel with
the same knocking characteristics as a mixture of 95% iso-octane
and 5% heptane would have an octane rating of 95. Of course, the
tested fuel does not contain just iso-octane and heptane in the
given proportions, but will have the same detonation resistance
characteristics. Since some fuels perform better than pure isooctane, octane numbers in excess of 100 are now accepted in a
modied denition. The most common rating is the RON (Research
Octane Number), determined in a test engine with a variable
compression ratio under controlled conditions (600 rpm). The
Motor Octane Number (MON) is also used, and often considered a
better measure of fuel behavior under load, as determined in a
900 rpm test engine. MON will normally be about 8e10 points
below RON. The low density of gasoline and ethanol are reected in
the higher mass than volume heat content of the fuel.
Ethanol has appropriate properties for spark ignition IC engines.
Its MON and RON are respectively 90 and 109, on average 99
compared to 91 for regular gasoline. The lower heating value (LHV)
of ethanol is two-thirds that of gasoline. Bio-ethanol fuel is used in
IC engines as 5e26 percent anhydrous ethanol blends to gasoline

Table 4
Literature (>2010) about algal bio-ethanol.
Reference Objectives

Main results

[70]

Glycerol as a feedstock in the production of biofuels

[71]
[72]

Bio-ethanol production from Gracilaria verrucosa (red alga)


Liquid biofuel production from algal feedstocks

[73]

Microalgae for third generation biofuel production

[74]

Bio-ethanol production of microalgae from coastal waters


of Pearl River Delta

[75]

Microalgal biofuels for sustainable development

[76]

Algal biomass conversion to bio-ethanol

[77]
[78]

Microalgal biomass as a fermentation feedstock for bio-ethanol


production
Analysis of microalgae biomass (Hawaii) for bio-ethanol production

[79]

Ethanol production potential of Thai seaweed species

[80]

Bio-ethanol from microalgal biomass (Marine Yeast)

[81]

Lignocellulosic biomass to produce ethanol as biofuel

[82]

Effect of glycerol and glucose on the enhancement of biomass

[83]

Two-stage hydrolysis of invasive algal feedstock for ethanol


fermentation

[84]

Simultaneous utilization of galactose and glucose for ethanol


production

- Glycerine is available in large volumes from the production of bio-ethanol;


- Glycerine-to-ethanol could rapidly change the market situation.
- The leftover pulp of enzymatic hydrolysis yielded 0.87 g sugars/g cellulose.
- Fermentation of microalgae to bio-ethanol is the most effective method of
producing biofuel;
- Cyanobacteria have been successfully engineered to produce ethanol.
- Microalgal feedstock reduces the cultivation area (per ha) compared to the
feedstock production of rst and second generation biofuel.
- Highest release of total sugars of 10.752 g L1 hydrolysate and glucose 5.730 g L1
hydrolysate (S. abundans PKUAC 12);
- S. abundans PKUAC 12 biomass was the best feedstock for fermentative bio-ethanol
production yielding 0.103 g of ethanol per gram of dry weight algae.
- The integrated algal biofuels production can potentially result in the improvement
of the economics of algal biorenery approach.
- Two different established fermentation routes, single-stage fermentation and twostage gasication/fermentation processes are discussed.
- The lipid extracted microalgae displayed a signicantly ethanol concentration of
3.83 g L1 (~38 wt%) compared with dried/intact microalgae.
- 49% of the total mass observed as sugars in the hydrolysate;
- The measured ethanol yield corresponded to approximately 90L per ton of dried
microalgae.
- Total sugar and yield of ethanol were found to be highest in G. tenuistipitata
(98.21%, 4.17 g ethanol/g sugar and 139.12 mg ethanol/g glucose, respectively).
- The hydrolysate at varying sugar from 2.7% to 5.5%, yielded 1.23e1.76 wt% ethanol;
- Utilization of marine yeast for converting algal sugar to ethanol would eliminate an
energy intensive step of electrodialysis.
- Acid pretreatment had a major inuence on the sugar release though enzymatic
hydrolysis (3 vol% of acid was found optimum);
- Alkaline pretreatment with 3% NaOH solutions was found optimum;
- Detoxication is needed to avoid the formation of inhibitors.
- Chlorella vulgaris can utilize glycerol as a sole carbon substrate;
- The effect of glycerol and glucose could enhance the Chlorella vulgaris growth,
biomass content and volumetric producitivity, and overcome the lower biomass
yield.
- Two-stage hydrolysis of the homogenates produced 13.8 g glucose/kg fresh algal;
- Batch fermentation analysis produced 79.1 g ethanol/kg dried invasive algae (using
Escherichia coli KO11).
- At 120 g/L of initial galactose concentration, the ethanol concentration reached
6.9 vol% with 88.3% of theoretical ethanol yield (0.51 g ethanol/g galactose);
- The ethanol fermentation efciency of the novel Saccharomyces cerevisiae strain
using the galactose base of red algae was superior.

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88


Table 5
Energy content and Octane Number of some fuels compared with ethanol.
Fuel type [87]
Dry wood (20% moisture)
Methanol
Ethanol
E85 (85% ethanol, 15% gasoline)
Liqueed natural gas
Autogas (LPG) (60% propane 40% butane)
Aviation gasoline (high-octane
gasoline, not jet fuel)
Gasohol (90% gasoline 10% ethanol)
Regular gasoline/petrol
Premium gasoline/petrol
Diesel

MJ/L

MJ/kg

17.9
21.2
25.2
25.3
26.8
33.5

~19.5
19.9
26.8
33.2
~55
50
46.8

33.7
34.8

47.1
44.4

38.6

45.4

RON
108.7
108.6
105

100/130
(lean/rich)
93/94
Min. 91
Max. 104
25

(5% maximum in Europe and in India, 10% in USA, 22e26%


mandatory blends in Brazil) or as pure fuel (100%) of dry ethanol in
dedicated vehicles, as illustrated in Table 6.
When anhydrous bio-ethanol is blended with gasoline in a small
proportion (up to 15%), the inuence of its LHV has no signicant
effect. For higher blend levels, more fuel is required compared to
when using conventional gasoline due to the lower LHV of ethanol.
Ethanol dedicated vehicles are adapted to improve the engine by
running at higher compression ratios to take advantage of the
better octane number of ethanol compared to gasoline: despite the
lower HV, such optimized vehicles running on pure ethanol still
have an operation range 75% of the range of gasoline fueled cars
on a volume basis. Finally, special Flexible Fuel Vehicles (FFV) are
equipped with line sensors, which measure ethanol levels and
adapt the airefuel ratio to maintain good combustion conditions.
These vehicles can burn fuel containing ethanol (0e85%) in gasoline and are becoming more and more frequent, with various
manufactures developing such vehicles on a commercial scale with
special modication of carburetor, fuel injector, fuel pump, fuel
tank and catalytic converter when using E10 to E25 blends, whist
additional modications of the engine itself, the intake manifold,
exhaust system and cold-start are incorporated for E25 to E100
(hydrous) blend. Apart from the advantages of using bio-ethanol in
IC engines, a few disadvantages should be considered: whereas low
levels of ethanol blended with gasoline increase the vapor pressure,
thus favoring evaporative emissions that contribute to smog formation, higher ethanol blend levels have a low vapor pressure that
can lead to difculties in cold weather conditions. Straight hydrous
ethanol as an automotive fuel has been widely used in Brazil for
neat ethanol vehicles and more recently for exible-fuel vehicles:
the ethanol fuel is distilled close to the azeotrope mixture of
95.63 wt% ethanol and 4.73 wt% water. At the end of 2012, there
were ~20 million exible-fuel vehicles running on Brazilian roads.
Hydrous ethanol imposes a limitation on normal vehicle operation,

Table 6
Common ethanolepetrol mixtures.
Code

Composition

Countries

Comments

E5

Max. 5% anhydrous ethanol,


min. 95% petrol
Max. 10% anhydrous ethanol,
min. 90% petrol
Max. 15% anhydrous ethanol,
min. 85% petrol
Max. 25% anhydrous ethanol,
min. 75% petrol
Max. 85% anhydrous ethanol,
min. 15% petrol
Hydrous ethanol (~5.3 wt% water)

Western Europe,
India
USA, Europe, China,
India, South Africa
USA, cars >2000,
South Africa
Brazil

Blends for
regular cars

USA, Europe

Flexefuel
vehicles

E10
E15
E25
E85
E100

Brazil

67

as the ethanol lower evaporative pressure (as compared to gasoline) causes problems when cold starting the engine at temperature
below 288 K. For this reason, both pure ethanol and exefuel vehicles are built with an additional small gasoline reservoir inside
the engine compartment to help in starting the engine when cold
by initially injecting gasoline. Once started, the engine is then
switched back to ethanol. An improved exefuel engine generation
was developed to eliminate the need for a gasoline tank by
warming the ethanol during starting and allowing them to start at
temperatures as low as 268 K. The use of ethanol blends in conventional gasoline vehicles is restricted to low mixtures, as ethanol
is corrosive and can degrade some of the materials in the engine
and fuel system. The hygroscopicity of ethanol should moreover be
considered in colder and humid climates, due to the possible
plugging of fuel lines by ice crystals, and due to the lower ashpoints at higher water contents.
Due to its low cetane number, ethanol does not burn efciently
by compression ignition. Moreover, ethanol is not easily miscible
with diesel fuel. To improve the use of ethanol in CI vehicles,
measures can be taken, such as the addition of an emulsier in
order to increase the ethanolediesel miscibility; or the addition of
ethylhexylnitrate or diterbutyl peroxide in order to enhance the
cetane number. Most blends of ethanol and diesel (E-diesel) are
limited to 15% ethanol and 5% emulsiers [88].
Alternative solutions are either a dual fuel operation in which
ethanol and diesel are introduced separately into the cylinder, or
the modication of diesel engines in order to adapt their characteristics of auto-ignition and make them capable to use high blends,
up to 95% ethanol [89].
2.2. The uses of bio-ethanol as fuel and feedstock in chemicals'
synthesis
The uses of bio-ethanol as a fuel or as raw material for the
chemical synthesis of other products have been dealt with in
numerous publications, some of the relevant and recent ones
(>2010) being summarized in Table 7.
When specically discussed in the respective publications, the
environmental effects of using bio-ethanol as fuel are also included
in the summaries. From the literature survey, it appears that the
main reasons of the success of bio-ethanol are the following: (i) its
use as renewable energy to partially substitute oil and increase
security of supply; (ii) its use as octane enhancer in unleaded
gasoline to replace metyl-tert-butylether (MTBE); (iii) its use as
oxygenated compound for clean combustion of the gasoline, thus
reducing the tailpipe emissions and improving the ambient air
quality and (iv) its use as renewable fuel to reduce CO2 emissions
and its contribution to a reduced effect on climate change.
In addition to the few publications concerning the conversion of
bio-ethanol to H2, there is a recent revival of its conversion into
various organic chemicals with either a higher number of carbonatoms in their molecules, or/and with an added-value. Fig. 4 summarizes the different options.
Although these conversion pathways offer possibilities to
upgrading bio-ethanol, their investigation is at an early stage of
research.
Specic recent literature about environmental effects has been
listed in Table 8.
The environmental advantages of using ethanol as a fuel are
recognized since it reduces harmful tailpipe emissions of carbon
monoxide, particulate matter, and nitrogen oxides. Argonne National Laboratory analyzed the greenhouse gas emissions of
different engine and fuel combinations. Comparing ethanol blends
with gasoline alone, they showed reductions of 8% with the biodiesel/petrodiesel blend known as B20, 17% with the conventional

68

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

Table 7
Literature (>2010) concerning the applications of bio-ethanol.
Ref.

Objectives

Bio-ethanol as fuel
[90]
Investigation of using blends of diesel and ethanol

[91]
[92]

Bio-ethanol injection in DISI engine for two wheels vehicles (2000


e4000 rpm)
Temperature dependent density of biodiesel-diesel-bioethanol blends

[93]

Bioethanolediesel fueled engine

[94]

Biofuels in HCCI engines

[95]

CHP using blends of gasohol in IC engine

[96]

Automotive spark ignition engine using bioethanol

[97]

Compression ignition engines using biofuels

[98]

China's bioethanol development and national application of E10

[99]

Bio-fuels for the gas turbine

[100]

Bioethanol/REM/diesel blend in Euro5 automotive diesel engine

[101]
[102]
[103]

Ethanol fuel from biomass


Alternative fuels for gas turbines
FT/Biodiesel/Bioethanol surrogate fuel oxidation in jet-stirred reactor

[104]

A comparison of motor fuels, related pollution and technologies

[105]

Increasing the stability of bio-ethanol/gas oil emulsions

[106]

Ethanol-unleaded gasoline blends in a spark-ignition engine

Bio-ethanol as chemical feedstock or for use in fuel cells


[23]
Effect of initial pH on thermophilic fermentative hydrogen production
from cassava ethanol wastewater
[107]

Production of hydrogen from Bio-ethanol

[108]

Vapor-fed PBI-based direct ethanol fuel cell

[109]

Techno-economic and environmental assessment of bioethanol-based


chemical process

[110]
[111]

Butadiene production from bioethanol and acetaldehyde


CoAlZn and NiAlZn mixed oxides in hydrogen production by bio-ethanol
partial oxidation

E85 ethanol blend, and a reduction of emissions by 64% when using


cellulosic ethanol. Despite these positive effects, literature also
draws our attention to the fact that ethanol combustion in an internal combustion engine can increase the emission of products of
incomplete combustion, leading to a signicantly larger photochemical reactivity that generates much more ground level ozone.

Main results
- Engine power increases (up to 8.5%);
- Fuel consumption increases (low caloric value of ethanol);
- Unburned hydrocarbon combustion emissions decrease.
- High injection pressure needed;
- Low tendency to soot formation.
- Modeling density;
- Recommendations on suggested blends.
- Reduced HC and CO emissions;
- Premixed combustion phasing decreases when bioethanol fraction increases.
- Biodiesel: reduction of NOx;
- Bioethanol: engine tolerance to exhaust gas recycle (EGR) was extended.
- E20 recommended;
- With increasing bioethanol, CO deceases, cylinder pressure and temperature
increase.
- The spark ingnition (SI) engine efciency and power performance increase by
using bioethanol;
- CO, HC, and NOx emissions decrease for bioethanol used in supercharged engine.
- Bioethanol: CO2 and smoke decrease;
- Diesel fuel (DF) and Rapeseed Methyl Ester (RME): NOx and CO2 decreases;
- Ethanol fraction in both fuels reaches 50%.
- Linear optimization model used to consider the economic cost of distributing
ethanol;
- Cassava, sweet potato, sweet sorghum and sugar beet are promising feedstocks for
bioethanol expansion.
- Low emissions and substantial fuel exibility obtained in lean, premixed, prevaporized (LPP) combustion with ethanol;
- NOx emission reduction from gas turbines.
- The closed loop combustion control (CLCC) adoption enables the performance
improvement with the blend.
- Lignocellulose-to-ethanol is the most viable pathway (environmental view).
- Brazilian ethanol has the lowest environmental performance.
- The kinetic modeling gave an overall good representation of the experimental
results.
- Ethanol exhaust 2.14 times as much as gasoline exhaust;
- The pollution contributing to smog is 1.7 for ethanol.
- High stability bioethanol/gas oil emulsion produced with 5.0 v/v% biodiesel and
4.0% new chemical structure additive (20  C) in the 5.0 vol% water containing
ethanol.
- Torque with E50 and E85 found to be higher than E0;
- The engine torque, power and fuel consumption increase; CO, NOx and HC
emissions decrease with ethanol;
- Compression ratio (CR) without knock occurrence increases with ethanol
egasoline blends
- Maximum hydrogen product;
- The total amount of VFA/ethanol and the proportion of acetic acid in the VFA
increase with the increase of pH.
- High hydrogen plant efciency and a highly efcient heat integration obtained
with typical ethanol feed.
- An improvement of the cell performance with an increase in temperature
(enhanced kinetics and electrolyte conductivity);
- The E/W weight ratio (0.25e0.5) was found to be suitable;
- Ethanol crossover currents increases with the temperature and the E/W;
- This cell can tolerate the use of the renewable bio-ethanol.
- Bioethanol-based processes have better cost saving and less global warming
potential;
- Bioethanol-based processes have great potential in ethyl acetate production.
- The pore size of SBA-15-100 is optimum for the ETB catalyst.
- Selectivity to H2 and CO on NiAlZn was lower than on CoAlZn (673e773 K), but the
selectivity trend reversed (>873 K);
- ZnAl2O4 spinel was less active in ethanol conversion and much less selective to H2
and CO;
- NiAlZn is a better catalyst.

3. Bio-ethanol production
3.1. Major raw materials
Bio-ethanol can be produced from a large variety of carbohydrates (mono-, di-, polysaccharides). Monosaccharides (e.g. xylose,

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

69

Fig. 4. Conversion pathways of ethanol to different organic chemicals.

glucose, fructose) consist of single sugars, represented by (CH2O)n


with n 3e7. Pentoses (n 5, xylose) and hexoses (n 6, glucose)
are the most common monosaccharides in nature. Whereas glucose
is the most widespread sugar transport form in animal organisms,
sugar is often transported in the form of disaccharides in plants
(sucrose, maltose, and lactose). A disaccharide is formed by dehydration coupling of two monosaccharides. Polysaccharides are
composed of similar subunits (monomers), with e.g. starch and
cellulose composed of monomers of glucose. Polysaccharides must
be hydrolyzed into disaccharides and/or monosaccharides before
ethanol fermentation.
Large-scale carbohydrate-to-ethanol plants mostly use sugarcane or sugar beet juice, corn or wheat. Ethanol is also commercially produced in the pulp and paper industry as a by-product.
These carbohydrates-to-ethanol crops include both multipurpose

crops, also devoted to food markets; and dedicated ethanol crops.


Whereas the latter are cultivated especially for ethanol production
on non-agricultural lands, the former provide almost all of the
feedstock now used to date for ethanol production (sugarcane in
Brazil and corn in the United States), where the development of
biomass-to-ethanol conversion emerged as alternative markets for
sugar, corn or wheat surpluses. In developing countries, the
possible competition with food is shifting from using agricultural
crops towards non-food feedstock, e.g. cassava, sweet sorghum.
Lignocellulosic biomass can be considered as feedstock for bioethanol production in the medium and long term due to its low cost
and high availability. Criteria and methodologies for assessing the
sustainability of energy crops have been developed [122,123]. As
illustrated in the literature survey of Table 3, the ethanol yields
when using lignocellulose biomass are only reported for lab-scale

Table 8
Literature (>2010) about the environmental effects.
Reference Objectives
[112]
[95]

[113]
[114]

[115]
[116]

[117]
[118]

[119]
[120]

[121]
[90]

Emission abatement by bio-ethanol and diesel/bio-ethanol blends

Main results

- Addition of bio-ethanol to diesel decreases particulate matter (PM) generation and


increase the diesel PM oxidation.
CHP using various blends of gasoline in an IC engine
- Cylinder pressure and fermentation increase at higher bio-ethanol blends;
- CO emissions are reduced by up to 50%;
- E20 provides maximum availability for heat recovery.
Integrated bio-ethanol fermentation and CHP
- All heat required in the plant can be generated from biogas produced by digesting
the stillage.
Combustion performance of bio-ethanol at various blend ratios in a
- Faster combustion, higher in-cylinder pressure;
gasoline direct injection engine
- Similar or reduced NOx emission;
- Reduced CO emission;
- Increased engine efciency.
Bio-ethanol affects combustion and emission reduction in an SI engine - CO and volatile HC decrease;
- NOx tends to increase as ambient air temperature increases.
Adding ethanol in a small capacity Diesel engine
- Optimum 15% of ethanol;
- Unburned hydrocarbon emissions decrease;
- Load smoke opacity is lowest at 14% ethanol.
Use of bio-ethanol fueled buses by air pollution screening and on-road - Higher emissions of acetaldehyde and acetic acid during driving conditions.
measurements
Wet ethanol in HCCI engines with exhaust heat recovery
- HCCI engines can use ethanol fuel with up to 20% H2O;
- Low NOx, CO emissions at high intake pressures, high equivalence ratios, and
delayed combustion timings.
NOx-PM trade-off in a single cylinder diesel engine by means of bio- Very low levels of NOx and PM (meeting 2009 Japanese Standards);
ethanol and exhaust gas recirculation (EGR)
- 50% ethanol blended diesel fuel and high EGR ratios are preferred.
DI diesel engine fueled with bio-ethanol diesel emulsions
- 5, 10 and 15% ethanol addition tested;
- NO reduced by 4%;
- Smoke emission cut by 20%;
- 5% blend has best performance.
Sulfur in bio-ethanol
- Sulfur content of ethanol fuel must be <10 mg/kg.
Tractor engine performance using diesel/ethanol blends
- Power and torque increase by ~5e10%;
- Fuel consumption increases;
- No engine modications needed;
- Volatile HC emissions decrease as rpm increases.

70

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

investigations. As of 2013, the rst conmercial-scale plants have


started operation. Multiple pathways for the conversion of different
feedstocks are being used. In the next few years, associated cost
data and their relative performance will become available. A preliminary assessment of the lignocellulosic bio-ethanol is given by
Limayem and Ricke [124].
Carbohydrate-to-ethanol processes have been used for several
decades, and conversion data are established. Per ton of bioethanol, the required amount of raw materials required ranges
from 5 to 6.6 ton for cassava, 3e3.2 ton for corn, and 12e13.5 ton
of sugarcane [125,126]. Similar feed stock requirements when
using lignocellulosics are not yet available. The ethanol production
cost is moreover scale sensitive, with also costs of feedstock
collection and transport to be considered. Optimal sizes currently
vary between 50 and 500 thousand ton per year, dening the
amount of feedstock required for given or known conversion
yields of the feedstock.
3.2. Main steps in biomass-to-ethanol processes
Once the feed stock is delivered to the ethanol plant, it needs to
be carefully stored and conditioned to prevent early fermentation
and bacterial contamination. Through pre-treatment, simple sugars
are made available in proportions depending on the type of
biomass used and the pre-treatment process. The main steps,
different according to the feedstock, are summarized in Fig. 5,
providing a general production owsheet for the different bioethanol generations.
The processes are equivalent from the fermenter onwards, and
then start with a dilute ethanol-in-water fermenter broth. The
hydrolysate, the yeasts, nutrients and other ingredients are added
from the beginning of the batch fermentation. Continuous processes, in which ingredients are constantly added and products
removed from the fermentation, are also used [127]. The cell densities are kept high by recycling or immobilizing the yeasts to

improve their activity and enhance the fermentation yield. The


fermentation reactions occur at 298e314 K and last 1 to ~3 days
depending on both the feedstock composition, and the type,
amount and activity of the yeasts. Distillation will separate ethanol
and water up to the azeotropic point. The residual ow from the
distillation column, known as vinasse or stillage, can either be used
to produce biogas by anaerobic digestion (to produce process steam
and electricity), or processed into animal food, fertiliser and other
valuable by-products.
3.3. Bioethanol from disaccharides- and starch
The most common disaccharide used for bio-ethanol production
is sucrose, i.e. the combined molecule of glucose and fructose. Its
fermentation is performed using commercial yeast such as S. cerevisiae in a successive hydrolysis of sucrose followed by fermentation of simple sugars: invertase (an enzyme present in the yeast)
catalyzes the hydrolysis of sucrose into glucose and fructose, and
another enzyme (zymase), also present in the yeast, subsequently
converts the glucose and the fructose into ethanol and CO2.
The practical efciency of fermentation is about 90e92%. Sucrose is mostly available in sugarcane and sugar beet, where the
production of sugar is the main objective, and bio-ethanol is produced from the molasses, the residue after sugar crystallization.
Starch consists of long chains of a-glucose monomers, 1000 for
one amylose molecule and 1000e6000 or more for amylopectin
[128]. To convert starch to ethanol, the polymer of a-glucose needs
to be hydrolyzed into glucose by the gluco-amylase enzyme.
glucoamylase

C6 H10 O5 n nH2 O ! nC6 H12 O6

(3.1)

The resulting sugar is known as dextrose or D-glucose, an isomer


of glucose. The enzymatic hydrolysis is then followed by fermentation and successive ethanol separation technique. Corn and cassava contain 60e70% starch. The milled raw material is hydrolyzed
and the resulting sugar contained in the hydrolysate is converted to
ethanol while the remaining residue containing bre, oil and protein is roasted and converted into a by-product known as Distillers
Dried Grains (DDG) or DDGS when it is combined to process syrup
and sold mainly as animal feed additive (as molasses are also
commonly used for). The CO2 produced can be sold for different
applications (carbonated beverages or dry ice).
3.4. Lignocellulosic biomass-to-ethanol

Fig. 5. Generations of bio-ethanol production.

Lignocellulose [128], the principal component of the plant cell


walls, is mainly composed of cellulose (40e60% of the total dry
weight), hemicellulose (20e40%) and ligin (10e25%). Cellulose
molecules consist in long chains of b-glucose monomers gathered
into micro-bril bundles. The hemicelluloses, mostly xyloglucans
or xylans, are linked to the micro-brils by hydrogen bonds. Lignins
are phenolic compounds which are formed by polymerization
different productions of three types of monomers (p-coumatyl,
coniferyl and synapyl alcohols). Lignin adds compressive strength
and stiffness to the cell wall [129]. Lignocellulose is abundant in
nature does not compete with food, is available as agricultural and
forestry residues, industrial wastes and dedicated woody crops
(willow, poplar). Once the lignocellulosic biomass is pre-treated
and hydrolysed, the released sugars are fermented and the downstream process is similar to that of rst generation feed stock.
Pretreatment involves delignication of the feedstock [130] in order to make cellulose more accessible in the hydrolysis step, using
physical, physico-chemical, chemical and biological treatment
(Table 9).

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88


Table 9
Assessment of selected pre-treatment processes.
Yield of
fermentable
sugars

Wastes

Investment

Physical or physico-chemical
Mechanical
Steam explosion
Ammonia ber explosion (AFEX)
Carbonic acid

Low
High
Moderate
Very high

Very low
Low
Very low
Very low

Low
High
High
Low

Chemical
Dilute acid
Concentrated acid
Alkaline extraction
Wet oxidation
Organosolv

Very high
Very high
Very high
High
Very high

High
High
High
Low
Low

Moderate
High
Low
Low
Very high

Pre-treatment process [132,133]

Carbonic acid and alkaline extraction have the best performance. However, the most common methods are steam explosion
and dilute acid pre-hydrolysis, which are followed by enzymatic
hydrolysis. Sulphuric acid or carbon dioxide are often added in
order to reduce the production of inhibitors and improve the solubilization of hemicellulose [131]. Steam explosion has a few limitations since the ligninecarbohydrate matrix is not completely
broken down; degradation products are generated that reduce the
efciency of hydrolysis and fermentation steps; and a portion of the
xylan fraction is destroyed.
The use of dilute sulphuric acid (0.5e1%; 433e463 K for 10 min)
has the preference of the US National Renewable Energy Laboratory
[134,135]: hemicellulose is largely hydrolysed releasing different
simple sugars (xylose, arabinose, mannose and galactose) but also
other compounds of the cellulosic matrix, that can however inhibit
the enzymatic hydrolysis and fermentation. Part of the acetic acid,
much of the sulphuric acid and other inhibitors produced during
the degradation of the materials need to be removed, and
neutralization is performed before fermentation.
Enzymatic hydrolysis of cellulose is achieved using cellulases,
usually a mixture of groups of enzymes such as endoglucanases,
exoglucanases and b-glucosidases acting in synergy for attacking
the crystalline structure of the cellulose, removing cellobiose from
the free chain-ends and hydrolysing cellobiose to produce glucose.
Cellulases are produced by fungi, mainly Trichoderma reesei, besides
Aspergillus, Schisophyllum and Penicillium. High concentrations of
cellobiose and glucose inhibit the activity of cellulase enzymes and
reduce the efciency of the saccharication. One of the methods
used to decrease this inhibition is to ferment the reduced sugars
along their release. This is achieved by simultaneous saccharication and fermentation (SSF), in which fermentation using yeasts (S.
cerevisiae) and enzymatic hydrolysis are achieved simultaneously
in the same reactor. The fermentation of the xylose released from
the pre-hydrolysis process can be carried out in a separate vessel or
in the SSF reactor using a genetically modied strain from the
bacterium Z. mobilis that can convert both glucose and xylose. The
latter method is named simultaneous saccharication and cofermentation (SSCF).
Compared to the sequential saccharication and fermentation
process, SSCF exhibits several advantages like lower requirement of
enzyme, shorter process time and cost reduction due to economy in
fermentation reactors (only one reactor compared to the multiple
sets). Disadvantages include the difference between optimal temperatures for saccharication (323e333 K) and fermentation
(303 K), inhibition of enzymes and yeast by ethanol and the
insufcient robustness of the yeast in co-fermenting C5 and C6
sugars. The efuent from the distillation column that contains most
of the lignin and other non-fermentable products is sent to a

71

combined heat and power (CHP) plant to produce process steam


and electricity required by the ethanol plant. Depending on the
proportion of lignin in the feedstock, excess electricity may be
available for export sale.
Contrary to the conversion of disaccharides and starch to
ethanol which are mature technologies, the modern lignocelluloseto-ethanol process is still at a pilot and demonstration stage,
although demonstration plants of NREL (USA) [135], Iogen Corporation (Canada) [69], and ETEK (Sweden) [136], have built pilot
plants capable of producing a few hundred thousand liters of
ethanol per year.

3.5. Fermentation of hexoses and pentoses


From the prevous review of rst and second generation bioethanol processes, it is clear that a variety of microorganisms
such as yeasts, bacteria or fungi are required to biochemically
convert hexoses (C6, rst generation, second generation) and
pentoses (second generation). The conversion reactions are
different:

C6 H12 O6 /2C2 H5 OH 2CO2

(3.2)

3C6 H10 O5 /5C2 H5 OH 5CO2

(3.3)

S. cerevisiae, a facultative anaerobic yeast, and Z. mobilis, a gramnegative bacterium are commonly used to convert C6 sugars. Both
are adapted to ethanol fermentation, with a high ethanol tolerance,
and amenability to genetic modications. Both are however unable
to ferment C5 sugars. Other yeasts and bacteria are under investigation to ferment xylose and other pentoses into ethanol, with
Candida shehatae and Pichia stiplis yeasts offering a good potential,
despite a low tolerance to ethanol, a low ethanol yield and inactivity at low pH. Pachysolen tannophilus and Escherichia coli, both
gram-negative bacteria, are able to use both pentose and hexose
sugars. Kluveromyces marxianus, a thermophilic yeast, is able to
grow at a temperature above 325 K; and is capable to ferment a
broad spectrum of sugars. Excess of sugars affect its alcohol yield. It
has however a low ethanol tolerance and poor xylose fermentation
yield. Thermophilic bateria, e.g. thermoanaerobacterium sacchaarolyticum, thermoanaerobacter ethanolicus, clostridium thermocellum, are extreme anaerobic bacteria, resistant to an
extremely high temperature of 343 K. These bacteria can ferment a
variety of sugars, display cellulolytic activity, but exhibit a low
tolerance to ethanol.
Genetically engineered microorganisms [69,137e139] that produce large volumes of cellulase, xylanase, and hemicellulase enzymes, are under investigation. These could convert agricultural
residues (e.g. corn stover, straw, sugar cane bagasse) and energy
crops (e.g. switchgrass) into fermentable sugars [69]. The development of genetically modied fermentative and cellulolytic microorganisms is recommended to increase the ethanol yield and
productivity under the stress conditions of high production bioethanol processes [140].
Genetic engineering has succeeded in altering the conventional
S. cerevisiae's capacity to ferment glucose and pentose sugars
simultaneously [141]. Almaida et al. [142] investigated a modied
S. cerevisiae, capable of co-fermenting saccharides but also generating less furfural inhibitors. Z. mobilis remains an attractive
candidate due to its high ethanol yield and resistance to temperature in the range of 313 K [137]. Numerous genes have been
introduced and heterologous expression has been incorporated
into Z. mobilis to extend its effectiveness toward other substrates
namely, xylose and arabinose [143]. Both the gene engineered

72

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

Z. mobilis and S. cerevisiae have proven a high ethanol yield and


adaptability [144].
Enhancing ethanol production by pre-treatment involving fungi
(e.g. T. reesei and Basidiomycetes) with appropriate lignocellulolytic
properties at low pH and high temperatures is also a promising and
added-value step in ethanol bioconversion. While fungi act slowly,
potential lignocellulolytic fungi have been produced by mutagenesis, gene expression and co-culturing [139]. Some genera, such as
Candida, Pichia and Dekkera were isolated from sugarcane molasses,
but resulted in low ethanol concentrations and produced acetic
acid, an inhibitor of the fermentative yeast. Some natural wild yeast
species appear capable of replacing S. cerevisiae in secondgeneration bioethanol, but their low bio-ethanol yield and poor
survival in the fermenter need further improvement.
3.6. The problem of ethanol-inhibition in the rst generation
feedstock fermentation
The yield of the fermentation xes the total energy needed
during the downstream separation processes, mostly by distillation. This yield is not only limited by the conversion rate, but also by
process inhibition, Inhibition occurs when the concentration of a
chemical, either produced during the reaction or present in the
reaction mix, reaches a toxic value, whereby the reaction rate is
hampered or possibly even stopped by deactivation of enzymes or
by the death of the microorganism. This phenomena is responsible
for the low achievable ethanol concentrations in industrial fermenters. The maximum concentration of ethanol attained in most
fermenters is around 11e12 wt%. Due to the fact that ethanolinhibition is a very complex and difcult phenomena, a lot of
unknows about the problem and its solution exist. Ethanol inhibits
the system in 3 ways, i.e. inhibition of the cell growth, inhibition of
the fermentation and cell death. The rst two effects are illustrated
below (Fig. 6), for fermentation of lactose by Candida pseudotropicalis (150 g/L) [145,146].
This inhibition is a consequence of the effect of ethanol on the
plasma membrane and the enzymes in the rst glycolysis step of
fermentation [146]. These enzymes, particularly hexokinase, are
very sensitive to high ethanol concentrations: their activity decreases and the fermentation rate drops. Furthermore the high
ethanol concentration hampers the plasma membrane uidity,
necessary for the transport of nutrients in and out of the microorganism, leading to reduced activity. A review of inhibition

investigations in the bio-ethanol fermentation is presented in


Table 10.
To increase the critical inhibition concentration, adapted yeasts
can be used. The most commonly used yeast is S. cerevisiae, with a
moderate yield of fermentation. Research has been done on more
promising yeasts and bacteria: Z. mobilis succeeds to survive higher
ethanol concentrations in the fermenter. Not only this advantage,
but also a high tolerance for acids and sugars, makes this a very
popular yeast for industrial application. The fermentation rate is also
higher with Z. mobilis in comparison to Saccharomyces cerevisiae
[146] The research on this baterium has lead to the development of
very high gravity (VHG) fermentation, as described in Section 5.2. Z.
mobilis is a bacterium belonging to the genus Zymomonas. Originally
isolated from alcoholic beverages like African palm wine or Mexican
pulque, it is also a contaminant of cider and beer in European
countries. The advantages of Z. mobilis over S. cerevisiae with respect
to producing bio-ethanol: (1) higher sugar uptake and ethanol yield
[152]; (2) lower biomass production; (3) higher ethanol tolerance,
up to 16 vol% [165]; (4) amenability to genetic manipulations. An
interesting characteristic of Z. mobilis is indeed that its plasma
membrane contains hopanoids, pentacyclic compounds similar to
eukaryotic sterols, thus providing an extraordinary tolerance to
ethanol in its environment, around 16 wt%.
However, in spite of these attractive advantages, its substrate
range is limited to glucose, fructose and sucrose. It cannot ferment
C5 sugars like xylose and arabinose which are important components of lignocellulosic hydrolysates. Unlike yeast, Z. mobilis cannot
tolerate toxic inhibitors present in lignocellulosic hydrolysates such
as acetic acid and various phenolic compounds [159]. Concentration of acetic acid in lignocellulosic hydrolysates can be as high as
1.5 wt%, well above the tolerance threshold of Z. mobilis.
Several attempts have been made to engineer Z. mobilis to
overcome its inherent deciencies by metabolic engineering,
mutagenesis or adaptive mutation to produce acetic acid resistant
strains of Z. mobilis [143,160,161]. However, when these engineered
strains metabolize mixed sugars in the presence of inhibitors, the
yield and productivity are much lower, thus preventing their industrial application.
As mentioned before, albeit outside the scope of our rst generation feedstock approach, other types of inhibition occur during
the processing of lignocellulosics. They include sugar degradation
products, lignin degradation products, compounds derived from
extractives and heavy metal ions. Sugar degradation products are

Fig. 6. Inhibition of the fermentation (a) and the cell growth (b).

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

73

Table 10
Literature review of inhibition studies (>2008).
Reference

Objectives

Main results

[147]

Tolerance of S. cerevisiae and Z. mobilis


to inhibitors of soybean meal

[148]

Enhancement of ethanol fermentation in


Saccharomyces cerevisiae sake yeast

[149]

Adaptive evolution of an industrial strain of


Saccharomyces cerevisiae for combined tolerance
to inhibitors and temperature

[150]

Investigation of hemicellulase inhibition in the


production of bioethanol

[151]

A Thermotolerant Saccharomyces cerevisiae Strain


(TT6) for bioethanol production
Microbial contamination of fuel ethanol fermentations

- Obtained both at 408 K, 2.0% H2SO4 and 45 min, and at 408 K, 1.25% H2SO4 and 45 min
showed inhibition in the growth of the tested microorganisms;
- Spiked yeast medium (YM) broths could withstand the highest levels of inhibitors.
- The maximum rate of CO2 production and nal ethanol concentration generated by the
atg32D laboratory yeast mutant were 7.50% and 2.12% higher than those of the parent
strain, respectively;
- The mutant produced ethanol at a concentration that was 2.76% higher
than the parent strain;
- The ethanol yield of the atg32D mutant was increased, although its biomass
yield was decreased.
- Saccharomyces cerevisiae isolate (ISO12) is capable of growing and fermenting the liquid
fraction of non-detoxied spruce hydrolysate at 312 K with an ethanol yield of
0.38 g ethanol/g hexoses;
- ISO12 shows a higher capacity to ferment hydrolysate at 312 K and higher viability
during heat-shock at 325 K than the industrial strain Ethanol Red (ER).
- A range of potential xylanase inhibitors observed in dilute-acid pretreatment slurries
and fermentation broths are dosed into washed insoluble solids (IS) and beechwood
xylan at varied concentrations prior to enzymatic hydrolysis.
- The ethanol yield from corn by simultaneous saccharication and fermentation (SSF)
with TT6 at 309 K was 91.7% of the theoretical yield.
- The contamination of bioethanol fermentations with lactic acid bacteria (LAB) and wild
yeasts is a signicant industrial problem causing production loss of anywhere from 2 to 22%;
- Dekkera bruxellensis has been cited as one of the main contaminant yeasts in
ethanol production.
- Fermentation yield of hydrolysates can be improved signicantly by
increasing pH of hydrolysates;
- For hydrolysates at 25% and 30% total solids, pH of 8.0e9.0 yields optimal fermentation
with almost no bacterial contamination.
- The bioethanol yield in the presence of furfural increased via mutation of a native
strain of Saccharomyces cerevisiae;
- A potent mutant was selected which produced 36.7% more bioethanol than the parent
strain at 0.2 vol% furfural.
- NFRI3225 produced ethanol from potato mash at the fastest rate and in the
highest volume (13.7 vol%), and the maximum productivity and ethanol yields
were 9.1 g L1h1 and 92.3% respectively during the VHG-SSF process.
- NFRI3225 would save cooling energy during the SSF process and heating energy
during distillation.
- Low pH activates the general stress response (GSR), and mainly the heat shock response;
- A ne regulatory protein kinase A (PKA) dependent mechanism might affect the cell
cycle in order to acquire tolerance to an acid environment.
- The volumetric productivities and ethanol yields were attained to 3.26 g L1h1 and
93.5% for brown rice, 2.62 g L1h1 and 90.4% for barley, and 3.28 g L1h1 and 93.7% for
cassava, respectively.
- The diffusion inside the biocatalyst particles avoids the inhibitory effects;
- The obtained values of Vmax and K0 m were found to be higher than those under
ethanol inhibition, but lower than those without inhibitory phenomena.

[152]

[153]

Crucial yeast inhibitors in bio-e


thanol and improvement of fermentation at high
pH and high total solids

[154]

Increasing the bioethanol yield in the presence of furfural


via mutation of a native strain of Saccharomyces cerevisiae

[155]

Selection of stress-tolerant yeasts for simultaneous


saccharication and fermentation (SSF) of very high
gravity (VHG) potato mash to ethanol

[156]

The stress response of Saccharomyces cerevisiae imposed by


strong inorganic acid with implication to
industrial fermentations
Ethanol production by Zymomonas mobilis CHZ2501
from industrial starch feedstocks

[157]

[158]

Kinetic studies on alcoholic fermentation under


substrate inhibition conditions

formed during hydrolysis and include mostly furfural from pentoses and 5-hydroxymethylfurfural (HMF) from hexoses.
The use of a detoxication methodology for a specic feedstock
is mandatory for attaining good results in 2nd generation bioethanol production.
4. Traditional processing routes
4.1. Integrated saccharication/fermentation processes versus twostep processes
Fig. 5 of Section 3.2 illustrated the successive steps of the overall
process, including the different enzymatic activities during
saccharication and fermentation. Whereas some processes, e.g.
Lurgi [162] separate both steps, other processes, e.g. Cofco [24]
combine the dual enzymatic activity within the fermenter solely.
Despite this difference, fermentation yields obtained are nearly
equal, and hence not affecting the subsequent bio-ethanol purication train and/or waste treatment. The fermentation yield depends upon the conversion of starch or monosaccharides content in
the rst place (mostly 60e70%, as function of the rst generation

raw material), coupled with the yield of fermentation itself,


generally >90%. The mass balance of the processes can hence be
established, and is illustrated in Fig. 7 for the cassava-based bioethanol production of the Cofco concept [24].
The combined Cofco process uses 6 fermenters in parallel, each
fermenter having a volume of 3025 m3, i.e. ~3025 ton. 865 ton of
cassava our is used per batch of 65 h. The production is based
upon a 70% starch-content of cassava and 90% starch-to-glucose
conversion. The amount of starch converted is hence 549 ton.
Fermenter broths of all 6 fermenters are collected and dealt with in
a single continuous distillation train.
The recycle to the fermenter is applied to keep the enzyme and
yeast balance at the required level.
For the total Cofco process, using 6 parallel fermerters, the total
production capacity is ~200,000 tpa of anhydrous ethanol. Since
this capacity is the average of current industrial capacities
(50,000e400,000 tpa), we have selected 200,000 tpa as target
capacity.
Several distillation owsheets have been presented in literature,
with minor process-specic defferences (2 or 3 distillation columns), but with an overall equivalent mode of operation to enrich

74

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

Fig. 7. Mass balance of the rst generation Cofco process of cassava-based bio-ethanol.

Fig. 8. Schematics of the Cofco distillation train.

the ethanol-lean broth to the ethanolewater azeotrope. Since eld


data for the Cofco application are available and published, Cofco is
more readily assessed than other applications where eld data are
not available.
The ow sheet of the current distillation process of the Cofco
cassava-based fuel ethanol production (200,000 tpa) is shown in
Fig. 8.
It consists of a sequence of distillation columns, heat exchangers, pumps and a ash tank. The feed stream enters the crude
column C1501. The distillate from C1501 is fed to column C1502.

The side stream 4 is fed to the stripping section C1503-1 of the


second column. C1503-2 is the rectifying section of the second
column. C1503-1 has no condenser and C1503-2 has no reboiler.
The bottom stream 10 enters C1502. Stream 15 is the ethanol-rich
phase. Simultaneously, the distillate from C1502 is separated via
the ash tank V1504, whose gas and liquid streams are both fed to
C1503. The bottom streams (5, 9, 16) are directed to the wastewater
treatment unit and should therefore be ethanol-lean (0.05 wt%
ethanol). Finally the distillate from C1503-2 is further dewatered
using molecular sieve adsorption to an anhydrous nished product

Fig. 9. The distillation process of Lurgi (C1501-Crude column; C1502-the polishing column).

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

75

of 99.7 wt% ethanol [163]. A reboiler saturated steam circuit at


443 K (8 bar) is available. Fig. 8 does not include condenser/reboiler
heat recovery, as well be dealt with in Section 5.1.
The distillation train of the Lurgi concept [162] is represented in
Fig. 9, and uses 2 distillation columns, 2 heat exchangers and
associated pumping. Distillation columns C1501 and C1502 use
steam of 373 K and 413 K, respectively.
The analogy between the illustrated Cofco and Lurgi designs is
outspoken, since the rst Cofco column (C1501) does not considerably
enrich the top and side streams, but mostly eliminates the ne suspended solids of the broth, further evacuated in the bottom stream.

Table 12
Operating parameters of each column.

4.2. Basic data and energy requirement of the process

QREB kW 9000 18; 222 11; 604 38; 826 kW

4.2.1. Fermentation broth


As example of the complete analysis, the large-scale cassavabased operation of Cofco [24,25] is examined as distillation feed.
The composition of the cassava fermentation broth is shown in
Table 11.
The feed ow rate is about 220,000 kg/h at atmospheric pressure, and 305 K. Since high boiling point components were not
individually analyzed, they are grouped as C20H16 for use in the
subsequent simulations.
Referring to Fig. 8 before and the different key-parts of the
distillation process, the number of the theoretical trays and operating pressure of each column are taken from the eld data, and
shown in Table 12.
The operating pressure increases along the distillation train, from
about 0.5 to about 6 bar. Number of trays and reux ratio are common
and important operation characteristics of the respective columns.
4.2.2. Simulation in Aspen Plus of the basic concept (no internal
energy recycle)
Aspen Plus V8.2 is a comprehensive chemical process modeling
system, used by the world's leading chemical and specialty chemical organizations, and related industries to design and improve
their process plants. To simulate the process operation, the software Aspen Plus V8.2 was used [164]. Due to the presence of the
highly polar components, ethanol and glycerol, both the nonrandom two liquid (NRTL) and universal quasi-chemical (UNIQUAC) thermodynamic/activity models were used to predict the
activity coefcients of the components in a liquid phase [165].
Some unavailable interaction parameter coefcients were estimated using the UNIQUAC liquideliquid equilibrium module. Results from the simulation towards operating temperatures in the
different columns were compared with the real values of Table 12.
The comparison of Table 13 shows an excellent agreement, with
simulated column temperatures and ethanol concentration of
simulation and plant measurements in fair agreement.

Column

Number
of trays

Feed tray
number

Reux
ratio

Top pressure
(kPa)

Bottom
pressure (kPa)

C1501
C1503
C1502

15
30
40

6
20
30

0.3

33.3
189.3
600.4

51.3
215.3
636.7

3.1

It conrms that the watereethanol distillation can adequately


be designed on the basis of Aspen Plus simulations.
The total reboiler duty amounts to:

(4.1)

For a stream feed at 443 K and the different reboiler temperatures, a conservative average sensible and latent enthalpy of steam
is taken at 2300 kJ/kg steam. For a total production of 25375 kg/h of
pure ethanol, the steam requirements are:

38; 826  3600


2:50 kg steam=kg bio  ethanol
25; 375  2200

(4.2)

A stream duty of about 0.5 kg steam/kg bio-ethanol must be


added for the molecular sieve separation (see Section 5.5).
This high steam requirement of the fundamental solution,
without in-process heat recovery should be reduced to as low value
as possible in order to improve the operation economy of the bioethanol production. This will be dealt with in Sections 5 and 6
hereafter.
5. Process improvements
5.1. Energy integration within the current production processes
As can be seen from Table 12, the pressure is progressively
increased along the distillation train. It is hence possible to recover
condenser heat to feed the reboilers of previous columns. The heat
duty difference between the top of C1502 and the bottom of its
neighbor C1503 is sufcient to modify the system into a combined
reboiler/condenser distillation: the top high pressure condensate of
condenser C1502 is used as the heat source for the reboiler of
C1503, and top high pressure condensate of C1503 is used as the
heat source for the reboiler of C1501. In this energy recycling
scheme and accounting for a required thermal driving force in the
reboilers of about 288e293 K, the remaining steam consumption
will only be due to the bottom reboiler of C1502.
Referring to Table 13, the total reboiler heat duty of the heatintegrated distillation train will be

QREB kW 18; 222 kW

(5.1)

At the average steam enthalpy of 2300 kJ/kg, the energy consumption of the integrated reboiler-condenser solution is
Table 11
Composition of the Cassava-fermentation broth.
Components

wt%

Water
Ethanol
Acetaldehyde
Methanol, isoamylalcohol, isobutanol,
propanol, butanol, pentanol, isopropanol
Methyl and ethyl acetate, ethyl formate
Ethyl isobutyrate
Succinate
Acetic, lactic, butyric, formic acid and propionic acid
Glycerol
C20H16

81.40
11.55
0.005
0.1249
0.0055
0.0005
0.25
0.26
0.4
6

18; 222  3600


1:18 kg steam=kg bio  ethanol
25; 375  2200

(5.2)

In the current Cofco operation, this reboiler-condenser integration is largely applied, with steam duties cited as ~1.3 kg steam/
kg bio-ethanol, in fair agreement with the Aspen-simulated results.
5.2. The use of VHG fermentation: principles and application results
5.2.1. Introduction of very high gravity (VHG) fermentation
In traditional starch fermentation, Saccharomyces cerevisiae is
used as yeast, however with the drawback of inhibition at ethanol
concentrations in excess of ~12 vol%. The use of specic yeasts or

76

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

Table 13
Comparison of simulation results with real industrial production data.
C1501
Plant value
Top temperature/K
Bottom temperature/K
Ethanol concentration of side stream (wt%)
Final ethanol concentration (wt%)
Heat duty (kW)
Reboiler
Condenser

329.10
355.70
16

C1502
Simulation value
326.99
355.16
16.9

Plant value
368.60
396.30

C1503
Simulation value
367.88
395.61

Plant value
404.20
432.90
93e94

9000
5694

bacteria, such as Z. mobilis, with higher inhibition levels, opened the


way to very high concentration fermentation, referred to as very
high gravity (VHG) fermentation, and capable of increasing the
total fermentation yield. In VHG, very high concentrations of sugars
(>250 g/L) are loaded into the fermenter when compared to classical fermentation. VHG is commonly referred to as the digestion of
a feed which contains more than 27 wt% of dissolved solids. Some
recent literature is reviewed in Table 14.
Much higher concentrations of ethanol can be maintained in the
fermenter by application of this technique, and ethanol concentrations of 23.8 vol% have been reached [174]. Besides this advantage, a lower risk of contamination and a saving of fermenter
volume space can be realised by using VHG fermentation [168], as
further illustrated in Section 5.4.
The principles of inhibition have been dealt with in Section 3.6,
with the high osmotic pressure and toxic effects of ethanol as main
causes. A lot of research has been done the past few years to understand these extreme conditions and to anticipate on them.
Research has been focusing on the micro-organisms themselves,
added nutrients and osmoprotectants. As a result, it became clear
that the fermentation rate can be raised [175] by addition of
ergosterol, oleic acid or polysaccharides, whilst, higher ethanol
concentrations can also be achieved when ureum, yeast-extracts or
ammonia salts are added [176]. The major drawback of applying
VHG fermentation is the high viscosity in the fermenter as a result
of the high sugar concentrations. This high viscosity hinders the
solid/liquid separation and the heat exchange during the distillation. To solve this problem Wang et al. suggest using pectine-

11,604
11,575

Simulation value
404.93
434.37
93.2
18,222
16,944

degrading enzymes or other enzymes like xylase, cellulase and


xylanase to lower the viscosity [177]. In the case of applying VHG
fermentation, the use of partial pervaporation (see 5.3) might be
very interesting: reducing the ethanol concentration in the VHG
fermerter will reduce the inhibition stress on the bacteria.
5.2.2. Effect of implementing VHG on the distillation thermal
requirements
The previous simulation (Table 13) accounted for 11.6 wt%
ethanol. If the ethanol content of the broth can be increased, the
required heat duty of the distillation can be reduced due to the
reduction of H2O content of the broth.
Although VHG results up to 23.8 vol% have been reported [174],
the authors conservatively examined situations between 15 and
19 vol%, at a constant production of 25,375 kg/h of ethanol.
Results are presented in Table 15, for the 15 and 19 vol% cases
only.
As expected, the lowest reboiler duty is obtained at VHG of
19 vol%. If again the condenser heat of C1502 recycled to the
reboilers of the lower-pressure column, the required heat duty is
reduced to:

QREB kW 15; 162 kW

(5.3)

with associated steam use of

15; 162  3600


0:94 kg steam
25; 375  2300


kg bio  ethanol

(5.4)

Table 14
Literature review of VHG studies (>2010).
Reference

Objectives

Main results

[166]

[168]

Construction of Saccharomyces
cerevisiae yeast strains for VHG fermentation
SSF of sweet potato for the production of
ethanol under VHG conditions
VHG ethanolic brewing and fermentation

[169]

Bio-ethanol production from mahula owers

[170]

Energy-saving direct ethanol production


from sweet potato using VHG
Improving ethanol performance in VHG fermentation
through chemical mutagenesis and meiotic recombination

- Increased ethanol yield by up to 8% and less glycerol with strain SZ3-1.


- Using cellulase to reduce the high viscosity of the liquid mash;
- The nal ethanol concentration reached 15 vol%, and the total sugar conversion
and ethanol yields were 96.5% and 87.8%, respectively.
- Fermentation temperature should not exceed 303K;
- Technically and economically viable; acceptable by industry.
- The S. cerevisiae strain (149 g/kg) showed 21.2% higher nal ethanol production in
comparison to Z. mobilis.
- A nal viscosity of 498.1 cp and ethanol yield of 135.1 g/kg was obtained (using
pretreated sweet potato mash).
- The strain improvement strategy is a powerful and high throughput method;
- Strain WS1D and WS5D obtained 15.12% and 15.59% increased ethanol yield
compared to the starting strain W303D.
- Over 30% increase in nal ethanol (17.1 vol%), 29% decrease in fermentation time
(60 h), 80% increase in biomass formation and 26% increase in glucose utilization;
- Both trehalose and plasma membrane ATPase contribute to the fermentation
enhancements.
- The initial yeast cell concentration was 5  107 cells ml1;
- An increase in sugar concentration in the inoculum preparation medium (from 10
to 100 g/L) improved the ability of the inoculum to produce ethanol;
- The cost of the sweet sorghum for ethanol production was US$ 0.63 per litre of
ethanol.

[167]

[171]

[172]

Medium optimization for ethanol production in VHG fermentation

[173]

The use of dried spent yeast in ethanol fermentation


from sweet sorghum juice under VHG conditions

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

77

Table 15
Heat duty of the distillation when using VHG fermentation.
VHG conditions
15 vol%

Reboiler duty (kW)


Condenser duty (kW)

19 vol%

C1501

C1502

C1503-1

5343
2264

15,094
14,356

18,163

C1503-2

C1501

C1502

C1503-1

10,062
9412

14,512

18,144

4163
2266

C1503-2
14,186

The comparison with the energy optimization by energy-pinch


alone (5.1) demonstrates that a further reduction in steam
requirement of ~0.3 kg steam/kg bio-ethanol is achieved.

given experimental results, and extrapolated to the large scale


operation of 220,000 kg/h, the following steady-state mass balance
of Fig. 10 prevails:

5.3. The development of hybrid (pervaporation) systems

FCOLUMN 0:7FFEED FRETENTATE 165; 000 kg=h;

(5.5)

FRETENTATE 47; 080 kg=h;

(5.6)

5.3.1. Introduction of hybrid operations


The use of water-ethanol separation membrane technology has
attracted an increasing attention in research, with some relevant
literature (>2010) summarized in Table 16.
Clearly, the hybrid operation mode is recognized as an important development, with high permeate uxes obtained, providing
additional energy savings, and overall cost reduction.
5.3.2. Effect of implementing pervaporation on the distillation
thermal requirements
The above literature survey already indicated that a hybrid
fermentation-distillation mode by integrating pervaporation, could
reduce the risk of fermenter ethanol inhibition by continuously
removing ethanol from the fermenter broth, whilst also reducing
the heat duty of the distillation through feeding the distillation
with ethanoleH2O streams at higher ethanol concentration.
Although literature provides indications concerning obtainable
membrane uxes, and ethanol selectivity, a conrmatory set of
experiments was carried out by the authors using a pervaporation
(hydrophobic) membrane of Pervatech PDMS. Experiments were
carried out with a ~12 vol% ethanoleH2O feed, at temperatures of
310, 318, 328 and 334 K. Detailed results are reported elsewhere
[190], but major results are presented hereafter. The permeate ux
and ethanol wt% are presented in Table 17.
From the presented data it is clear that the permeate ux
signicantly increases as the temperature increases, whilst the
ethanol membrane selectivity remains the same. At 334 K, the
permeate ux reaches 5.8 kg/m2 h.
From the observations of ethanol concentration in the permeate,
a higher feed temperature does not result in a higher ethanol
membrane selectivity. Since ethanol molecules have a larger size
than water, the assumption can be made that diffusion effects are
dominant over sorption effects, and are thus expected to be the
determining parameter in the transport mechanism description
through the PDMS-based membrane.
It is expected that the minor increase of the selectivity at higher
temperature is explained by a dependent decrease of the hydrogen
bonding between water and alcohol molecules. When this is the
case, less water is stimulated to permeate through the membrane.
Hence, not only membrane swelling seems to play an important
role during ethanol/water pervaporation but also the coupling effects between molecules, which tend to be lower at higher
temperatures.
To reduce the ethanol concentration in the fermenter below the
inhibition threshold limit of about 12 vol%, a hybrid system should
continuously at least remove 20e35% of the total fermenter broth
ow rates through the membrane. Although it is expected that the
continuous removal of ethanol will slightly increase the yield of the
fermenter, this was not accounted for in the calculations. At the

FPREMEATE 7920 kg=h:

The required membrane surface area at 334 K is 7920/


5.8 1366 m2. Although this surface area is considerable, it can be
reduced (i) by operating the pervaporation at as close a temperature as possible to the azeotropic point (~351 K), hence e.g. 348 K.
Extrapolating the T-dependent ux to 348 K, yields values of ~2.2 g/
m2s or 8 kg/m2h, leading to a reduction in membrane surface by
27.5%; (ii) by using a different membrane, as suggested in the
literature review of Table 15, where uxes up to 10 kg/m2h at
comparable or higher selectivities were indicated, again indicating
a further decrease in membrane surface area of about 25%; or (iii)
by reducing the extracted ow of current 33% to a lower value,
however high enough to reduce the inhibition risk.
It is not reconsidered to operate the pervaporation membranes
at higher ethanol concentrations, hence e.g. in association with
VHG, since it was demonstrated by Chovau [191] that the selectivity
of all examined membranes is strongly dependent on the alcohol
content in the feed: a higher concentration of ethanol in the feed
led to a decrease in membrane selectivity. Similar observations
were found by Chen et al. [192], who investigated the performance
of unsupported lled silicone rubber membranes in the separation
of ethanol/water mixtures ranging between 4 and 91 wt% of
ethanol in the feed. The authors observed the highest selectivity
factor at 4 wt% ethanol (a 28.5), while the factors decreased
considerably at higher ethanol concentrations in the feed (a of 7.7
and 2.3 at 26 and 91 wt% of ethanol, respectively). This decrease
was explained by an extensive swelling of the PDMS rubber at
higher feed ethanol concentrations. According to Jia et al. [193],
incorporation of silicalite llers in the polymeric network results in
a reduction of membrane swelling and moreover improves the
mechanical strength of the active layer.
The mass and heat balances around the complete unit, when
integrating pervaporation in a 33% side stream, as illustrated in
Fig. 10 before, revealed the following effect on the distillation plant
for a broth feed at 11.6 wt%.
The required steam amount is:

18; 222  3600


1:13 kg steam
25; 375  2300


kg bio  ethanol

(5.7)

5.4. Overall assessment


The effect of the different possible improvements can be summarized in a mass balance, as presented in Fig. 11.
It is evident that the use of VHG is recommended, since a
fermenter of the same volume can deal with about 30% more raw
material. The same applies to a lesser extent when pervaporation is
used to treat the fermenter broth. The combination of partial

78

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

Table 16
Literature review (>2010) of applying pervaporation in Bio-ethanol production.
Reference

Objectives

Main results

[178]

Membrane technology for bio-ethanol production

[179]

Signicant improvement of bio-ethanol recovery using a pervaporation


(silicone rubber-coated membrane)

[180]

Energy demand of biofuel production applying distillation and/or


pervaporation
Concentrating lignocellulosic ethanol through polydimethylsiloxane
(PDMS) membrane

- A highly selective and energy-saving separation process, with ~15%


lower costs in comparison to the batch process.
- The recovered ethanol concentration in the permeate was 67 wt%, and
the amount of recovered ethanol from the broth was more than 10
times higher than when using a non-coated membrane.
- Multi-stage pervaporation is needed to reach the fuel-grade quality of
the bio-ethanol.
- Yeasts, solid particles, and salts can increase ethanol ux and
selectivity;
- Integrated process can effectively eliminate product inhibition,
improve ethanol productivity, and enhance the glucose conversion rate.
- Pervaporative separation of bio-ethanol from the fermentation of
waste newspaper is possible without signicantly decreasing process
performance.
- Possible but dependent on the feed concentration and operating
temperature for ethanol concentrations of 1e10%;
- Relatively good pervaporation performance with a total ux of 231
e252 g/m2 h.
- Increase of cell density, decreasing ethanol inhibition, improved
productivity and yield, and resumption of clean and concentrated
ethanol;
- Ethanol productivity increased by 26.83% over conventional batch
fermentation;
- Ethanol concentration in permeated side was 6e7 times higher than
that of the broth.
- Zeolite 13X lled CA membrane has the better pervaporation
performance;
- Application of CA membranes in industrial scale pervaporation units
may be feasible for separation of ethanol/water mixtures;
- Bio-ethanol can be produced economically (Using distillation and
pervaporation by hybrid systems).
- As a highly selective and energy-saving unit operation, membranes
have a great potential in the bio-ethanol industry, both for starch-based
and 2nd generation technology of bio-ethanol.
- Pervaporation membrane bioreactors (PVMBRs) and Membrane
distillation bioreactors (MDBRs) lead to higher ethanol productivity and
minimize the inhibition;
- Membrane hybrid processes further improve biorening and
bioenergy production by decreasing energy consumption, reducing the
number of processing steps, and producing high quality nal products.
- Pervaporation performance was strongly dependent on the feed
concentration and operating temperature;
- Ethanol uxes were increased with the temperature increase;
- Lactic acid increased the hydrophilicity of the membrane.
- A separation factor of 40 and a ux of 1288 g/m2h with high long term
stabilitiy.

[181]

[182]

Pervaporative separation of bio-ethanol produced from the


fermentation of waste newspaper

[183]

Pervaporative separation of bio-ethanol using a polydimethylsiloxane/


polyetherimide composite hollow-ber membrane

[184]

Enhanced bio-ethanol production by pervaporation using a PDMS


membrane bioreactor

[185]

Bio-ethanol recovery using the pervaporation separation technique

[186]

Membrane process opportunities and challenges in the bio-ethanol


industry

[187]

Membrane technologies for biorening and bioenergy production

[188]

Pervaporation of ethanol produced from banana waste

[189]

Pervaporation membranes with modied polyvinylidene uoride


substrate for ethanol dehydration

pervaporation and VHG fermentation would moreover reduce the


ethanol concentration in the fermenter, thereby reducing the inhibition stress on the bacteria.
As a conclusion, the steam use per kg of bio-ethanol is reduced
to 1.17, 0.94 and 1.13 kg steam/kg bio-ethanol for options VHG
(15%), VHG (19%) and pervaporation, respectively, a considerable
saving with respect to the basic and to the energy-integrated solution, as summarized in Table 19.
5.5. The process of nal ethanol dewatering to fuel grade
5.5.1. Introduction
Since ethanol enrichment beyond the azeotropic mixture is
impossible by traditional distillation, specic alternatives need to
be employed.
There are 3 major techniques available to separate the ethanolewater azeotrope. Heterogeneous azeotropic distillation has
been widely studied in many papers and textbooks and widely
applied in the alcohol industry to dehydrate ethanol (e.g. 60% of
dehydration plants in Brazil are azeotropic distillation based).
However, heterogeneous azeotropic distillation reports some disadvantages associated with the high degree of nonlinearity,

multiple steady states, distillation boundaries, long transients, and


heterogeneous liquideliquid equilibrium, limiting the operating
range of the system under different feed disturbances [194e198].
Extractive distillation is based on the introduction of a selective
solvent that interacts differently with each of the components
[196e199] of the original mixture and which generally shows a
strong afnity with one of the key components [197,200,201]. The
addition of an entraining agent, such as benzene, cyclohexane, or
heptane, forms a new ternary azeotrope comprising ethanol, water,
and the entraining agent. Since this lower-boiling ternary azeotrope is removed preferentially, water-free ethanol is obtained [10].
Extractive distillation is seldom used nowadays.

Table 17
Permeate ux and ethanol concentration at different temperature.
Temperature (K)

Ethanol in permeate (%)

Permeate ux (g/m2s)

310
318
328
334

28.7
33
33
34

0.32
0.91
1.24
1.61

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

79

Table 18
Simulation results of the hybrid operation mode.

Ethanol (wt%)

To column

PV permeate

16

Product

11.6

30

<108

<104

91.7

Reboiler duty (kW)


Condensor duty (kW)

C1501

C1502

C1503-1

8676
5410

18,222
16,944

15,390

C1503-2
15,081

Table 19
Summary results of the Cofco distillation process.

Fig. 10. The mass balance of the hybrid operation.

Adsorption on molecular sieves takes advantage of the difference of the molecular size of ethanol and water molecules to
adsorb water molecules in a selective way and allowing ethanol
separation. The use of e.g. synthetic zeolite in pellet form, can
selectively adsorb water from the 95.6% ethanol solution. The
zeolite can be regenerated an unlimited number of times by drying.
Although a variety of plant-derived sorbents, including cornmeal,
straw, and sawdust can also absorb water, they cannot readily be
regenerated, but where ethanol is made from corn, cornmeal is
available at low cost.
In addition to these major technologies, other alternatives are
drawing increased attention, i.e:
 the use of vacuum distillation: the composition of the ethanolewater azeotrope shifts to ethanol-rich mixtures at pressures below atmospheric pressure, whilst the azeotrope
disappears at pressures below 9.333 kPa, making it possible to
distill absolute ethanol from an ethanolewater mixture. Such
high vacuum operation is presently not economical.
 applying pressure-swing distillation: in this technique, a
reduced-pressure distillation rst yields an ethanolewater
mixture of more than 95.6% ethanol. Then, fractional distillation
of this mixture at atmospheric pressure distills off the 95.6%
azeotrope, leaving anhydrous ethanol at the bottom.
 using membranes to separate ethanol and water: since membrane separations are not subject to the limitations of the
watereethanol azeotrope because separation is not based on
vaporeliquid equilibria, the selective permeation of one of the
components is possible by using either hydrophobic membranes
(ethanol permeates) or hydrophilic membranes (water

System

Steam consumption
(kg steam/kg ethanol)

Basic
Integrating reboilers and condensers
VHG (15%)
VHG (19%)
Pervaporation

2.50
1.18
1.17
0.94
1.13

permeates). Membranes are often used in the so-called hybrid


membrane distillation process, where distillation and membrane separation are combined, and the membrane operates
either in vapor permeation or in pervaporation mode, using
either a vapor membrane feed or a liquid membrane feed,
respectively.
 applying a variety of other techniques, as discussed in literature,
i.e. a liquideliquid extraction of ethanol from an aqueous solution, the extraction of ethanol from the fermenter broth by supercritical CO2 [202,203], or the use of Pressure Swing
Adsorption [204].
These alternatives are only at the early stages of experimental
investigation.
The three main ethanol dehydration technologies will be
reviewed hereafter in order to establish the main operating conditions required to obtain high purity ethanol.
5.5.2. Major processes
5.5.2.1. Azeotropic distillation with cyclohexane. Azeotropic distillation uses a solvent with an intermediate boiling point to introduce new azeotropes to the mixture and at the same time to
generate two liquid phases that allow, in a combined way, separating ethanol from water. This technique although being widely
used has lost acceptance due to its poor stability and high energy
consumption. The rst heterogeneous azeotropic distillation

Fig. 11. The mass balance of basic process and different improvements: (a) traditional process with energy recovery; (b) VHG fermentation; and (c) partial pervaporation.

80

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

column is designed to obtain high-purity ethanol product at the


column bottom while obtaining a minimum boiling ethanolewaterecyclohexane azeotrope at the top of the column. The
azeotrope obtained at the top is heterogeneous and the top vapor
stream is then condensed to form two liquid phases in the decanter
[200,205]. The organic phase containing mainly cyclohexane is
reuxed back to the heterogeneous azerotropic distillation column.
The aqueous phase is drawn from the decanter to be sent to the
entrainer recovery column where essentially pure water is obtained
as bottom stream, and at the top cyclohexane is removed to be
recycled to the rst column. The organic reux ow rate and the
recycle ow rate increase the ethanol concentration at the bottom
of the dehydration column, thus improving the separation performance but also increasing the heat duties.
5.5.2.2. Extractive distillation with ethyleneglycol. Extractive distillation is a partial vaporization process in the presence of a nonvolatile and high boiling point entrainer which does not form any
azeotropes with the original components of the azeotropic mixture.
The process has two columns: the extractive distillation column
and the entrainer recovery column. The entrainer is continuously
fed in one of the top stages of the extractive column while the
azeotropic feed is entered in a middle stage lower down the column. At the top of the extractive distillation column anhydrous
ethanol is obtained, and at the bottom stream a mixture of watereethyleneglycol is removed and sent to the second entrainer recovery column. In the recovery column at the top water is
withdrawn with some traces of ethanol, and at the bottom highpurity ethyleneglycol is recycled back to the extractive distillation
column [200]. The extractive distillation process with ethyleneglycol shows some important advantages over azeotropic distillations, with makeup entrainer and quantity of entrainer being
lower, thus reducing the diameter of the columns. Additionally, the
energy consumption in the columns is reduced. On the other hand,
the most important variables used to achieve the desired ethanol
concentration are the entrainer to feed molar ratio and the reux
ratio. The former has a limited effect on the energy consumption
compared with the reux ratio impact on the reboiler duty.
5.5.2.3. Molecular sieves to dewater the azeotropic ethanoleH2O
mixture. A molecular sieve is a material with very small pores of
precise and uniform size: these holes are small enough to block
large molecules while allowing small molecules to pass. Many
molecular sieves are used as desiccants, silicagel being a major
example. Uniform cavities are present, which selectively adsorb
molecules of a specic size. The diameter of a molecular sieve is
measured in Angstroms () or nanometres (nm 10 ). According
to the IUPAC notation, microporous materials have pore diameters
less than 2 nm (20 ) and macroporous materials have pore diameters of exceeding 50 nm (500 ). The mesoporous category
thus lies in the middle with pore diameters between 2 and 50 nm
(20e500 ) [206].
Molecular sieves are used as adsorbent for gases and liquids,
where e.g. a water molecule may be small enough to enter the
pores while larger molecules are not: water is trapped and retained
into the pores, thus acting as a desiccant. A molecular sieve can
adsorb water up to 22% of its own weight [207].
Synthetic molecular sieves have improved desiccant capacities
[208], and novel sieves can also be used in synthetic organic processes, frequently allowing isolation of desired products from
condensation reactions that are governed by generally unfavorable
equilibria [209]. They are also used in the ltration of air supplies
for breathing equipment [210], removing particulates and
compressor exhaust products from the breathing air supply.

Methods for regeneration of molecular sieves include a pressure


change (as in oxygen concentrators), heating and purging with a
carrier gas (as when used in ethanol dehydration), or heating under
high vacuum. Regeneration temperatures range from 448 K to
588 K depending on the molecular sieve type [209]. In contrast,
silicagel can be regenerated by heating at 393 K. Sufcient heat
must be applied to raise the temperature of the adsorbate, the
adsorbent and the vessel to vaporize the liquid and offset the heat
of wetting the molecular-sieve surface. The bed temperature is
sieve-specic towards regeneration. Bed temperatures in the
448e533 K range are usually employed for type 3 , whilst 4 , 5
and 13X sieves require temperatures in the 473e588 K range. After
regeneration, a cooling period is necessary to reduce the molecular
sieve temperature to within 288 K of the temperature of the stream
to be processed. This is most conveniently done by using the same
gas stream as for heating, but without heat input. Small quantities
of molecular sieves may be dried in the absence of a purge gas by
oven heating followed by slow cooling in a closed system, such as a
desiccator [209].
A 4e8-mesh sieve is normally used in gas phase applications,
while the 8e12-mesh type is common in liquid phase applications.
The powder forms of the 3 , 4 , 5 and 13X sieves are suitable for
specialized applications [209]. The characteristics of common molecular sieves are given in Table 20. 3 molecular sieves do not
adsorb chemicals whose molecular diameters exceed 3 . The
characteristics of these molecular sieves include fast adsorption
speed, frequent regeneration ability, good crushing resistance and
pollution resistance. The characteristics of common molecular
sieves are given in Table 20, with full details provided in Ref. [209].
In the traditional molecular sieve process, as illustrated in
Fig. 12, the azeotropic feed is superheated to >393 K, and sent to
one of the sieve columns, while the other column is being regenerated (release of adsorbed water).
The column regeneration is performed using superheated dry
ethanol, condensed and returned to the nal distillation column of
the bio-ethanol process as light ethanol.
Both columns adsorb and regenerate alternately, and ensure the
uninterrupted continuous production. The recovered heat in the
condensers can be used in the low temperature reboilers (distillation) or for preheating.
The mass and energy balances of the molecular sieve process for
a typical 200,000 tpa bio-ethanol plant [211] are given in Table 21.
5.5.2.4. Comparison. Within the pre-cited technologies of extractive distillation, ethyleneglycol is preferred above of the other
solvents such as benzene, glycerol or others. Azeotropic distillation
with hexane or other solvents has been abandoned because of high
investment and operation costs (30% above glycol extractive
distillation) [212].
Molecular sieves are now commonly used to produce pure
ethanol (>99.5%), despite the investment, the high tonnage of
molecular sieve needed, and the steam consumption being major
drawbacks.
5.5.3. Production of anhydrous ethanol using hydrophilic
membranes
Initial experiments were conducted according to the procedures
described by Dotremont et al. [213]. A Dow Filmtech membrane
(BW30XLE) was installed with a surface of 20 cm2. Details are given
in Kang et al. [163]. The data, as represented in Table 22, demonstrate that the permeate ux is signicantly inuenced by the
operating temperature, whilst the ethanol selectivity remains
nearly constant for T > 318 K.
The T-enhanced ux values are inherently due to the solutiondiffusion action of the membrane, facilitated by the enhanced

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

81

Table 20
Illustration of properties and characteristics of common Molecular Sieves.
Type [205]

Form

Bead or particle size

Pore diameter ()

Bulk density (kg/m3)

Moisture (%)

Eq. H2O capacity (%, theory)

Regeneration temp. (K)

3
5
13X

Bead
Bead
Bead

8e12 mesh
8e12 mesh
8e12 mesh

3
5
10

720e736
704
688

1.5
1.5
1.5

21
21.7
29.5

448e533
473e588
473e588

The adsorption heat is ~4187 kJ/kg H2O.

Fig. 12. Current molecular sieve dehydration of azeotropic ethanoleH2O mixtures.

molecular mobility at higher temperatures. To further increase the


ux, it is recommended to increase the temperature close to the
boiling point of the azeotropic mixture.
Applied as example to the Cofco plant [24,25], a mass balance
determines the required membrane surface area. According to the
data of Table 21, the nal ethanolewater distillation produces an
average 27,150 kg/h at 94.4 wt% ethanol, hence containing 1512 kg/
h of water and 25,638 kg/h of ethanol. Molecular sieves dewater the
ethanol to a purity of ~99.7wt% ethanol, i.e. removing 1314 kg/h of
water.
For this amount of water to be eliminated (1314 kg/h), the total
permeate ow (at 44.4 wt% ethanol) will be 2363 kg/h. The anhydrous ethanol retentate (>99.7 wt%) will be 24,786 kg/h. The
permeate, at 44.4 wt% of ethanol, needs to be recycled to the nal
distillation column, where it consists less than 2% of the feed mass
ow rate. For operation at 335 K, the required surface area of the
membrane is hence xed at 2363 (kg/h)/23.78 (kg/m2h) or 92 m2.

For the desiccation of e.g. 200,000 ton/year of ethanol, the investment of the molecular sieve unit exceeds 1 MV, including 160
ton of 3 molecular sieve. The annual steam consumption will be
96,000 ton, requiring ~7000 ton of fuel in the steam boiler. For the
membrane application, the investment including feeding pump,
vacuum pump and condenser, is estimated at ~0.3 MV [214,215].
Since no steam is required (the feed is delivered between 333 and
363 K after the nal watereethanol distillation column), only
~50 kWh needs to be accounted for as operating costs, representing
a signicant energy saving in comparison with the current molecular sieve unit.
It is clear that hydrophilic membranes can be a valid alternative,
without steam addition and with a minimum electrical power
required to overcome the transmembrane pressure and create the
permeate vacuum.
5.6. Additional developments
5.6.1. Cross-ow microltration of bio-ethanol fermentation broth

Table 21
Mass and energy balance of the molecular sieve process.
Mass balance
Feed water concentration (wt%)
Product water concentration (wt%)
Total feed ow rate (kg/h)
Feed ethanol ow rate (kg/h)
Product mass ow rate (kg/h)
Water absorption (kg/h)

5.6
0.8
27,160
25,638
25,845
1314

Absorption column
Absorption column
Regeneration column
Regeneration column
Total steam (kg/h)
(kg steam/kg ethanol)

Table 22
Permeate characteristics as function of operating time.
Temperature (K)

Energy balance
Evaporator heat

5.6.1.1. Introduction. Bio-ethanol fermentation broth consists of


mainly water and ethanol, together with solid residues of unreacted feedstock and additives (mostly yeast). Prior to further
processing (distillation), and to avoid fouling of heat exchangers
and distillation columns, the solid residues of the broth need to be

Product heat (kJ/kg)


Heat content of steam (kJ/kg)
Steam consumption (kJ/h)
Steam consumption (kg/h)
Steam consumption (kJ/h)
Steam consumption (kg/h)

868
2061
23,563,127
11,434
1,452,207
705
12,139
0.48

Time (min)
30

295
318
323
335

60

90

120

150

EF

PF

EF

PF

EF

PF

EF

PF

EF

PF

51
42
43
44

1.01
2.59
3.78
6.30

51
40
44
45

0.84
2.72
4.15
6.98

51
42
45
44

0.83
2.44
3.96
6.77

51
44
45
44

0.77
2.48
3.61
7.62

51
44
44
45

0.82
2.58
3.83
8.13

EF: Ethanol Fraction (%); PF: Permeate Flux (g/m2s).

82

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

Table 23
Characteristics of enhanced surface ltration medium.
Type
SFM
SFM
SFM
SFM
SFM
SFM

1.3
1.5
2.3
4.3
5.7
6.4

Fiber
diameter (mm)

Thickness
(mm)

Weight
(g/m2)

Porosity (%)

Air permeability
(l/dm2min)

1.25
1.57
2.28
4.31
5.70
6.35

0.1
0.12
0.12
0.12
0.28
0.12

400
400
400
400
900
400

50
60
60
60
60
60

2.47
6.6
7
27
29
57

Fig. 13. Size analysis of broth solids.

removed to as low a concentration as possible. The current mechanical separation (belt lter or centrifuge) can only
remove 10 mm particles representing about 90% of the total solids
content. The remaining 10% is usually recovered in the bottom
stream of the rst distillation column, and forms the stillage. To
avoid fouling and even eliminate the rst distillation column where
the ethanol fraction is only increased from 12% (feed) to 16% (top), a
better solids removal is required. The sintered metal ber (SFM)
eeces are highly efcient for microltration and the removal of
suspended solids largely exceeds 99% [216]. Properties of SFM lters are illustrated in Table 23, indicating the range of very ne bers used, the low thickness of the eece, and the fair to excellent
permeability.
The major applications of microltration are situated at different
levels, i.e. (i) in the harvesting and pre-treatment of algal biomass,
hence in item 1 in Fig. 2 [217]; (ii) to ltrate the broth before the

distillation (and pervaporation, in a hybrid system), hence after


item 3 in Fig. 2; and (iii) to perform a nal ltration step of the
wastewater treatment, hence integrated in item 6 of Fig. 2. The
benets of introducing membrane techniques are now recognized
[186,218e220], although extensive research is still needed to
scientically and economically justify their application.
5.6.1.2. Solids' characteristics of the fermentation broth. The broth
solid residue mainly consists of non-converted feedstock and yeast.
Typical particle size distributions are illustrated in Fig. 13 for a cornbased fermentation. Traditional particle separation techniques,
with a cut-size of 5e10 mm, can indeed only separate about 90% of
the suspended particles.
The residual 10% of nes is usually separated in the bottom
stream of the rst distillation column, contains mostly yeasts, and
forms the stillage. If microltration were to be used, it would need
to also separate the <5e10 mm particles.
5.6.1.3. Microltration results. Irrespective of the operating conditions, the membranes tested were nearly total barriers for the
suspended solids, with a >99% retention [221]. The permeate ux
was a strong function of TMP and values are given at a backpulse
every 20 s. Fluxes up to 22.5 m3/m2 h were measured at low concentrations (~1 g/L), but reduced to between 10 and 5 m3/m2 h at
yeast concentrations of 3e30 g/L respectively [221].
For a large-scale application, the concentration of solids in the
broth suspension after 90% removal by traditional mechanical
separation, will be about 35 T in ~2400 T of H2O/ethanol. Since the
treatment of the fermenter broth is spread over a period of about
60 h, the total broth ow of 2400 m3 requires a ltration capacity of
40 m3/h. The microltration feed concentration will be around 15 g/
L. At such a concentration, the microltration ux will be around
5 m3/m2h. Using the obtained ux values, an SFM of 8 m2 is
therefore required [221].
To enable the use of the experimental results for different suspensions and in scaling-up, a model was developed, based upon the
stagnant lm theory [221]. The model predicts the steady-state
permeate ux of cross-ow microltration with a fair accuracy
(error << 10%).
5.6.2. Novel distillation concepts
Considering that the partial pressure of ethanol is higher than
that of water, ethanol vapor can transfer preferentially through the
membrane pores, thus favoring the use of membrane distillation to
separate ethanol and water. The selectivity is determined by
ethanol volatility. There are very few studies about separating the
ethanolewater solution using membrane distillation [222,223],
although initial results also reveal that it can improve the ethanol
productivity due to the simultaneous removal of trace-inhibitors.

Table 24
Literature about advanced distillation.
Reference Objectives

Main results

[227]

Enhanced bioethanol dehydration in Dividing-Wall Columns (DWC)

[228]

Design and optimization of an ethanol dehydration process using


stochastic methods
Innovative single step bioethanol dehydration in an Extractie
Dividing -Wall Column

- Lower investment and operating costs, and reducing the equipment and
carbon footprint;
- Energy savings of 10e20%.
- Savings in total annual cost (32%) for a feed stream with 22 wt% ethanol

[229]

[230]
[231]

Breaking azeotropes in a Dividing-Wall Column


Efcient bioethanol dehydration in azeotropic and extractive DWC

- Only one column shell is used in combination (one condenser and two
reboilers);
- Savings on energy total investment and annual costs are about 17%.
- Energy savings of 10e20% and less equipment units.
- DWC as the key separation unit for large scale dehydration of pre-concentrated
bioethanol;
- Energy savings of 10e20%, and less equipment units.

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

Direct membrane distillation and air gap membrane distillation


coupled with fermentation are possible congurations for ethanol
production. Udriot et al. [224] integrated direct membrane distillation with fermentation for the separation of ethanol from the
culture medium. The results showed an 87% increase in ethanol
productivity from 0.99 to 1.85 g L1 h1. Lewandowicz et al. [225]
reported that the ethanol production increased by 15.5% due to
facilitation of the continuous process, more complete fermentation
of sugars, lowering the osmotic pressure in the fermentation broth,
decreasing the glycerol synthesis level and increasing yeast cells
number and viability. In addition to improve the ethanol productivity, membrane performance is another research focus. Studies
indicate that the temperature has an important inuence on
membrane ux [226].
Current studies have been conned to lab-scale as illustrated in
Table 24, and larger scale performance needs to be further
investigated.
Moreover the membrane selectivity is relative low, ranging from
1 to 7 only, thus making the development of high performance
membrane material an important focus of research in membrane
distillation. Pilot-scale applications are certainly required to
conrm the lab-scale ndings.
5.6.3. The improved bio-ethanol production plant
From the results reported in the paper, it is clear that the current
ow sheet and mass balance, as presented in Figs. 5 and 7, can be
substantially modied to take advantages of recent improvements.
Fig. 14 summarizes the authors' views on an improved production
process, integrating different unit operations.
The batch fermentation will be performed in VHG-mode (1),
with ethanol concentration between 12 and 15 wt%. To reduce the
stress on the bacteria in the fermenter, a fraction of the ethanol will
be continuously removed during fermentation by a pervaporation
membrane (3). Ethanol-rich permeate will be fed to the 2nd
distillation stage, whist the ethanol-lean fraction will be recycled to
the fermenter. To prevent fouling of the pervaporation membrane,
suspended solids of the fermentation broth will be removed by
microltration (2) and recycled to the fermenter. Suspended solids
of the fermentation broth will be removed by a combined mechanical (>10 mm) and microltration (<10 mm) step, respectively
indicated as (4) and (5).
Solids' residues will be as currently produced, both into DDGS
(6), and as wastewater to be anaerobically digested (7), with its
biogas feeding a CHP unit to provide heat and electricity (8). Due to
the use of a microltration, the rst Cofco-like distillation column

83

can be eliminated, since the stillage is removed by microltration. A


2-stage distillation (9.1 and 9.2) will be sufcient to produce the
azeotropic ethanolewater mixture. Anhydrous ethanol can then be
produced by hydrophilic membranes (10) and stored (11). As will
be shown in the following tentative economic assessment, the
process sequence as described above, offers both investment benets (e.g. less fermenter volume, less distillation needs,) and
reduced production costs.
6. Preliminary economic assessment
The full production cost of bio-ethanol at Cofco in 2013 is reported at 820 V/ton bio-ethanol. Based upon the predictions of the
previous sections, a tentative cost calculation per ton bio-ethanol
can be made, for a plant depreciation of 25 years, and other costs
calculated on the basis of common ratios or literature data [232].
The efciency of the steam boiler is estimated at 85%. For sake of
comparison, the use of corn of sugarcane is also included.
The required raw materials cassava, corn and sugarcane are
considered at 6.6, 3.2 and 13.5 ton/ton bio-ethanol respectively,
with current costs (delivered at the plant) of respectively 38.6
[232], 220 [125] and 34.5 [233] V/ton of raw material. Through a
similar approach, and based upon the estimated investment costs
of the membrane modules, the impact of both hybrid fermentation
and substitution of molecular sieves by membrane dehydration can
be calculated. The life span of the membrane modules is estimated
at 10 years, with a replacement of the membranes every 4 years.
Results are illustrated in Fig. 15.
Since the calculated production cost for cassava-based bioethanol (830 V/ton) is comparable with the Cofco-cited 820 V/ton,
the proposed calculation procedure appears fair and acceptable.
It is therefore evident that the production cost by both applying
a VHG fermentation, and substituting molecular sieves by a
membrane process, can reduce the total production cost by ~90 V/
ton bio-ethanol, i.e. about 11%.
Further research of the different alternatives, preferably on pilot
scale, appears economically and technically justied. Cellulosic
biomass could be a valid feed stock provided its cost of supply and
required pretreatment can compete with the equivalent costs for
cassava.
7. Conclusions
Bio-ethanol can be produced from different kinds of renewable
feedstock. Whereas the rst generation of processes (saccharide-

Fig. 14. Improved rst generation bio-ethanol production.

84

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

This work was supported by the National Basic Research Program of China (973 program) (2013CB733600, 2012CB725200), the
National Nature Science Foundation of China (21436002,
21390202), the National High-Tech R&D Program of China (863
Program) (2012AA021404, 2014AA021903, 2014AA021904), and
Key Projects in the National Science & Technology Pillar Program
during the 12th Five-year Plan Period (2011BAD22B04).
References

Fig. 15. Cost calculations for a 200,000 tpa plant (with internal energy recovery), and
economic impact of Molecular Sieve (MS) substitution and using VHG ( : Cost of feedstock; Pre-treatment of feedstock; : Treatment of residues; : Depreciation, Plant
Overherads and insurance, etc; : Labour; : Cost of electricity; : Cost of heavy fuel oil).

based) have been well documented and are largely applied, the
second and third generation of bio-ethanol processes (cellulose- or
algae-based) need further research and development since bioethanol yields are still too low to be economically viable. Applications of bio-ethanol mostly target the combined use with petrol.
The conversion pathways for the production of bio-ethanol from
disaccharides, from starches, and from lignocellulosic biomass
were analyzed. The yield of the fermentation xes the total energy
needed during the downstream separation processes, mostly by
distillation. Mass and energy balances of the common processing
routes were presented. To compare eld data with common engineering design simulations, the software Aspen Plus V8.2 was used.
Field and simulation results showed an excellent agreement.
Without energy recovery, 2.5 kg steam/kg ethanol are required.
Possible energy improvements through 5 modications were
examined and demonstrated that an internal energy re-use, the
application of VHG, or the application of a hybrid operation mode
(pervaporation) could reduce the steam use per kg of bio-ethanol to
1.18, 1.17, 0.94 and 1.13 kg steam/kg bio-ethanol for options of
energy-pinch, VHG (15%), VHG (19%) and pervaporation, respectively. A further energy saving can be achieved in the ethanol
dewatering to fuel grade. Whilst molecular sieves are now
commonly used, the use of hydrophilic membranes was highlighted and conrmed by literature and experimental results,
yielding a signicant energy saving in comparison with the molecular sieve units. Additional technologies to improve the operation of the ethanolewater separation, include microltration of the
broth before the distillation (and pervaporation, in a hybrid system), and membrane distillation. Whereas the benets of introducing these techniques are recognized, extensive research is still
needed to scientically and economically justify their application. A
tentative economic assessment demonstrated that the application
of process improvements can lead to production cost savings of
~11%. Further research of the different alternatives, preferably on
pilot or large scale, appears economically and technically justied.
Acknowledgments
The assistance of Jannis Huybrechts (MEng) during experimental research and Aspen Plus simulation is greatly appreciated.

[1] David RL. CRC handbook of chemistry and physics. 81st ed. Maryland: CRC
Press; 2000.
[2] Windholz M. The Merck index: an encyclopedia of chemicals and drugs. 9th
ed. Rahway New Jersey: Merck & Co., Inc.; 1976.
[3] Costigan MJ, Hodges LJ, Marsh KN, Stokes RH, Tuxford CW. The isothermal
displacement calorimeter: design modications for measuring exothermic
enthalpies of mixing. Aust J Chem 1980;33:2103e19.
[4] Lide DR, Kehiaian HV. CRC handbook of thermophysical and thermochemical
data. Boca Raton: CRC Press; 1994.
[5] Pemberton RC, Mash CJ. Thermodynamic properties of aqueous nonelectrolyte mixtures II. Vapour pressures and excess Gibbs energies for
water ethanol at 303.15 to 363.15 K determined by an accurate static
method. J Chem Thermodyn 1978;10:867e88.
[6] Flash points of ethanol-based water solutions. Retrieved from: http://www.
engineeringtoolbox.com/ethanol-water-d_989.html [04.06.14].
[7] Wolke RL. Combustible combination. Retrieved from: http://www.
washingtonpost.com/wp-dyn/content/article/2006/07/04/
AR2006070400283.html [04.06.14].
[8] Roberts JD, Caserio MC. Basic principles of organic chemistry. 2nd ed. Menlo
Park, CA: W. A. Benjamin, Inc.; 1977.
[9] Naim K, Zdravko D, Adalbert F, Hermann S, Stephanie BM, Otto G, et al.
Ethanol, Ullmann's encyclopedia of industrial chemistry. Weinheim: WileyVCH Verlag GmbH & Co., KGaA; 2011.
[10] Raymond EK, Donald FO, Jacqueline IK, Mary HG. Ethanol. Encyclopedia of
chemical technology. New York: Wiley; 1991.
[11] Akponah E, Akpomie OO, Ubogu M. Bio-ethanol production from cassava
efuent using Zymomonas mobilis and Saccharomyces cerevisiae isolated from
raa palm (Elaesis guineesi) SAP. Eur J Exp Biol 2013;3:247e53.
[12] Fisher K, Mckaskle K. Illinois state geological survey evaluation of CO2 capture options from ethanol plants. Trimeric Corporation; 2013.
[13] Bradshaw MJ. Global energy dilemmas: a geographical perspective. Geogra J
2010;176:275e90.
[14] Balat M. Production of bioethanol from lignocellulosic materials via the
biochemical pathway: a review. Energy Convers Manage 2011;52:858e75.
[15] Sanchez OJ, Cardona CA. Trends in biotechnological production of fuel
ethanol from different feedstocks. Bioresour Technol 2008;99:5270e95.
[16] Renewable Fuels Association. World fuel ethanol production. Retrieved
from: http://ethanolrfa.org/pages/World-Fuel-Ethanol-Production; 2013
[04.06.14].
[17] Marina OSD, Tassia LJ, Carlos EVR, Rubens MF, Antonio B. Evaluation of
process congurations for second generation integrated with rst generation
bioethanol production from sugarcane. Fuel Process Technol 2013;109:
84e9.
[18] Liu SH, Liu XF. Technological development of non-grain based fuel ethanol
production. Liquor Making 2010;37:9e11.
[19] Akihiko K, Akihiko K, Mitsuyoshi U, Yoshinori M, Pilanee V, Warunee T.
Production of ethanol from cassava pulp via fermentation with a surface
engineered yeast strain displaying glucoamylase. Renew Energy 2009;34:
1354e8.
[20] Choi GW, Um HJ, Kim Y, Kang HW, Kim M, Chung BW. Isolation and characterization of two soil derived yeasts for bioethanol production on Cassava
starch. Biomass Bioenergy 2010;34:1223e31.
[21] Shanavas S, Padmaja G, Moorthy SN, Sajeev MS, Sheriff JT. Process optimization for bioethanol production from cassava starch using novel ecofriendly enzymes. Biomass Bioenergy 2011;35:901e9.
[22] Rattanachomsri U, Tanapongpipat S, Eurwilaichitr L, Champreda V. Simultaneous non-thermal saccharication of cassava pulp by multi-enzyme activity and ethanol fermentation by Candida tropicalis. J Biosci Bioeng
2009;107:488e93.
[23] Luo G, Xie L, Zou ZH, Zhou Q. Effect of initial pH on thermophilic fermentative hydrogen production from cassava ethanol wastewater. J Tongji Univ
2010;38:1036e40.
[24] Zhang HL, Baeyens J, Tan TW. Mixing phenomena in a large-scale fermenter
of starch to bio-ethanol. Energy 2012;48:380e91.
[25] Zhang HL, Baeyens J, Tan TW. The bubble-induced mixing in starch-toethanol fermenters. Chem Eng Res Des 2012;90:2122e8.
[26] Zhang NN, Steven Green V, Ge XM, Savary BJ, Xu JF. Ethanol fermentation of
energy beets by self-occulating and non-occulating yeasts. Bioresour
Technol 2014;155:189e97.
[27] Kim SH, Kim CH. Evaluation of whole Jerusalem artichoke (Helianthus
tuberosus L.) for consolidated bioprocessing ethanol production. Renew Energy 2014;65:83e91.

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88


[28] Eshtiaghi NM, Kuldiloke J, Yoswathana N, Ebadi AG. Application of ultrasound and technical enzymes during bioethanol production from fresh
cassava root. J Agric Environ 2012;10:905e9.
[29] Popov S, Rankovi
c J, Dodi
c J, Dodi
c S, Joki
c A. Bioethanol production from raw
juice as intermediate of sugar beet processing: a response surface methodology approach. Food Technol Biotechnol 2010;48:376e83.
[30] Nikoli
c S, Mojovi
c L, Rakin M, Vukasinovi
c-Sekuli
c M, Pejin D, Pejin J.
Improvement of bioethanol production from corn by ultrasound and microwave pretreatments. Chem Eng Trans 2010;21:1327e32.
itog
lu M. Production of bioethanol by immobilized Saccharomyces
[31] Inal M, Yig
Cerevisiae onto modied sodium alginate gel. J Chem Technol Biotechnol
2011;86:1548e54.
[32] Fujiwara E, Onoa E, Yamakawab CK, Ienczakb JL, Rossell CEV, Suzukia CK.
Real-time monitoring of fermentation process applied to sugarcane bioethanol production. Proc SPIE 2012;8421:161e4.
[33] McHenry MP, Doepel D, De Boer K. Rural African renewable fuels and fridges:
cassava waste for bioethanol, with stillage mixed with manure for biogas
digestion for application with dual-fuel absorption refrigeration. Biofuels
Bioprod Bioref 2014;8:103e13.
[34] Dewan A, Li Z, Han BB, Karim MN. Saccharication and fermentation of waste
sweet potato for bioethanol production. J Food Process Eng 2013;36:739e47.
[35] Nikoli
c S, Mojovi
c L, Rakin M, Pejin D, Pejin J. Utilization of microwave and
ultrasound pretreatments in the production of bioethanol from corn. Clean
Technol Environ Policy 2011;13:587e94.
[36] Quarterman JC, Jin YS, Price ND, Kim P. Optimal strain design of Saccharomyces cerevisiae for bioethanol production. In: AIChE annual meeting: food,
Pharmaceutical and Bioengineering Division; 2011.
[37] Rankovi
c J, Dodi
c J, Dodi
c S, Popov S. Bioethanol production from intermediate products of sugar beet processing with different types of Saccharomyces cerevisiae. Chem Ind Chem Eng Q 2009;15:13e6.
[38] Cui WH, Xu JL, Cheng JJ. Duckweed starch accumulation for bioethanol
production. Louisville: The American Society of Agricultural and Biological
Engineers; 2011.
[39] Taherzadeh MJ, Karimi K. Acid-based hydrolysis processes for ethanol from
lignocellulosic materials: a review. BioResources 2007;2:472e99.
[40] Gonela V, Zhang J. Design of the optimal industrial symbiosis system to
improve bioethanol production. J Clean Prod 2014;64:513e34.
[41] Sindhu R, Kuttiraja M, Binod P, Sukumaran RK, Pandey A. Bioethanol production from dilute acid pretreated Indian bamboovariety (Dendrocalamus
sp.) by separate hydrolysis and fermentation. Ind Crop Prod 2014;52:
169e76.
[42] Marx S, Ndaba B, Chiyanzu I, Schabort C. Fuel ethanol production from sweet
sorghum bagasse using microwave irradiation. Biomass Bioenergy 2014;65:
145e50.
[43] Ofori-Boateng C, Lee KT. Ultrasonic-assisted simultaneous saccharication
and fermentation of pretreated oil palm fronds for sustainable bioethanol
production. Fuel 2014;119:285e91.
[44] Li JH, Li SZ, Han B, Yu MH, Li GM, Jiang Y. A novel cost-effective technology to
convert sucrose and homocelluloses in sweet sorghum stalks into ethanol.
Biotechnol Biofuels 2013;6:174. http://dx.doi.org/10.1186/1754-6834-6-174.
[45] Zhang J, Osmani A, Awudu I, Gonela V. An integrated optimization model
for switchgrass-based bioethanol supply chain. Appl Energy 2013;102:
1205e17.
[46] Shaheen M, Choi M, Ang W, Zhao YP, Xing J, Yang R, et al. Application of lowintensity pulsed ultrasound to increase bio-ethanol production. Renew Energy 2013;57:462e8.
z F, Ballesteros I, Ballesteros M, et al.
[47] Manzanares P, Negro MJ, Oliva JM, Sae
Different process congurations for bioethanol production from pretreated
olive pruning biomass. J Chem Technol Biotechnol 2011;86:881e7.
[48] Takagi T, Uchida M, Matsushima R, Ishida M, Urano N. Efcient bioethanol
production from water hyacinth Eichhornia crassipes by both preparation of
the saccharied solution and selection of fermenting yeasts. Fish Sci
2012;78:905e10.
 pez-Abelairas M, Lu-Chau TA, Lema JM. Enhanced saccharication of bio[49] Lo
logically pretreated wheat straw for ethanol production. Appl Biochem
Biotechnol 2013;169:1147e59.
[50] Onuki S, Koziel JA, Van Leeuwen JH, Jenks WS, Grewell D, Cai L. Ethanol
production, purication, and analysis technology: a review. In: Agricultural
and biosystems engineering conference papers, posters and presentations;
2008.
 pez-Abelairas M, Lu-Chau TA, Lema JM. Fermentation of biologically pre[51] Lo
treated wheat straw for ethanol production: comparison of fermentative
microorganisms and process congurations. Appl Biochem Biotechnol
2013;170:1838e52.
[52] Phillips RB, Jameel H, Chang HM. Integration of pulp and paper technology
with bioethanol production. Biotechnol Biofuels 2013;6:13. http://
dx.doi.org/10.1186/1754-6834-6-13.
[53] Minu K, Kurian Jiby K, Kishore VVN. Isolation and purication of lignin and
silica from the black liquor generated during the production of bioethanol
from rice straw. Biomass Bioenergy 2012;39:210e7.
[54] Macedo AL, Santos RS, Pantoja L, Santos AS. Pequi cake composition, hydrolysis and fermentation to bioethanol. Braz J Chem Eng 2011;28:9e15.
s-Pejo
 E, Ballesteros M, Negro MJ. Pretreatment technologies
[55] Alvira P, Toma
for an efcient bioethanol production process based on enzymatic hydrolysis: a review. Bioresour Technol 2010;101:4851e61.

85

pez-Go
 mez A. Production of bioethanol by fermentation
[56] Boluda-Aguilar M, Lo
of lemon (Citrus limon L.) peel wastes pretreated with steam explosion. Ind
Crop Prod 2013;41:188e97.
[57] Velmurugan R, Muthukumar K. Sono-assisted enzymatic saccharication of
sugarcane bagasse for bioethanol production. Biochem Eng J 2012;63:1e9.
[58] Cheng JJ, Timilsina GR. Status and barriers of advanced biofuel technologies:
a review. Renew Energy 2011;36:3541e9.
[59] De Souza AC, Carvalho FP, Silva e Batista CF, Schwan RF, Dias DR. Sugarcane
bagasse hydrolysis using yeast cellulolytic enzymes. J Microbiol Biotechnol
2013;23:1403e12.
[60] Keshwani DR, Cheng JJ. Switchgrass for bioethanol and other value-added
applications: a review. Bioresour Technol 2009;100:1515e23.
[61] Lu J, Li XZ, Yang RF, Zhao J, Qu YB. Tween 40 pretreatment of unwashed
water-insoluble solids of reed straw and corn stover pretreated with liquid
hot water to obtain high concentrations of bioethanol. Biotechnol Biofuels
2013;6:159.
[62] Cheng YS, Zheng Y, Labavitch JM, VanderGheynst JS. Virus infection of
Chlorella variabilis and enzymatic saccharication of algal biomass for bioethanol production. Bioresour Technol 2013;137:326e31.
[63] Prasetyo J, Park EY. Waste paper sludge as a potential biomass for bioethanol production. Korean J Chem Eng 2013;30:253e61.
nez-Gutie
rrez A, El-Halwagi MM. Assessment of com[64] Conde-Meja C, Jime
binations between pretreatment and conversion congurations for bioethanol production. Sustain Chem Eng 2013;1:956e65.
[65] Hong SH, Lee JT, Lee S, Wi SG, Cho EJ, Singh S. Improved enzymatic hydrolysis of wheat straw by combined use of gamma ray and dilute acid for
bioethanol production. Radiat Phys Chem 2014;94:231e5.
[66] Kuhad RC, Gupta R, Khasa YP, Singh A. Bioethanol production from Lantana
camara (red sage): pretreatment, saccharication and fermentation. Bioresour Technol 2010;101:8348e54.
[67] Ojeda K, S
anchez E, Kafarov V. Sustainable ethanol production from
lignocellulosic biomass-application of exergy analysis. Energy 2011;36:
2119e28.
[68] Taherzadeh MJ, Karimi K. Enzymatic-based hydrolysis processes for ethanol
from lignocellulosic materials: a review. BioResources 2007;2:707e38.
[69] Mousdale DM. Biofuels: biotechnology, chemistry and sustainable development. Chapter 4.1: the Iogen Corporation process as a template and paradigm. New York: CRC Press, Taylor & Francis Group; 2008.
[70] Bauer F, Hulteberg C. Is there a future in glycerol as a feedstock in the production of biofuels and biochemicals? Biofuel Bioprod Bioref 2013;7:43e51.
[71] Kumar S, Gupta R, Kumar G, Sahoo D, Kuhad RC. Bioethanol production from
Gracilaria verrucosa, ared alga, in a biorenery approach. Bioresour Technol
2013;135:150e6.
[72] Daroch M, Geng S, Wang GY. Recent advances in liquid biofuel production
from algal feedstocks. Appl Energy 2013;102:1371e81.
[73] Maity JP, Bundschuh J, Chen CY, Bhattacharya P. Microalgae for third generation biofuel production, mitigation of greenhouse gas emissions and
wastewater treatment: present and future perspectives-a mini review. Energy 2014. http://dx.doi.org/10.1016/j.energy.2014.04.003.
[74] Guo H, Daroch M, Liu L, Qiu GY, Geng S, Wang GY. Biochemical features and
bioethanol production of microalgae from coastal waters of Pearl River Delta.
Bioresour Technol 2013;127:422e8.
[75] Zhu LD, Hiltunen E, Antila E, Zhong JJ, Yuan ZH, Wang ZM. Microalgal biofuels: exible bioenergies for sustainable development. Renew Sust Energy
Rev 2014;30:1035e46.
[76] Harun R, Yip JWS, Thiruvenkadam S, Ghani WAWAK, Cherrington T,
Danquah MK. Algal biomass conversion to bioethanol-a step-by-step
assessment. Biotechnol J 2014;9:73e86.
[77] Harun R, Danquah MK, Forde GM. Microalgal biomass as a fermentation
feedstock for bioethanol production. J Chem Technol Biotechnol 2010;85:
199e203.
[78] Yoza BA, Masutani EM. The analysis of macroalgae biomass found around
Hawaii for bioethanol production. Environ Technol 2013;34:1859e67.
[79] Chirapart A, Praiboon J, Puangsombat P, Pattanapon C, Nunraksa N. Chemical
composition and ethanol production potential of Thai seaweed species.
J Appl Phycol 2014;26:979e86.
[80] Khambhaty Y, Upadhyay D, Kriplani Y, Joshi N, Mody K, Gandhi MR. Bioethanol from macroalgal biomass: utilization of Marine Yeast for production
of the same. Bioenergy Res 2013;6:188e95.
[81] Singh DP, Trivedi RK. Acid and alkaline pretreatment of lignocellulosic
biomass to produce ethanol as biofuel. Int J Chem Tech Res 2013;5:727e34.
[82] Kong WB, Yang H, Cao YT, Song H, Hua SF, Xia CG. Effect of glycerol and
glucose on the enbhancement of biomass, lipid and soluble carbohydrate
production by Chlorella vulgaris in mixotrophic culture. Food Technol Biotechnol 2013;51:62e9.
[83] Wang X, Liu XH, Wang GY. Two-stage hydrolysis of invasive algal feedstock
for ethanol fermentation. J Interg Plant Biol 2011;53:246e52.
[84] Park JH, Kim SH, Park HD, Kim JS, Yoon JJ. Simultaneous utilization of
galactose and glucose by Saccharomyces cerevisiae mutant strain for ethanol
production. Renew Energy 2014;65:213e8.
[85] Brief history of rockets. Retrieved from: http://quest.arc.nasa.gov/space/
teachers/rockets/history.html [04.06.14].
[86] The brain from top to bottom-alcohol (intermediate, molecular). Retrieved
from:
http://thebrain.mcgill.ca/ash/i/i_03/i_03_m/i_03_m_par/i_03_m_
par_alcool.html [04.06.14].

86

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

[87] RACQ. Ethanol blends and octane number. Retrieved from: http://www.racq.
com.au/motoring/cars/car_advice/car_fact_sheets/ethanol_blends_and_
octane_number [09.06.14].
[88] Hansen AC, Lyne PWL. Ethanol-diesel fuel blends-a review. Bioresour
Technol 2005;96:277e85.
[89] Lei J, Bi Y, Shen L. Performance and emission characteristic of diesel engine
fueled with ethanol-diesel blends in different altitude regions. J Biomed
Biotechnol 2011;2011:417421.
[90] Behdad S, Talal Y, Hossein HAA, Barat G. Experimental investigation of the
tractor engine performance using diesohol fuel. Appl Energy 2014;114:874e9.
[91] Francesco C, Paolo S, Bianca MV. Optical characterization of bio-ethanol injection and combustion in a small DISI engine for two wheels vehicles. Fuel
2013;106:651e66.
[92] Istv
an B. Predicting the temperature dependent density of biodiesel-dieselbioethanol blends. Fuel 2013;109:563e74.
[93] Su HP, Seung HY, Chang SL. HC and CO emissions reduction by early injection
strategy in a bioethanol blended diesel-fueled engine with a narrow angle
injection system. Appl Energy 2013;107:81e8.
[94] Tsolakisa A, Megaritisb A, Yap D. Application of exhaust gas fuel reforming in
diesel and homogeneous charge compression ignition (HCCI) engines fuelled
with biofuels. Energy 2008;33:462e70.
[95] Mohammad A, Barat G, Iman B. Technical comparison of a CHP using various
blends of gasohol in an IC engine. Renew Energy 2008;33:1469e74.
[96] Constantin P, Niculae N, Alexandru C. Improvement of the automotive spark
ignition engine performance by supercharging and the bioethanol use. In:
SAE-China, FISITA, editors. Proceedings of the FISITA 2012 world automotive
congress. Lecture notes in electrical engineering, vol. 191; 2012. http://
dx.doi.org/10.1007/978-3-642-33777-2-3.
[97] Kowalewicz A. Application of biofuels to compression ignition engines. New
and renewable energy technologies for sustainable development. World
Scientic; 2007. p. 105e16.
[98] Tao J, Yu SR, Wu TX. Review of China's bioethanol development and a case
study of fuel supply, demand and distribution of bioethanol expansion by
national application of E10. Biomass Bioenergy 2011;35:3810e29.
[99] Gupta KK, Rehman A, Sarviya RM. Bio-fuels for the gas turbine: a review.
Renew Sust Energ Rev 2010;14:2946e55.
[100] Chiara G, Carlo B, Pierpaolo N. Application of bioethanol/RME/diesel blend in
a Euro5 automotive diesel engine: potentiality of closed loop combustion
control technology. Appl Energy 2013;102:13e23.
[101] Edgard G, Arnaud D. Ethanol fuel from biomass: a Review. J Sci Ind Res
2005;64:809e21.
[102] Surveyer A. Alternative fuels for gas turbines: a consequential LCA for
 De Montre
al; 2012.
electricity generation in 2020. Universite
 C, Dagaut P. Experimental
[103] May-Carle JB, Pidol L, Nicolle A, Anderlohr J, Togbe
and numerical study of FT/biodiesel/bioethanol surrogate fuel oxidation in
jet-stirred reactor. Combust Sci Technol 2012;184:901e15.
[104] Jones TTM. The clean fuels report: a quantitative comparison of motor fuels,
related pollution and technologies. 2008.
 k J, Eller Z, Marsi G, Nagy G. Increasing the stability of bioethanol/gas
[105] Hancso
oil emulsions by a new emulsion additive. Petrol Coal 2011;53:106e14.
[106] Mustafa K, Yakup S, Tolga T, Hseyin SY. The effects of ethanol-unleaded
gasoline blends on engine performance and exhaust emissions in a sparkignition engine. Renew Energy 2009;34:2101e6.
re F, Sanger R, His S, Boyer C, Doshi K, Xu JJ. Production of hydrogen
[107] Giroudie
from bio-ethanol. In: 16th World hydrogen energy conference; 2006.
~ izares P, Rodrigo MA, Linares JJ. Testing a vapour-fed PBI-based
[108] Lobato J, Can
direct ethanol fuel cell. Fuel Cells 2009;5:597e604.
[109] Nguyen THT, Yasunori K, Hirokazu S, Masaru N, Masahiko H. Techno-economic and environmental assessment of bioethanol-based chemical process:
a case study on ethyl acetate. Environ Prog Sust Energy 2011;30:675e84.
[110] Chae HJ, Kim TW, Moon YK, Kim HK, Jeong KE, Kim CU, et al. Butadiene
production from bioethanol and acetaldehyde over tantalum oxidesupported ordered mesoporous silica catalysts. Appl Catal B: Environ
2014;150e151:596e604.
[111] Elka K, Sergey S, Giorgio N, Ursula B, Heike E. Catalytic performance of
CoAlZn and NiAlZn mixed oxides in hydrogen production by bio-ethanol
partial oxidation. Int J Hydrogen Energy 2014;39:209e20.
[112] Grisel C, Edgar A. Bioethanol and diesel/bioethanol blends emissions
abatement. Fuel 2008;87:3537e42.
[113] Wukovits W, Pfeffer M, Liebmann B, Friedl A. Integration of the bio-ethanol
process in a network of facilities for heat and power production from
renewable sources using process simulation. In: 17th European symposium
on computer aided process engineering; 2007.
[114] Turner D, Xu HM, Cracknell RF, Natarajan V, Chen XD. Combustion performance of bio-ethanol at various blend ratios in a gasoline direct injection
engine. Fuel 2011;90:1999e2006.
[115] Yoon SH, Lee CS. Effect of undiluted bioethanol on combustion and emissions
reduction in a SI engine at various charge air conditions. Fuel 2012;97:
887e90.
[116] Bhupendra SC, Naveen K, Shyam SP, Yong DJ. Experimental studies on
fumigation of ethanol in a small capacity diesel engine. Energy 2011;36:
1030e8.
pez-Aparicio S, Hak C. Evaluation of the use of bioethanol fuelled buses
[117] Lo
based on ambient air pollution screening and on-road measurements. Sci
Total Environ 2013;452e453:40e9.

[118] Saxena S, Schneider S, Aceves S, Dibble R. Wet ethanol in HCCI engines with
exhaust heat recovery to improve the energy balance of ethanol fuels. Appl
Energy 2012;98:448e57.
[119] Masahiro I, Shohei Y, Hironobu U, Daisaku S. Remarkable improvement of
NOxePM trade-off in a diesel engine by means of bioethanol and EGR. Energy
2010;35:4572e81.
[120] Hansdah D, Murugan S, Das LM. Experimental studies on a DI diesel engine
fueled with bioethanol-diesel emulsions. Alex Eng J 2013;52:267e76.
[121] Kitamaki Y, Zhu YB, Inagaki S, Matsuo M, Taniguchi S, Numata M, et al.
Determination of sulfur in bioethanol certied reference material. J Jpn
Petrol Inst 2013;56:171e5.
[122] Biewinga EE, Van der Bijl G. Sustainability of energy crops. A methodology
developed and applied. Report no. 234. Utrecht, The Netherlands: Centre for
Agriculture and Environment (CLM); 1996.
[123] Rasul G, Thapa GB. Sustainability of ecological and conventional agricultural
systems in Bangladesh: an assessment based on environmental, economic
and social perspectives. Agric Syst 2004;79:327e51.
[124] Limayem A, Ricke SC. Lignocellulosic biomass for bioethanol production:
current perspectives, potential issues and future prospects. Prog Energ
Combust Sci 2012;38:449e67.
[125] Jansson C, Westerbergh A, Zhang J, Hu X, Sun C. Cassava, a potential biofuel
crop in the People's Republic of China. Appl Energy 2009;86:595e9.
[126] Indexmundi.
Retrieved
from:
www.indexmundi.com/commodities/?
commoditycorn; 2014 [14.05.14].
[127] Wyman CE. Ethanol fuel. In: Encyclopaedia of energy. New York: Elsevier
Inc.; 2004.
[128] Raven PH, Evert RF, Eichhorn SE. Biology of plants. 6th ed. New York:
Freeman and Company/Worth Publishers; 1999.
[129] Shleser R. Ethanol production in Hawaii. USA: State of Hawaii; 1994.
[130] Sun Y, Cheng J. Hydrolysis of lignocellulosic materials for ethanol production: a review. Bioresour Technol 2002;83:1e11.
[131] Morjanoff PJ, Gray PP. Optimization of steam explosion as method for
increasing susceptibility of sugarcane bagasse to enzymatic saccharication.
Biotechnol Bioeng 1987;29:733e41.
[132] Wooley RR, Sheehan J, Ibsen K. Lignocellulosic biomass to ethanol process
design and economics utilizing co-current dilute acid prehydrolysis and
enzymatic hydrolysis: current and futuristic scenarios. NREL Report, NREL/TP
580-26157. 1999.
[133] Hendriks ATWM, Zeeman G. Pretreatments to enhance the digestibility of
lignocellulosic biomass. Bioresour Technol 2009;100:10e8.
[134] Aden A, Ruth M, Ibsen K, Jechura J. Lignocellulosic biomass to ethanol process
design and economics utilizing co-current dilute acid prehydrolysis and
enzymatic hydrolysis for corn stover. NREL Report, NREL/TP 510-32438. 2002.
[135] The DOE. Bioethanol pilot plant [DOE leaet GO-10200-1114]. 2000.
[136] Lindstedt J. Alcohol production from lignicellulosic feedstock. FVS Fachtag
2003:228e37.
[137] Dien BS, Cotta MA, Jefferies TW. Bacteria engineered for fuel ethanol production: current status. Appl Microbiol Biotechnol 2003;63:258e66.
[138] Hahn-H
agerdal B, Karhumaa HBK, Fonseca C, Spencer-Martins I, GorwaGrauslund MF. Toward industrial pentose-fermenting yeast strains. Appl
Microbiol Biotechnol 2007;74:937e53.
[139] Dashtban M, Schraft H, Qin W. Fungal bioconversion of lignocellulosic residues; opportunities & perspectives. Int J Biol Sci 2010;5:578e95.
[140] Chen YCB. Initial investigation of xylose fermentation for lignocellulosic
bioethanol production. Thesis. AL: Auburn University; 2009.
[141] Ho NWY, Chen ZD, Brainard AP. Genetically engineered Saccharomyces yeast
capable of effective cofermentation of glucose and xylose. Appl Environ
Microbiol 1998;64:1852e9.
[142] Almaida JRM, Modig T, Roder A, Liden G, Gorwagrauslund MF. Pichia stiplis
xylose reductase helps detoxifying lignocellulosic hydrolysate by reducing 5hydroxymethyl-furfural (HMF). Biotechnol Biofuels 2008;11:1e12.
[143] Deanda K, Zhang M, Eddy C, Picataggio S. Development of an arabinosefermenting Zymomonas mobilis strain by metabolic pathway engineering.
Appl Environ Microbiol 1996;62:4465e70.
[144] Weber C, Farwick A, Benisch F, Brat D, Dietz H, Subtil T, et al. Trends and
challenges in the microbial production of lignocellulosic bioalcohol fuels.
Appl Microbiol Biotechnol 2010;87:1303e15.
[145] Nagodawithana TW, Steinkraus KH. Inuence of the rate of ethanol production and accumulation on the viability of Saccharomyces cerevisiae in
rapid fermentation. Appl Environ Microbiol 1976;31:158e62.
[146] Karsch T, Stahl U, Esser K. Ethanol production by Zymomonas and Saccharomyces, advantages and disadvantages. Appl Microbiol Biotechnol 1983;18:
387e91.
[147] Lujan-Rhenals DE, Morawicki RO, Ricke SC. Tolerance of S. cerevisiae and Z.
mobilis to inhibitors produced during dilute acid hydrolysis of soybean meal.
J Environ Sci Health B 2014;49:305e11.
[148] Shiroma S, Jayakody LN, Horie K, Okamoto K, Kitagaki H. Enhancement of
ethanol fermentation in Saccharomyces cerevisiae sake yeast by disrupting
mitophagy function. Appl Environ Microbiol 2014;80:1002e12.
[149] Wallace-Salinas V, Gorwa-Grauslund MF. Adaptive evolution of an industrial
strain of Saccharomyces cerevisiae for combined tolerance to inhibitors and
temperature. Biotechnol Biofuels 2013;6:151.
[150] Kuhn E, Chen X, Dibble CJ, Shekiro J, Nagle NJ, Elander RT. Investigation of
hemicellulase inhibition in the production of bioethanol. In: AIChE annual
meeting, conference proceedings; 2012.

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88


[151] Sung LJ, Jang YR, Lim YH, Kim K. Construction of a thermotolerant Saccharomyces cerevisiae strain for bioethanol production with reduced fermentation time and saccharifying enzyme dose. J Microbiol Biotechnol 2012;22:
1401e5.
[152] Beckner M, Ivey ML, Phister TG. Microbial contamination of fuel ethanol
fermentations. Lett Appl Microbiol 2011;53:387e94.
[153] Huang H, Guo X, Li D, Liu M, Wu J, Ren H. Identication of crucial yeast inhibitors in bio-ethanol and improvement of fermentation at high pH and
high total solids. Bioresour Technol 2011;102:7486e93.
[154] Rahimian ZB, Azin M. Increasing the bio-ethanol yield in the presence of
furfural via mutation of a native strain of Saccharomyces cerevisiae. Afr J
Microbiol Res 2011;5:651e6.
[155] Takashi W, Sathaporn S, Mitsuhiro A, Seiji T, Masaru Y, Itsuki W, et al. Selection of stress-tolerant yeasts for simultaneous saccharication and
fermentation (SSF) of very high gravity (VHG) potato mash to ethanol. Bioresour Technol 2010;101:9710e4.
~es DA, Morais Jr MA. Physiological
[156] De Melo HF, Bonini BM, Thevelein J, Simo
and molecular analysis of the stress response of Saccharomyces cerevisiae
imposed by strong inorganic acid with implication to industrial fermentations. J Appl Microbiol 2010;109:116e27.
[157] Choi GW, Kang HW, Kim YR, Chung BW. Ethanol production by Zymomonas
mobilis CHZ2501 from industrial starch feedstocks. Biotechnol Bioproc Eng
2008;13:765e71.
[158] Galaction AI, Lup
asteanu AM, Cascaval D. Kinetic studies on alcoholic
fermentation under substrate inhibition conditions using a bioreactor with
stirred bed of immobilized yeast cells. Open Syst Biol J 2010;3:9e20.
[159] Rogers P, Lee K, Skotnicki M, Tribe D. Microbial reactions: ethanol production
by Zymomonas mobilis. New York: Spinger-Verlag; 1982.
[160] Joachimsthal EL, Rogers PL. Characterization of a high-productivity recombinant strain of Zymomonas mobilis for ethanol production from glucose/
xylose mixtures. Appl Biochem Biotechnol 2000;84e86:343e56.
[161] Chen R, Wang Y, Shin HD, Agrawal M, Mao ZC. Strains of Zymomonas mobilis
for fermentation of biomass. US Patent Application no. 20090269797; 2009.
[162] Lurgi. Retrieved from: http://gep-france.com/biocarb/Bioethanol-Lurgi.pdf;
2014 [05.24.14].
[163] Kang Q, Huybrechts J, Van der Bruggen B, Baeyens J, Tan TW, Dewil R. Hydrophilic Membranes to replace Molecular Sieves in Dewatering the bioethanol/water azeotropic mixture. Sep Purif Technol 2014;136:144e9.
[164] Kang Q, Appels L, Baeyens J, Dewil R, Tan TW. Energy-efcient production of
cassava-based bio-ethanol. Adv Biosci Biotechnol 2014 [on-line as 7300924].
[165] Luyben WL. Control of a multiunit heterogeneous azeotropic distillation
process. AlChE J 2005;52:623e7.
[166] Tao XY, Zheng DQ, Liu TZ, Wang PM, Zhao WP, Zhu MY, et al. A novel
strategy to construct yeast Saccharomyces cerevisiae strains for very high
gravity fermentation. PLoS ONE 2012;7(2):e31235.
[167] Cao YX, Tian HC, Yao K, Yuan YJ. Simultaneous saccharication and
fermentation of sweet potato powder for the production of ethanol under
conditions of very high gravity. Front Chem Sci Eng 2011;5:318e24.
[168] Puligundla P, Smogrovicova D, Obulam VSR, Ko S. Very high gravity (VHG)
ethanolic brewing and fermentation: a research update. J Ind Microbiol
Biotechnol 2011;38:1133e44.
[169] Behera S, Mohanty RC, Ray RC. Comparative study of bio-ethanol production
from mahula (Madhuca latifolia L.) owers by Saccharomyces cerevisiae and
Zymomonas mobilis. Appl Energy 2010;87:2352e5.
[170] Zhang L, Chen Q, Jin YL, Xue HL, Guan JF, Wang ZY, et al. Energy-saving direct
ethanol production from viscosity reduction mash of sweet potato at very
high gravity (VHG). Fuel Process Technol 2010;91:1845e50.
[171] Liu JJ, Ding WT, Zhang GC, Wang JY. Improving ethanol fermentation performance of Saccharomyces cerevisiae in very high-gravity fermentation
through chemical mutagenesis and meiotic recombination. Appl Microbiol
Biotechnol 2011;91:1239e46.
[172] Hu CK, Qin Q, Gao PP. Medium optimization for improved ethanol production in very high gravity fermentation. Chin J Chem Eng 2011;19:1017e22.
[173] Sridee W, Laopaiboon L, Jaisil P, Laopaiboon P. The use of dried spent yeast as
a low-cost nitrogen supplement in ethanol fermentation from sweet sorghum juice under very high gravity conditions. Electron J Biotechnol
2011;14(6). http://dx.doi.org/10.2225/vol14-issue6-fulltext-5.
[174] Thomas KC, Hynes SH, Jones AM, Ingledew WM. Production of fuel alcohol
from wheat by VHG technology: effect of sugar concentration and fermentation temperature. Appl Biochem Biotechnol 1993;43:211e26.
[175] Pham TNL, Doan NHD, Le VVM. Using fed-batch fermentation in very high
gravity brewing: effects of Tween 80 and ergosterol supplementation on
fermentation performance of immobilized yeast in calcium alginate gel. Int
Food Res J 2010;17:995e1002.
[176] Srichuwong S, Fujiwara M, Wang XH, Seyama T, Shiroma R, Arakane M, et al.
Simultaneous saccharication and fermentation (SSF) of very high gravity
(VHG) potato mash for the production of ethanol. Biomass Bioenergy
2009;33:890e8.
[177] Wang FQ, Gao CJ, Yang CY, Xu P. Optimization of an ethanol production
medium in very high gravity fermentation. Biotechnol Lett 2007;29:233e6.
[178] Wei P, Cheng LH, Zhang L, Xu XH, Chen HL, Gao CJ. A review of membrane
technology for bioethanol production. Renew Sust Energy Rev 2014;30:
388e400.
[179] Ikegami T, Kitamoto D, Negishi H, Haraya K, Matsuda H, Nitanai Y, et al.
Drastic improvement of bioethanol recovery using a pervaporation

[180]
[181]

[182]

[183]

[184]

[185]
[186]
[187]

[188]

[189]

[190]
[191]

[192]

[193]
[194]

[195]
[196]
[197]
[198]

[199]
[200]
[201]

[202]
[203]
[204]

[205]
[206]

[207]

[208]

[209]
[210]
[211]

87

separation technique employing a silicone rubber-coated silicalite membrane. J Chem Technol Biotechnol 2003;78:1006e10.
Nagy E, Boldyryev S. Energy demand of biofuel production applying distillation and/or pervaporation. Chem Eng Trans 2013;35:265e70.
Chen JW, Zhang HM, Wei P, Zhang L, Huang H. Pervaporation behavior and
integrated process for concentrating lignocellulosic ethanol through polydimethylsiloxane (PDMS) membrane. Bioprocess Biosyst Eng 2014;37:183e91.
Trinh LTP, Cho EJ, Lee YJ, Bae HJ, Lee HJ. Pervaporative separation of bioethanol produced from the fermentation of waste newspaper. J Ind Eng
Chem 2013;19:1910e5.
Lee HJ, Cho EJ, Kim YG, Choi IS, Bae HJ. Pervaporative separation of bioethanol using a polydimethylsiloxane/polyetherimide composite hollowber membrane. Bioresour Technol 2012;109:110e5.
Esfahanian M, Ghorbanfarahi AH, Ghoreyshi AA, Najafpour G, Younesi H,
Ahmad AL. Enhanced bioethanol production in batch fermentation by pervaporation using a PDMS membrane bioreactor. Int J Eng Trans B: Appl
2013;25:249e58.
Hilmioglu ZD. Bioethanol recovery using the pervaporation separation
technique. Manage Environ Qual 2009;20:165e74.
Lipnizki F. Membrane process opportunities and challenges in the bioethanol
industry. Desalination 2010;250:1067e9.
He Y, Bagley DM, Leung KT, Liss SN, Liao BQ. Recent advances in membrane
technologies for biorening and bioenergy production. Biotechnol Adv
2012;30:817e58.
Bello RH, Linzmeyer P, Franco CMB, Souza O, Sellina N, Medeiros SHW, et al.
Pervaporation of ethanol produced from banana waste. Waste Manage
2014;34:1501e9.
Zhang Y, Le NL, Chung TS, Wang Y. Thin-lm composite membranes with
modied polyvinylidene uoride substrate for ethanol dehydration via
pervaporation. Chem Eng Sci 2014;118:173e83.
Kang Q. Improvement of cassava-based bio-ethanol production. PhD
dissertation-under review. Beijing University of Chemical Technology; 2014.
Chovau S, Gaykawad S, Straathof AJJ, Van der Bruggen B. Inuence of
fermentation by-products on the purication of ethanol from water using
pervaporation. Bioresour Technol 2011;102:1669e74.
Chen X, Ping ZH, Long YC. Separation properties of alcohol-water mixture
through silicalite-1-lled silicone rubber membranes by pervaporation.
J Appl Polym Sci 1998;67:629e36.
Jia MD, Peinemann KV, Behling RD. Preparation and characterization of thinlm zeolite-PDMS composite membranes. J Membr Sci 1992;73:119e28.
Barba D, Brandani V, Giacomo GD. Hyperazeotropic ethanol salted-out by
extractive distillation theorical evaluation and experimental check. Chem
Eng Sci 1985;40:2287e92.
Black C. Distillation modeling of ethanol recovery and dehydration process
for ethanol and gasohol. Chem Eng Prog 1980;76:78e85.
Meirelles A, Weiss S, Herfurth H. Ethanol dehydration by extractive distillation. J Chem Technol Biotechnol 1992;53:181e8.
Chianese A, Zinnamosca F. Ethanol dehydration by azeotropic distillation
with mixed solvent entrainer. J Chem Eng 1990;43:59e65.
s AF, Asensi J. Dehydration of ethanol using
Gomis V, Pedraza R, France
azeotropic distillation with iso-octane. Ind Eng Chem Res 2007;46:
4572e6.
Hanson N, Lynn F, Scott D. Multi-effect extractive distillation for separating
aqueous azeotropes. Ind Eng Chem Process Des Dev 1988;25:936e41.
Widagdo S, Seider W. Azeotropic distillation. AIChE J 1996;42:96e130.
Black C, Distler D. Dehydration of aqueous ethanol mixtures by extractive
distillation. Extractive and azeotropic distillation. Adv Chem Ser 1972;115:
1e15.
Tai CY, Wu SY. Kinetics of supercritical uid extraction of ethanol from
aqueous solution. Chem Eng Commun 2005;192:1347e60.
lu , alimli A. Supercritical CO2 extraction of ethanol.
Gven A, Mehmetog
Turk J Chem 1999;23:285e91.
Sino M. Dehydration of ethanol using pressure swing adsorption. In: Ramaswamy S, Huang HJ, Ramarao BV, editors. Separation and purication
technologies in bioreneries. Chichester, UK: John Wiley & Sons, Ltd. doi:10.
1002/9781118493441 [chapter 19].
Doherty M, Malone M. Conceptual design of distillation systems. New York:
McGraw Hill; 1972.
Rouquerol J, Avnir D, Fairbridge CW, Everett DH, Haynes JM, Pernicone N,
et al. Recommendations for the characterization of porous solids. Pure Appl
Chem 1994;66:1739e58.
Amarasekara AS. Dehydration to fuel grade ethanol. In: Amarasekara AS,
editor. Handbook of cellulosic ethanol, Chapter 15: dehydration of fuel grade
ethanol. New Jersey, Massachusetts: J. Wiley & Sons and Scrivener Publishing LLC; 2014.
Diaz JC, Gil-Chaves ID, Giraldo L, Moreno-Pirajan JC. Separation of ethanolwater mixture using type-a zeolite molecular sieve. Eur J Chem 2010;7:
483e95.
Green DW, Perry RH. Perry's chemical engineers' handbook. 8th ed. New
York: McGraw Hill Professional; 2007.
Wikipedia. Molecular Sieve. Retrieved from: http://en.wikipedia.org/wiki/
Molecular_sieve [05.06.14].
Essalhi M, Khayet M. Application of a porous composite hydrophibic/hydrophilic membrane in desalination by air gap and liquid gap membrane
distillation: a comparative study. Sep Purif Technol 2014;133:176e86.

88

J. Baeyens et al. / Progress in Energy and Combustion Science 47 (2015) 60e88

[212] Bastidas PA, Gil ID, Rodrguez G. Comparison of the main ethanol dehydration technologies through process simulation. In: 20th European symposium
on computer aided process engineering; 2010.
[213] Dotremont C, Van den Ende S, Vandommele H, Vandecasteele C. Concentration polarization and other boundary layer effects in the pervaporation of
chlorinated hydrocarbons. Desalination 1994;95:91e113.
[214] Dow.
Retrieved
from:
www.dowwaterandprocess.com/en/products
[05.06.14].
[215] Sihi Vacuum pumps. Retrieved from: www.sterlingsihi.com/cms/vacuumpumps.html [05.06.14].
[216] Bekaert Advanced Filtration SA. Developing new media based on short metal
bres. Retrieved from:www.bekaert.com/baf, ; 2014 [09.06.14].
[217] Amarasekara AS. Handbook of cellulosic ethanol. New Jersey, Massachusetts:
J. Wiley & Sons and Scrivener Publishing LLC; 2014.
 A, Se
chet V, Sabiri NE, Pontie
 M, Haure J, et al.
[218] Castaing JB, Masse
Immersed hollow bres microltration (MF) for removing undesirable
micro-algae and protecting semi-closed aquaculture basins. Desalination
2011;276:386e96.
[219] Christenson L, Sims R. Production and harvesting of microalgae for wastewater treatment, biofuels, and bioproducts. Biotechnol Adv 2011;29:
686e702.
[220] Brennan L, Owende P. Biofuels from microalgae-a review of technologies for
production, processing, and extractions of biofuels and co-products. Renew
Sustain Energ Rev 2010;14:557e77.
[221] Kang Q, Huybrechts J, Baeyens J, Tan TW, Dewil R. Cross-ow sintered metal
ber microltration of bio-ethanol fermentation broth. Korean J Chem Eng
2014. KJCE-D-14-00698.
[222] Franken ACM, Nolten JAM, Mulder MHV, Smolders CA. Ehtanol-water separation by membrane distillation, effect of temperature polarization. Berlin:
Walter de Gruyter; 1987.

[223] Gostoli C, Sarti GC. Separation of liquid mixtures by membrane distillation.


J Membr Sci 1989;41:211e24.
[224] Udriot H, Ampuero S, Marison JW, Stockar U. Extractive fermentation of
ethanol using membrane distillation. Biotechnol Lett 1989;11:509e14.
[225] Lebwandowicz G, Bialas W, Marczewski B, Szymanowska D. Application of
membrane distillation for ethanol recovery during fuel ethanol production.
J Membr Sci 2011;375:212e9.
[226] Izquierdo-Gil MA, Jonsson G. Factors affecting ux and ethanol separation
performance in vacuum membrane distillation (VMD). J Membr Sci
2003;214:113e30.
[227] Kiss AA, Suszwalak DJ-PC. Enhanced bioethanol dehydration by extractive
and azeotropic distillation in dividing-wall columns. Sep Purif Technol
2012;86:70e8.
zquez-Ojeda M, Segovia-Herna
ndez JG, Herna
ndez S, Hern
[228] Va
andezAguirre A, Kiss AA. Design and optimization of an ethanol dehydration
process using stochastic methods. Sep Purif Technol 2013;105:90e7.
[229] Kiss AA, Ignat RM. Innovative single step bioethanol dehydration in an
extractive dividing-wall column. Sep Purif Technol 2012;98:290e7.
[230] Kiss AA, Suszwalak DJ-PC, Ignat RM. Breaking azeotropes by azeotropic and
extractive distillation in a dividing-wall column. Chem Eng Trans 2013;35:
1279e84.
[231] Kiss AA, Suszwalak DJ-PC. Efcient bioethanol dehydration in azeotropic and
extractive DWC. In: Coference Paper, 19th International congress of chemical
and process engineering; 2011.
[232] Poramacom N. Cassava production, prices and related policy in Thailand. Am
Int J Contemp Res 2013;5:43e50.
[233] China sugar annual report. Retrieved from: http://www.thebioenergysite.
com/articles/628/china-sugar-annual-report-2010; 2010 [14.05.14].

You might also like