You are on page 1of 20

Renewable and Sustainable Energy Reviews 65 (2016) 1179–1198

Contents lists available at ScienceDirect

Renewable and Sustainable Energy Reviews


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

A sustainable integrated in situ transesterification of microalgae for


biodiesel production and associated co˗products ˗ A review
Kamoru A. Salam a,n, Sharon B. Velasquez-Orta b, Adam P. Harvey b
a
Department of Chemical Engineering, Faculty of Engineering, University of Abuja, P.M.B. 117, Airport Road, Main campus, FCT-Abuja, Nigeria
b
School of Chemical Engineering and Advanced Materials (CEAM), Newcastle University, NE1 7RU, United Kingdom

art ic l e i nf o a b s t r a c t

Article history: Microalgae has large scale cultivation history particularly in aquaculture, pigments and nutraceutical
Received 7 February 2016 production. Despite the advantages of microalgal oil as feedstock for biodiesel production, algal biodiesel
Received in revised form is still at laboratory scale due to technical challenges required to be overcome to make it economical and
25 June 2016
sustainable. Indeed, complete drying of microalgae is energy intensive and significantly increases the
Accepted 21 July 2016
cost of algae pre-treatment. In situ transesterification is more water tolerant due to excess methanol to oil
molar ratio required by such production route. However, the need to remove unreacted methanol (494%
Keywords: of it) from the product streams certainly requires distillation heat load which increases the operating
Microalgae cost. This article reviews the key process variables affecting efficiency of in situ transesterification. These
Biodiesel
include alcohol to oil molar ratio, moisture, stirring rate, reaction time, temperature, microalgal cell wall
In situ transesterification
and catalyst type. Potential solutions of improving the efficiency/economy are discussed. Overall, an
Dimethyl ether (DME)
Biogas integrated approach of in situ dimethyl ether (DME) production along with the desired biodiesel
Bio-digestate synthesis during in situ transesterification would substantially reduce the volume of unreacted methanol
thereby reduces operating cost. Use of resulting microalgal residue for biogas (methane) production can
provide energy for biomass production/separation from the dilute algae˗water mixture. Use of bio˗di-
gestate as nutrients for supporting microalgal growth is among the probable solutions suggested for
reducing the production cost of in situ transesterification.
& 2016 Published by Elsevier Ltd.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1180
2. Alternative fuels. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1180
2.1. Energy content of gasoline and alternative fuels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1180
2.2. Comparative cost of gasoline and alternative fuels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1182
3. Microalgae compositions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1182
3.1. Lipids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1182
3.1.1. Effect of lipid composition on fuel quality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1182
3.1.2. Effect of algae cell wall on bioprocesses. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1182
3.1.3. Microalgae cell disruption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1183
3.1.4. Energy requirement of algae cell disruption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1183
4. The transesterification reaction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1184
4.1. Kinetics of reactive extraction (“in situ transesterification”) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1185
4.2. A comparison of two step transesterification and reactive extraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1185
4.3. Overview of In situ transesterification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1186
4.4. Key process variables in in situ transesterification of microalgae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1186
4.4.1. Solvent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1187
4.4.2. Temperature and reaction time. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1187
4.4.3. Agitation rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1188

n
Corresponding author.
E-mail address: kamorusalam@gmail.com (K.A. Salam).

http://dx.doi.org/10.1016/j.rser.2016.07.068
1364-0321/& 2016 Published by Elsevier Ltd.
1180 K.A. Salam et al. / Renewable and Sustainable Energy Reviews 65 (2016) 1179–1198

4.4.4. Catalyst. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1188


4.4.5. Acid-catalysed reactive extraction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1188
4.4.6. Alkali-catalysed reactive extraction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1190
4.4.7. Heterogeneous-catalysed reactive extraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1191
4.4.8. Non-catalysed reactive extraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1191
4.4.9. Moisture content . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1191
5. Microalgae species. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1191
5.1. Protein and carbohydrate content of algae. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1192
6. Probable ways of making a sustainable in situ transesterification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1192
6.1. Anaerobic digestion of algal residue after reactive extraction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1192
6.1.1. Causes of inhibition during biogas production. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1192
6.2. marine species for biofuel production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1192
6.3. In situ co-production of diethyl-ether during in situ transesterification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1192
6.3.1. Fuel properties of DME or DME-Biodiesel blends . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1194
7. Concluding remarks. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1195
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1196

1. Introduction translated to commercialisation of algal biodiesel. The problem is


that there are still a numbers of constrains required to be over-
Biodiesel is commonly produced from refined vegetable oil come to achieve commercial reality of algal biodiesel.
made from soybean, rapeseed, sunflower and palm oil. In Europe For instance, microalgal cell consists of phospholipid bilayer
and America, it is largely produced from rapeseed and soybean, membranes (about 7–8 nm thick) embedded with glycolipids,
respectively, due to their availability and the cold flow properties glycoproteins and polysaccharides [9]. This configuration provides
of the resulting biodiesel [1]. Edible oil crops account for  95% of structural rigidity for the cells to adapt to their environments but
worldwide biodiesel feedstocks [1]. However, using these crops for adversely affects the efficiency of algal bioprocesses including
fuel would require significant amounts of freshwater [2] and ara- fermentation, anaerobic digestion, oil extraction and biodiesel
ble lands [3]. This would create undue competition between using production. Indeed, it leads to significant high solvent require-
these resources for energy crop and agricultural or domestic pur- ments and energy load during the extraction processes [10]. A
poses. In Europe and America, the increased cost of food oil crops direct biodiesel production requires a significant methanol to oil
has been strongly influenced by their use as energy crops [4]. High molar ratio, usually in the range of 100:1–1000:1 methanol to oil
prices may be beneficial to farmers, but this might lead to food molar ratio but can be as high as 1570:1 [11]. Dhar and Kirtania
shortages in many developing countries, particularly where almost [12] reported that, for a transesterification process operating at 6:1
50% of their earnings are used for food [4]. It could also lead to and 40:1 methanol to oil molar ratios, 662 and 6450 kW respec-
global food insecurity. Competition for land between food and fuel tively were required for a 95% methanol recovery. By extrapola-
can also have a negative environmental implication. Recently, vast tion, 16,658–171,898 kW will be required for the same methanol
areas of rainforests particularly in Malaysia, Indonesia and Thai- recovery by a reactive extraction process operating at 100:1–
land have been exploited for cultivation of palm oil as it is in high 1000:1 methanol to oil molar ratio. This means the downstream
demand for food and fuel [5]. This has caused deforestation and distillation heat load required to recycle unreacted methanol
negatively affects the forest ecosystems. In order to sustainably (494% of it) would significantly increase the operating cost and
provide raw materials for large-scale production of biodiesel, has been one main reason why the process is not economically
non˗edible feedstocks that require marginal land and insignificant viable. In addition, FAME yield was reported to increase with rising
freshwater are preferable. Such feedstocks include waste oils and acid concentration during in situ transesterification [6,7,13] but the
microalgae particularly the marine species [2]. Non-food oil crops need to neutralise the unreacted acid in the product streams will
and waste oils can only supply limited quantities of biofuels, so increase the operating cost to some extent. Additionally, limited
cannot meet world transport fuels requirements. Biodiesel can be supply of concentrated CO2, nonfully utilisation of nitrogen or
produced from microalgae through reactive extraction (“in situ phosphorous nutrients, adverse effect of using small quantity of
transesterification”) or a two˗step transesterification. In a two˗step fresh water even if marine algae is used, and inefficient utilisation
transesterification, pre-extracted oil from microalgae can be con- of algal residue after oil extraction hinder commercialisation of
verted into fatty acid methyl ester (FAME) (“biodiesel”) with al- algal biofuels [2]. This review provides ways of reducing the excess
kalised or acidified methanol. Production of algal biodiesel via a methanol to oil molar ratio and efficient utilisation of residual
two-step transesterification usually includes dewatering, conven- algae produced during in situ transesterification. Factors affecting
tional drying, solvent extraction, oil degumming, trans/esterifica- in situ transesterification were also critically reviewed, and a new
tion, neutralisation and product purification. Alternatively, algal integrated process to improve the economy of the in situ trans-
biodiesel can be produced via in situ transesterification by con- esterification is discussed.
tacting the algal biomass directly with an alcohol containing a
catalyst [6,7]. It is potentially a cost-effective alternative way of
producing algal biodiesel due to its elimination of the solvent 2. Alternative fuels
extraction step and its higher water tolerance [6,7] as the solvent
extraction and drying steps account for  90% of the process en- 2.1. Energy content of gasoline and alternative fuels
ergy in a two-step transesterification of algal oil to biodiesel [8].
Though microalgae has short growing time, high lipid productivity Irreversible depletion of fossil fuel and the climatic change due
and can be use for capturing concentrated CO2, they can be cul- to its combustion result into a significant global interest in re-
tivated on non-arable land using wastewater; and are adaptable to newable and sustainable transport energy sources such as biofuels
harsh environments but these advantages are still yet to be (biodiesel and bioethanol). Similarly, use of hydrogen as energy
K.A. Salam et al. / Renewable and Sustainable Energy Reviews 65 (2016) 1179–1198 1181

carrier in fuel cell vehicles (FCVs) has attracted public 80000

Oil Equivalent (Thousand Tonnes)


attention because water and hot air are the end products of elec-
tricity generation using this approach [14]. So its emissions are not 70000
negatively affecting human health. However, there are still some
60000
challenges blocking wide applications of hydrogen as energy car-
rier. These include safety, high cost of production and transpor- 50000
tation, storage problems and high cost of fuel cells. Considering
the lingering effect of hydrogen explosion during Hindenburg 40000
tragedy, good safety techniques of hydrogen transportation and
storage are required to allay public fear and to encourage sig-
30000
nificant sponsoring of hydrogen production as transport fuel. Most 20000
of the well-established methods of storing hydrogen such as
compression and cryogenic or liquefaction are energy intensive, 10000
costly and require extra safety measures. For instance liquid hy-
drogen requires cryogenic storage around  252 °C [15]. Extra cost 0
would be required to insulate the hydrogen tank to prevent energy
loss at this temperature.
The energy density and specific energy of gasoline and alter- Production year
native fuels are shown in Table 1.
As can be seen in the Table, liquid hydrogen contains energy Fig. 1. Global biofuels production from 2005 to 2015.
Source: British Petroleum (2016).
density of 0.01 MJ L  1 which is significantly lower than energy
density of gasoline (34 MJ L  1), biodiesel (34 MJ L  1) and that of
ethanol (24 MJ L  1). This means a substantial bigger tank is re- consumption of electricity. These include hydrogen generation
quired to store liquid hydrogen having the same energy contents from solar energy [27]; photolysis, involving exploiting of chlor-
as the alternative fuels. On mass basis, the specific energy content ophyll and enzyme in microalgae to chemically splitting sea water
of hydrogen is 142 MJ kg  1 which is significantly greater than that into hydrogen and oxygen [28] or production of hydrogen from
of gasoline (46 MJ kg  1), biodiesel (42 MJ kg  1) and ethanol starch and water [29]. While all these processes are technically
(30 MJ kg  1). However, storing hydrogen as gas typically requires feasible, they are still at laboratory scale meaning that use of hy-
compressing and maintaining it at about 350–700 bar [18]. drogen as energy carrier for transport fuel may not meet the
Besides extra cost of compression, the public needs to be con- world's energy need in the near term.
vinced on the safety in carrying such high pressure tank on board. Consideration of biofuels such as biodiesel and bioethanol as
More research efforts are now being focused on efficient and safe transport fuels has attracted a global interest. Between 2005 and
techniques of storing or producing hydrogen on board. These in- 2015, there has been a continued increase in global production of
clude metal hydrides, particularly the magnesium based one, as biodiesel and bioethanol, as shown in Fig. 1.
they have potential of producing hydrogen capacity up to 7.6 wt% Global biofuels production in 2015 shows that 42.9% was ac-
[19,20]; glass microspheres [21] and activated carbons [22,23]. counted for by North America, 27.9% by South and Central America,
Fossil fuel is still the major source of industrial hydrogen. Ap- 18.3% by Europe and Eurasia, 10.8% by Asia while Africa contributed
proximately 96% of commercial scale hydrogen is being produced 0.1% to the global biofuels production [30].
from fossil fuel by steam reforming while about 4% is being pro- Based on the highest biofuels country producer in each con-
duced through water electrolysis [24,25]. tinent, the United State of America has the highest share (41.4%) in
Steam reforming of hydrocarbons produces significant large the North America, Brazil produced the highest (23.6%) in the
amount of CO2 rendering zero emission criteria of hydrogen ques- South and Central America, Germany produced the highest (4.2%)
tionable. Use of water electrolysis for hydrogen production requires in the Europe and Eurasia while China was the highest producer
significant electricity consumption, causing high production cost. For (3.2%) in the Asia pacific [30]. However, bioethanol is produced
instance use of electrolysis for hydrogen generation in the United mainly from food crops in these countries. For instance, bioethanol
States requires building 1800 new power plants that emit CO2 and has been produced on a commercial scale in Brazil, United State of
about $1 trillion would be needed to execute the project [26]. Pre- America and European countries solely from food crops such as
sently, 168,000 gasoline fueling stations are in the United States and cane sugar, corn and wheat [31]. This approach is unsustainable as
a cost of $2.8 million has been projected by National Renewable every hectare of land used for cultivating these energy crops leads
Energy Laboratory (NREL) for each hydrogen fueling station capable to clearing of the same amount of land for food production.
of handling 100 cars per day [26]. Therefore, a total capital invest- Clearing rainforests, peatlands, savannas or grass lands to produce
ment required to replicate the number of gasoline stations with food crop-based biofuels increases annual greenhouse gas emis-
hydrogen fuel stations would be $470 billion which is quite high. sions [32,33] and world food prices [4,34].
A more appealing procedure would be hydrogen production Alternatively, bioethanol can be produced from agricultural by-
that require no fossil fuel and which does not require significant products such as wheat straw, sugarcane bagasse and corn stover;
forest residues including sawdust, thinning forests; energy crops
Table 1 such as Salix and switch grass [31]. These feedstocks have high
Energy density and specific energy of different fuels. content of cellulose and hemicellulose but they are difficult to
Source: Gaur and Reed [16]; ETB [17]. hydrolysis in a high yield into monomer sugars than obtaining
sugar from carbohydrate- or sugar-containing substances such
Fuel Energy density, MJ L  1 Specific energy, MJ kg  1
corn or cane sugar for bioethanol production [31].
Gasoline (Petrol) 34 46 In Europe biodiesel is largely produced from oilseed, but in-
Hydrogen liquid at 20 K 10 142 sufficient supply from this feedstock has increased demand for
Hydrogen gas 0.01 143 palm oil from Malaysia and Indonesia causing significant clearing
Biodiesel 34 42
Ethanol 24 30
of rainforest for palm plantation [34] which can negatively affect
the ecosystem. Biodiesel from microalgae could be alternative
1182 K.A. Salam et al. / Renewable and Sustainable Energy Reviews 65 (2016) 1179–1198

source of transport fuel. Although it has not been be commercia- common algal substrate known for biodiesel synthesis. In contrast,
lised for some technical reasons which will be discussed later in- polar lipids contain polar groups, such as choline, ethanolamine,
side the text. Nevertheless, this can be overcome. Previous studies serine, water, glycerol and phosphatidyglycerol in phospholipids.
by Chisti [35] has shown that a 2.8 energy ratio (i.e. the renewable Glycolipids are other forms of polar lipid but are less polar than
energy produced per unit of fossil fuel input) was obtained from phospholipids. They are simple sugar-containing lipids. They in-
raceway production of microalgae biomass containing 20% (w/w), clude monogalactosyl diglyceride, digalactosyl diglyceride and
biomass productivity of 0.025 kg m  2 d  1[36–38] and concentra- sulpholipids. They are located in the cell walls. They are, as might
tion of 1 kg m  3 in the ponds. An energy output of 1444 GJ ha  1 be expected, more soluble in polar solvents, such as methanol and
year  1 based on 90% operational factor for the production facility water. They give cell walls their structural rigidity. Structures of
has been estimated [35]. In comparison, a total energy output of non-polar and polar lipids are shown in Table 4.
163.9 GJ ha  1 year  1 was estimated for cane sugar having total Where R1, R2 and R3 in Table 4 denote fatty acids moieties
energy yield (bioethanol and bagasse) of 218.5  10  2 GJ per tonne which could be involved in transesterification process as com-
of cane sugar [39] assuming a cane productivity of  75 t ha  1 monly observed in triglycerides. The symbol “X” in phospholipids
[35]. In addition microalgae can be cultivated on marginal land or can be any of the substituent group listed in Table 5.
wastewater. It is also non-food oil substrate, so cannot cause food
shortage. 3.1.1. Effect of lipid composition on fuel quality
Microalgae are extremely diverse, with substantially different
2.2. Comparative cost of gasoline and alternative fuels fatty acid profiles, which means that the resulting biodiesel will
also vary substantially [19]. It has been shown by Ben-Amotz et al.
The latest national average price of transport fuels obtained [47] that a wide range of microalgae species can synthesize C14:0,
from alternative fuels data center of United State Department of C16:0, C18:1, C18:2 and C18:3 free fatty acids, while other fatty
Energy [40] and the price of hydrogen reported by Dears [26] are acids including C16:2, C16:3, C16:4, C18:4, C20:5, C22:6 were
listed Table 2. All the prices were converted to $/gasoline gallon strain-specific. In addition, deficiency in Nitrogen caused sig-
nificant increase in the lipid content in Ankistrodesmus sp. Thomas,
equivalent (GGE) to allow comparison with the price of the
Isochrysis sp., Nannochloris sp. Thomas, Nitzchia sp., Nitzchia sp.
equivalent energy content of gasoline. As can be seen in the Table,
Chapman, Botryococcus braunii Kutz except Dunaliella species that
hydrogen fuel has the highest price, followed by biodiesel and
accumulated higher carbohydrate content. Other investigations
then ethanol while gasoline has the least price. While gasoline has
have shown that culture, environmental conditions, habitat and
the lowest price, its combustion contributes significantly to cli-
growth phase all have a significant effect on the fatty acid profile
mate change.
of microalgae [48,49]. Table 6 shows the effect of growth media,
Presently there is no industrial scale production of biodiesel
environmental stress and catalyst type on fatty acid profile of
from microalgae. Therefore, algal biodiesel production cost is ob-
different microalgae.
tained from techno-economic analysis reported in the literatures
As can be seen in the Table, growth media [50,52] and varying
[41–44]. An updated cost estimated for in situ transesterification of
catalyst [51] have significant effect on fatty acid profile of algal
wet microalgae was carried out by Nagarajan et al. [41]. They
lipids and cause difference in FAME compositions. Variation in
obtained cost of production ranged from $1.69–3.90/GGE which is
fatty acid compositions due to catalyst type was attributed to
comparable to that of gasoline. As can be seen in the Table 3, they
differences exhibited by alkali or acid catalyst to free fatty acids
obtained lower biodiesel production cost than other investigators.
(FFA), cell wall lipids and triglyceride during in situ transester-
This can be attributed to the fact that their process is wet-in situ
ification [51]. It is well known that microalgae accumulate more
transesterification. Lardon et al. [8] reported that solvent extrac-
neutral lipids (triglycerides) rather that structural lipids when
tion and drying of algae are energy intensive and contributes
cultivated on nitrogen deplete media. Since both lipids have dif-
significantly to the process energy. Similarly, as shown in the Table
ferent fatty acid composition, the resulting total lipids is sig-
the cost of biodiesel produced from microalgae cultivated in open
nificantly affected by growth media [50].
pond is generally greater than that produced from photo-bior-
Regardless of the factors, the lipids contain large amount of
eactor. This is because capital and operating cost of photo-bior- polyunsaturated FAME. Significant high amount of polyunsaturated
eactor is significantly greater than open pond [44]. (PUFA) reduces oxidative stability and cetane number of the resulting
algal biodiesel. Although, the oxidative stability of the lipid/FAME is
strongly dependent on the structure of the PUFA as bisallylic sites in
3. Microalgae compositions the PUFA are more prone to oxidation than the allylic site [53] The
bis˗allylic sites are a methylene groups (CH2) adjacent to two double
3.1. Lipids bond, while allylic sites are the one adjacent to a double bond [53].
EN 14,214 recommends that the maximum contents of linolenic
Lipids are broadly classified as neutral or polar. A neutral lipid acid methyl ester should not be more than 12% (m/m) while
has no overall polarity. It is located inside the cells in the form of polyunsaturated methyl esters (with four or more double bonds)
triglycerides, mono- and diglycerides or free fatty acids. They are should not be more than 1% (m/m) [54]. Probable solution to re-
more soluble in non-polar solvents such as hexane and chloro- duce PUFAs include addition of fuel additive containing 0.03% tert-
form. Neutral lipids are energy storage products. They are the butyl hydroquinone (TBHQ) [55] or harvesting of microalgae using
ozone floatation [56] .
Table 2
Price of transport fuels.
3.1.2. Effect of algae cell wall on bioprocesses
Fuels Price ($/GGE) Reference Microalgal cell wall accounts for  10% of dry algal biomass [57] It
consists of phospholipid bilayers membrane embedded with glyco-
Gasoline 2.06 US DoE (2016) lipids, glycoproteins and polysaccharides [9]. Eukaryotic cell's wall
Ethanol (E 85) 2.56 US DoE (2016) typically contains fibril and mucilage such as polysaccharides, lipid
Biodiesel (B99-B100) 3.00 US DoE (2016)
Hydrogen 14.3 Dears (2015)
and protein [9,57,58]. Analysis of the cell wall compositions of few
microalgae conducted by Abo-shady et al. [58] are shown in Table 7.
K.A. Salam et al. / Renewable and Sustainable Energy Reviews 65 (2016) 1179–1198 1183

Table 3
Cost of biodiesel production from microalgae by techno-economic analysis.

Cultivation system Cost of biodiesel ($/GGE) Remarks Reference

Open pond (400 ha) 1.69–3.90 Wet in situ transesterification [41]


Ultra-sonication assisted process
Open pond (333 ha) 6.80 Monte Carlo Sampling method [42]
PBR 11.30 Monte Carlo Sampling method
Open pond 15.80 Monte Carlo Financial feasibility model [43]
PBR 40.00 Monte Carlo Financial feasibility model [43]
Open pond (1951 ha) 11.00 [44]
PBR 23 [44]

Table 4 Table 5
Structures of non-polar and polar lipids. Substituent group in phospholipids.
Source: Wood [45] Source: Wood [45]

Lipid Structure Name Structure Abbreviation for lipid

þ
Triglyceride Choline .O.CH2.CH2.N (CH3)3 P.C
Ethanolamine .OCH2.CH2.NH2 P.E
Serine P.S

Water .H P.A
Sulpholipid Glycerol .CH2(OH)CHCH2OH PG
Phosphatidyglycerol .CH2CH(OH)CH2. D.D

P.A: Phosphatidic acid; P.G: phosphatidyl glycerol; D.D: diphosphatidyl glycerol; P.


C: phosphatidyl choline; P.E: phosphatidyl ethanolamine; P.S: phosphatidyl serine.

Monogalactosyl diglyceride
wall [65]. Botryococcus braunnii is not an attractive option for
biodiesel production because it is difficult to cultivate [65]. Be-
cause microalgal neutral lipids are contained within cells they are
not readily available for extraction, so require disruption prior to
Digalactosyl diglyceride transesterification [66].

3.1.3. Microalgae cell disruption


Disruption techniques are broadly classified into mechanical,
physical, chemical and enzymatic [66,67] as shown in Fig. 2.
Phospholipids
Cell disruption enhances lipid extraction from microalgae. Lee
et al. [68] while using different cell disruption for enhancing lipids
extraction from Botryococcus sp., Chlorella vulgaris and Scenedesmus
sp. obtained for autoclaving (5.4–11.9% lipids), bead-beating (7.9–
8.1% lipids), microwave irradiation (10–28.6% lipids), sonication
As can be seen in the Table, different microalgae contain
(6.1–8.1% lipids) and 10% NaCl osmotic shock (6.8–10.9% lipids).
varying amount of carbohydrate, protein and lipids in the cell wall.
Microwave irradiation produced the highest lipid enhancement.
Depending on the microalgae species, a hard surface such as silica,
Surendhiran and Vijay [69] compared the effect of different cell
calcium carbonate, algaenan or sporopollelin can be present. Other
wall disruption (enzymatic, ultra-sonication, osmotic shock and
products found in the cells are listed in the Table 8.
acid lysis) on lipid extraction from Nitrogen replete and deplete
As shown in the Table, cellulose is commonly found in the cell
Nannochloropsis occulata. Acid lysis at a pH of 2 for 2 h was the
wall of Chlorella vulgaris. Nannochloropsis occulata cell wall contained
most effective, producing 54.3% lipid in Nitrogen depletion but
polysaccharide with sugar unit as glucose (68.8%), fucose (4.4%), ga-
33.2% in replete samples. This is contrary to microwave irradiation
lactose (3.8%), mannose (6.1%), rhamnose (8.3%), ribose (4.6%), and
reported to be most efficient in disrupting Chlorella vulgaris,
xylose (4.4%) [64]. In addition, algaenan (a chemical-resistant, bio- Botrycoccus sp. and Scenedemus sp. reported by Lee et al. [68].
macromolecule) are found in the cell wall of Nannochloropsis occulata Microalgae cell wall chemistries vary, consequently cell disruption
[63]. Some Chlorella species contain algaenan, albeit with a different techniques should be strain and species dependent.
chemical composition to those in Nannochloropsis occulata [61,62].
Sporopollenin (a chemical-resistant bio-macromolecule) was found 3.1.4. Energy requirement of algae cell disruption
in Chlorella sp.[60]. Cell disruption enhances lipid extraction, but can be very energy
Microalgal cell wall as described above provides structural rigidity intensive. In addition, algal cells are unsuitable for mechanical
for the cells to adapt to their environments but the cell wall resistance presses as they are too small, and can pass through unchanged [66].
adversely affects efficiency of algal bioprocessing. These includes ge- An overview of energy consumptions of microalgal cell dis-
netic transformation, fermentation, anaerobic digestion, oil extraction ruptions that are considered suitable for commercial scale pro-
and biodiesel production. Indeed, it causes significant large solvent duction [66] is shown in Table 9.
requirement and energy load during extraction processes [10]. A 33 MJ kg  1 (dry cells) was the lowest energy consumed,
There are, however, exceptions, such as Botryococcus braunnii, which is significantly greater than 27 MJ kg  1 (the estimated
that excretes oil as less oxygenated isoprenoids outside the cell combustion energy from a typical algae biomass). Furthermore,
1184 K.A. Salam et al. / Renewable and Sustainable Energy Reviews 65 (2016) 1179–1198

Table 6
Effect of media and catalyst type on fatty acid profile of microalgal lipids.

a b c d e f
wt% C. vulgaris C. vulgaris N. occulata Z. zofingiensis D. tertiolecta I. galbana

C12:0 – – 2 (1) – – –
C14:0 2 () 2 (0.1) 6 (6) 2 () 2 (1) 17 (22)
C14:1 – 1 (0.3) 0.1 (0.3) – – –
C15:0 – – 0.2 (2) – – –
C16:0 20 (17) 6 (5) 26 (30) 18 (15) 18 (25) 9 (14)
C16:1 1 (1) 16 (11) 22 (21) 1 (1) 1 (1) 4 (2)
C16:2 11 (3) – 1 (3) 7 (4) 2 (1) –
C16:3 14 (6) – – 9 (2) 4 (2) 2 (˗)
C16:4 – – – 2 (˗) 20 (12) –
C18:0 1 (2) 10 (11) 4 (2) 2 (3) 1 (2) 1 (1)
C18:1 4 (47) 27 (25) 12 (7) 18 (47) 5 (14) 5 (27)
C18:2 19 (10) 0 (9) 5 (3) 20 (17) 11 (9) 13 (3)
C18:3 28 (14) 21(22) 0.2 (1) 18 (8) 35 (31) 9 (4)
C18:4 – – 3 (0.4) 2 () 1 (1) 10 (10)
C20:0 – –  (0.2) – – 1 (2)
C20:1 – – 4 (5) – – –
C20:4 – – – – – –
C20:5 – – 13 (18) – – –
C22:6 – –  (0.2) – – 15 (12)
∑saturated 23 (19) 18(16) 38(41) 22(18) 21(28) 28(37)
∑monounsaturated 5 (48) 44(36) 38(33) 19(48) 6 (15) 34(29)
∑polyunsaturated 67 (33) 21(31) 21(25) 51(31) 73 (56) 34(29)

ND: The conditions inside the bracket refers to the FFA inside the bracket listed in the Table.
a,d-f
Breuer et al. [50]: Nitrogen replete medium (Nitrogen deplete medium).
b
Velasquez˗orta et al. [51]: acid catalysed reactive extraction (alkali-catalysed reactive extraction).
c
Renaud et al. [52]: NO3˗containing medium (NH4˗containing medium).

Table 7 4. The transesterification reaction


Composition of the microalgae cell wall (% of cell wall dry matter).
Biodiesel is usually made by transesterification of triglycer-
Microalgae Carbohydrate (%) Protein (%) Lipids (%) ND (%)
ides or esterification of free fatty acids derived from plants or
Chlorella vulgaris (F) 30.00 2.46 15.00 52.54 animals with low molecular mass alcohols containing catalyst.
Chlorella vulgaris (S) 35.00 1.73 10.00 53.27 Transesterification involves reacting a triglyceride with an al-
Kirchneriella lunaris 75.00 3.96 12.50 8.54 cohol in the presence of alkali or acid catalyst to form alkyl
ND: Non-detected substances.
esters of the corresponding alcohol and glycerol. Methanol is
commonly used because is the cheapest alcohol [46]. If the
Table 8
reaction goes to completion, three molecules of alkyl ester
Typical cell wall composition and storage products in few algae. (biodiesel) of the corresponding alcohol are formed. For in-
stance, if methanol is used, three molecules of fatty acid methyl
Substance Chlorophyta Chor- Ochrophyta Nanno- Reference ester and one molecule of glycerol are formed as illustrated in
ella vulgaris chloropsis occulata
Fig. 3.
Phycobilins Absent Absent [9] The stoichiometric ratio of alcohol to oil needed for this reac-
Storage Starch Chrysolaminarin [9,59] tion for an alkali catalyst is 3:1, but in practice, since it is an
carbohydrate (α˗1,4˗glucan) (β˗1,3˗glucan) equilibrium reaction, an excess of alcohol of 6:1 is typically re-
Cell wall cellulose polysaccharides [9,59] quired, to increase the rate and conversion [76]. A typical alkali
a c a
Additional cell wall Sporopollelin Algaenan [60],c[63]
components b
Algaenan b
[61,62]
transesterification takes at least 3 min for two step transester-
ification [77] but 10 min for in situ transesterification of Chlorella
vulgaris[78] to reach completion. In contrast, an acid-catalysed
reaction can be  4000 times slower [79].
the energy consumption is greater than the estimated minimum The alkali reaction is faster because it involves a strong nu-
theoretical energy consumption by a factor of 105. cleophile alkoxide species. Once it is formed, it directly attacks the
There would be economic justification for this high energy carbonyl moieties in the triglyceride to form the corresponding
consumption for algal cell disruption if the lipids or pigments alkyl esters as illustrated in the Fig. 4.
The steps involved in homogeneous alkali catalysis are as
were extracted for high value commodities, such as pharma-
shown in Fig. 4 [75]. They are:
ceutical or nutraceutical products. However, it becomes difficult
to sustain if the lipids are extracted for bulk fuels such as bio-
a) formation of active alkoxide catalyst species, RO  ;
diesel [66]. In order to make algal oil economically competitive
b) nucleophilic attack of RO  to the carbonyl group on trigly-
biodiesel feedstock, a less energetic disruption technique is re- ceride (TG) producing a tetrahedral intermediate;
quired. For instance chemical pre-treatment such as surfactant/ c) breaking down of the tetrahedral intermediate;
surfactant catalyst, alkali or acid catalyst for disrupting algae d) regeneration of the alkoxide (RO  ) species. The sequences are
cells are less explored. In addition, research effort should be repeated for both diglyceride and monoglyceride formation.
directed towards the effect of these chemical pre-treatment
in depolymerising the polysaccharides and algaena on the algal It should be noted that a detailed alkali catalysed two step
cell wall. transesterification schemes contains side reactions including free
K.A. Salam et al. / Renewable and Sustainable Energy Reviews 65 (2016) 1179–1198 1185

Fig. 2. Classification of cell disruption techniques. Modified from [66].

fatty acid neutralisation. saponification of triglycerides and bio- a two-step transesterification. At 600:1 methanol to oil molar ra-
diesel (FAME) [77]. tio, the process was not adversely affected by up to 20% moisture
In contrast, as shown in Fig. 5, acid catalysis involves formation in the microalgae. A derived numerical model fitted with experi-
of an electrophilic species, which reacts with the alcohol to form a mental data from their study showed that other side reactions
tetrahedral intermediate. including FAME and triglyceride saponification; free fatty acid
As shown in Fig. 5, the homogeneous acid-catalysed reaction neutralisation occur alongside the desired biodiesel synthesis in a
scheme for triglyceride transesterification [75] is: NaOH˗catalysed reactive extraction. Their findings is summarised
in the Fig. 7.
a) Protonation of the carbonyl group by the acid catalyst to create
an electrophilic species; 4.2. A comparison of two step transesterification and reactive
b) Nucleophilic attack by the alcohol to generate tetrahedral extraction
intermediate;
c) Proton migration and breaking down of the intermediate. The Two step transesterification requires refined oil from either ve-
sequences are repeated for both diglycerides and monoglycerides. getable or other oil seeds such as canola, rapeseed or soymeal. As
much as 88% of total production cost of a two˗step biodiesel pro-
Acid catalysis takes considerably longer than alkali catalysis. duction is ascribed to the refined oil feedstock [81] . It is also im-
However, acid catalysts can catalyse esterification of free fatty portant during two˗step transesterification to control the water
acids to biodiesel, as shown in Fig. 6. This is why it is applicable for content in the feedstock, catalyst or methanol particularly for alkali-
high free fatty acid (FFA) substrates. catalysed process. Typically, the maximum tolerable water content
in oil is 0.3 wt% [76]. Beyond this value, there could be saponification
4.1. Kinetics of reactive extraction (“in situ transesterification”) of the oil to soap, which reduces the biodiesel yield [82] and causes
difficulty in product separation [82,83]. In addition two step bio-
Reactive extraction (“In situ transesterification”) is a direct diesel production involves hexane extraction steps that are energy-
production of fatty acid methyl ester (FAME) from oil-bearing intensive and time-consuming. Up to 90% of the process energy can
biomass, achieved by contacting the material directly with an al- be accounted for in the hexane extraction and drying steps [8]. Al-
cohol containing a catalyst. Zakariah and Harvey [80] studied the ternatively, a reactive extraction (“in situ transesterification”) could
kinetics of reactive extraction of rapeseed to FAME with metha- be used. In this process, the biomass is fed directly into the reaction
nolic NaOH. Their simulated and experimental data suggested that system. This eliminates the extraction steps, biomass pre-treatment
the process could be either mass transfer or kinetically controlled and degumming steps, and tolerates some level of water [6,7,84]
depending on the concentration of the catalyst. At higher catalyst due to excess methanol used in the process. The basic differences
concentrations (40.1 mol/kg-solvent) the process was controlled between the two processes are shown in the Fig. 8.
by diffusion rate, but when the concentration was lower, it was Previous studies have demonstrated the feasibility of obtaining
kinetically controlled. However, they did not consider in their greater FAME conversion from such in situ transesterification than
model the competing reactions which occur alongside the desired from a two-step approach [85–87]. It is equally effective in making
FAME production. In addition, the model did not include the effect biodiesel from pure algal strains [87,88] and mixed cultures of
of moisture on the FAME conversion. Complete drying of micro- microalgae [6,84]. However, microalgae are aquatic species. Their
algae prior to biodiesel (FAME) production is energy intensive and dewatering usually results into 5–35% solids [89,90]. Approxi-
has been a critical factor in blocking commercial production of mately 20–30% of algal biomass production cost is accounted for
algal FAME production. Salam et al. [78] reported the kinetics of by algal recovery [91]. Therefore the cost associated with drying of
in situ transesterification of microalgae that included side reac- wet microalgae, if required, is prohibitively high [8].
tions. Both their experiments and model showed that a maximum Besides, another drawback of this method is the requirement
96% biodiesel yield can be obtained in 10 min using excess me- for large amounts of methanol to oil molar ratio which usually in
thanol to oil molar ratio of 925: 1;excess catalyst concentration of the range of 100:1–1000:1. This is necessary since methanol plays
96% NaOH w/w oil before saponification FAME losses become a dual role: it acts as an oil extractor and as a reactant. Triglycides
significant. The process shows higher level of water tolerance than are non-polar lipids, so not readily soluble in methanol. In
1186 K.A. Salam et al. / Renewable and Sustainable Energy Reviews 65 (2016) 1179–1198

addition, microalgal cell wall's resistance causes large solvent re-

References
quirement during bioprocessing [10]. The main problem of such
high excesses of methanol is the capital and running (energy)

[68]

[68]
[70]
[70]

Laboratory, industrial [72]


Laboratory, industrial [73]
[71]

Laboratory, pilot scale [74]


(F)
costs associated with its recovery from the product streams, which

Laboratory, industrial
Laboratory, industrial
Laboratory, industrial

Laboratory, industrial

Laboratory, industrial
would almost certainly involve a substantial distillation column.
Scale of use

4.3. Overview of In situ transesterification

The feasibility of in situ transesterification of sunflower seed oil


(E)

was first reported in 1985 by Harrington and D’Arcy˗Evans [85].


They identified the following advantages:
Energy consumption MJ (kg dry

1. Esterification of the oil embedded in the hull, which could im-


prove the overall yield of the alkyl ester.
2. Reduction of the oil losses from the hull/kernel separation.
9.6 (Microwave only)

3. Esterification of the lipid, which may not be extracted by the


hexane due to its different solubility from triglyceride.
140 (modelled)

4. Improvement of carbohydrate digestibility of the residue by acid


or alkali catalyst interaction.
mass)  1

529

504

420
132

In addition, they observed the approach produced higher


(D)

68

33

conversion of alkyl ester than the two step transesterification. Both


processes produced the same quality of alkyl ester. The technical
Calculated energy consumption GJ m  3 cell

feasibility of making biodiesel from microalgae via in situ trans-


esterification has also been demonstrated by various investigators
[6,51,78,84].

4.4. Key process variables in in situ transesterification of microalgae


1.2 (Microwave only)

Some of the process variables that determine the FAME yield/


rate during in situ transesterification of microalgae for FAME are
1.4 (modelled)

shown in Table 10. The results of some in situ studies on plants are
suspension

also included for comparison.


Generally, the following observations can be made from the
4.5
0.7

2.5

0.3
(C)

2.1
1.1

Table:
Botryococcus, Chlorella, Scenedesmus (100 mL, 5 kg m  3, 700 W,
Botryococcus, Chlorella, Scenedesmus (100 mL, 5 kg m  3, 840 W,

1. FAME can be produced via in situ transesterification from both


Chlorococcum sp. (200 mL, 8.5 kg m  3, 2.5 kW, 6 min, High)

Hydrodynamic cavitation Saccharomyces cerevisiae (50 L, 10 kg m  3, 5.5 kW, 50 min,


Chlorococcum sp. (200 mL, 8.5 kg m  3, 750 W, 5 min, low)

Saccharomyces cerevisiae (0.8 L, 10 kg m  3, 600 W, 15 min,

freshwater and marine microalgae and other oil-bearing feedstocks.


2. The process requires either homogeneous alkali, acid, or het-
Scenedesmus (100 mL, 75 kg m  3, 1.2 kW, 1 min, high)
Overview of experimental cell disruption techniques and their energy consumptions.

erogeneous catalyst to proceed at a reasonable rate.


3. The process requires high mixing rate (150–500 rpm) to achieve
Mathematical modelling on an industrial scale

a reasonable FAME conversion.


4. When a heterogeneous catalyst is used a larger amount of
HPH: High pressure homogenization; HSH: High speed homogenization.

methanol is needed than for either homogeneous alkali or acid


Substrate & experimental conditions

catalyst, which is probably due to phase transfer limitations.


However, inclusion of co solvents such as hexane or methylene
dichloride helps to reduce the amount of methanol.
5. The reaction can proceed without catalyst with supercritical
methanol.
Column D¼ Column C/(Concentration in column B).

6. Microalgae in situ transesterification can occur at room tem-


perature particularly with alkali catalyst. Whalen et al. [6]
5 min, high)

5 min, high)

observed significant increase in FAME conversion rate with


medium)

medium)

increase in temperature from 60 to 80 °C but no significant


change in the rate was observed by increasing the temperature
(B)

from 80 to 110 °C strongly due to evaporation of methanol.


Similarly, Ehimen et al. [93] recorded significant in FAME
Microwave þ solvent

conversion rate when the temperature was increased from 30


to 60 °C, but observed no significant change in FAME conversion
Disruption technique

rate between 60 and 90 °C.


Freeze drying
Microwave
Bead mills
1. Sonication

7. FAME yield during in situ transesterification depends on a number


of variables, which include microalgae species, temperature, cat-
2. HPH
3. HSH

alyst/oil molar ratio, methanol/oil molar ratio, agitation rate,


Table 9

moisture content of the reactants or the feedstock, reaction time,


(A)

4.

5.

6.

8.
7.

phase and type of the catalyst and co solvent.


K.A. Salam et al. / Renewable and Sustainable Energy Reviews 65 (2016) 1179–1198 1187

Fig. 3. Overall transesterification reaction.

Fig. 4. A simplified homogeneous alkali transesterification schemes.

4.4.1. Solvent microalgae) with 6 mL hexane as co-solvent. Zeng et al. [96] ob-
The solvent plays a dual role during in situ transesterification. It tained 98% FAME yield during alkali catalysed reactive extraction
functions as an extractant and a reactant. Methanol is the most of sunflower at 20 °C , 101:1 methanol to oil molar ratio with 58 ˗
commonly employed solvent because it is cheaper than all other 1 diethoxyl methane (DEM) to methanol molar ratio as co˗solvent.
aliphatic alcohols. It is also less expensive to recover than ethanol The two investigations did not include the FAME yield that would
because it does not form an azeotrope with water [46]. Ordinarily, be produced in the absence the co-solvents (i.e. hexane and DEM).
methanol is the poorest extractant of triglycerides among aliphatic However, from the overview of parameters influencing the effi-
alcohols [98]. This is because the dissolution of triglyceride in- ciency of reactive extraction shown in Table 10, the effect of the
creases with increase chain length of the alcohol [98]. Ester yield co˗solvent can be clearly seen, in that they have the lowest me-
during acid-catalysed in situ transesterification of soy bean oil thanol to oil ratio. However, the co˗solvent should be carefully
increased with decrease in the polarity of the alcohol [98]. In screened for health and environmental hazards. Many of these
contrast, the ester yields during acid˗catalysed in situ transester- co˗solvents have significant environmental impacts, and would
ification of C. gracilis with methanol, ethanol, 1˗butanol, 2˗me- significantly adversely affect the processes’ life cycle carbon
thyl˗1˗propanol, and 3˗methyl˗1˗butanol did not vary significantly emissions. Co˗solvents that are difficult to separate from other
[6]. This could be because soybean oil and C. gracilis are not similar species in the reaction mixture can reduce the purity of biodiesel
in terms of cell wall chemistry and transesterifiable lipid. Thus the [84]. Furthermore, there would be an added process cost for the
activity of the alcohol during the reactive extraction of the bio- separation of the co-solvent.
diesel should not be expected to be the same. In situ transester-
ification is always characterized by a large amount of methanol to 4.4.2. Temperature and reaction time
oil ratio, between 100:1–1000:1 methanol to oil molar ratio Temperature affects the yield and/or rate of alkyl ester depending
[11,92,93]. on the type of catalyst. It has been reported that there was no sig-
Co-solvents can be used to reduce the methanol molar ex- nificant difference during alkali in situ transesterification of Jatropha
cesses. For instance, Li et al. [88] obtained 95% FAME yield during curcas and soybean oil between 30 and 65 °C [94,97].
acid-catalysed reactive extraction of Chlorella pyrenoidosa at 90 °C, However, acid-catalysed in situ esterifications increase in rate
methanol to oil molar ratio of 165:1 (4 mL methanol to 1 g with temperature. Wahlen et al. [6] observed an increase in FAME
1188 K.A. Salam et al. / Renewable and Sustainable Energy Reviews 65 (2016) 1179–1198

Fig. 5. Homogeneous acid catalysed transesterification scheme.

continuous stirring, stirring for only 1 h, stirring for 1 h on and 1 h


off and not stirring at all on the FAME conversion during in situ
transesterification. They observed that mixing is positively affect-
ing FAME conversion and not stirring at all caused significantly
low FAME conversion. Stirring for only 1 h produced in 8 h, a
maximum 91% FAME conversion of the amount obtained in con-
Fig. 6. Acid˗catalysed esterification of free fatty acid to alkyl ester. tinuous stirring for 1 h. They obtained no significant difference in
FAME conversion using 1 h on and 1 h off stirring treatment and
yield/rate during acid˗catalysed in situ transesterification of C- continuous stirring. This mixing treatment is therefore suggested
gracilis with increase in temperature from 20 to 150 °C. They found for energy savings during in situ transesterification of microalgae.
that most significant changes occurred between 60 and 80 °C. Si- Kasim and Harvey [94] studied the effect of mixing intensity (100–
milarly, Ehimen et al. [93] reported an increase in biodiesel rate 400 rpm) on alkali-catalysed reactive extraction of FAME from
during acid-catalysed in situ transesterification of Chlorella oil from Jatropha curcas at 60 °C and catalyst concentration (0.1 N). They
23 to 90 °C. They also found that no significant changes occurred observed that the FAME yield increased with increase in mixing
in the range 60–90 °C. The reaction time during in situ transes- rate. The process was independent of mixing speed at 300 rpm as
terification depends also on the nature of catalyst. Just like two  90% FAME yield was obtained at that condition which was not
step transesterification in which alkali catalyst is  4000 times significantly different with that of 400 rpm. At 100 rpm, the seed
faster than acid catalysed transesterification [79], alkali in situ settled down at the bottle of the reaction vessel while the FAME
transesterification is faster than its acid-catalysed counterpart. For yield significantly reduced to 37.2%.
instance, Velasquez˗orta et al. [51] reported a FAME yield (97%) at
60 °C during acid-catalysed reactive extraction of Chlorella vulgaris 4.4.4. Catalyst
at 20 h. They obtained 78% FAME yield in 1.25 h using alkali cat- In situ transesterification of lipid-bearing feedstocks requires a
alyst at the same conditions. Salam et al. [78] used 925:1 methanol catalyst for it to proceed at “mild” conditions [94,95]. Various
to oil molar ratio, 96% w/w oil alkali concentration at 60 °C to catalysts have been used for in situ transesterification, including
achieve 96% FAME yield in 10 min during in situ transesterification homogeneous alkali acid catalysts, and heterogeneous catalysts.
of Chlorella vulgaris. In practice, it is economical to operate at re-
action temperatures close to the boiling point of the alcohol [93]. 4.4.5. Acid-catalysed reactive extraction
Harrington and D’Arcy˗Evan [85] in their pioneering research
4.4.3. Agitation rate demonstrated the feasibility of acid-catalysed in situ transester-
A high mixing rate should be used during in situ transester- ification of sunflower oil. They obtained a 40% yield with H2SO4
ification to obtain a high FAME yield. As can be seen in Table 5, concentration of 1.2% v/v of methanol in  4 h reaction. In contrast
mixing rate of 150–500 rpm [51,84,93–95] are considered to be a 30% yield was obtained with hexane-extracted oil from sun-
enough to achieve a high FAME yield but some researchers always flower with the same process conditions. Kildiran et al. [98] con-
ensure that a high mixing rate which facilitate a complete sus- ducted an extraction and acid-catalysed in situ transesterification
pension of particles in the reaction vessel is sufficient to achieve a of soybean oil using methanol, ethanol, n-propanol and n˗butanol.
high yield [93,97] . Ehimen et al. [93] studied the effect of They found that in situ transesterification sequentially proceeds
K.A. Salam et al. / Renewable and Sustainable Energy Reviews 65 (2016) 1179–1198 1189

Fig. 7. Proposed reaction scheme for NaOH  catalysed reactive extraction of microalgae for FAME production. FAME: fatty acid methyl ester (Biodiesel); TAG: triacylgly-
cerides; FFA: free fatty acid; OH  : hydroxide species; CH3O  : alkoxide species (the actual catalyst), MeOH: methanol.
Source: Salam et al. [78].

Fig. 8. Comparison between in situ and two˗step transesterification.

through oil dissolution and transesterification of triglyceride and moisture content at 65 °C; 308:1 alcohol to oil molar ratio. Zhao and
the triglyceride dissolution increased with increasing alkyl chain Liu [92] conducted acid-catalysed in situ transesterification on a di-
length of the alcohol. verse species of oil-bearing feed stocks (Lipomyces starkey, Mortierella
In microalgae research, acid˗catalysed in situ transesterification of isabella, Rhodosporidium toruloides) and obtained a maximum FAME
microalgae oil at high yields have been demonstrated by many au- conversion of 98% at 70 °C, 868:1 alcohol:oil molar ratio at 20 h.
thors. Li et al. [88] reported a 95% FAME conversion in 2 h during Similarly, Wahlen et al.[6] reactively extracted FAME from dif-
acid-catalysed in situ transesterification of Chlorella pyrenoidosa at ferent cultures of microalgae biomass with acid catalysts and ob-
90 °C, 154:1 methanol: oil molar ratio, 0.234: 1 H2SO4: oil molar with tained a maximum FAME conversion of 77% at 0.33 h, 80 °C, and
hexane as co-solvent. They reported hexane to be an effective co- 1831:1 M ratio of methanol to oil. They found that the water tol-
solvent for reducing methanol-oil molar ratio. Velasquez-orta et al. erance of the process increased to as high as 400% dry weight of
[51] obtained a 97% FAME conversion in 20 h during reactive ex- Chaetoceros gracilis. However, the molar ratio of methanol to oil, at
traction of Chlorella vulgaris at 60 °C, 600:1 methanol:oil molar ratio, 3460:1 was very high. Downstream methanol removal would be a
0.35:1 acid to oil molar ratio. A maximum FAME conversion of 96% significant running cost at such molar ratios. This would have to be
was reported by Haas and Wagner [84] who performed acid catalysed weighed against the running cost savings due to the reduced
in situ transesterifications with algae biomass containing different feedstock drying duty.
1190 K.A. Salam et al. / Renewable and Sustainable Energy Reviews 65 (2016) 1179–1198

Table 10
Process conditions: in situ transesterifications of algal biomass.

Feedstock Temp. (°C) Solvent catalyst (oil Mixing Molar ratio Water con- Reaction Conversion (oil Remarks References
basis) (mol/ rate (rpm) (solvent: oil) tent % (w/w) time (hr) basis) (%)
mol) dry algae

Nannochloropsis Spe- 65 Methanol Mg˗Zr – 1569:1 0 4 60 Methylene [11]


cies (biomass) oxide dichloride
1.65:1 co˗solvent
Nannochloropsis Spe- 65 Methanol Mg˗Zr – 592:1 0 4 47 Methylene [11]
cies (Oil) oxide dichloride co-
1.65:1 solvent
Chlorella pyrenoidosa 90 Methanol H2SO4 – 154:1 0 2 95 Hexane [88]
0.234:1 co˗solvent
Chlorella vulgaris 60 Methanol H2SO4 380 600:1 0 20 97 [51]
0.35:1
Chlorella vulgaris 60 Methanol NaOH 0.15:1 380 600:1 0 1.25 78 [51]
Algae biomass 65 H2SO4 150 308:1 8 2 80 [84]
0.678:1
Algae biomass 65 Methanol H2SO4 150 308:1 1 2 86 [84]
0.678:1
Algae biomass 65 Methanol H2SO4 150 308:1 0.2 2 96 [84]
0.797:1
L.starkeyi 70 Methanol H2SO4 – 868:1 0 20 97 – [92]
0.093:1
HCl
0.186:1
M. isabellina 70 Methanol H2SO4 – 868:1 0 20 91 – [92]
0.093:1
R. toruloides 70 Methanol H2SO4 – 868:1 0 20 98 – [92]
0.093:1
HCl
0.186:1
Chaetoceros gracilis 80 Methanol H2SO4 – 988:1 0 0.33 82 [6]
0.158:1
Chaetoceros gracilis 80 Methanol H2SO4 – 1,977:1 100 0.33 67 [6]
0.158:1
Chaetoceros gracilis 80 Methanol H2SO4 – 3,460:1 400 0.33 57 [6]
0.158:1
Chlorella sorokiniana 80 Methanol H2SO4 – 1,831:1 dry 0.33 77 – [6]
0.158:1
Synechococcus 80 Methanol H2SO4 – 2,354:1 dry 0.33 40 – [6]
elongatus 0.158:1
80 Methanol H2SO4 – 3,013:1 dry 0.33 74 – [6]
0.158:1
60 Methanol H2SO4 500 314:1 dry 8 92 [93]
8.49:1
Jatropha curcas 30 Methanol NaOH 300 400: 1 0.5 88 o0.71 mm [94]
2.4:1 Particle size
Rapeseed 30–60 Methanol NaOH 200 600 :1 o 6.7 wt% 1 85 [95]
2.1:1
Sunflower 20 Methanol NaOH – 101:1 4.6% 0.2 98 DEM co˗solvent [96]
0.5:1
Soybean NaOH – [97]
23 Methanol 2:1 543:1 8
60 Methanol 1.6:1 226:1 8 84

DEM: Diethoxymethane.

It is interesting to note that the times to reach a maximum significant difference when 23 °C or 60 °C was used. However, they
FAME conversion vary between the microalgae species, as shown found that operating at 23 °C required larger methanol oil molar
in Table 10. The times range from 0.33 to 20 h. This is expected ratios (543:1) than (226:1), which was needed for 60 °C. Other
since microalgae are very diverse (more diverse than plant and researchers also observed no significant change in FAME conver-
animal kingdoms), and in particular have a wide range of cell wall sion between 23 °C and 60 °C in alkali-catalysed reactive extrac-
compositions. The difference in their cell wall compositions could tion of FAME from Jatropha curcas [94] and rapeseed [95]. An
have a significant effect on the time to reach optimum FAME 87.8% FAME conversion was obtained by Kasim and Harvey [94] for
conversion. Another major advantage of using acid catalysts for Jatropha curcas L. Similarly, Zakaria and Harvey [72] who worked
in situ transesterification is that they are more tolerant to high free on rapeseed/NaOH/methanol system obtained an 85% FAME con-
fatty acid concentrations. version at optimum conditions. Zeng et al. [96] observed a 97.7%
FAME conversion with sunflower/NaOH/methanol/DEM. However,
4.4.6. Alkali-catalysed reactive extraction this entailed the inclusion of an extra solvent, which will lead to
A number of researchers have obtained high FAME conversions increased complexity in the separations train. While several pub-
in alkali˗catalysed in situ transesterification of oil-bearing feed- lications have been conducted on the alkali reactive extraction of
stocks. Haas et al. [97] reported catalysed reactive extraction of oil seeds as indicated above, few studies have been conducted
soybean oil, where he obtained an 84% FAME conversion with no using alkali catalyst for reactive extraction of microalgae, which
K.A. Salam et al. / Renewable and Sustainable Energy Reviews 65 (2016) 1179–1198 1191

has created a gap in generalising the schemes of reaction exhibit two step transesterification. Velasquez˗orta et al. [7] found that a
by alkali reactive extraction. Velasquez-orta et al. [51] reported a reactively extracted wet Nannochloropsis cell at 1.5% moisture
78% FAME at 60 °C during alkali-catalysed reactive extraction of content has equal FAME yield or higher than a dried cell with both
Chlorella vulgaris in 1.25 h. Salam et al. [78] used 925:1 methanol acid and methoxide catalyst. However, they observed a decrease in
to oil molar ratio, 96% w/(w oil) alkali concentration at 60 °C to the FAME yield at 10% moisture. They found similarly moisture
achieve 96% FAME yield in 10 min during in situ transesterification tolerances with Chlorella cells for a moisture content which was
of Chlorella vulgaris. They achieved this high FAME yield despite not greater than 1.5%. Salam et al. [78] observed during alkali-
high levels of free fatty acid (6% lipid) in Chlorella vulgaris. This catalysed in situ transesterification of Chlorella vulgaris that at
shows high catalyst concentration and excess methanol to oil 600:1 methanol to oil molar ratio, the process was not adversely
molar ratio but short reaction time can be used to avoid saponi- affected by up to 20% moisture in the microalgae. Similarly, during
fication FAME losses and to produce enhanced rate during alkali- acid˗catalysed in situ transesterification of C. gracilis, Wahlen et al.
catalysed in situ transesterification of microalgae. [6] achieved a 57% FAME yield with 3460:1 methanol: oil mole
ratio at 400 wt% moisture content compared to a 82% FAME yield
4.4.7. Heterogeneous-catalysed reactive extraction with 988:1 methanol: oil mole for fully dried cells. The difference
A heterogeneous catalyst has also been used to promote in situ in the moisture tolerance obtained by these authors is possibly
transesterification [11]. They compared the FAME yield obtained due to techniques used to quantify the moisture content. While
during in situ transesterification of Nannochloropsis and two step Velasquez-orta et al. [7] used bound water in the algae, Salam et al.
transesterification of Nannochloropsis oil with a heterogeneous [78] and Wahlen et al. [6] used free water. Zakariah and Harvey
(Mg2Zr5O12) catalyst using methylene dichloride as a co-solvent. A [95] also observed some level of moisture tolerance during alkali-
maximum FAME yield of 60% was observed at 65 °C, 4 h, 10 wt% catalysed reactive extraction of rapeseed oil for FAME. However,
catalyst and 45 mL of mixed solvent (3:1 v/v methanol/methylene they observed a drastic reduction in the FAME yield when the
chloride). A lower FAME yield of 47% was obtained with two step moisture content was greater than 6.7 wt%. Water tolerance of
transesterification using the same process conditions as the in situ in situ transesterification is caused by excess methanol to oil molar
transesterification. The authors found that the FAME yield increased ratio but removal of the unreacted methanol ( 94% of it) usually
with increased catalyst concentration and volume of mixed solvent by distillation would increase operating costs. It has been shown
but further increases in these two parameters lead to reduced yields. that reducing the water content reduces the excess methanol re-
Recently, micro-algal biodiesel production via a two-step in situ quired: Haas and Scott [102] reported that a 60% reduction in
transesterification was reported by Dong et al. [99]. This process in- methanol and a 56% reduction in NaOH were achieved when fully
volved a two-step in situ transesterification, where the algae free fatty dried soybean was used than when the bean contained 2.6 wt%
acid was reduced with Amberlyst-15 before alkali in situ transester- moisture content. No researchers have investigated the effect of
ification. They obtained a maximum FAME conversion of 94.9%. including SDS in H2SO4 on the water tolerance of the in situ
transesterification of microalgae. This is important, as the sig-
4.4.8. Non-catalysed reactive extraction nificant amounts of energy required to completely dry microalgal
It is possible for two step and in situ transesterifications to biomass or microalgal oil to the levels required in two steps bio-
proceed without catalyst in supercritical water or alcohol. Saka diesel production render the process uneconomic. This is currently
and Kusdiana [100] reported transesterification of rapeseed oil in one of the major technical challenges to microalgal biodiesel
supercritical methanol. They obtained a  95% FAME yield in 4 h at production. Complete drying of algae is energy intensive, which
350 °C, 45 Mpa and 1:42 oil to methanol molar ratio. Similarly, Lim significantly increases the cost of algae pre-treatment.
et al. [101] conducted in situ transesterification of Jatropha curcas L
seeds with supercritical methanol with the aid of hexane as co-
solvent. A maximum FAME yield of 103% (w/w) was obtained at 5. Microalgae species
300 °C, 240 MPa, 10 mL/g methanol to solid ratio and 2.5 mL/g
hexane to seed ratio. This FAME yield was greater than theoretical In a practical biodiesel production, selection of a suitable algal
FAME yield obtained from n-hexane Soxhlet extraction. This in- species is an important factor, since there are differences between
dicates that FAME was extracted from components of the biomass lipid content and biomass productivity among different species and
besides triacylglycerides possibly polar lipids such phospholipids. even within the same algal species. It can be seen from Table 11, that
The advantages of this method over catalysed transesterification, the Nannochloropsis species is a competitive biodiesel candidate
as reported by Saka and Kusdiana [100], include: among marine microalgae, as it is relatively high-yielding and
productive. Nannochloropsis occulata have been reported to accu-
1. Shorter time to completion. mulate as much as 60% lipid (w/w dry algae) but this high lipid yield
2. Simpler process, requiring fewer purification steps. was achieved only in a nitrogen-depleted medium [103]. However,
3. The ester yield is greater than catalysed process. lipid content of 8–35% lipid (w/w dry algae) is common in this
species if the cells are not environmentally stressed [103] Amongst
The major drawbacks of the process are that it operates at high freshwater microalgae, Chlorella species are suitable due to their
temperature and pressure, leading to significant increases in capital substantial lipid accumulation. Chlorella species have been reported
cost, and increased costs associated with safety and monitoring. to accumulate 19–66% lipid content (w/w dry algae) but this high
yield is achieved under heterotrophic conditions using urea [104] or
4.4.9. Moisture content rice straw hydrolysate [103] as carbon source. Marine microalgae
One major challenge in biodiesel production is the need for dry have additional advantages over their freshwater counterparts as
feedstocks. The moisture limit for common biofuel feedstocks is they do not require fresh water, so could not compete with food
0.5 wt% [83], and in practice lower moisture limits are preferred. crops for this resource. Some types of microalgae, such as Nanno-
Water, in the feedstock or the methanol causes a significant re- chloropsis occulata, Dunaliella tertiolecta [105], Pavlova lutheri, Ske-
duction in the yield of biodiesel [83]. It also results in soap for- letonema sp, Chaetoceros calcitrans, Tetraselmis sp. and Isochrysis sp.
mation in alkali-catalysed transesterification, leading to increased [103] as can be seen in Table 11, can be cultivated on brackish or sea
complexity in the product separation train [83]. However, in situ water, thereby posing no threat to freshwater for human con-
transesterification has been shown to be more water-tolerant than sumption or for agricultural use.
1192 K.A. Salam et al. / Renewable and Sustainable Energy Reviews 65 (2016) 1179–1198

5.1. Protein and carbohydrate content of algae in anaerobic digestion. Approximately 20–30C/N ratio is considered
as optimum for anaerobic digestion process [123]. A significant high
In addition to lipids, microalgae contain substantial levels of C/N ratio leads to rapid consumption of nitrogen by methanogens for
proteins and carbohydrates [57]. Results of the study conducted by its metabolic protein requirement resulting into low biogas produc-
Becker [57] on the biochemical compositions of some microalgae tion while a low C/N causes accumulation of ammonium ion (NH4)
are illustrated in the Table 12. It shows that significant portions of [108]. Excess NH4 raises the pH of the bio˗digestate above the opti-
algal biomass were carbohydrate and protein, which means if mum (6.7–7.5 pH) causes inhibition of methane production [108].
utilise after the reactive extraction as added value products could Co˗digestion of low C/N substrate with a high C/N substrate is one
improve the process's economy. probable solution to reduce its inhibitory effect. Yen and Brune [124]
achieved a significant increase in methane production from 573775
to 1170775 mL CH4/day when algal sludge was co-digested with
6. Probable ways of making a sustainable in situ waste paper.
transesterification When whole algae is used for anaerobic digestion, effect of long
chain free fatty acids (LCFAs) should be considered as they inhibit
6.1. Anaerobic digestion of algal residue after reactive extraction biodegradability of hydrolytic bacteria, acidogens, acetogens, and
methanogens community resulting into low biogas production
Algal residue after reactive extraction may contain carbohy- [116–120]. Their inhibitory effects increase with increase in the
drate, protein, and unconverted lipids. Therefore, it can be used as chain length [125], presence of double bond [126,127], and in-
feed for either bioethanol or biogas production. If the residue crease in the number of double bonds [125,128]. Use of adequate
contains no toxic substances, it could serve as supplement for amount of inoculum to substrate (I/S) ratio has been effective for
animal feeds. Biogas production through anaerobic digestion is reducing LCFAs inhibitory effects [111,129,130]. A I/S ratio of 1–2 is
considered as better option due to its efficient conversion of all the considered to be optimum [111]. Similarly, addition of calcium ion
residual carbohydrate, proteins and unconverted lipids to final [111,130] has been shown to partially reduce the LCFAs inhibition
products [108] as shown in the Fig. 9. particularly when the I/S ratio met the threshold level [130].
Anaerobic digestion is a biological breakdown (bio˗depoly- However, inclusion of calcium ion has no significant effect on
merisation) of complex organic compounds by consortium of LCFAs when residual algae was used [111] obviously due to low
bacteria in the absence of oxygen to produce biogas. It occurs level of LCFAs in the algal residue. Therefore when spent algal
naturally in a landfill or artificially in a digester under a control residue after reactive extraction serve as feed in anaerobic diges-
conditions. The resulting biogas usually contains methane tion, LCFAs inhibition should not be a major concern as both the
(  60%), CO2 (  40%) and other trace gases [109]. Due to occur- neutral lipids (TAG) and part of the membrane lipids would have
rence of a naturally induced anaerobic digestion in a landfill, been converted into biodiesel causing significant low level of
spent microalgae or degradable organic compounds should be LCFAs in the residue.
disposed of in an environmentally sustainable way.as methane's
greenhouse gas effect was  20 times greater than CO2 [110]. 6.2. marine species for biofuel production
Whole algae and spent algal residue of lipid extraction can be
digested anaerobically to produce biogas [111,112] but little or no Using fresh water microalgae are still not economically attrac-
information are available in the literature on the use of algal tive and cannot support a large scale biodiesel production [2] due
residue resulting from in situ transesterification of microalgae for to a huge fresh water footprint required [131]. A water footprint of
biogas production. Apparently algal residues obtained from lipid-  3700 kg kg  1 of biodiesel is estimated for algae cultivated in
extracted microalgae and reactive extracted (in situ transester- open pond in fresh water without water recycling but recycling
ification) microalgae would be different due to varying process reduced the water requirement to 600 kg.kg–1 of biodiesel. In
conditions. One major difference is that in situ transesterification contrast, the water footprint plus the compensated fresh water for
involved methanol containing catalyst (acid or alkali) which evaporation for algae cultivated in open pond in seawater was
might cause disruption of the algae cell wall resulting in en- 370 kg kg–1[131]. Since there was a significant decrease in the
hancement of biogas production. More research effort should be freshwater footprint when marine species was used, this species is
directed towards utilisation of such residue for biogas produc- a good alternative to freshwater species to achieve commercial
tion. It could contribute to the improvement of the economy of reality of algal fuel. Apparently, residue from marine species will
in situ transesterification. Besides, the biodigestate can be re- contain high salt concentration resulting in salinity inhibition
covered as manure to support algal cultivation [113] and the during anaerobic digestion [121,122]. This causes bacteria cells
biogas (methane) produced can be burnt to supply the electrical death and inhibition of reaction pathways in anaerobic digestion
energy required for separating or concentrating the dilute algal resulting to low biogas production [132]. Probable techniques of
biomass-water mixture [114,115]. Burning of methane can also reducing salinity inhibition include using municipal bio-solids
supply the energy required for drying the biomass particular in a digestate as inoculum [133], use of halophilic microorganisms or
region where sun˗drying is not a feasible option. halo-tolerant microorganism obtained from a long acclimation as
inoculum [132] or washing of the biomass with fresh water to
6.1.1. Causes of inhibition during biogas production reduce its salinity [2]. Although utilisation of residue for biogas
The yield of methane produced during anaerobic digestion of production will improve the economy of the process, significant
microalgae strongly depends on a number of inhibitory agents and cost during in situ transesterification resulting from algae drying
conditions. These inhibitors include carbon/nitrogen (C/N) ratio [108], and recycling of the excess unreacted methanol.
long chain free fatty acids (LCFAs), [116–120], extraction solvents
[111], salinity [121,122] particularly for marine species. Becker [57] 6.3. In situ co-production of diethyl-ether during in situ
showed that most microalgae contained protein more than carbo- transesterification
hydrate. Considering the fact that spent residue after reactive ex-
traction (in situ transesterification) would certainly have low lipid In situ transesterification of microalgae for biodiesel production
content due to high percentage of it being converted to biodiesel, C/N requires excess methanol due to mass transfer limitation resulting
ratio is therefore a major concern when such residue is used as feed from solid (algae) and liquid (catalyst plus methanol) involved.
Table 11

K.A. Salam et al. / Renewable and Sustainable Energy Reviews 65 (2016) 1179–1198
Different microalgae species, habitat, lipid content, biomass productivity and applications.

Algae species Habitat Lipid content % w/ Biomass productivity mg/L/ Applications Remarks References
w day

Nannochloropsis species Marine, fresh and brackish 60 300 Biofuel and feed supplement Nitrogen deprived medium promoted lipid content [103]
water
Nannochloropsis oculata Marine, fresh and brackish 7.90–15.86 – Biofuel Nitrogen deprived medium and temperature in- [103]
water fluence lipid content
Nannochloropsis species Marine , fresh and brackish 29.6–35.5 170–210 Biofuel CO2 enriched air; continuous illumination for [103]
water cultivation
Dunaliella tertiolecta Marine 15.20 28 CO2 capturing; Biofuel/ wastewater treatment – [105]
Pavlova lutheri Marine 35.50 140 Biofuel CO2 enriched air; continuous illumination for cul- [103]
tivation 2
Skeletonema species Marine 31.80 90 Biofuel CO2 enriched air; continuous illumination [103]
Chaetoceros Calcitrans Marine 39.80 40 Biofuel CO2 enriched air; continuous illumination for [103]
cultivation
Tetraselmis sp. species Marine 12.90–14.70 280–300 Biofuel CO2 enriched air; continuous illumination for [103]
cultivation
Isochrysis species Marine 22.90–27.40 140–170 Biofuel CO2 enriched air; continuous illumination for [103]
cultivation
Scenedesmus species Freshwater 12.80–21.10 126.54–260 Waste treatment/Biofuel;CO2 mitigation – [103,106]
Chlorella protothecoides Freshwater 50.30–55.20 2020–7300 Biofuel Heterotrophic condition enhanced lipid [107]
accumulation
Chlorella species Freshwater 19.30–66.10 230 Biofuel, food supplement, sorption of toxic chemical Urea was used as a source of low cost Nitrogen [103,104]
Chlorococcum species Freshwater 19.30 280 CO2 mitigation CO2 enriched air; continuous illumination for [103]
cultivation
Chlorella pyrenoidosa Freshwater 56.30 1100 Biofuel Rice straw hydrolysate served as carbon source [103]

1193
1194 K.A. Salam et al. / Renewable and Sustainable Energy Reviews 65 (2016) 1179–1198

After completion of the reaction the unreacted methanol (  94% of pressure (  5 MPa) there could be additional capital cost and in-
it) needs to be removed from the product streams and this re- creased cost associated to safety and monitoring. However, this
quires substantial distillation heat load. Removal of this excess cost will be substantially less than producing biodiesel from pre-
methanol has been a major drawback of in situ transesterification. extracted oil at supercritical condition of methanol such as 350 °C
It significantly contributes to the high operating cost of this pro- and 45 MPa reported by Saka and Kusdiana [100] and in situ
cess rendering it uneconomical. transesterification of the algae at supercritical methanol condi-
A more appealing process will be to co-produce dimethyl ether tions such as 300 °C and 240 MPa reported by Lim et al. [101].
(DME) in situ along with biodiesel during in situ transesterification
of microalgae. This would require conducting the in situ transes- 6.3.1. Fuel properties of DME or DME-Biodiesel blends
terification at conditions favourable to methanol etherification. Apart from potential improve in the economy of in situ trans-
DME can be produced in good yield at 260 °C and 5 MPa [134]. esterification, blending DME with biodiesel would also improve
Conducting the reaction at this conditions can also serve as pre- the fuel quality. For instance, blending DME (5–15% by volume)
treatment for the residue when used in biogas production. Com- with biodiesel obtained from Jatropha oil has been shown to in-
bination of acid/alkali with thermal pre-treatment [135] as well as crease the brakes thermal efficiency (BTE), decreased the brakes
thermal pre-treatment alone has been shown to cause enhance- specific fuel consumption (BSFC) and decrease the emission levels
ment in biogas production [135,136]. This enhancement occur due of all blends as shown in Table 13 by Loganathan et al. [137].
to significant increase in carbohydrate dissolution (  5 fold) with DME is a green fuel. It is non corrosive to any metals, contains
thermal pre-treatment alone and (  7 fold) with combination of neither sulphur non nitrogen, non-harmful to human body and
thermal and acid. Pre-treatment. Although the use of thermal pre- has physical and combustion characteristics comparable to other
treatment for enhancing biogas production alone is energy in- fuels. As can be seen in Table 14, it has high cetane number
tensive and resulted in negative energy balance [135], considering (55  60) comparable to diesel cetane number (40 55) so it can be
the fact that dimethyl ether would be produced along with bio- used as perfect alternative for diesel or as blends [134].
diesel might improve the energy gain as there will not be the need Handling of DME does not pose serious environmental concern.
to recycle significant amount of unreacted methanol. Since the Studies have shown that DME has low toxicity potential, no gen-
process would operate at high temperature (  260 °C) and high otoxicity potential in vitro and sub˗chronic and no significant
toxicity effect as shown in cancerogenicity inhalation studies by
Table 12 European Food Safety Authority [138].
Major chemical composition of different algae. Fig. 10. shows a schematic diagram of the proposed new in-
tegrated process of co˗producing DME in situ along with biodiesel
Algae Protein (%) Carbohydrate (%) Lipids (%) Nucleic
acid (%) production via in situ transesterification as described in Section 6.3.
Fig. 10: shows a proposed new way of efficient utilising algae
Anabaena cylindrica 43–56 25–30 4–7 – residue for biogas generation and bio˗digestate for supporting al-
Aphanizomenon flos- 62 23 3 –
gae cultivation. As shown in the Fig. DME would be produced
aquae
Chlamydomonas 48 17 21 –
rheinhardii Table 13
Chlorella pyrenoidosa 57 26 2 – Engine performance and emission level of DME- Biodiesel blends.
Chlorella vulgaris 51–58 12–17 14–22 4–5
Dunaliella salina 57 32 6 – Blends BTE in- BSFC in- HC reduc- CO reduc- NOx reduc-
Euglena gracilis 39–61 14–18 14–20 – crease (%) crease (%) tion (%) tion (%) tion (%)
Porphyridium 28–39 40–57 9–14 –
cruentum BDE 5 6 8 42 67 17
Scenedesmus obliquus 50–56 10–17 12–14 3–6 BDE 10 10 12 36 51 28
Spirogyra sp. 6–20 33–64 11–21 – BDE 15 3 17 26 35 35
Arthrospira maxima 60–71 13–16 6–7 3–4.5
Spirulina platensis 46–63 8–14 4–9 2–5 BDE 5 (DME, 5% and biodiesel, 95%); BDE 10 (DME, 10% and biodiesel, 90%); BDE 15
Synechococcus sp. 63 15 11 5 (DME, 15% and biodiesel, 85%); BTE: Brake thermal efficiency; BSFC: Brake specific
fuel consumption.

Fig. 9. Methane and alcoholic route of conversion of algal residue. Modified from Chandra et al. [108].
K.A. Salam et al. / Renewable and Sustainable Energy Reviews 65 (2016) 1179–1198 1195

Table 14
Physical properties of DME and other fuels.
Source: Ogawa et al. [134].

Properties Chemical Boiling Liquid Specific Heat of vapor- Vapour Ignition Explosion Cetane nob Net calorific va- Net calorific
formula Point (K) Density Gravity ization (kJ kg  1) Pressure Temp. (K) limit lue value
(g cm  3)a (VS. air) (atm.)a 106 J (Nm)  2 106 J kg  1

DME CH3OCH3 247.9 0.67 1.59 467 6.1 623 3.4–17 55–60 599.44 28.90
Propane C3H8 231.0 0.49 1.52 426 9.3 777 2.1–9.4 (5)b 91.25 46.46
Methane CH4 111.5 – 0.55 510 – 905 5–15 0 36.0 50.23
Methanol CH3OH 337.6 0.79 – 1097 – 743 5.5–36 5 – 21.10
Diesel – 180.4 0.84 – – – – 0.6–6.5 40–55 – 41.86

a
(293 K).
b
(estimated value).

Fig. 10. Proposed in situ co-production of DME during in situ transesterification.

in situ along with biodiesel during in situ transesterification of the co-solvent itself and extra downstream separation duties.
microalgae using conditions favourable to biodiesel and DME Furthermore, dimethoxymethane, which has been evaluated for
synthesis. Therefore, there may be less or no methanol remain in this application is not a “green” solvent. It should also be borne
the product streams resulting in significant reduction in distilla- in mind that co-solvents can alter the range of products.
tion heat that might be require to remove the unreacted methanol.  Microalgae lipids are bound by a cell wall that inhibits FAME
Apparently acid catalyst would be used as it promotes etherifica- extraction. Cell disruptions have been shown to be effective in
tion. The high temperature required would help in disrupting the enhancing lipids extraction but are prohibitively energy in-
microalgal cells, and it will also shorten the reaction time. Me- tensive. The resistance provided by the cell wall causes addi-
thane produced can be burnt to provide energy for algal biomass tional excess requirement of solvent which translates to extra
processing such as concentration and separation from dilute algae- production cost. A cost effective pre-treatment technique will
water mixture [2,114]. The bio˗digestate can serve as manure to substantially improve the process economy.
support microalgae cultivation [113]. This approach could produce  Sulphuric acid is the most common homogeneous catalyst for
a sustainable integrated in situ transesterification with a potential reactive extraction of microalgae. Heterogeneous catalysts have
improve in the process economy. also been shown to be effective, but require co-solvents or even
greater excesses of methanol. When H2SO4 is used in reactive
extraction, a high concentration of the catalyst is always required
7. Concluding remarks to achieve high yield [6,7]. However, the need to neutralise the
unreacted acid in the product streams will increase operating
costs. Inclusion of SDS in H2SO4 catalyst for FAME enhancement
 In situ transesterification of microalgae to produce biodiesel has has not been explored in reactive extraction of microalgae.
been demonstrated to be technically feasible for a range of Considering the fact that surfactants can disrupt cell walls, their
marine and freshwater species. However, one major drawback is use could lead to enhancement of FAME yield and/or rate.
the relatively high molar ratio of methanol to oil (range: 100:1  Unlike acid catalysts, there are handful publications on use of
to 1000:1) required for desired yield. In addition complete alkali catalyst for reactive extraction of microalgae. Salam et al.
drying of microalgae to a level required by two step transes- [78] has shown that a 96% FAME yield can be obtained during
terification is energy intensive and significantly hinders the alkali-catalysed in situ transesterification in 10 min using excess
commercial production of large scale algae biofuels. methanol and alkali concentration but short reaction time. This
 The possible cost savings due to the increased water tolerance significant reduction in reaction time can save operating cost.
of reactive extraction of microalgae oil for FAME are consider-  Overall, an integrated approach of producing dimethyl ether
able [8], but they must be weighed against the costs of (DME) in situ along with the desired biodiesel and utilisation of
regenerating the alcohol (almost certainly by distillation). algae residue after reactive extraction for methane production
 Co-solvent use can reduce the molar excess to as low as 101:1 and the biodigestate for supporting algal growth may further
(dimethoxymethane). However, this introduces extra costs for reduce the cost of in situ transesterification. Becker (1994)
1196 K.A. Salam et al. / Renewable and Sustainable Energy Reviews 65 (2016) 1179–1198

reported that microalgae contained substantial carbohydrate Productivity and Sustainability, Sao Paulo: Editora Edgard Blücher, p. 697–
and protein besides the lipids used as the transesterification 716.
[32] Searchinger T, Heimlich R, Houghton RA, Dong F, Elobeid A, Fabiosa J, Tokgoz
substrate. This makes microalgae residue after reactive extrac- S, Hayes D, Yu T-H. Use of U.S. croplands for biofuels increases greenhouse
tion an attractive candidate for biogas production. Reactive ex- gases through emissions from land-use change. Science 2008;319(1):238–40.
traction/in situ transesterification of microalgae to biodiesel [33] Fargione J, Hill J, Tilman D, Polasky S, Hawthorne P. Land Clearing and the
biofuel carbon debt. Science 2008;319(1):235–8.
may well be a good basis for a microalgae-based bio˗refinery. [34] Licht FO. World Fuel Ethanol Production, op. cit. note 92, p. 365; Bill Guerin,
European Blowback for Asian Biofuels, Asia Times, 8 February 2007.
[35] Chisti Y. Response to Reijnders: do biofuels from microalgae beat biofuels
from terrestrial plants? Trends Biotechnol 2008;26(7):351 252.
References
[36] Becker EW, Venkataraman V. Production and processing of algae in pilot
plant scale experiences of the Ido-German project. In: Shelef G, Soeder CJ,
[1] Gui MM, Lee KT, Bhatia S. Feasibility of edible oil vs. non–edible oil vs. waste editors. Algae Biomass: Production and Use. Elsevier; 1980. p. 35–50.
oil as biodiesel feedstock. Energy 2008;33:1646–53. [37] Kawaguchi K. Microalgae production systems in Asia. In: Shelef G, Soeder CJ,
[2] Chisti Y. Constraints to commercialization of algal fuels. J Biotechnol editors. Algae Biomass: Production and Use. Elsevier; 1980. p. 25–33.
2013;167:201–14. [38] Pulz O. Photobioreactors: production systems for phototrophic micro-
[3] Chisti Y. Biodiesel from microalgae’. Biotechnol Adv 2007;25(3):294–306. organisms. Appl Micro Biotechnol 2001;57:287–93.
[4] D. Mitchell (2008). A Note on Rising Food Prices-World Bank Search; avail- [39] Macedo I, Seabra JEA, Silva JEAR. Greenhouse gas emissions in the produc-
able at: 〈http://search.worldbank.org/all? tion and use of ethanol from sugarcane in Brazil: the 2005/2006 averages
qterm¼ a þnoteþ on þrising þ food þ prices&title ¼&file type ¼〉 (Accessed: and prediction for 2020. Biomass Bioenergy 2008;32:582–95.
11/04/2015). [40] US DoE (2016) United State Department of Energy, alternative fuels data
[5] RA. Butler (2006). Why is oil palm replacing tropical rainforests? Why are center. 〈http://www.afdc.energy.gov/fuels/prices.html〉. Accessed June 2016.
biofuels fuelling deforestation? Available from: 〈http://news.mongabay.com/ [41] Nagarajan S, Chou SK, Cao S, Zhou Z. An updated comprehensive techno-
2006/0425-oil_palm.html〉 (Accessed: 11/04/2015). economic analysis of algae biodiesel. Bioresour Technol 2013;145:150–6.
[6] Wahlen BD, Willis RM, Seefeldt LC. Biodiesel production by simultaneous [42] Delrue F, Setier PA, Sahut C, Cournac L, Rouband A, Peltier G, Froment AK. An
extraction and conversion of total lipids from microalgae, cyanobacteria, and economic, sustainability, and energetic model of biodiesel production from
wild mixed-cultures. Bioresour Technol 2011;2011(102):2724–30. microalgae. Bioresour Technol 2012;111:191–200.
[7] Velasquez-orta SB, Lee JGM, Harvey A. Evaluation of FAME production from [43] Richardso JW, Johnson MD, Outlaw JL. Economic comparison of open pond
wet marine and freshwater microalgae by in situ transesterification. Biochem raceways to photo-bioreactors for profitable production of algae for trans-
Eng J 2013;76:83–9. portation fuels in the Southwest. Algal Res 2012;1(1):93–100.
[8] Lardon L, Sialve B, Steyer J, Bernard O. Life-Cycle assessment of biodiesel [44] Davis R, Aden A, Pienkos PT. Techno-economic analysis of autotrophic mi-
production from microalgae. Environ Sci Technol 2009;17:6475–81. croalgae for fuel production. Appl Energy 2011;88(10):3524–31.
[9] Barsanti L, Gualtieri P. Algae anatomy, biochemistry, and biotechnology. 2nd [45] Wood BJB. Fatty acids and saponifiable lipids. In: Stewart WDP, editor. Algal
ed. Boca Raton: CRC Press, Taylor and Francis Group; 2014. Physiology and Biochemistry. Oxford, London: Blackwell Scientific Publica-
[10] Gerken HG, Donohoe B, Knoshaug EP. Enzymatic cell wall degradation of tions; 1974. p. 236–64.
Chlorella vulgaris and other microalgae for biofuels production. Planta [46] Demirbas A. Biodiesel: A Realistic Fuel Alternative For Diesel Engines.Lon-
2012:1–15. don: Springer-Verlag; 2008.
[11] Li Y, Lian S, Tong D, Song R, Yang W, Fan Y, Qing R, Hu C. One-step production [47] Ben-Amotz A, Tornabene TG, Thomas WH. Chemical profile of selected
of biodiesel from Nannochloropsis sp. on solid base Mg–Zr catalyst. Appl species of microalgae with emphasis on lipid. J Phycol 1985;21(1):72–81.
Energy 2011;88:3313–7. [48] Patil V, Kallqvist T, Olsen E, Vogt G, Gislerod HR. Optimization of direct
[12] Dhar BR, Kirtania K. Excess methanol recovery in biodiesel production pro- conversion of wet algae to biodiesel under supercritical methanol condi-
cess using a distillation column: a simulation study. Chem Eng Res Bull tions. Bioresour Technol 2011;102(1):118–22.
2009;13:55–60. [49] Valeem EE, Shameel M. Influence of aquatic environment on the composi-
[13] El-Shimi HI, Attia NK, El-Sheltawy ST, El-Diwani GI. Biodiesel production tion of fatty acids in algae growing in Sindh. Pak Proc Pak Acad Sci 2009;46
from Spirulina-platensis microalgae by in-situ transesterification process. J (3):109–16.
Sustain Bioenergy Syst 2013:224–33; [50] Breuer G, Lamers PP, Martens DE, Draaisma RB, Wijffels RH. The impact of
British Petroleum (BP). Statistical Review of World Energy; 2011. nitrogen starvation on the dynamics of triacylglycerol accumulation in nine
[14] Fontes E, Nilsson E. Modeling the fuel cell. Ind. Phys. 2001;7(4):14–7. microalgae strains. Bioresour Technol 2012;124:217–26.
[15] Zuttel A, Remhof A, Borgschulte A, Friedrichs O. Hydrogen: the future energy [51] Velasquez-orta SB, Lee JGM, Harvey A. Alkaline in situ transesterification of
carrier. Philos Trans R Soc 2010;368:3329–42. Chlorella vulgaris. fuel 2012;94:544–50.
[16] Gaur S, Reed T. Thermal data for natural and synthetic fuels. New York: [52] Renaud SM, Parry DL, Thinh LV, Kuo C, Padovan A, Sammy N. Effect of light
Marcel Dekker; 1998. intensity on the proximate biochemical and fatty acid composition of Iso-
[17] ETB (2011). Engineering Tool Box. 〈http://www.engineeringtoolbox.com〉. chrysis sp. and Nannochloropsis oculata for use in tropical aquaculture. J Appl
Accessed June 2016. Phycol 1991;3:43–53.
[18] US DoE (2016).United State Department of Energy. 〈http://energy.gov/eere/ [53] Knothe G. Structure indices in FA chemistry: how relevant is the iodine va-
fuelcells/hydrogen-storage〉. Accessed June 2016. lue? J Am Oil Chem Soc 2002:847–54.
[19] Sakintuna B, Lamari-Darkrim F, Hirscher M. Metal hydride materials for solid [54] Mittelbach M, Remschmidt C. Biodiesel: the comprehensive handbook. 3rd
storage: a review. Int J Hydrog Energy 2007;32:1121–40. ed. Austria: Martin Mittelbach (Publisher); 2006.
[20] Jain IP, Lal C, Jain A. Hydrogen storage in Mg: a most promising material. Int J [55] Bucy HB, Marc E, Baumgardner ME, Anthony J, Marchese AJ. Chemical and
Hydrog Energy 2010;35:5133–44. physical properties of algal methyl ester biodiesel containing varying levels
[21] Kojima Y, Kawai Y. IR characterizations of lithiumimide and amide. J Alloy of methyl eicosapentaenoate and methyl docosahexaenoate. Algal Res
Compd 2005;395(1–2):236–9. 2012;1:57–69.
[22] Dillon AC, Jones KM, Bekkedahl TA, Kiang CH, Bethune DS, Heben MJ. Storage of [56] Komolafe O, Velasquez-orta SB, Monje-Ramirez I, Noguez IY, Harvey AP, Orta
hydrogen in single walled carbon nanotubes. Nature 1997;386(6623):377–9. Ledesma MT. Biodiesel production from indigenous microalgae grown in
[23] Costa PMFJ, Coleman KS, Green MLH. Influence of catalyst metal particles on wastewater. Bioresour Technol 2014;2014(154):297–304.
the hydrogen sorption of single walled carbon nanotube materials. Nano- [57] Becker W. Microalgae in human and animal nutrition. In: Richmond A,
technology 2005;16(4):512–7. editor. Handbook of microalgal culture: biotechnology and applied phycol-
[24] Ogden JM. Prospects for building a hydrogen energy infrastructure. Annu Rev ogy. Oxford, UK: Blackwell Science Ltd.; 2004. p. 312–51.
Energy Environ 1999;24:227–79. [58] Abo-Shady AM, Mohamed YA, Lasheen T. Chemical composition of the cell
[25] Press RJ, Massoud KSVS, Miri J, Bailey AV, Takacs GA. Introduction to hy- wall in some green algae species. Biol Plant 1993;35(4):629–32.
drogen technology.USA: John Wiley & Sons, Inc.; 2008. [59] Richmond A. Handbook of microalgal culture: biotechnology and applied
[26] Dears D (2015). 〈https://dddusmma.wordpress.com/2015/11/03/the-mas phycology.UK: Blackwell Science Ltd.; 2004.
sive-cost-of-fuel-cell-vehicles-part-1/〉. Accessed June 2016. [60] Atkinson A,W, Gunning Jr. BES, John PCL. Sporopollenin in the cell wall of
[27] Yerga RMN, Galvan MCA, Valle F, Mano JAV, Fierro JLG. Water splitting on Chlorella and other algae: ultrastructure, chemistry, and incorporation of
semiconductor under visible-light irradiation. Chem Sus Chem 2009;2 14C-acetate studied in synchronous cultures. Planta (Berl.) 1972;107:1–32.
(6):471–85. [61] Allard B, Templier J. Comparison of neutral lipid profile of various trilaminar
[28] Anja H, Anastasios M, Thomas H. Analytical approaches to photo biological outer cell wall (TLS)-containing microalgae with emphasis on algaenan oc-
hydrogen production in unicellular green algae. Photosynth Res 2009;102(2– currence. Phytochem 2000;54:369–80.
3):523–40. [62] Allard B, Templier J. High molecular weight lipids from the trilaminar outer
[29] Zhang YHP. A sweet out-of-the-box solution to the hydrogen economy: is wall (TLS) –containing microalgae Chlorella emersonii, Scenedesmus commu-
the sugar-powered car science fiction? Energy Environ Sci 2009;2(3):272– nis and Tetraedron minimum. Phytochem 2001;57:459–67.
82. [63] Gelin F, Volkman JK, Largeau C, Derenne S, Sinningh JS, Damste A, De Leeuw
[30] British Petroleum (BP). Statistical Review of World Energy; 2016. JW. Distribution of aliphatic, nonhydrolyzable biopolymers in marine mi-
[31] Galbe M, Zacchi G (2014). Production of ethanol from lignocellulosic mate- croalgae’. Org Geochem 1999;30:147–59.
rials. In Luis Augusto Barbosa. Cortez (Coord.), Sugarcane bioethanol R&D-for [64] Brown MR. The amino-acid and sugar composition of 16 species of
K.A. Salam et al. / Renewable and Sustainable Energy Reviews 65 (2016) 1179–1198 1197

microalgae used in mariculture. J Exp Mar Biol Ecol 1991;145:79–99. [100] Saka S, Kusdiana D. Biodiesel fuel from rapeseed oil as prepared in super-
[65] Wijffels RH, Barbosa MJ, Eppink MHM. Microalgae for the production of bulk critical methanol. Fuel 2001;80:225–31.
chemicals and biofuels. Biofuel Biop Bioref 2010;4(3):287–95. [101] Lim S, Hoog SS, Teong LK, Bhatia S. Supercritical fluid reactive extraction of
[66] Lee AK, Lewis DM, Ashman PJ. Disruption of microalgal cells for the ex- Jatropha curcas L. seeds with methanol: a novel biodiesel production method.
traction of lipids for biofuels: processes and specific energy requirements. Bioresour Technol 2010;101(18):7169–72.
Biomass Bioenergy 2012;46:89–101. [102] Haas MJ, Scott KM. Moisture removal substantially improves the efficiency of
[67] Middelberg APJ. Process-scale disruption of microorganisms. Biotechnol Adv in situ biodiesel production from soybeans. JAOCS 2007;84:197–204.
1995;13(3):491–551. [103] Rodolfi L, Zittelli GC, Bassi N, Padovani G, Biondi N, Bonini G, Tredici MR.
[68] Lee J–Y, Yoo C, Jun S–Y, Ahn C–Y, Oh H-M. Comparison of several methods for Microalgae for oil: Strain selection, induction of lipid synthesis and outdoor
effective lipid extraction from microalgae. Bioresour Technol 2010;101:S75–7. mass cultivation in a low-cost photobioreactor. Biotechnol Bioeng 2009;102
[69] Surendhiran D, Vijay M. Effect of various pre-treatment for extracting in- (1):100–12.
tracellular lipid from Nannochloropsis oculata under nitrogen replete and [104] Wu W, Hsieh C. Cultivation of microalgae for oil production with a cultivation
depleted conditions. ISRN Chem Eng 2014. http://dx.doi.org/10.1155/2014/ strategy of urea limitation. Bioresour Technol 2009;100:3921–6.
536310. [105] Chinnasamy S, Bhatnagar A, Hunt RW, Das KC. Microalgae cultivation in a
[70] Halim R, Harun R, Danquah MK, Webley PA. Microalgal cell disruption for wastewater dominated by carpet mill effluents for biofuel applications.
biofuel development. Appl Energy 2012;91(1):116–21. Bioresour Technol 2010;101(9):3097–105.
[71] Shirgaonkar IZ, Lothe RR, Pandit AB. Comments on the mechanism of mi- [106] Voltolina DB, Cordero B, Nievesc M, Soto LP. Growth of Scenedesmus sp. in
crobial cell disruption in high-pressure and high-speed devices. Biotechnol artificial wastewater. Bioresour Technol 1998;68(3):265–8.
Prog 1998;14(4):657–60. [107] Xiong W, Gao C, Yan D, Wu C, Wu Q. Double CO2 fixation in photosynthesis–
[72] Balasubramanian S, Allen JD, Kanitkar A, Boldor D. Oil extraction from Sce- fermentation model enhances algal lipid synthesis for biodiesel production.
nedesmus obliquus using a continuous microwave system-design, optimiza- Bioresour Technol 2010;101(7):2287–93.
tion, and quality characterization. Bioresour Technol 2011;102(3):3396–403. [108] Chandra R, Takeuchi H, Hasegawa T. Methane production from lignocellulosic
[73] Ratti C. Hot air and freeze-drying of high-value foods: a review. J Food Eng agricultural crop wastes: a review in context to second generation of biofuel
2001;49(4):311–9. production. Renew Sustain Energy Rev 2012;16(13):1462–76.
[74] Balasundaram B, Pandit AB. Selective release of invertase by hydrodynamic [109] Rodriquez C, Alaswad A, Mooney J, Prescott T, Olabi AG. Pre-treatment
cavitation. Biochem Eng J 2001;8(3):251–6. Techniques used for anaerobic digestion of algae. Fuel Process. Tech.
[75] Lotero E, Liu Y, Lopez DE, Suwannakarn K, Bruce DA, Goodwin JG. Synthesis 2015;138:765–79.
of biodiesel via acid catalysis. Ind Eng Chem Res 2005;44:5353–63. [110] D Rutz, R. Janssen Biofuel Technology Handbook, Second WIP Renewable
[76] Freedman B, Pryde EH, Mounts TL. Variables affecting the yields of fatty Energies, Munchen; 2008.
esters from transesterified vegetable oils. JAOCS 1984;61(10):1638–43. [111] Zhao B, Ma J, Zhao Q, Laurens L, Jarvis E, Chen S, Frear C. Efficient anaerobic
[77] Eze VC, Phan AN, Harvey AP. A more robust model of biodiesel reaction, digestion of whole microalgae and lipid extracted microalgae residues for
allowing identification of process conditions for enhanced rate and water methane energy production. Bioresour Technol 2014;161:423–30.
tolerance. Bioresour Technol 2014;156:222–31. [112] Ehimen EA, Connaughton S, Sun Z, Carrington GC. Energy recovery from lipid
[78] Salam KA, Velasquez-orta SB, Harvey AP. Kinetics of fast alkali reactive ex- extracted, transesterified and glycerol codigested microalgae biomass. GCB
traction/in situ transesterification of Chlorella vulgaris that identified process Bioenergy 2009;1(6):371–81.
condition for a significant enhanced rate and water tolerance. Fuel Proc Tech [113] Bohutskyi P, Bouwer E. Biogas production from algae and cyanobacteria
2016;144:212–9. through anaerobic digestion- a review, analysis and research needs. In: Lee
[79] Srivastava A, Prasad R. Triglycerides-based diesel fuels. Renew Sust Energy JW, editor. Advance biofuels and bioprod, 2013; 2013. p. 873–975.
Rev 2000;4(2):111–33. [114] Chisti Y. Biodiesel from microalgae beats Bioethanol. Trends Biotechnol.
[80] Zakaria R, Harvey AP. Kinetics of reactive extraction/in situ transesterifica- 2008;26:126–31.
tion of rapeseed oil. Fuel Proc Tech 2014;125:34–40. [115] Harun R, Davidson M, Doyle M, Gopiraj R, Danquah M, Forde G. Technoeco-
[81] Haas MJ, McAloon AJ, Yee WC, Foglia TA. A process model to estimate bio- nomic analysis of an integrated microalgae photobioreactor, biodiesel and
diesel production costs. Bioresour Technol 2006;97:671–8. biogas production facility. Biom Bioenergy 2011;35(1):741–7.
[82] Canakci M, Gerpen VJ. Biodiesel production via acid catalysis. Am Soc Agric [116] Angelidaki I, Ahring BK. Effects of free long-chain fatty acids on thermophilic
Biol Eng 1999;42(5):1203–10. anaerobic digestion. Appl Micro. Biotechnol. 1992;37(6):808–12.
[83] Ma F, Hanna MA. Biodiesel production: a review. Bioresour Technol [117] Alves MM, Mota VJA, Álvares PRM, Pereira MA, Mota M. Effects of lipids and
1999;70:1–15. oleic acid on biomass development in anaerobic fixed-bed reactors. Part II:
[84] Haas MJ, Wagner K. Simplifying biodiesel production: The direct or in situ oleic acid toxicity and biodegradability. Water Res 2001;35(1):264–70.
transesterification of algal biomass. Eur J Lipid Sci Technol 2011;113:1219–29. [118] Hanaki K, Matsuo T, Nagase M. Mechanism of inhibition caused by long-chain
[85] Harrington KJ, D’Arcy-Evans C. Transesterification in situ of sunflower seed fatty acids in anaerobic digestion process. Biotechnol Bioeng 1981;23
oil’. Ind Eng Chem Prod Res Dev 1985;24(2):314–8. (7):1591–610.
[86] Lewis T, Nichols PD, McMeekin TA. Evaluation of extraction methods for [119] Lalman J, Bagley DM. Effects of C18 long chain fatty acids on glucose, butyrate
recovery of fatty acids from lipid-producing microheterotrophs. J Microbiol and hydrogen degradation. Water Res 2002;36(13):3307–13.
Methods 2000;43(2):107–16. [120] Nielsen HB, Ahring BK. Responses of the biogas process to pulses of oleate in
[87] Vicente G, Bautista LF, Rodríguez R, Gutiérrez FJ, Sadaba I, Ruiz-Vázquez RM, reactors treating mixtures of cattle and pig manure. Biotechnol Bioeng
Torres–Martínez S, Garre V. Biodiesel production from biomass of an olea- 2006;95(1):96–105.
ginous fungus. Biochem Eng J 2009;48:22–7. [121] McCarty PL. Anaerobic waste treatment fundamentals. Public Works 1964;95
[88] Li Y, Lian S, Tong D, Song R, Yang W, Fan Y, Qing R, Hu C. One-step production (9):107–12.
of biodiesel from Nannochloropsis sp. on solid base Mg–Zr catalyst. Appl [122] Lakaniemi A-M, Tuovinen OH, Puhakka JA. Anaerobic conversion of micro-
Energy 2011;88:3313–7. algal biomass to sustainable energy carriers – a review. Bioresour Technol
[89] Grima MM, Acien Fernandez FG, Medina AR. Downstream processing of cell- 2013;135:222–31.
mass and Products’. In: Richmond A, editor. Handbook of Microalgal culture: [123] Mital KM. Biogas systems: principle and applications.New Delhi, India: New
Biotechnology and Applied Phycology. UK: Blackwell Science Ltd.; 2004. Age International (P) Limited; 1996.
p. 215–51. [124] Yen H-W, Brune DE. Anaerobic co-digestion of algal sludge and waste paper
[90] Show KY, Lee DJ, Chang JS. Algal biomass dehydration. Bioresour Technol to produce methane. Bioresour Technol 2007;98(1):130–4.
2013;135:720–9. [125] Desbois A, Smith V. Antibacterial free fatty acids: activities, mechanisms of
[91] Gudin C, Therpenier C. Bioconversion of solar energy into organic chemicals action and biotechnological potential. Appl Microbiol Biotechnol 2001;85
by microalgae. Adv Biotechnol Proc 1986;6:73–110. (6):1629–42.
[92] Zhao Z, Liu O. Biodiesel production by direct methanolysis of oleaginous [126] McCracken MD, Middaugh RE, Middaugh RS. A chemical characterization of
microbial biomass. J Chem Technol Biotechnol 2007;82:775–80. an algal inhibitor obtained from Chlamydomonas. Hydrobiologia
[93] Ehimen EA, Sun ZF, Carrington CG. Variables affecting the in situ transes- 1980;70:271–6.
terification of microalgae lipids. Fuel 2010;89(3):677–84. [127] Zheng CJ, Yoo J-S, Lee T-G, Cho H-Y, Kim Y-H, Kim W-G. Fatty acid synthesis is
[94] Kasim FH, Harvey AP. Influence of various parameters on reactive extraction a target for antibacterial activity of unsatu- rated fatty acids. FEBS Lett
of Jatropha curcas L. for biodiesel production. Chem Eng J 2011;171:1373–8. 2005;579(23):5157–62.
[95] Zakaria R, Harvey AP. Direct production of biodiesel from rapeseed by reactive [128] Kakisawa H, Asari F, Kusumi T, Toma T, Sakurai T, Oohusa T, Hara Y, Chihara
Extraction /in situ transesterification. Fuel Proc Technol 2012;102:53–60. M. An allelopathic fatty acid from the brown alga Cladosiphon okamuranus.
[96] Zeng J, Wang X, Zhao B, Sun J, Wan Y. Rapid in situ transesterification of Phytochem 1988;27:731–5.
sunflower oil. Ind Eng Chem Res 2009;48(2):850–6. [129] Palatsi J, Laureni M, Andrés M, Flotats X, Nielsen H, Angelidaki I. Strategies for
[97] Haas MJ, Scott KM, Marmer WN, Foglia TA. In situ alkaline transesterification: recovering inhibition caused by long chain fatty acids on anaerobic ther-
an effective method for the production of fatty acid esters from vegetable mophilic biogas reactors. Bioresour Technol 2009;100(20):4588–96.
oils. JAOCS 2004;81(1):83–9. [130] Ma J, Zhao Q-B, Laurens LLM, Jarvis EE, Nagle NJ, Chen S, Frear CS. Me-
[98] Kildiran G, Ozgul-Yucel S, Turkay S. In-situ alcoholysis of soybean oil. JAOCS chanism, kinetics and microbiology of inhibition caused by long-chain fatty
1996;73:225–8. acids in anaerobic digestion of algal biomass. Biotechnol Biofuels 2015;8
[99] Dong T, Wang J, Miao C, Zheng Y, Chen S. Two-step in situ biodiesel pro- (141):1–12.
duction from microalgae with high free fatty acid content. Bioresour Technol [131] Yang J, Xu M, Zhang X–Z, Hu Q-A, Sommerfeld M, Chen Y-S. Life-cycle analysis
2013;136:8–15. on biodiesel production from microalgae: Water footprint and nutrients
1198 K.A. Salam et al. / Renewable and Sustainable Energy Reviews 65 (2016) 1179–1198

balance. Bioresour Technol 2011;102:159–65. pretreatments. Bioresour Technol 2013;149:136–41.


[132] Mottet A, Habouzit F, Steyer JP. Anaerobic digestion of marine microalgae in [136] Cho S, Park S, Seon J, Yu J, Lee T. Evaluation of thermal ultrasonic and alkali
different salinity levels. Bioresour Technol 2014;158:300–6. pretreatments on mixed microalgal biomass to enhance anaerobic methane
[133] Robert KP, Heaven S, Banks CJ. Quantification of methane losses from accli- production. Bioresour Technol 2013;143:330–6.
matisation of anaerobic digestion to marine salt concentrations. Renew En- [137] Loganathan M, Anbarasu A, Velmurugan A. Emission characteristics of ja-
ergy 2016;86:497–506. tropha - dimethyl ether fuel blends on a DI diesel engine. Int J Sci Technol.
[134] Ogawa T, Inoue N, Shikada T, Ohno Y. Direct dimethyl ether synthesis. J Nat Res 2012;1:28–32.
Gas Chem 2003;12:219–27. [138] European Food Safety Authority (EFSA). Safety in use of dimethyl ether as
[135] Mendez L, Mahdy A, Timmers RA, Ballesteros M, Gonzalez-Fernandez C. extraction solvent. EFSA J 2008;984:1–13.
Enhancing methane production of Chlorella vulgaris via thermochemical

You might also like