You are on page 1of 19

Energy for Sustainable Development 43 (2018) 143–161

Contents lists available at ScienceDirect

Energy for Sustainable Development

Review

Biodiesel production in micro-reactors: A review


Appurva Tiwari, V.M. Rajesh ⁎, Sanjeev Yadav
Department of Chemical Engineering, Shiv Nadar University, Tehsil Dadri, Gautam Buddha Nagar, Greater Noida, Uttar Pradesh 201314, India

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

Article history: In recent years, biodiesel, as a renewable and environment-friendly fuel has emerged as one of the most investi-
Received 14 September 2017 gated bio-fuel with a goal to decrease our dependence on petroleum fuels and reduce environmental pollution.
Revised 9 January 2018 The current commercial technique of biodiesel production via transesterification is constrained by high operating
Accepted 9 January 2018
cost and energy requirements, long residence time and limited conversion due to equilibrium limitations. On the
Available online 3 February 2018
other hand, the process intensification using the micro-reactor technology demonstrated an excellent perfor-
Keywords:
mance ascribed to short diffusion distance and high surface area-to-volume ratio that can lead to high heat
Renewable energy and mass transfer rates and improved mixing compared to the conventional reactors. This review provides an
Micro-reactors overview of the current status of biodiesel production in micro-reactors. It includes various types of micro-
Biodiesel reactors used in the production of biodiesel, factors affecting the biodiesel production (i.e., temperature, resi-
Transesterification dence time, alcohol to oil molar ratio, micro-channel size, inlet mixer type, internal geometries, co-solvent, cat-
Bio-fuel alyst type and concentration). This review also includes the factors affecting the liquid-liquid flow patterns
and application of micro-reactor technology in the purification of biodiesel. Some of the critical observations
from this review are, 1) inlet mixer type, channel size, and internal channel geometry (zig-zag, omega, and
tesla shaped channels) have shown a significant effect on mixing in micro-channels. 2) In case of base-catalyzed
transesterification, the biodiesel yield was found to increase up to the reaction temperature of 60–65 °C. 3) Ho-
mogeneous alkaline catalyst (NaOH, KOH, CH3ONa) was preferred for the feedstock with low free fatty acid con-
tent (b1 wt%). However, an acid catalyst with high concentration, a significant amount of methanol and long
reaction time were required for high free fatty acid feedstock (N1 wt%). Therefore, the current research is more
focused on the investigation of heterogeneous catalysts in micro-reactors to develop an ecologically friendly pro-
cess for the production of biodiesel. 4) Also, the reaction temperature and inlet mixer type had shown a signifi-
cant effect on liquid-liquid flow patterns. This review also addressed the following literature gaps; a numbering-
up technique to increase the throughput; catalyst development for high free fatty acid feedstock; continuous pro-
duction of biodiesel in micro-reactors with in-line purification step to meet the energy demand and quality
standards.
© 2018 International Energy Initiative. Published by Elsevier Inc. All rights reserved.

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
Present state of the art: micro-reactors used in biodiesel synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
Factors affecting biodiesel production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
Effect of reaction temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
Residence time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
Alcohol to oil molar ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
Effects of geometrical configurations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
Effect of channel size . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
Effect of inlet mixer type . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
Effect of internal geometries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
Effect of co-solvent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
Effects of catalyst type and concentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
Homogeneous catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
Heterogeneous catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153

⁎ Corresponding author.
E-mail address: vm.rajesh@snu.edu.in (V.M. Rajesh).

https://doi.org/10.1016/j.esd.2018.01.002
0973-0826/© 2018 International Energy Initiative. Published by Elsevier Inc. All rights reserved.
144 A. Tiwari et al. / Energy for Sustainable Development 43 (2018) 143–161

Catalyst-free process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154


Liquid-liquid two-phase flows in micro-reactors for biodiesel production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
Factors affecting flow patterns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
Effect of temperature on flow patterns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
Effect of alcohol on oil molar ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
Effect of inlet mixer type . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
Purification of biodiesel in micro-reactors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
Scope for future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159

Introduction exothermic reactions safely with precise control over the operating con-
ditions (Benke, 2014; Burns & Ramshaw, 1999; Kobayashi, Mori, &
With the rapid depletion of fossil fuels and increasing environmental Kobayashi, 2006). However, the production capacity of micro-reactors
concerns, the need for an alternative fuel that is cheaper, renewable and can be increased by the scale-out approach (Aran & Lammertink,
environment-friendly is growing consistently. Biodiesel has emerged as 2013; Geyer & Seeberger, 2008).
a potential alternative fuel to fossil fuels, which has similar properties to Some researchers have successfully applied micro-reactors for
fossil diesel with some advantages (Freedman, Pryde, & Mounts, 1984; biodiesel synthesis via transesterification (Aghel, Rahimi, Sepahvand,
Van Gerpen, 2005b). Biodiesel is a renewable fuel, less-toxic than con- Alitabar, & Ghasempour, 2014; Buddoo, Siyakatshana, & Pongoma,
ventional diesel and also biodegradable (Al-Zuhair, 2007; Bozbas, 2008; Guan et al., 2009a; Martínez Arias, Fazzio Martins, Jardini
2008; Knothe, Sharp, & Ryan, 2006; Koonin, 2006; Meher, Vidyasagar, Munhoz, Gutierrez-Rivera, & Maciel, 2012; Sun, Wang, Yao, Zhang, &
& Naik, 2006). Biodiesel blends with conventional diesel up to B20 Xu, 2009; Wen, Yu, Tu, Yan, & Dahlquist, 2009). It was demonstrated
(20% biodiesel and 80% diesel) can be used in a diesel engine without that as compared to conventional batch reactors, higher conversion in
engine modification. However, biodiesel N20% in the blend required shorter residence time could be achieved in micro-reactors (Dai, Li,
few engine modifications (Canakci, Erdil, & Arcaklioğlu, 2006; Sharma, Zhao, & Xiu, 2014; Elkady, Zaatout, & Balbaa, 2015; Guan et al., 2009a;
Singh, & Upadhyay, 2008; Van Gerpen, 2005a; Xue, Grift, & Hansen, Martínez Arias et al., 2012; Rahimi, Aghel, Alitabar, Sepahvand, &
2011). Biodiesel is produced via transesterification of vegetable oils Ghasempour, 2014; Santacesaria et al., 2012; Santana, Tortola, Reis,
or fats with alcohol, typically methanol, in the presence of a catalyst Silva, & Taranto, 2016; Schürer, Thiele, Wiborg, Ziogas, & Kolb, 2014;
(Demirbaş, 2003; Van Gerpen, 2005b). The transesterification reaction Sun, Ju, Ji, Zhang, & Xu, 2008). It should also be noted that as compared
takes place in a liquid-liquid two-phase system and the reaction to conventional reactors, biodiesel could be produced 10 to 100 times
rate is limited by mass transfer due to immiscibility of oil and alcohol faster in a micro-reactor. Moreover, the requirements of high energy
(Crawford et al., 2008; Guan, Kusakabe, Moriyama, & Sakurai, in the mixing of reactants, ample floor space and standing time for the
2009a). Therefore, an efficient contact between the feed oil and the separation of products are eliminated using micro-reactors (Buddoo
alcohol-catalyst mixture becomes crucial for achieving the high rate of et al., 2008; Jovanovic, 2006). Although the amount of biodiesel pro-
reaction. duced using a single micro-reactor is a trickle, each unit can be num-
Several studies on biodiesel production were carried out using batch bered up to increase the throughput from lab to commercial scale
reactors (Dorado, Ballesteros, López, & Mittelbach, 2004; Lotero et al., (Billo et al., 2015; Jovanovic, 2006; Schürer et al., 2016). The process
2005; Shah, Sharma, & Gupta, 2004; Vicente, Martınez, & Aracil, 2004; of transesterification in micro-reactors is observed to be affected by
Xie, Peng, & Chen, 2006; Zhang, Dube, McLean, & Kates, 2003). These the operating conditions as well as the type and geometrical param-
reactors were also commercially used for the production of biodiesel eters of micro-reactor (Xie et al., 2012). With this background, the
at industrial scale (Haas, McAloon, Yee, & Foglia, 2006). However, present review has been organized as following sections: 1) the
unfavourable economics and other design and operational issues at present state of the art: the types of micro-reactors used in
large-scale batch reactors pose challenges to the commercialization of biodiesel synthesis with different feedstock have been discussed.
biodiesel (Maddikeri, Pandit, & Gogate, 2012; Qiu, Zhao, & Weatherley, 2) The factors affecting biodiesel production in micro-reactors, in
2010; Santacesaria, Vicente, Di Serio, & Tesser, 2012). The critical which, the effect of operating parameters such as temperature,
challenges are, the long residence time of several hours, considerable alcohol to oil molar ratio, the effect of co-solvent addition, catalyst
investment in equipment and manufacturing floor space, high energy type and concentration as well as the effect of the geometrical
consumption and low production efficiency (Jachuck, Pherwani, & configuration of micro-reactors have been discussed. 3) Liquid-
Gorton, 2009; Qiu et al., 2010; Slinn & Kendall, 2009; Warabi, liquid two-phase flows in micro-reactors for biodiesel production:
Kusdiana, & Saka, 2004). On the other hand, with these limitations on flow patterns observed during production of biodiesel and the
a conventional batch process, several alternative highly energy efficient factors affecting the flow patterns have been discussed. 4) Purifica-
technologies to produce biodiesel in a continuous mode of operation tion of biodiesel in micro-reactors: the current research techniques
were investigated (Maddikeri et al., 2012; Qiu et al., 2010). In recent applied for biodiesel separation from glycerol at micro-scale are
years, micro-reactor technology has emerged as a promising technique mentioned. Finally, the current research gaps that need further
for high-efficiency continuous biodiesel production (Xie, Zhang, & Xu, investigations into full-scale commercialization of biodiesel produc-
2012). Micro-reactors are miniaturized reaction systems with reduced tion using micro-reactors are also addressed.
internal characteristic dimensions of 1000 μm–10 μm. Micro-reactors
offer small diffusion distances and large surface area to volume ratio Present state of the art: micro-reactors used in biodiesel synthesis
(10,000–50,000 m2 m−3) and thus lead to high heat and mass transfer
rates. As a result higher conversion is achieved in shorter residence Micro-reactors can be categorized by geometries, fabrication mate-
time (Kashid & Kiwi-Minsker, 2009). Mixing time in micro-reactors is rial, and techniques. For the production of biodiesel, mainly micro-
also considerably reduced due to their reduced internal dimensions. tubular and micro-channel reactors have been applied. In general, the
Moreover, small volume favours micro-reactors to perform highly micro-reactor system is composed of a mixer for intensifying mixing
A. Tiwari et al. / Energy for Sustainable Development 43 (2018) 143–161 145

and a micro-channel or micro-tube for completion of the reaction. pre-mixed rapeseed oil and methanol using KOH as a catalyst in quartz
Various combinations of mixers and reaction loop (micro-channel or capillary of 0.25 mm inner diameter and length of 30 m. At a catalyst con-
micro-tube), fabricated with different materials have been used in bio- centration of 1% by weight and methanol to oil molar ratio of 6:1, a fatty
diesel synthesis as shown in Table 1. To demonstrate that the production acid methyl ester (FAME) yield of 98.80% was obtained in 6 min at 60
of biodiesel is feasible using micro-reactor technology Al-Dhubabian °C. Instead of using a microfluidic mixer the reactants were a premixed
(2005) developed a micro-reactor with channel dimensions; 23.3 mm in a tank reactor with constant stirring where the yield reached 81% be-
length, 10.5 mm width, and 0.1 mm height as shown in (Fig. 1). Under fore injecting the mixture into the capillary. Hence, the merits of using a
operating conditions of 7.2:1 methanol/soya bean oil molar ratio, 1 wt% micro-tube capillary as a reactor were not clearly observed.
NaOH catalyst, at atmospheric pressure and 25 °C, a conversion of 91% Guan et al. (2008) carried out continuous production of biodiesel
was obtained in 10 min. Similarly, Canter (2006) reported a micro- from sunflower oil using a micro-reactor system consisting of Fluori-
reactor with parallel micro-channels cut into a thin plastic plate could nated ethylene propylene (FEP) tube with an inner diameter of 1 mm
be used for making biodiesel under mild operating conditions. Biodiesel and 160 mm in length and a mixer (split and recombine mixer or a T-
yield above 90% was obtained at a residence time of about 4 min. type) to create dispersion. Oil conversion reached 100% in 112 s, using
However, the details of the reactor and operating parameters were not methanol to oil molar ratio of 23.9:1 and 1 wt% KOH concentration
described completely. These works proved that micro-reactors could and split and recombine mixer. It was noted that for complete conver-
be potentially applied for the production of biodiesel commercially. sion of oil, the residence time decreased drastically from 600 s to
The findings of Al Dhubabian (2005) and Canter (2006) were sup- b240 s when lab scale batch reactor was replaced by a micro-reactor,
ported by Sun et al. (2008), where they performed transesterification of under same operating conditions. In their progressive work, Guan

Table 1
Micro-reactors used in the production of biodiesel.

Mixer type Reaction loop Reactants Catalyst T (°C) Alcohol/oil Residence Biodiesel Ref.
amount molar ratio time yield (%)
(wt%)

Stirred tank i.d. = 0.25 mm Rapeseed oil + methanol 1% KOH 60 °C 6:1 6 min 98.80% Sun et al. (2008)
l = 30 m
Quartz capillary
Stirred tank i.d. = 0.25 mm Cottonseed oil + methanol 1% KOH 60 °C 6:1 5.89 min 99.40% Sun et al. (2008)
l = 30 m
Stainless-steel capillary
Split and recombine i.d. = 1 mm Sunflower oil + methanol 1% KOH 60 °C 23.9:1 112 s 100% Guan, Kusakabe,
micromixer l = 160 mm Moriyama, and Sakurai
FEP tube (2008)
T-type i.d. = 0.8 mm Sunflower oil + methanol 4.5% KOH 60 °C 23.9:1 100 s 100% Guan et al. (2009a)
l = 300 mm
FEP tube
T-mixer i.d. = 1.5 mm Canola oil + methanol 1% NaOH 60 °C 6:1 3 min 99.80% Jachuck et al. (2009)
l = n.i.
PTFE tube
T-shaped junction dh = 0.24 mm Soya bean oil + methanol 1.2% NaOH 56 °C 17:1 28 s 99.5% Wen et al. (2009)
l = 1.07 m
Stainless-steel
zig-zag channel
Split interdigital i.d. = 3 mm Cottonseed oil + methanol 1% KOH 70 °C 8:1 17 s 99.5% Sun et al. (2009)
micromixer l = n.i.
(SIMM-V2) PTFE tube packed with Dixon
rings
T-Type dh = 0.5 mm Castor oil + ethanol 1% NaOH 50 °C 12:1 10 min 96.70% Martínez Arias et al.
l=1m (2012)
Tesla-shaped PDMS
micro-channel
T-Type dh = 0.5 mm Castor oil + ethanol 1% NaOH 50 °C 12:1 10 min 95.30% Martínez Arias et al.
l=1m (2012)
Omega-shaped PDMS
micro-channel
T-Type dh = 0.5 mm Castor oil + ethanol 1% NaOH 50 °C 12:1 10 min 93.50% Martínez Arias et al.
l=1m (2012)
T-shaped PDMS micro-channel
Zig-zag micro-channel dh = 1.5 mm Soya bean oil + methanol 1.20% KOH 59 °C 8.5:1 14.9 s 99.50% Dai et al. (2014)
mixer l = n.i.
Stainless-steel zig-zag reaction
channel
T-type i.d. = 0.90 mm Soya bean oil + methanol 1.20% KOH 60 °C 9:1 180 s 99% Aghel et al. (2014)
l = n.i.
Stainless micro-tube embedded
with wire coil (30 cm length,
0.5 mm pitch)
Four-way micromixer i.d. = 1.58 mm Soya bean oil + methanol + 1% KOH 57.2 °C 6:1 9.05 s 98.80% Rahimi, Mohammadi,
with 45° confluence l = n.i. hexane (co-solvent) Basiri, Parsamoghadam,
angle Stainless-steel tube and Masahi (2016)
KM micromixer Waste vegetable oil + 1% NaOH 70 °C 12:1 n.i. 97% Elkady et al. (2015)
methanol + THF (co-solvent)

n.i. = no information, i.d. = inner diameter, l = length, dh = hydraulic diameter.


146 A. Tiwari et al. / Energy for Sustainable Development 43 (2018) 143–161

Fig. 1. Schematic diagram of a micro-reactor system with a digital image of the side view of the micro-reactor. Reprinted with permission from Al-Dhubabian (2005). Copyright 2005
Oregon State University.

et al. (2009a) performed transesterification of sunflower oil using a FEP zig-zag-9. The micro-channel reactors were fabricated on stainless-steel
tube of 0.8 mm in diameter with T-shaped junction. A complete conver- sheet by electric spark processing and assembled with a T-shaped
sion of oil was obtained at 300 mm tube length, 4.5 wt% KOH concentra- three-way mixer. The micro-channels of all reactors were of rectangular
tion, in a residence time of 100 s using the same alcohol to oil molar ratio cross-section with the same length of 1.07 m but a different number
and operating conditions as their previous work (Guan et al., 2008). of periodic turns (10, 50, 100, 200 & 350) and hydraulic diameter
Jachuck et al. (2009) used a PTFE (Polytetrafluoroethylene) narrow chan- (240–900 μm) as shown in Table 3. Using the micro-channel reactor
nel reactor with an internal diameter of 1.5 mm, coupled with a T-mixer with a smallest hydraulic diameter (240 μm) and most number of turns
to intensify the biodiesel production process. Conversions higher than (350), a maximum yield of 99.5% was reached with methanol to soya
98% were obtained in a residence time of 3 min with NaOH loading of 1 bean oil molar ratio of 9:1, NaOH concentration of 1.2 wt% in a very
wt% and methanol to canola oil molar ratio of 6:1 at 60 °C. short residence time of 28 s. Based on the same idea, Dai et al. (2014) de-
In the previous works, simple T-flow structures were applied to gen- veloped a micro-reactor in which the conventional T-type mixer used by
erate dispersion in micro-reactors (Canter, 2006; Guan et al., 2009a; Sun Wen et al. (2009) was replaced by zig-zag micro-channel mixers
et al., 2008). To investigate another high efficiency passive micro- patterned on stainless-steel sheet along with zig-zag reaction channels.
structured mixers for biodiesel production, Wen et al. (2009) The mixing and reacting channels were linked through a connecting
developed nine different zig-zag micro-channel reactors, zig-zag-1 to channel as shown in Fig. 2. Using a micro-channel of hydraulic diameter

Fig. 2. The configuration of zig-zag micro-channel reactor with zig-zag mixers. Reprinted with permission from Dai et al. (2014). Copyright 2014 American Chemical Society.
A. Tiwari et al. / Energy for Sustainable Development 43 (2018) 143–161 147

(A) (B)

Fig. 3. (A) Mixing forms and channel dimensions (width × height) of various micromixers: (a) T-mixer, (b) J-mixer, (c) rectangular interdigital micromixer (RIMM), and (d) split
interdigital micromixer (SIMM-V2). M, an inlet of the methanol phase; O, an inlet of oil. “Reprinted with permission from Sun et al. (2009). Copyright 2010 American Chemical Society.”
(B) PDMS micro-reactor internal geometries: (a) Omega-shaped, (b) Tesla shaped, and (c) T-shaped. Reprinted with permission from Martínez Arias et al. (2012). Copyright 2012
American Chemical Society.

1.5 mm, a biodiesel yield of 99.5% was obtained at a residence time of The micro-channels were of the quadratic cross-section with width and
14.9 s, methanol to soya bean oil molar ratio of 8.5:1 and KOH concentra- height of 500 μm and length of 1 m. For T-, Tesla and Omega shaped
tion of 1.2 wt% at 59 °C. micro-reactors, ethyl ester conversions of 93.5%, 95.3%, and 96.7%
Sun et al. (2009) performed experiments on the fast synthesis of were attained respectively, at alcohol to oil molar ratio of 25:1, catalyst
biodiesel from cottonseed oil using high throughput micro-structured re- loading of 1 wt% and a reaction temperature of 50 °C in 10 min. Aghel
actor with various mixer types as shown in Fig. 3(A). The FAME yields ob- et al. (2014) studied biodiesel synthesis using a modified micro-
tained using the micromixers were almost twice of the yields obtained reactor consisting of a silver wire coil inserted in a stainless-steel tube
using T- and J-mixers due to more intense mixing in micromixers. Using of 0.9 mm inner diameter and a micromixer. At T- shaped junction
rectangular interdigital micromixer (RIMM) a maximum yield of 99.5% of micromixer, soya bean oil and methanol were mixed in a molar
was achieved under the conditions of methanol to oil molar ratio of 8:1, ratio of 1:9 with 1.2 wt% of KOH as a catalyst. Using a wire coil with
a reaction temperature of 70 °C, the residence time of 17 s and a flow 0.8 mm average diameter, 30 cm length and 0.5 mm average pitch
rate of 10 mL/min. Similarly, Bhoi et al. (2014) investigated KOH catalyzed length, a FAME conversion of 99% was obtained at 60 °C in residence
biodiesel production using three different reaction systems having differ- time of 180 s.
ent microfluidic junctions, namely, T-type, cross-type and split and Recently, Santana et al. (2016) performed experiments in a PDMS
recombine mixer followed by serpentine mixing and reaction channels based T-junction micro-channel (l × w × h; 41.1 cm × 1.5 cm × 0.2 cm)
etched on a glass chip. It was observed that using methanol to sunflower reactor. In their experiments, sunflower oil was used as feedstock with
oil molar ratio of 10.3:1, conversion above 90% could be obtained in a ethanol to produce biodiesel through transesterification with sodium hy-
residence time of 0.5 min, 1 min, and 5 min with, T-type, cross-type and droxide as a catalyst. It was observed that minimum residence time of 1
split and recombine mixer respectively. min was needed at room temperature for a conversion of 95.8% as com-
Avellaneda and Salvadó (2011) demonstrated a helicoidal system pared to the batch process in which 94.1% conversion was achieved in
comprising a T-mixer for aggregation of reactants and a series of spirals 180 min. In order to increase the conversion and reduce the residence
(i.d. 6 mm) connected consecutively by a fixture, for the continuous time, the same group (Santana, Tortola, Silva, & Taranto, 2017) performed
production of biodiesel using pre-treated recycled oil. It was determined experiments in a PDMS micro-channel (l × w × h; 35.1 cm × 1.5 cm ×
that 89% yield of fatty acid methyl ester could be attained in 13 min at 0.2 cm) reactor with internal static elements (1000 μm × 100 μm)
60 °C, methanol to oil molar ratio of 6:1 and using 0.6 g of NaOH as cat- for the improved mixing. It was found that maximum yield
alyst per 100 g of oil. For the same yield in the batch reactor, a residence (99.53%) was obtained with 1 wt% catalyst concentration, residence
time of 75 min was required. Santacesaria et al. (2011) proposed the time of ~12 s and ethanol to oil molar ratio of 9:1 at 50 °C.
use of a tubular stainless-steel reactor of 10 mm inner diameter and Guan, Sakurai, and Kusakabe (2009b) and Rahimi et al. (2016) stud-
20 cm in length with three different packing configurations, that were, ied the production of biodiesel in the presence of co-solvent at room tem-
(a) spheres of 2.5 mm diameter, (b) spheres of 2.5 and 1 mm diameters perature using micro-tube reactors. Guan, Sakurai, and Kusakabe (2009b)
and (c) spheres of 2.5 and 0.39 mm diameters, forming micro-channels reported biodiesel production at room temperature in the presence of co-
of about 1000 μm, 500 μm and 300 μm respectively, for studying KOH solvent in a FEP tube (0.96 mm i.d) micro-reactor with the T-shaped junc-
catalyzed methanolysis of soya bean oil. Using a methanol to oil ratio tion. Using diethyl ether (DEE) as co-solvent in a molar ratio of 0.73 to
of 6:1 at 60 °C, very high oil conversions were observed in residence methanol, 1 wt% KOH as catalyst and methanol to sunflower oil molar
time of b1 min. In their subsequent work, Santacesaria, Di Serio, et al. ratio of 8:1, a conversion of 93.2% at a tube length of 36 cm was
(2012) used the same cylindrical tube filled with stainless-steel wool observed in 93 s. Later, Rahimi et al. (2016) performed transesterification
to obtain a higher void fraction which resulted in increased of soya bean oil in the presence of hexane as a co-solvent using three dif-
productivity. ferent four-way micromixers with 0.8 mm inner diameter followed by a
Martínez Arias et al. (2012) employed different channel geometries stainless-steel tube with inner diameter of 1.58 mm. The micromixers
(Omega, Tesla, and T-shaped), as shown in Fig. 3(B) made of polydi- were designed with various confluence angles of 45o, 90o and 135o
methylsiloxane (PDMS) using soft lithography process for producing namely E1, E2 and E3 respectively on a polymethylmethacrylate
biodiesel from canola oil and ethanol in presence of NaOH as a catalyst. (PMMA) flat plate by milling as shown in Fig. 4. The highest FAME
148 A. Tiwari et al. / Energy for Sustainable Development 43 (2018) 143–161

Fig. 4. The experimental setup used by Rahimi et al. (2016). Reprinted from Rahimi et al. (2016), with permission from Elsevier.

yield of 98.8% was observed using E1 mixer under optimal conditions and Yoo (2011) developed a continuous bio catalytic membrane
of methanol to oil volumetric ratio of 3:1, hexane to methanol micro-reactor using an asymmetric polyether sulfone membrane with
volumetric ratio of 0.45, reaction temperature of 57.2 °C and 1 wt% a nominal molecular weight limit of 300 kDa (PES 300) as an enzyme
KOH concentration. carrier. Each pore of membrane acted as a particular micro-system in
Several investigations were performed on the continuous catalyst- which biocatalytic transesterification of triolein with methanol oc-
free production of biodiesel using micro-reactors in the presence of curred. With a membrane diameter of 63.5 mm and 0.28 mm thickness,
co-solvent and heterogeneous catalyst under supercritical conditions 80% conversion of triolein to methyl oleate (biodiesel) was acquired in a
(Bertoldi et al., 2009; Silva, Castilhos, Oliveira, & Filho, 2010; Schürer reaction time of 19 min. Kurayama et al. (2010) reported that micro-
et al., 2014, 2016; Trentin et al., 2011). Bertoldi et al. (2009) carried capsules could be used as a micro-reactor for biodiesel production.
out transesterification of soya bean oil using supercritical ethanol with The transesterification of rapeseed oil was carried out by adding meth-
carbon dioxide as a co-solvent in a tubular micro-reactor which was anol to Calcium oxide (CaO) loaded micro-capsules and oil mixture in
made of stainless-steel with an internal diameter of 3.2 mm or 0.76 the batch-type reactor. A maximum FAME yield was reached under op-
mm. The reactions were performed in the absence of a catalyst at the timum conditions of methanol to oil molar ratio of 8:1, 20 wt% of CaO
temperature and pressure ranges from 300 to 350 °C and 7.5 to 20 content in micro-capsule and the temperature of 65 °C.
MPa respectively. In addition, the ranges of oil to methanol molar ratio From the above literature, it is evident that biodiesel production
and co-solvent to substrate mass ratio were also maintained as 1:10 to process is highly intensified applying micro-reactor technology. The
1:40 and 0:1 to 0.5:1 respectively. It was found that the FAEE (fatty
acid ethyl ester) yield was improved at the optimum design and operat- Table 2
ing conditions such as 0.76 mm diameter tube, CO2 to substrate mass Comparison of batch versus micro-reactor plant for production of biodiesel. Reprinted
ratio of 0.05:1, oil to ethanol molar ratio of 1:40 and at 350 °C and 10 with permission from Buddoo et al. (2008).
MPa. Production of biodiesel Batch Micro-reactor Comments
Sootchiewcharn, Attanatho, and Reubroycharoen (2015) produced plant plant
biodiesel from palm oil with supercritical ethylacetate in a stainless- Plant output (tonne/yr) 20,000 20,000
steel micro-reactor with 0.762 mm i.d. and 15.14 m in length. At a Reactor volume (m3) 10 2.4 × 10−3 4167× smaller
temperature of 350 °C, oil to the ethyl acetate molar ratio of 1:50, and Plant footprint (m2) 149 60 60% smaller
ethyl acetate pressure of 20 MPa, biodiesel yield of 78.3% was obtained Surface area to volume ratio (m2/m3) 14.9 2.5 × 104 1678× higher
Productivity (kg/h/m3) 250 10.4 × 105 4167× higher
in 20 min. Recently, Schürer et al. (2014) reported a stainless-steel
Energy input (kJ per kg) 7.1 0.4 18× lower
micro-channel reactor coated with alumina as a catalyst to synthesize bio- Mass transfer coefficient kla (s−1) 10−2–10 10–100 104 higher
diesel under supercritical conditions. A complete conversion of tricaprin Heat transfer coefficient (kJ/m3) 628 2.86 × 106 4554× higher
in residence time of b30 s was observed at 375 °C and 10 MPa pressure. Mixing efficiency (Re) 7 × 105 10 7 × 104 higher
Unlike the conventional micro-tubular and micro-channel reactors Capital cost (Rm) 8.6 6.5 24.4% saving
Manufacturing costs (R/L) 6.60 5.87 11.1% saving
used for biodiesel production, Machsun, Gozan, Nasikin, Setyahadi,
A. Tiwari et al. / Energy for Sustainable Development 43 (2018) 143–161 149

residence time of several hours in a conventional stirred tank reactor can respectively as shown in Fig. 5. It can be ascribed to enhanced mass
be minimized up to several seconds using a micro-reactor. A comparison transfer caused by an improvement in alcohol–oil miscibility and in-
between batch and micro-reactor plant for biodiesel production (Table 2) creased rate of reaction at elevated temperature (Čerče, Peter, &
exhibited a superior performance of the micro-reactors over the batch Weidner, 2005; Zhou, Lu, & Liang, 2006). Moreover, it was observed
plant regarding productivity, energy input and capital cost for same that on increasing the temperature from 46 to 60 °C the percentage in-
plant capacity (Buddoo et al., 2008). However, in most of the studies appli- crease in conversion was higher at lower catalyst loading in a micro-
cation of this technology is confined to lab scale, producing biodiesel only channel reactor. At a catalyst loading of 0.25 wt%, the conversion in-
in small quantities. Billo et al. (2015) developed a cellular manufacturing creased by 14.80% while only 2.5% increase in conversion was
process for fabrication and assembly of a full-scale micro-reactor to pro- obtained for 1 wt% catalyst loading (Jachuck et al., 2009).
duce biodiesel at the rate of 2.47 L/min and a capacity of 1.2 million L Most of the works were carried out at temperatures below the
per year. The scale-up of micro-reactor was done through numbering- normal boiling point of methanol. Sun et al. (2009) examined the effect
up of individual micro-channel laminae into subassemblies and assem- of reaction temperature on transesterification of cottonseed oil both
blies. Schürer et al. (2016) designed a micro-reactor plant with annual below and above the boiling point of methanol using a micro-reactor
capacity of 40 t for production of biodiesel under super critical conditions with rectangular interdigital micromixer (RIMM). The fatty acid methyl
in the presence of a heterogeneous catalyst. The micro-reactor consisted ester yield increased as the reaction temperature was increased from 60
of three stainless-steel plates on which the micro-channels were etched to 80 °C at methanol to oil molar ratio b9:1. The two-phase slug flow at
by wet chemical etching. 30 wt% La2O3 supported by γ-Al2O3 was coated 60 °C was transformed to gas-liquid two-phase slug-annular flow when
on to the micro-channels as a catalyst. The plates were stacked and sealed the temperature was raised to 80 °C. However, Aghel et al. (2014) and
by laser welding. At a reaction conditions of 375 °C and 175 bar and meth- Sun et al. (2008) found that when the reaction temperature increases
anol to oil molar ratio of 40:1, a complete conversion of rapeseed oil was beyond the optimum level, the biodiesel yield decreases. The fatty acid
obtained within a few seconds. methyl ester yield increased from 96.2% to 99.4% as the temperature in-
creased from 30 to 60 °C, but further increase in temperature to 70 °C
resulted in a decrease of FAME yield to 99.1% (Sun et al., 2008). Aghel
Factors affecting biodiesel production
et al. (2014) observed an increase in the percentage of FAME yield
with increase in temperature from 55 °C to 60 °C but the yield decreased
The process of biodiesel production is affected by various factors
rapidly when the temperature was increased from 60 °C to 65 °C. It may
depending on the reaction conditions and the geometrical parameters
be because of the change in flow pattern from slug to bubble flow and
of micro-reactor used. The effects of these factors are discussed below.
the accelerated saponification reaction of triglycerides in the presence
of an alkaline catalyst at higher temperatures (Aghel et al., 2014;
Effect of reaction temperature Leung & Guo, 2006; Sun et al., 2008). To perform continuous biodiesel
production by two-step acid catalyzed process using acid oil, tempera-
The effect of temperature on the transesterification of vegetable oils tures N100 °C were needed. Both esterification and transesterification
has been studied over a wide range of reaction temperatures in micro- steps were observed to be positively affected by the reaction tempera-
reactors. It was noticed that the temperature of reaction clearly influ- ture (Sun, Sun, Yao, Zhang, & Xu, 2010).
ences the yield of the biodiesel product (Aghel et al., 2014; Jachuck Even higher reaction temperatures were required for production of
et al., 2009; Martínez Arias et al., 2012). Using a micro-tube reactor, it biodiesel under supercritical conditions in micro-reactors (Gonzalez
was found that the conversion of sunflower oil increased with increase et al., 2013; Silva et al., 2014; Silva & Oliveira, 2014; Sootchiewcharn
in reaction temperature (Guan et al., 2009a). Martínez Arias et al., 2012 et al., 2015). An increase in the FAEE yield was reported with increase
studied the effect of temperature on biodiesel yield using three different in the reaction temperature. However, at prolonged residence times,
micro-reactors. As the temperature was increased from 30 to 70 °C the too high temperatures led to thermal decomposition of the product
biodiesel yield increased from 67.3 to 89.0%, from 73.9 to 92.2% and (Gonzalez et al., 2013; Sootchiewcharn et al., 2015). Compared to higher
from 75.9 to 92.6% for T-, Omega-, and Tesla-shaped micro-reactors, reaction temperatures, when the reactions were carried out at room
temperature, a longer residence time was required to obtain a FAME
yield above 90% (Canter, 2006).

Residence time

The residence time for biodiesel production in micro-reactors varies


with the type of micro-reactor applied. In recent years, several studies
on biodiesel production in micro-reactors have reported that an
increase in residence time is favourable for higher biodiesel yield.
Martínez Arias et al. (2012) transesterified castor oil under the condi-
tions of ethanol to oil molar ratio of 9:1, catalyst amount of 1 wt% at
60 °C in three different micro-reactors. It was observed that for first 4
min of the reaction, with increase in the residence time the biodiesel
yield increased, and after that, the yield attained almost a constant
value as shown in Fig. 6. After a residence time of 15 min, biodiesel
yield of 75.9%, 94.1%, and 93.7% was acquired for T-, Omega and Tesla
shaped micro-reactors respectively. Aghel et al. (2014) observed a sim-
ilar effect of residence time on transesterification of soya bean oil with
methanol using micro-reactor with a wire coil. The percentage of
methyl ester conversion increased from 87.9% to 99% with an increase
in residence time from 20 to 180 s. The conversion increased rapidly
Fig. 5. Influence of reaction temperature on yield of ethyl ester using an ethanol/oil molar
ratio of 9:1, NaOH concentration of 1 wt%, and residence time of 10 min in T-, omega-, and
for first 20 to 60 s and then very slightly as the residence time was
Tesla-shaped micro-reactors. Reprinted with permission from Martínez Arias et al. (2012). prolonged from 60 to 180 s. However, in the study of Sun et al. (2008)
Copyright 2012 American Chemical Society. methyl ester yield first increased with increase in residence time from
150 A. Tiwari et al. / Energy for Sustainable Development 43 (2018) 143–161

Fig. 6. Ethyl ester yield of castor oil transesterification carried out in T-, omega-, and Tesla-
shaped micro-reactors at different residence times and using an ethanol/oil molar ratio of Fig. 7. Effect of molar ratio (methanol/oil) on FAME%. Reprinted from Aghel et al. (2014),
9:1, catalyst amount of 1 wt% and temperature of 50 °C. Reprinted with permission from with permission from Elsevier.
Martínez Arias et al. (2012). Copyright 2012 American Chemical Society.
could cause a part of glycerol to remain in biodiesel phase (Leung &
3.68 to 5.89 min then on the further prolongation of residence time the Guo, 2006).
yield decreased to about 92%. It may have resulted due to weakened Compared to alkali-catalyzed transesterification higher alcohol to oil
mass transfer (decrease in internal circulations) at prolonged residence molar ratios were required under acid catalyzed transesterification of
time resulted by a decrease in average velocity for a fixed length of capil- oils with a significant amount of free fatty acid content in oil. A molar
lary (Dummann et al., 2003; Sun et al., 2008). Under supercritical condi- ratio of 30:1 and 20:1 was used for esterification and transesterification
tions, the residence time for transesterification of oils in micro-reactors respectively, with sulfuric acid as a catalyst to obtain a yield of 99.5% in a
can be shortened up to a few seconds. Using supercritical methanol in ab- stainless-steel capillary micro-reactor. In the esterification step, the
sence of catalyst, a complete conversion of triglyceride was achieved in a conversion of FFA (free fatty acid) increased with increase in methanol
micro-channel reactor at a residence time of 5 min (Schürer et al., 2014). to FFA molar ratio from 10:1 to 30:1. Whereas, in the transesterification
However, when the supercritical transesterification was performed in step the yield first increased from 96.1 to 99.9% as the methanol to tri-
presence of a heterogeneous catalyst coated on the micro-channels, a glyceride molar ratio was increased from 6 to 20 and then decreased
complete conversion was obtained in a very short residence time of on further increase in the molar ratio (Sun et al., 2010). Under supercrit-
b15 s (Schürer et al., 2014; Schürer et al., 2016). ical conditions, much higher alcohol to oil molar ratios were required in
comparison to subcritical transesterification (Gonzalez et al., 2013; Silva
& Oliveira, 2014; Trentin et al., 2011). Hence, transesterification in micro-
Alcohol to oil molar ratio reactors is positively affected by an increase in alcohol to oil molar ratio,
but too high alcohol to oil molar ratios may lead to a decrease in biodiesel
The molar ratio of alcohol to oil is one of the most important param- yield. Moreover, the alcohol to oil molar ratio depends on the type of
eters affecting the yield of alkyl ester. Stoichiometrically, 3 mol of alco- micro-reactor applied, reaction conditions (subcritical or supercritical)
hol is required for transesterification of 1 mol of oil to yield 3 mol of and catalyst used for biodiesel production.
alkyl ester and 1 mol of glycerol. However, an excess of alcohol to oil
molar ratio results in greater oil conversions (Freedman et al., 1984). Effects of geometrical configurations
A linear increase in fatty acid methyl ester yield was observed as the
methanol to cottonseed oil molar ratio was raised from 3:1 to 9:1 be- Effect of channel size
tween 60 and 80 °C with 1 wt% KOH catalyst, in a micro-reactor assem- The micro-channel size in a micro-reactor has a strong influence on
bly consisting of rectangular interdigital micromixer (RIMM) connected the yield of methyl ester. It could be attributed to different mass transfer
to a stainless-steel delay loop (Sun et al., 2009). Aghel et al. (2014) ex- distance and rate between the oil and alcohol phases in various micro-
amined the effect of alcohol to oil molar ratio on ester yield with alkali channel sizes. Using 0.25 mm inner diameter capillary micro-reactor the
catalyst using methanol to oil molar ratios of 6:1, 9:1 and 12:1 in two methyl ester yield of 98.8% was obtained in a residence time of 6 min at
different micro-reactors, one with an inserted wire coil and other with- methanol to oil ratio of 6:1 and 1 wt% KOH concentration while the
out the coil. The results indicated that increase in a molar ratio from 6:1 yield reached only 96.7% in a residence time of 8.2 min when 0.53 mm
to 9:1 led to 4.6% and 7.2% improvement in the percentage of methyl inner diameter capillary micro-reactor was used under same operating
ester for micro-reactor with and without coil respectively as shown in conditions (Sun et al., 2008). Guan et al. (2009a) observed a linear in-
Fig. 7. However, as the molar ratio was increased above 9:1, a slight de- crease in sunflower oil conversion as the micro-tube diameter was de-
crease in the percentage of methyl ester was observed. This behavior creased from 1 mm to 0.8 mm, 0.6 mm and 0.4 mm. A similar relation
between methanol to oil molar ratio and ester yield agreed with Sun between micro-channel size and biodiesel yield was noticed when zig-
et al. (2008) observations. Sun et al. (2008) concluded that in a capillary zag micro-channel reactors with a hydraulic diameter between 240 and
micro-reactor when the methanol to soya bean oil molar ratio was in- 900 μm were used for transesterification of soya bean oil. The micro-
creased from 3 to 6, the methyl ester yield increased from 82.1% to channel with smallest hydraulic diameter gave the maximum biodiesel
99.3% at 60 °C and 1 wt% KOH concentration. However, a further in- yield as shown in Table 3 (Wen et al., 2009). Kalu, Chen, and Gedris
crease in the molar ratio resulted in a decrease in yield to 95.3%. This un- (2011) studied the effect of channel depth on the performance of the re-
expected decrease in fatty acid methyl ester yield was probably because actor using slit channel reactor for biodiesel synthesis. The result indicated
at too high methanol to oil ratio; methanol acts as an emulsifier that that with a decrease in channel depth from 10, 5 and 2 to 1 mm the
A. Tiwari et al. / Energy for Sustainable Development 43 (2018) 143–161 151

Table 3 slit interdigital micromixer (SIMM-V2) followed by stainless-steel cap-


Geometric parameters and performances of micro-channel reactors. Reprinted from Wen illary with an i.d. of 0.6 mm. The ester yields obtained using T and J
et al. (2009), with permission from Elsevier.
mixer were almost equal and half of those achieved with the RIMM
Name Section (μm × μm) Hydraulic diameter (μm) Turns Yield (%)a and SIMM-V2 as represented in Fig. 8(a). It may be due to smaller
Zig-zag-1 200 × 300 240 350 97.3 mixer dimensions and more intensified mixing in RIMM and SIMM-V2
Zig-zag-2 300 × 500 375 350 91.3 micromixers. Similarly, three different mixers such as T-type, cross(†)
Zig-zag-3 500 × 500 500 350 81.1 type and split and recombine were used for methanolysis of sunflower
Zig-zag-4 500 × 900 643 350 80.9
oil. Micro-reactor with †-type mixer appeared to be better than the
Zig-zag-5 900 × 900 900 350 77.8
Zig-zag-6 500 × 500 500 10 60.0 micro-reactor with a T-type mixer and the micro-reactor with split
Zig-zag-7 500 × 500 500 50 63.2 and recombine micromixer (Bhoi et al., 2014). Rahimi et al. (2016) de-
Zig-zag-8 500 × 500 500 100 70.2 signed three unlike micromixers with different confluence angles of 45°,
Zig-zag-9 500 × 500 500 200 79.8 90° and 135° namely E1, E2 and E3 respectively (Fig. 4) for biodiesel
a
Reaction conditions: molar ratio of methanol/oil 6, catalyst amount 1% (based on oil production from soya bean oil. The results demonstrated that the reac-
weight), temperature 60 °C, and residence time 28 s. tion was most enhanced in E1 mixer followed by E3 and E2 respectively.
Hence from the above review, it can be concluded that efficiency the
conversion of soya bean oil increased. With increasing depth, the chan- mixer is an effective parameter on the transesterification of oil to fatty
nels approach the batch system and hence are less efficient than shallow acid alkyl ester. It can be attributed to varying dimensions of different
channels. mixers and the flow behavior at the outlet of the mixers. Apart from
the mixer type, Shaaban et al. (2015) investigated the effect of mixer
Effect of inlet mixer type inner diameter and geometry on the yield of fatty acid ethyl ester at
The mixing mechanism has a strong influence on the reaction when same reactor volume, flow rate and residence time using three different
it is carried out between two immiscible liquids. Therefore, the biodiesel inlet mixers T1, T2 and T3 as shown in Fig. 8(b). The T1 mixer recorded
production in micro-reactor is affected by the configuration of mixer ap- highest conversion of 97% in 80 s using 9:1 methanol to oil molar
plied to enhance the mixing of alcohol and oil phase. The microfluidic ratio. This was ascribed to its smaller internal dimensions which led to
systems for biodiesel synthesis consist of an integrated mixer to gener- formation of smaller slugs and increase in interfacial area.
ate dispersion, along with the reaction channel. A FAME yield of 98.80%
was obtained in a residence time of 6 min when instead of a microfluidic Effect of internal geometries
mixer, the reactants were mixed in a batch reactor then injected into Apart from diffusion, advection is another important form of mass
0.25 mm capillary micro-reactor. The yield already reached 81% before transfer in a microfluidic system. The so-called chaotic advection can
the mixture was pumped into the micro-reactor (Sun et al., 2008). significantly improve mixing of oil and alcohol. For passive mixing in
Guan et al. (2008) carried out transesterification of sunflower oil using micro-channels, the chaotic advection can either be introduced by mod-
a T-type and a split and recombine micromixer. The oil conversion ifications in channel geometry or by inserting an obstacle in the channel
was higher when the micromixer was used instead of a T-type mixer. (Nguyen & Wu, 2005). Several studies have been conducted for the pro-
Applying the split and recombine micromixer it was noticed that sun- duction of biodiesel in micro-reactors with simple Y and T structures
flower oil could be completely converted to biodiesel in a residence (Canter, 2006; Guan et al., 2009a; Jachuck et al., 2009). Moreover,
time of as short as 112 s. Even shorter residence time of 14.9 s was ob- there are few efforts on modified channel configurations for biodiesel
served to reach a yield of 99.5% when zig-zag mixer channels with a synthesis. Using a zig-zag micro-channel reactor, the efficiency of biodie-
zig-zag reaction channel (Fig. 2) were used for the production of biodie- sel production process improved significantly. The FAME yield reached
sel using soya bean oil (Dai et al., 2014). 99.5% in few seconds under mild operating conditions (Dai et al., 2014;
Sun et al. (2009) studied the effect of mixer type on the biodiesel Wen et al., 2009). Martínez Arias et al. (2012) investigated the effect of
yields using four different mixers as shown in Fig. 3(A), namely, a T- channel geometries on alkali-catalyzed transesterification of castor oil
mixer, a J-mixer, a rectangular interdigital micromixer (RIMM) and a using a T-shaped, Tesla shaped and Omega shaped micro-reactor. Under

(a) (b)

Fig. 8. (a) FAME yields of different micromixers at 60 °C for a methanol to oil molar ratio of 8:1, a KOH concentration of 1 wt%, a total flow rate of 1 mL/min, and a residence time of 44 s:
(T) T-mixer, (J) J-mixer, (RIMM) rectangular interdigital micromixer, (V2) (SIMM-V2) split interdigital micromixer. Reprinted with permission from Sun et al. (2009). Copyright 2010
American Chemical Society. (b) Specifications of the micromixers used by Shaaban, El-Shazly, Elkady, and Ohshima (2015). Reprinted with permission from Shaaban et al. (2015).
152 A. Tiwari et al. / Energy for Sustainable Development 43 (2018) 143–161

Table 4
Ethyl ester yields for different volumetric rates using ethanol/oil molar ratio of 9:1, catalyst
amount of 1 wt% and temperature of 50 °C. Reprinted with permission from Martínez
Arias et al. (2012). Copyright 2012 American Chemical Society.

Geometry Volumetric rate Residence Ethyl ester conversion (%)


(mL/h) time
(min)
Castor Ethanol T-shaped Omega Tesla
oil shaped shaped

Transsection (μm × μm), 15.0 8.1 1 58.9 67.1 69.2


500 × 500 hydraulic 7.5 4.1 2 69.4 79. 3 83.8
dia, 500 μm 5.0 2.7 3 73.1 87.5 88.8
3.8 2.1 4 74.7 88.3 91.5
3.0 1.6 5 71.1 89.7 91.1
2.1 1.1 7 75.1 88.5 91.7
1.5 0.8 10 73.4 87.4 92.2 Fig. 9. Effect of co-solvent volumetric ratio (THF/methanol) on percentage biodiesel yield.
1.0 0.5 15 75.9 91.4 93.7 Reprinted with permission from Elkady et al. (2015).

than that of methanol. Therefore, at the end of the reaction, the


unreacted methanol and THF can be co-distilled and recycled (Elkady
same operating conditions, the Omega and Tesla shaped micro-reactors et al., 2015). Rahimi et al. (2016) studied the influence of co-solvent
resulted in higher ethyl ester conversions, followed by T-shaped micro- (hexane) to methanol volumetric ratio on the FAME content. Using an
reactor as shown in Table 4. It can be ascribed to chaotic flow in Tesla E1 mixer (see Fig. 4), an increase in the percentage of FAME from 81%
and Omega micro-reactor which favoured better contact between the to 88% was reported as the co-solvent to methanol volume ratio was in-
oil and alcohol phase. creased from 0.1 to 0.4 at 45 °C, 2:1 oil to methanol volumetric ratio and
Recently, Aghel et al. (2014) developed a micro-reactor with an 9 s residence time. However, on further increasing the ratio, the FAME
inserted wire coil to improve the mixing efficiency. Compared to a sim- content decreased. It may be due to dilution of reactants at large
ple T-shaped micro-reactor used in their previous work (Rahimi et al., amounts of co-solvent (Mohammed-Dabo, Ahmad, Hamza, Muazu, &
2014), the micro-reactor equipped with wire coil produced a higher Aliyu, 2012; Pardal, Encinar, González, & Martínez, 2010; Zhang, Jin,
percentage of FAME at the same reaction conditions. Similarly, the effi- Zhang, Huang, & Wang, 2012).The addition of co-solvent to alcohol to
ciency of micro-channel reactor with static elements (MSE) in terms of overcome the immiscibility improves the contact between the alcohol
the mixing and reaction was also compared with plain T-shaped micro- and oil phase but increases the number of downstream purification
channel reactor without static elements. It was found that the micro- steps.
channel with static elements showed superior performance of mixing
index and oil conversion due to induced changes in the direction
of flow, disturbances in boundary layer and formation of vortex Effects of catalyst type and concentration
(Santana et al., 2017). From the above review, it is noticeable that better
conversion in micro-reactors with modified micro-channel structure Depending on the characteristics of feedstock, the nature of catalyst
can be obtained as a result of chaotic mixing in comparison to laminar employed plays an important role for transesterification of triglycerides
mixing in simple T and Y micro-channels. to biodiesel. The production of biodiesel can be carried out by acid, alkali
or enzyme-catalyzed processes or in the absence of catalyst using super-
Effect of co-solvent critical methanol. The alkali and acid catalysts include both homoge-
neous and heterogeneous catalysts. Homogenous alkaline catalysts are
The biggest obstacle in the transesterification of vegetable oil is most commonly used to catalyze transesterification due to their low
immiscibility of alcohol and oil phase that slows down the reaction cost, shorter reaction times and high conversion under mild operation
significantly. In order to overcome slow reaction rates caused by the ex- conditions (Casas, Fernández, Ramos, Pérez, & Rodríguez, 2010;
tremely low solubility of the alcohol in the triglyceride phase, addition Demirbas, 2007; Freedman, Butterfield, & Pryde, 1986; Wang, Ou, Liu,
of an organic co-solvent has been recommended by Meher et al. & Zhang, 2007). However, the process involving homogeneous alkaline
(2006). Several works on transesterification of triglycerides to fatty catalysts needs feedstock with free fatty acid content b1 wt%. In the case
acid alkyl ester using an organic co-solvent have been performed at of high FFA levels, a considerable amount of alkaline catalyst reacts with
micro-scale to take the advantages of micro-reactors simultaneously FFAs to form soaps which hinder biodiesel purification process and dis-
(Elkady et al., 2015; Guan, Sakurai, & Kusakabe, 2009b; Rahimi et al., turbs the flow patterns in micro-reactors (Guan et al., 2010).
2016). An improved process was investigated for methanolysis of sun- Although alkali catalyzed, reactions are much faster compared to
flower oil with co-solvent in a micro-tube reactor. At a methanol to oil acid catalysts, to overcome the difficulties due to soap formation acid
molar ratio of 8:1 and 25 °C, oil conversion was considerably high in catalyzed transesterification is preferable when high content of FFA
the presence of co-solvent (Diethyl ether, DEE) as compared to conver- (N1 wt%) is present in the oil. The acidic catalysts yield high biodiesel
sion obtained without co-solvent. It was observed that the flow was ho- content for high FFA feedstock, but it is greatly ignored due to long reac-
mogenous at the entrance of the micro-tube and with the formation of tion times, high catalyst concentration and a large amount of methanol
immiscible glycerol the homogenous flow was disturbed (Guan, requirement (Sun et al., 2010). The heterogeneous catalysts are still
Sakurai, & Kusakabe, 2009b). under development, and only a limited number of investigations have
Using Tetrahydrofuran (THF) as co-solvent, transesterification of been done on heterogeneous catalysts for the production of biodiesel
waste vegetable oil was observed at methanol volumetric ratio ranges in micro-reactors (Kurayama et al., 2013; Kurayama et al., 2010;
of 0.2 to 1 in a KM micromixer. As shown in Fig. 9, an increase in the bio- Machsun et al., 2011; Schürer et al., 2014; Schürer et al., 2016). The het-
diesel yield was seen with an increase in the THF to methanol volumet- erogeneous catalyzed process exhibits some advantages, such as good
ric ratio. A maximum yield of 97.30% was obtained at THF to methanol tolerance to free fatty acids, simple product purification, no waste
volumetric ratio of 0.3:1. However, while increasing the ratio above water treatment and mild operating conditions but because of high pro-
0.3:1, no noticeable increase in the yield was found. In particular, THF duction cost and short lifetime, their application for biodiesel produc-
was chosen because its boiling point of 67 °C is only two degrees higher tion is limited (Gorji, 2015; Sun et al., 2010).
A. Tiwari et al. / Energy for Sustainable Development 43 (2018) 143–161 153

Apart from catalyzed process in micro-reactors, research has been concentration led to higher triglyceride conversion. However, the rate
done using supercritical alcohols for biodiesel synthesis (Silva et al., of change of conversion on catalyst loading was negligible at catalyst con-
2012; Farobie, Sasanami, & Matsumura, 2015; Gonzalez et al., 2013; centrations above 0.75% (Jachuck et al., 2009). Martínez Arias et al. (2012)
Schürer et al., 2014; Schürer et al., 2016; Silva et al., 2014; carried out biodiesel synthesis in three different micro-reactors i.e. a T-,
Sootchiewcharn et al., 2015). However, the high cost of the production Tesla, and Omega shaped micro-reactors using different NaOH concentra-
process in the supercritical process due to strict demand of equipment tions (0.5, 1.0, and 1.5 wt%) for each micro-reactor. The residence time,
and high temperature and pressure requirements is a major limitation temperature and ethanol to castor oil molar ratio were held constant at
(Chen et al., 2008). Besides the type of catalyst, the concentration of cata- 10 min, 50 °C and 9:1 respectively. It was observed that with an increase
lyst used in the reaction has an optimum value below which the reaction in catalyst concentration from 0.5 to 1.5 wt% the yield of ethyl ester in-
remains incomplete, and an excess catalyst amount may lead to reinforc- creased from 50.6 to 79.1%, from 54.3 to 96.2% and from 56.3 to 98.9%
ing the lateral saponification reaction (Dunn, 2001). The optimum con- for T-, Omega and Tesla shaped micro-reactors, respectively. Likewise,
centration of catalyst depends on the operating parameters as well as NaOH was tested as a catalyst in a range of 0.5 to 2 wt% for biodiesel pro-
catalyst type (Gorji, 2015). Several investigations on the effect of catalyst duction from waste vegetable oil in a KM micromixer. A maximum yield
concentration on biodiesel yield in micro-reactors have been performed of 95% was achieved with 1 wt% catalyst concentration. Using 0.5 wt% cat-
(Guan et al., 2009a; Kaewchada, Pungchaicharn, & Jaree, 2016; Rashid, alyst concentration the oil conversion was observed to be incomplete
Uemura, Kusakabe, Osman, & Abdullah, 2014). A similar influence as ob- while biodiesel yield decreased significantly as the NaOH concentration
served for the conventional batch process was observed in micro- was increased above 1 wt% (Elkady et al., 2015). Santana et al. (2016)
reactors. studied the relation between fatty acid ethyl ester percentage and catalyst
concentration. The results indicated that the percentage of FAEE (fatty
Homogeneous catalysts acid ethyl ester) yield increased slightly from 89.13% to 89.89% when
The majority of research for biodiesel production using micro-reactor the NaOH concentration was increased from 0.2 to 0.85 wt%, but it dimin-
technology has been done using low FFA feedstock or pre-treated feed- ished when NaOH concentration was higher than 0.85%.
stock in the presence of homogeneous alkaline catalysts. On the contrary, Crawford et al. (2008) performed transesterification of triolein cata-
transesterification of high acid value oils catalyzed by homogeneous acid lyzed by sodium methoxide in a commercial Syrris 250 μL micro-reactor.
catalyzed has not been explored much. It may be due to simplicity, and Using a 0.1 M catalyst concentration, a complete conversion of triolein to
strong catalytic activity of an alkali-catalyzed process in which the reac- methyl oleate was obtained at room temperature in 2.5 min. It has been
tion proceeded in a single step while using acid catalyst required removal reported that compared to sodium hydroxide (NaOH), sodium methoxide
of water as it may hinder the transesterification process to some extent (NaOCH3) is more efficient because it is disintegrated into CH3O− and
(Guan, Teshima, et al., 2009; Sun et al., 2010). However, use of alkali cat- Na+ and does not form water as in the case of NaOH/KOH. Additionally,
alyst is limited to oils with low acid content because of the formation of in contrast to sodium hydroxide the required amount of sodium
soaps when a great amount of free fatty acid is present which disturbs methoxide is about 50% lower, but due to its high cost it is not used widely
the flow behavior in micro-channels. (Shahid & Jamal, 2011).

Alkaline catalysts. In micro-reactors, potassium hydroxide and sodium Acid catalysts. Despite the fact that alkaline catalysts are very efficient
hydroxide are the most investigated homogeneous base catalysts for and common for production of biodiesel, these catalysts do not exhibit
transesterification of vegetable oils. Sun et al. (2008) used KOH as a cat- good results when the acid value and water content in oil are high
alyst, with methyl alcohol to convert the cottonseed oil into biodiesel. (N1 wt%). The water contents and acid values of most of the non-
By varying the concentration of catalyst at a constant methanol to oil edible and cheap feedstock are higher than performance range of base
molar ratio of 6, residence time of 6 min and 60 °C reaction temperature, catalysts. In such cases, acid catalysts are preferred (Fukuda, Kondo, &
they concluded that an increase in KOH concentration from 0.40 wt% to Noda, 2001). Sulfuric acid is the most widely used acid catalyst, which
1 wt% led to an increase in methyl ester yield from 86% to a maximum of can be used to conduct both esterification and transesterification
99.3% but with further increase in KOH concentration the FAME yield simultaneously.
decreased to 94.80%. This reduction in FAME yield was ascribed to sa- Acid catalyzed transesterification of high FFA oils in micro-reactors
ponification of oil with KOH since an increase in soap concentration has not been examined much. Sun et al. (2010) carried out production
was observed from 0.2 to 0.34 wt%, when catalyst concentration was in- of biodiesel with acid oil as a feedstock by a two-step sulfuric acid cata-
creased from 1.0 to 1.2 wt%. Similarly, the influence of catalyst amount lyzed process in a micro-reactor assembled with a split interdigital
was studied on transesterification of soya bean oil with three different micromixer (SIMM-V2) and a 0.6 mm i.d. stainless-steel capillary
KOH concentrations of 0.6, 1.2 and 1.8 wt%. The results indicated that delay loop. Esterification of oleic acid and transesterification of cotton-
the percentage of methyl ester increased when the catalyst concentra- seed oil with methanol were investigated as model reactions to explore
tion was raised from 0.6 to 1.2 wt% while a slight decrease in the propor- optimum condition for the two-step process. Based on these reactions,
tion of methyl ester was observed when a KOH concentration of 1.8 wt% an acid catalyzed two-step process was developed with the first and
was used (Aghel et al., 2014). second steps conducted separately under the optimized esterification
Dai et al. (2014) performed series of experiments using KOH as a cat- and transesterification conditions. The results indicated that the
alyst in the range of 0.8–1.5 wt% for converting soya bean oil to methanol acid value of the acid oil was reduced from 160 to 1.1 mg KOH/g,
in a zig-zag micro-channel reactor. A maximum yield of 99.5% was ob- with a methanol to the acid molar ratio of 30, the H2SO4 concentration
tained at 59 °C using 1.2 wt% KOH and 8.5:1 methanol to oil molar ratio of 3 wt%, and a residence time of 7 min at 100 °C in the first step. The
in 14.9 s. Kaewchada et al. (2016) and Azam, Uemura, Kusakabe, and final FAME yield reached 99.5% with a methanol to triglyceride molar
Bustam (2016) carried out KOH catalyzed transesterification of palm oil. ratio of 20, the H2SO4 concentration of 3 wt%, and a residence time of
It was noted that an increase in the concentration of catalyst positively af- 5 min at 120 °C in the second step. Thus, biodiesel can be produced in
fected the biodiesel yield. When NaOH is used as a catalyst, the reaction a residence time of b15 min by an acid catalyzed process in a micro-
temperature applied is usually low to avoid emulsification and saponifica- reactor. However, higher concentration of catalyst is required for acid cat-
tion (Xie et al., 2012). Experiments were conducted using four different alyzed transesterification compared to base catalyzed transesterification.
NaOH concentrations from 0.25 to 1 wt%, for transesterification of canola
oil in a small channel reactor. The residence time, methanol to oil molar Heterogeneous catalysts
ratio (6:1), temperature and pressure of reaction were kept constant for Homogeneous catalysts are very efficient and are most frequently
each catalyst loading. The results indicated that the use of higher catalyst used in biodiesel production. However, these catalysts require excessive
154 A. Tiwari et al. / Energy for Sustainable Development 43 (2018) 143–161

washing for their removal from the final product and subsequent treat- Schürer et al. (2014) performed transesterification of tricaprin in
ment of waste water. As a result, a lot of water, energy and time are con- stainless-steel micro-channels. A complete conversion was achieved in
sumed. Besides that, these catalysts cannot be reused (Fukuda et al., residence time of 5 min at 40-fold excess of methanol, 200 bar reaction
2001). The use of heterogeneous catalysts offers several advantages pressure and 375 °C reaction temperature. However, by application of
over homogeneous catalysts, such as; simple product separation, simple heterogeneous catalyst under supercritical conditions the required
catalyst recovery, catalyst reusability, less energy requirements and less pressure was reduced to 100 bar and a residence time of b30 s was re-
added cost of purification (Atadashi, Aroua, Aziz, & Sulaiman, 2013). quired for 100% conversion.
Biodiesel synthesis in micro-reactors applying heterogeneous catalysts
is still under investigations. To the best of our knowledge only heteroge- Liquid-liquid two-phase flows in micro-reactors for
neous alkaline and enzyme catalysts have been examined to catalyze biodiesel production
transesterification of oils in micro-reactors. Kurayama et al. (2013)
used CaO loaded micro-capsules as a micro-reactor for biodiesel It is crucial to study hydrodynamics in the multiphase reactor be-
production from rapeseed oil. The micro-capsules were prepared by cause different flow patterns influence mass transfer and axial disper-
encapsulation of CaO particles with butanol modified alginate as a sion, which affect the conversion and selectivity of the reactions.
shell material using co-extrusion technique. It was suggested that Depending on the volumetric flow ratio of the two phases and total
CaO-loaded micro-capsules could be successfully reused for three flow rate, different flow patterns can be attained in micro-channels. In
times without the loss of the catalytic activity. Under supercritical con- two phase liquid-liquid system the conventional flow regimes include
ditions, the application of heterogeneous catalyst coatings in micro- as droplet (Fig. 10A), parallel (Fig. 10B), or slug flow (Fig. 10C). For
channels for biodiesel production was first investigated by Schürer liquid-liquid two-phase mass transfer limited reactions, application of
et al. (2014). They performed experiments on transesterification of slug flow has been suggested (Burns & Ramshaw, 2001; Dummann
tricaprin under supercritical conditions using alumina as a catalyst, et al., 2003; Kashid et al., 2005). In slug flow, the mass transfer between
coated onto stainless-steel micro-channels by applying wash coating two immiscible phases is significantly intensified due to the presence of
process. A complete conversion was obtained in a very short residence internal circulations (Burns & Ramshaw, 1999, 2001; Kashid et al., 2005;
time of the 30 s. In their successive work, Schürer et al. (2016) designed Khan, Günther, Schmidt, & Jensen, 2004). These internal circulations are
a small-scale heterogeneous alkali catalyzed (30 wt% La2O3 supported induced as a result of shear force between the two liquid phases as well
by γ-Al2O3) plant with annual capacity of 40 t for biodiesel synthesis as by liquid/wall friction (Malsch et al., 2008). It has been reported that
under supercritical conditions. The implementation of heterogeneous enhanced mixing and rapid reaction rates can be achieved using seg-
catalyst to micro-reactors minimized the residence time significantly. mented flow with internal circulations (Burns & Ramshaw, 1999; Tice,
Machsun et al. (2011) used a bio catalytic membrane micro-reactor Song, Lyon, & Ismagilov, 2003).
with immobilized lipase in the pores of an asymmetric polyether sul- Using a quartz capillary micro-reactor, Sun et al. (2008) observed
fone (PES) membrane for transesterification of triolein with methanol. that the methanol and oil phases were separated from each other
Each membrane pore acted as a micro-reactor in which the reaction oc- forming a slug flow due to high interfacial forces between the two
curred. A maximum conversion of 80% was observed in a residence time phases. An alternating long oil slug and a short methanol slug was
of 19 min. No activity decay of immobilized lipase was observed over a formed at the capillary inlet. They also reported a presence of a thin
period of 12 days of continuous operation. methanol wall film surrounding the oil slug due to superior wettability
of methanol phase. The dimensions of oil slugs decreased with a de-
Catalyst-free process crease in capillary diameter suggesting an increased mass transfer at
In the catalyst-free process, transesterification of oils is performed in smaller channel diameters due to an increase in interfacial area. As the
the absence of catalyst using supercritical alcohol at extremely high reaction between the two phases progressed, a similar alternative slug
pressure and temperature (Demirbaş, 2002; Kusdiana & Saka, 2001; flow of methyl ester and glycerol was seen at the outlet part of the cap-
Saka & Kusdiana, 2001; Schürer et al., 2014; Warabi et al., 2004). Biodie- illary. However, the flow patterns in the middle of capillary were not
sel production by the catalyst-free supercritical process has several clearly identified and the effect of reaction conditions and geometrical
advantages over catalytic processes, including higher reaction rates, configurations of micro-reactor were not investigated. Assuming droplet
easier separation, and purification of products, environment friendli- flow could occur in tubular micro-reactor, Lopez-Guajardo et al. (2017)
ness and more tolerance to the presence of water and free fatty acids suggested application of the droplet flow to achieve high interfacial area
(Kusdiana & Saka, 2004; Rathore & Madras, 2007). and intense local mixing for higher conversions. At a constant methanol
Bertoldi et al. (2009) and Trentin et al. (2011) carried out continuous volume of 3.97 × 10−7 m3, the highest interfacial area was reported for
biodiesel production in a stainless-steel micro-tube reactor through droplet flow as shown in Table 5. Yeh, Huang, Cheng, Cheng, and Yang
catalyst-free transesterification of soya bean oil in supercritical ethanol (2016) devised a millimetrically scaled device that employed a droplet-
in presence of CO2 as a co-solvent. Bertoldi et al. (2009) performed ex- based co-axial fluidic system to complete alkali-catalyzed
periments in temperatures between 300 and 350 °C, pressure from 7.5 transesterification for biodiesel production. The large surface area-to-
to 20 MPa, oil to ethanol molar ratio of 1:6 to 1:40, co-solvent to sub- volume ratio of the droplet-based system, and the internal circulation in-
strate mass ratio from 0:1 to 0.5:1. An appreciable yield was obtained duced inside the moving droplets, significantly enhanced the reaction
at 350 °C and 10 MPa pressure, using oil to ethanol molar ratio of 1:40 rate of immiscible liquids i.e. soya bean oil and methanol. The droplet-
and CO2 to substrate mass ratio of 0.05:1. Trentin et al. (2011) used based co-axial fluidic system performed better than other methods of
the similar experimental reaction system previously used by Bertoldi continuous-flow production. This study demonstrated the high potential
et al. (2009) but a different range of operating parameters. In the inves- of droplet-based fluidic chips for energy production.
tigated experimental range it was observed that temperature, pressure,
and co-solvent to substrate mass ratio had a positive effect on FAEE Factors affecting flow patterns
(fatty acid ethyl ester) yield and significant yields were achieved at
325 °C, 20 MPa, oil to ethanol molar ratio of 1:20, 0.8 mL/min substrates Effect of temperature on flow patterns
flow rate and CO2 to substrate mass ratio of 0.2:1. Sootchiewcharn et al. Temperature clearly influences the methanol-oil two-phase flow be-
(2015) investigated the production of FAEE from refined palm oil with havior in micro-channels as the miscibility of oil and methanol increases
supercritical ethyl acetate in a micro-reactor. A biodiesel yield of 78.3% with increase in temperature (Čerče et al., 2005; Zhou et al., 2006). Guan
was obtained at oil to ethyl acetate molar ratio of 1:50, the temperature et al. (2009a) used a transparent fluorinated ethylene propylene (FEP)
of 350 °C and 20 min of residence time. Using supercritical methanol, tube to observe the flow pattern during biodiesel production under
A. Tiwari et al. / Energy for Sustainable Development 43 (2018) 143–161 155

Fig. 10. Scheme of flow patterns of the reactants during the synthesis of biodiesel in the tubular micro-reactor: (A) droplet flow; (B) parallel flow; (C) slug flow. Reprinted with permission
from López-Guajardo et al. (2017). Copyright 2017 Taylor & Francis.

different reaction conditions. In order to obtain a clear view of flow, liquid two-phase slug-annular flow, similar to that at 80 °C during all re-
methanol was dyed with inert red phloxine B which is also soluble in action times, was formed. Thus, the reaction temperature has a significant
glycerol but insoluble in oil. The flow patterns were observed at differ- influence on the flow behavior.
ent temperatures for two different alcohol to sunflower oil molar ratios
of 4.6 and 23.9. A gradual change in flow patterns was observed with in- Effect of alcohol on oil molar ratio
crease in temperature from 20 to 60 °C. When the methanol to oil molar It has been reported that alcohol to oil molar ratio strongly affects
flow ratio was 4.6, at 20 °C a clear segmented flow was formed 400 mm the flow patterns (Guan et al., 2009a; Sun et al., 2009). At a methanol
apart from the tube inlet and then fine red droplets composed of meth- to oil molar ratio of 4.6, short methanol segments dispersed in long oil
anol and glycerol were observed in oil slugs which agglomerated and segments were observed. When the methanol to oil molar ratio was in-
merged into the methanol segment to form larger red segments near creased to 23.9, the number of methanol segments per unit length of the
the exit region of tube (Fig. 11a). On the increase in temperature, the ag- tube slightly decreased compared to the number of segments at 4.6
glomeration of segments began closer to the tube inlet (Fig. 11b). At 60 °C methanol to oil molar ratio as shown in Fig. 11. Moreover, the length
the agglomeration started at 100 mm from the inlet of the tube, and a of methanol segments increased, while the length of oil segments de-
quasi-homogenous phase was formed at the outlet region of the micro- creased. The formation of red droplets containing glycerol and methanol
tube (Fig. 11c). When the tube outlet was cooled to room temperature, and their aggregation was more intense for methanol to oil molar ratio
the quasi-homogenous segments were separated into a segmented flow of 23.6, and a complete conversion was obtained at 300 mm from tube
of red segments containing methanol and glycerol, and oil segments con- inlet (Guan et al., 2009a). Sun et al. (2009) reported that on increasing
taining unreacted oil and FAMEs. For methanol to oil molar ratio of 23.9, the methanol to oil molar ratio from 3:1 to 15:1 at 60 °C, the droplets
agglomeration of the red droplets appeared at 100 mm and 50 mm of glycerol at the outlet part of the tube are enlarged gradually. Thus,
from reaction inlet at 20 °C and 60 °C respectively. Hence, the time re- depicting higher conversion at higher methanol to oil molar ratio.
quired to reach the quasi-homogenous phase became shorter as the tem-
perature is increased. Effect of inlet mixer type
Previous research on biodiesel synthesis studied flow patterns up to The mixer type also has a strong influence on the flow pattern in
temperatures below the boiling point of methanol. Sun et al. (2009) micro-reactors. Sun et al. (2009) observed the flow patterns at the out-
used a rectangular interdigital micromixer (RIMM) coupled with poly- let of four different types of micromixers, namely, a T-mixer, a J-mixer, a
vinyl chloride (PVC) tube to examine the flow behavior both below rectangular interdigital micromixer (RIMM) and a slit interdigital
and above the boiling point of methanol. At 60–65 °C a fully developed micromixer (SIMM-V2). As shown in Fig. 12, a uniform slug flow with
droplet flow of methanol dispersed in oil phase was seen at the outlet of methanol slug dispersed in oil phase was formed at the outlet T-mixer
the mixer. As the temperature was increased, the flow transformed into a and J mixer, whereas a mixture of some slugs and many methanol drop-
mixture of many droplets and a small number of bubbles at 70 °C and to a lets appeared at the outlet of RIMM and SIMM-V2. Hence, the contact
mixture of many bubbles and a smaller number of droplets at 75 °C. Fi- area between the methanol and oil phase was significantly increased
nally, at 80 °C a gas-liquid two-phase slug-annular flow (Yue, Chen, using RIMM and SIMM-V2 resulting in higher FAME yields. Similarly,
Yuan, Luo, & Gonthier, 2007) was formed. At the outlet part of the tube, Bhoi et al. (2014) investigated the flow in micro-reactors with three dif-
a two-phase slug flow was observed between 60 and 70 °C and however, ferent mixers, a T-type, a cross (†) type, and a split and recombine mixer
some methanol bubbles were seen in the slugs at 70 °C. At 75 °C, a gas- at various flow rates and constant methanol to oil molar ratio of 10.3:1
during methanolysis of sunflower oil. The flow pattern for micro-reactor
Table 5 with T-mixer was mainly a parallel flow with intermittent droplets
Estimated total interfacial areas for different flow patterns with a constant methanol vol- which were formed due to interface instabilities. This parallel flow
ume of 3.97 × 10−7 m3. Reprinted with permission from López-Guajardo et al. (2017). existed for a complete range of flow rates depicting no change in the
Copyright 2017 Taylor & Francis.
interfacial area with the flow rate. For micro-reactor with the †-type
Flow Volume of single N° droplets-lugs Total interfacial area junction, a mixture of slug and droplet flow was observed at a flow rate
pattern droplet-slug (m3) in TMR (m2/m3) of 0.5 mL/min. When the flow rate was decreased gradually, short
Drop flow 4.2 × 10−15 9.45 × 107 6.00 × 104 slugs were seen at 0.25 mL/min which continued up to a flow rate
Slug flow 4.82 × 10−10 822 1.46 × 103 of 0.10 mL/min. As the flow rate was decreased further to 0.07 and
Parallel flow N.A.a N.A.a 6.31 × 102 0.05 mL/min annular flow was observed. Hence, in the flow rate range
a
N.A.: not available. of 0.05 to 0.5 mL/min, three different flow patterns were observed for a
156 A. Tiwari et al. / Energy for Sustainable Development 43 (2018) 143–161

Fig. 11. Total flow patterns in a transparent FEP micro-tube reactor (inner diameter = 0.8 mm, length = 1000 mm). Total flow rate =8.2 cm3/h. Arrows stand for the flow direction.
Reaction temperature: (a, d) 20, (b, e) 40, and (c, f) 60 °C. Methanol/oil molar ratio: (a–c) 4.6 and (d–f) 23.9. Reprinted with permission from Guan et al. (2009a). Copyright 2009
American Chemical Society.

cross-type junction. For micro-reactor with split and recombine mixer, at Hessel, Lowe, & Baerns, 2004; Naleini, Rahimi, & Heydari, 2015; Sahoo,
flow rates equal to 0.05, 0.07 and 0.20 mL/min a slug flow was seen while Kralj, & Jensen, 2007; Singh, Renjith, & Shenoy, 2015) but only limited
parallel flow with intermittent droplets was observed at a flow rate equal work has been performed on purification of biodiesel product in
to 0.10, 0.15, 0.25 and 0.5 mL/min. micro-reactors (Xie et al., 2012). Crawford et al. (2008) used a Syrris
Flow Liquid-Liquid Extraction (FLLEX) module for in-line continuous
Purification of biodiesel in micro-reactors separation of fatty acid methyl esters and glycerol. The FLLEX module
consisted of a microfluidic channel fitted with a polytetrafluoroethylene
Transesterification of oils in the presence of methanol in micro- (PTFE) membrane designed by Kralj, Sahoo, and Jensen (2007). The
reactors involves the formation of FAME and glycerol as the reaction separation was achieved by the addition of water to the mixture leaving
progresses. Separation of glycerol phase from biodiesel is a major step the micro-reactor chip at a rate of 100 μL/m in the FLLEX module. The
to meet the quality standards. As both FAME and glycerol are immisci- hydrophobic biodiesel phase was able to traverse the membrane
ble, they have to be separated from each other by density difference in pores, while the aqueous phase containing glycerol and methanol did
batch process. In the conventional batch process, an emulsion of biodie- not. According to a review article by Xie et al. (2012), Sun (2010)
sel and glycerol is formed which is hard to break and consequently the performed the separation of biodiesel from glycerol by washing in a
separation of the two phases is considerably slow. However, traces of micro-separator with channel dimensions of 500 × 500 × 500,000 μm.
glycerol remain in the biodiesel phase after phase separation. Conven- A sandwich flow pattern between the water and raw biodiesel was
tionally, this small amount of glycerol is removed from the biodiesel observed in the separator. The glycerol content in the final biodiesel
phase through washing followed by drying to remove the moisture product was decreased to 0.02%. Purification of biodiesel in micro-
(Perez et al., 2014; Shahid & Jamal, 2011). For microfluidics devices reactors is still under development and new technologies are needed
with small dimensions the separation of two phases on the basis of den- for continuous separation of biodiesel from glycerol. Several phase sep-
sity difference is inappropriate due to negligible gravitational forces. aration techniques have been developed and implemented for separa-
Separation and purification in microfluidic devices have gained tion of two liquid phases (Fig. 13) which may be investigated for
much importance in recent years (Hartman & Jensen, 2009; Jahnisch, separation of biodiesel from glycerol at micro-scale (Tsaoulidis, 2015).

Scope for future work

It is evident from the reported literature that production of biodiesel


through transesterification of oils with alcohol in the presence of a
catalyst have been investigated experimentally using various types of
micro-reactors, micro-tubes and micro-channels. Moreover, the advan-
tages of micro-reactors were also demonstrated and the potential of
micro-reactors for small-scale production of biodiesel fuel, rather than
Fig. 12. Flow patterns in the transparent PVC tubes (inner diameter 1.2 mm) at the outlets
of different micromixers: (a) T-mixer, (b) J-mixer, (c) rectangular interdigital micromixer
large-scale centralized production using batch reactors. Through the re-
(RIMM), (d) split interdigital micromixer (SIMM-V2). Reprinted with permission from duced footprint of the entire production system which can be made
Sun et al. (2009). Copyright 2010 American Chemical Society. possible through the application of micro-reactor, the portable biodiesel
A. Tiwari et al. / Energy for Sustainable Development 43 (2018) 143–161 157

Fig. 13. A schematic of separation devices during liquid-liquid plug or parallel flow. (a) Phase separation using a Y-splitter Xue et al. (2011), (b) phase separation by capillary forces
(Angelescu, Mercier, Siess, & Schroeder, 2010), (c) phase separation during parallel flow (Aota, Mawatari, & Kitamori, 2009) and (d) rhase separation by wettability combined with
pressure balance (Scheiff, Holbach, & Agar, 2013). Reprinted with permission from Tsaoulidis (2015). Copyright 2015 Springer.

system could be easily installed in the fields where oil-producing feed- highly desirable and this can be achieved using the concept of
stock exists rather than transporting feedstock. Thus, avoiding the numbering-up.
transportation cost and loss of oil due to inevitable drying and spilling Numbering-up is the multiple, parallel repetitions of micro-reactor
of the feedstock oil while in transit. However, in reviewing the pub- units to increase the throughput without altering the microfluidic flow
lished research on biodiesel production in micro-reactors, most of the properties (Schenk, Hessel, Hofmann, Löwe, & Schönfeld, 2003). The
production scale was relatively small and limited effort has been done to- concept of numbering-up becomes important to prove the applicability
wards the development and scale-up of micro-reactors for production of of micro-reactor at industrial scale. Kashid, Gupta, Renken, and Kiwi-
biodiesel at pilot scale (Billo et al., 2015; Schürer et al., 2016). The Minsker (2010) demonstrated the different numbering-up techniques
throughput of a single unit of micro-reactor is too low to meet the produc- such as external numbering-up and internal numbering-up as shown
tion rates at industrial scale. Hence, the scale-up of micro-reactors units is in Fig. 14(a & b). In internal numbering-up the two liquid phases are

(c)

Fig. 14. Numbering-up concept for two-phase micro-reactor. (a) Internal numbering-up and (b) external numbering-up. Reprinted with permission from Kashid et al. (2010). Copyright
2010 Elsevier Ltd. (c) Splitting distributor developed by Adamson, Mustafi, Zhang, Zheng, and Ismagilov (2006). Reprinted with permission from Adamson et al. (2006). Copyright 2006
Royal Society of Chemistry.
158 A. Tiwari et al. / Energy for Sustainable Development 43 (2018) 143–161

mixed at the junction and the mixture is distributed in several parallel additional studies are required for in-line continuous purification of
channels. The advantage of internal numbering-up is that only a single biodiesel to meet the industrial quality standards.
pumping system and one mixing element is required. However, equal
distribution of two phase mixture in parallel channels is a challenging Conclusions
task. In external numbering-up several mixing elements are arranged
in parallel and the two phases are mixed in each mixing element inde- • In recent years, biodiesel has gained significant importance with an
pendently. Hence, in external numbering-up performance of single aim to replace fossil diesel fuel. It is evident from the reported litera-
micro-reactor is ensured along with increased throughput but require- ture that many nations of the world are commercially producing bio-
ment of multiple independent reaction systems limits the use of this diesel using conventional reactors. However, conventional technique
strategy. In recent years, a new numbering-up strategy of designing a for biodiesel production faces several challenges of long residence
splitting distributor has emerged in which a single micro-channel is fab- time, high investment in equipment and manufacturing floor space,
ricated on a plate through etching and branched into multiple micro- high cost of operation and energy requirements and low production
channels to form a tree like structure as shown in Fig. 14(c). Splitting efficiency. Micro-reactor technology can be successfully applied to
distributors ensure better flow uniformity as compared to external overcome these challenges. Attributed to small diffusion distances
and internal numbering-up solutions (Adamson et al., 2006; Al- and large surface area to volume ratio in micro-reactors, heat, and
Rawashdeh et al., 2012). mass transfer are significantly intensified, thus resulting in higher
To overcome the disadvantages of internal and external numbering- conversions in shorter residence times. Moreover, the energy require-
up, Schenk et al. (2004) developed a first liquid-flow splitting unit ments in the mixing of reactants, ample floor space and standing time
which ensured the flow equidistribution. The flow equipartition was for the separation of products are also eliminated.
achieved by building up a pressure barrier. It was equipped with six • Production of biodiesel through transesterification in micro-reactors is
docking stations and three dampening elements for six micro-devices. observed to be affected by operating parameters such as temperature,
Separation-layer mixers were chosen as micro-devices, since the carry- alcohol to oil molar ratio, the effect of co-solvent addition, catalyst
ing out of fouling-sensitive reactions was considered as one major appli- type and concentration. Additionally, the geometrical configuration of
cation of the flow splitting unit. Kashid et al. (2010) developed a micro-reactors as well as the flow regime have significant effect on
capillary micro-structured reactor, numbered-up for six capillaries to the biodiesel yield.
investigate the effect of interfacial tension on the mass transfer perfor- • Production of biodiesel in micro-reactors has been investigated for a
mance for various operating conditions. To reduce the equipment cost wide range of temperatures. Most of the reported work has been
and size, they integrated internal numbering-up approach for distribut- performed below the boiling temperature of methanol (60–65 °C).
ing single phase fluids and external numbering-up for two-phase For alkali catalyzed transesterification the biodiesel yield increases
contacting. Yue et al. (2010) developed an external numbering-up ge- with increase in temperature up to an optimum temperature beyond
ometry of a parallel micro-channel contactor with two constructal dis- which the yield decreases. This can be ascribed to accelerated
tributors to study the flow distribution and mass transfer saponification reaction of triglycerides in the presence of an alkaline
characteristics during CO2-water flow. Each distributor comprised of a catalyst at higher temperatures. However, to perform acid catalyzed
dichotomic tree structure that fed 16 micro-channels. It was found transesterification and transesterification under supercritical
that constructal distributors could ensure a nearly uniform distribution condition, temperature above 100 °C and 300 °C were required
of the two phases when the ideal flow pattern was slug-annular flow. respectively.
Hoang et al. (2014) designed a splitting distributor for uniform distribu- • Apart from the temperature it is important to optimize the residence
tion of bubbles over the exit channel. They observed that all bubbles at time for the reaction. The residence time for biodiesel production
the junctions will be broken if the value of capillary number is greater varies with the type of micro-reactor applied. However, at prolonged
than a critical value i.e. Ca N Cacritical. Kriel, Woollam, Gordon, Grant, residence time resulted by a decrease in average velocity for fixed
and Priest (2016) performed experimental numbering-up of microfluidic length micro-channel may cause a decrease in internal circulations
chips from one to five and then ten in an extraction module to extract leading to weakened mass transfer. Hence, a decrease in the yield
Platinum (IV) Chloride using a secondary amine. The extraction perfor- may be noticed at decreased flow velocities. Similarly, under super-
mance was observed to remain unchanged with increasing throughput. critical conditions the biodiesel yield decreases due to decomposition
Al-Rawashdeh et al. (2012) developed barrier-based distributor which as- of the reactants and products at too long residence time.
sured flow uniformity and no channelling between the two phases. With • Alcohol to oil molar ratio is another important factor that affects the
large constrictions in the upstream a flow uniformity of N90% was ob- yield of biodiesel during transesterification. Methanol and ethanol
served. Hence, it is evident from the reported literature that several at- are most investigated for biodiesel production in micro-reactors.
tempts were made to numbering-up the channels using different Under base catalyzed transesterification, most of the researchers
techniques. However, application of these techniques for production of obtained the maximum yield at alcohol to oil molar ratio of 6:1.
biodiesel were not explored. However, the yield varied with the type of micro-reactor used.
Most of the current researches on biodiesel production in micro- When the alcohol to oil molar ratio was increased above the optimum
reactors were performed through transesterification of low free fatty value a decrease in the biodiesel yield decreases. This unexpected
acid (FFA) edible oils using homogenous base catalysts to avoid the decrease in yield is probably because at too high alcohol to oil molar
formation of soaps. A confined work has been done on biodiesel produc- ratio; the alcohol acts as an emulsifying agent which causes a part of
tion using low-cost non-edible feedstock with high FFA content. The glycerol to remain in biodiesel phase. Compared to base catalyzed
conventional process of two-step acid-base transesterification for transesterification a high alcohol to oil molar ratio of up to 20:1 may
biodiesel production using high FFA feedstock is yet to be applied at the be required for acid catalyzed transesterification. Even higher excess
micro-scale. Alternative design and modifications of the microfluidic de- alcohol to oil molar ratio (up to 40:1) is required for transesterification
vices for continuous removal of soap have to be carried out for biodiesel under supercritical conditions.
production using oil with high FFA, which may minimize the cost of • The reaction temperature, residence time and alcohol to oil molar ratio
production. Moreover, to overcome the difficulties due to soap formation depend upon type of catalyst, and catalyst concentration. For biodiesel
development of alternative catalysts with good tolerance to FFA and production in micro-reactors generally homogenous base catalysts are
adaptability at micro-scale needs to be investigated. From the above used. Most of the researchers have used Potassium hydroxide (KOH)
review, it is evident that the current research is focused on increasing or Sodium hydroxide (NaOH). For feedstock with high free fatty con-
the production efficiency through application of micro-reactors. However, tent acid catalyst (H2SO4) is preferred over base catalyst. However,
A. Tiwari et al. / Energy for Sustainable Development 43 (2018) 143–161 159

high reaction temperature, residence time and alcohol to oil molar Azam, N. A. M., Uemura, Y., Kusakabe, K., & Bustam, M. A. (2016). Biodiesel production
from palm oil using micro tube reactors: Effects of catalyst concentration and
ratio are required for acid catalyst. The application of heterogeneous residence time. Procedia Engineering, 148, 354–360.
catalyst makes the catalyst separation obsolete but requires high dura- Benke, J. N. -S. M. (2014). Micro-reactors: A new concept for chemical synthesis and tech-
bility of the catalyst. Heterogeneous catalyst micro-reactors are still nological feasibility. Materials Science and Engineering, 39, 89–101.
Bertoldi, C., da Silva, C., Bernardon, J. P., Corazza, M. L., Filho, L. C., Oliveira, J. V., et al.
under development. (2009). Continuous production of biodiesel from soybean oil in supercritical ethanol
• Catalyst free process using supercritical alcohol can also be used to and carbon dioxide as cosolvent. Energy & Fuels, 23, 5165–5172.
accelerate the transesterification but the high cost of the production Bhoi, R., Sen, N., Singh, K. K., Mahajani, S. M., Shenoy, K. T., Rao, H., et al. (2014).
Transesterification of sunflower oil in micro-reactors. International Journal of
process in the supercritical process due to strict demand of equipment Chemical Reactor Engineering, 12, 47–62.
and high temperature and pressure requirements is a major limitation Billo, R. E., Oliver, C. R., Charoenwat, R., Dennis, B. H., Wilson, P. A., Priest, J. W., et al.
for this method. The conversion efficiency is further improved when (2015). A cellular manufacturing process for a full-scale biodiesel micro-reactor.
Journal of Manufacturing Systems, 37, 409–416.
supercritical transesterification is performed in presence of a co-
Bozbas, K. (2008). Biodiesel as an alternative motor fuel: Production and policies in the
solvent or a heterogeneous catalyst. European Union. Renewable and Sustainable Energy Reviews, 12, 542–552.
• The geometrical configurations of micro-reactor i.e. channel size, inlet Buddoo, S., Siyakatshana, N., & Pongoma, B. (2008). Micro-reactors-a marvel of modern
mixer type and internal geometry of the channels also affect the ester manufacturing technology: Biodiesel case study.
Burns, J. R., & Ramshaw, C. (1999). Development of a micro-reactor for chemical produc-
yield significantly. Shorter dimensions of channel lead to shorter diffu- tion. Chemical Engineering Research and Design, 77, 206–211.
sion distances between the reactant molecules which enhances the Burns, J. R., & Ramshaw, C. (2001). The intensification of rapid reactions in multiphase
mass transfer. Hence, with decrease in the channel size (hydraulic di- systems using slug flow in capillaries. Lab on a Chip, 1, 10–15.
Canakci, M., Erdil, A., & Arcaklioğlu, E. (2006). Performance and exhaust emissions of a
ameter) the yield of product increases. However, small dimension of biodiesel engine. Applied Energy, 83, 594–605.
channels may lead to low throughput therefore apart from diffusion, Canter, N. (2006). Making biodiesel in a micro-reactor. Tribology & Lubrication Technology,
advection can be used to intensify the transesterification reaction 62, 15.
Casas, A., Fernández, C. M., Ramos, M. J., Pérez, Á., & Rodríguez, J. F. (2010). Optimization
through improved mixing of oil and alcohol. of the reaction parameters for fast pseudo single-phase transesterification of sun-
• Micro-reactors in which chaotic advection is induced by modifications flower oil. Fuel, 89, 650–658.
in channel geometry like application of zig-zag, omega, tesla shaped Čerče, T., Peter, S., & Weidner, E. (2005). Biodiesel-transesterification of biological oils
with liquid catalysts: Thermodynamic properties of oil−methanol−amine mixtures.
channels or by inserting an obstacle in the channel tend to provide Industrial and Engineering Chemistry Research, 44, 9535–9541.
higher yields. Moreover, the yield of biodiesel also depends on the Chen, W., Wang, C., Ying, W., Wang, W., Wu, Y., & Zhang, J. (2008). Continuous production
inlet mixer type. It can be attributed to varying dimensions of different of biodiesel via supercritical methanol transesterification in a tubular reactor. Part 1:
Thermophysical and transitive properties of supercritical methanol. Energy & Fuels,
mixers and the flow behavior at the outlet of the mixers. The nature of
23, 526–532.
flow between the two phases determines the surface area to volume Crawford, Elizabeth, Duquette, Douglas, Grant, Daniel, Gray, Richard, Musselman, Brian,
ratio. The droplet-based flow performs better than other methods of Peacock, Martin, et al. (2008). Continuous biodiesel production with continuous
continuous-flow production. liquid-liquid extraction and online MS analysis.
Dai, J. -Y., Li, D. -Y., Zhao, Y. -C., & Xiu, Z. -L. (2014). Statistical optimization for biodiesel
• Optimization of operating parameters and design of micro-reactor is production from soybean oil in a microchannel reactor. Industrial and Engineering
necessary to gain higher productivity. However, for biodiesel produc- Chemistry Research, 53, 9325–9330.
tion in micro-reactors, there are some issues still needed to be tackled Demirbaş, A. (2002). Biodiesel from vegetable oils via transesterification in supercritical
methanol. Energy Conversion and Management, 43, 2349–2356.
with, including the numbering-up of micro-reactor from lab to pilot Demirbaş, A. (2003). Biodiesel fuels from vegetable oils via catalytic and non-catalytic su-
scale, design, and development of micro-reactors for handling high percritical alcohol transesterifications and other methods: A survey. Energy
free fatty acids feedstock and heterogeneous catalyst micro-reactors. Conversion and Management, 44, 2093–2109.
Demirbas, A. (2007). Importance of biodiesel as transportation fuel. Energy Policy, 35,
Furthermore, continuous biodiesel purification using an integrated 4661–4670.
microfluidic device is still under development. More design and sur- Dorado, M. P., Ballesteros, E., López, F. J., & Mittelbach, M. (2004). Optimization of alkali-
face modification of the microfluidic devices have to be carried out catalyzed transesterification of Brassica C arinata oil for biodiesel production. Energy
& Fuels, 18, 77–83.
to meet the quality standards. Dummann, G., Quittmann, U., Gröschel, L., Agar, D. W., Wörz, O., & Morgenschweis, K.
(2003). The capillary-micro-reactor: A new reactor concept for the intensifica-
tion of heat and mass transfer in liquid–liquid reactions. Catalysis Today, 79,
Acknowledgement
433–439.
Dunn, R. (2001). Alternative jet fuels from vegetable oils. Transactions of the American So-
The authors would like to thank Shiv Nadar University for financial ciety of Agricultural Engineers, 44, 1751–1758.
assistance to carry out this work. Elkady, M. F., Zaatout, A., & Balbaa, O. (2015). Production of biodiesel from waste vegeta-
ble oil via KM micromixer. Journal of Chemistry, 2015, 1–9.
Farobie, O., Sasanami, K., & Matsumura, Y. (2015). A novel spiral reactor for biodiesel
References production in supercritical ethanol. Applied Energy, 147, 20–29.
Freedman, B., Butterfield, R. O., & Pryde, E. H. (1986). Transesterification kinetics of soy-
Adamson, D. N., Mustafi, D., Zhang, J. X., Zheng, B., & Ismagilov, R. F. (2006). Production of bean oil 1. Journal of the American Oil Chemists' Society, 63, 1375–1380.
arrays of chemically distinct nanolitre plugs via repeated splitting in microfluidic Freedman, B., Pryde, E., & Mounts, T. (1984). Variables affecting the yields of fatty esters
devices. Lab on a Chip, 6, 1178–1186. from transesterified vegetable oils. Journal of the American Oil Chemists Society, 61,
Aghel, B., Rahimi, M., Sepahvand, A., Alitabar, M., & Ghasempour, H. R. (2014). Using a 1638–1643.
wire coil insert for biodiesel production enhancement in a micro-reactor. Energy Fukuda, H., Kondo, A., & Noda, H. (2001). Biodiesel fuel production by transesterification
Conversion and Management, 84, 541–549. of oils. Journal of Bioscience and Bioengineering, 92, 405–416.
Al-Dhubabian, A. A. (2005). Production of biodiesel from soybean oil in a micro scale reactor. Geyer, K., & Seeberger, P. H. (2008). Micro-reactors as the key to the Chemistry Laboratory
Al-Rawashdeh, M., Fluitsma, L., Nijhuis, T., Rebrov, E., Hessel, V., & Schouten, J. (2012). De- of the Future. VDI Berichte, 2039, 91.
sign criteria for a barrier-based gas–liquid flow distributor for parallel microchannels. Gonzalez, S. L., Sychoski, M. M., Navarro-Díaz, H. J., Callejas, N., Saibene, M., Vieitez, I., et al.
Chemical Engineering Journal, 181, 549–556. (2013). Continuous catalyst-free production of biodiesel through transesterification
Al-Zuhair, S. (2007). Production of biodiesel: Possibilities and challenges. Biofuels, of soybean fried oil in supercritical methanol and ethanol. Energy & Fuels, 27,
Bioproducts and Biorefining, 1, 57–66. 5253–5259.
Angelescu, D., Mercier, B., Siess, D., & Schroeder, R. (2010). Microfluidic capillary separa- Gorji, A. (2015). A review on the biodiesel production, key parameters in transesterification
tion and real-time spectroscopic analysis of specific components from multiphase reaction, its effects on the environment and human health.
mixtures. Analytical Chemistry, 82, 2412–2420. Guan, G., Kusakabe, K., Moriyama, K., & Sakurai, N. (2008). Continuous production of
Aota, A., Mawatari, K., & Kitamori, T. (2009). Parallel multiphase microflows: Fundamental biodiesel using a microtube reactor. Chemical Engineering Transactions, 14,
physics, stabilization methods and applications. Lab on a Chip, 9, 2470–2476. 237–244.
Aran, H., & Lammertink, R. (2013). Porous ceramic and metallic micro-reactors. Nanotechnology Guan, G., Kusakabe, K., Moriyama, K., & Sakurai, N. (2009a). Transesterification of sun-
for water and wastewater treatment (pp. 301). flower oil with methanol in a microtube reactor. Industrial & Engineering Chemistry
Atadashi, I., Aroua, M., Aziz, A. A., & Sulaiman, N. (2013). The effects of catalysts in biodie- Research, 48, 1357–1363.
sel production: A review. Journal of Industrial and Engineering Chemistry, 19, 14–26. Guan, G., Sakurai, N., & Kusakabe, K. (2009b). Synthesis of biodiesel from sunflower oil at
Avellaneda, F., & Salvadó, J. (2011). Continuous transesterification of biodiesel in a helicoi- room temperature in the presence of various cosolvents. Chemical Engineering
dal reactor using recycled oil. Fuel Processing Technology, 92, 83–91. Journal, 146, 302–306.
160 A. Tiwari et al. / Energy for Sustainable Development 43 (2018) 143–161

Guan, G., Teshima, M., Sato, C., Son, S. M., Irfan, M. F., Kusakabe, K., et al. (2010). Two‐ Pardal, A. C., Encinar, J., González, J., & Martínez, G. (2010). Transesterification of rapeseed
phase flow behaviour in microtube reactors during biodiesel production from oil with methanol in the presence of various co-solvents.
waste cooking oil. AIChE Journal, 56, 1383–1390. Perez, V. H., Silveira Junior, E. G., Cubides, D. C., David, G. F., Justo, O. R., MPP, Castro, et al.
Haas, M. J., McAloon, A. J., Yee, W. C., & Foglia, T. A. (2006). A process model to estimate (2014). Trends in biodiesel production: Present status and future directions, 281–302.
biodiesel production costs. Bioresource Technology, 97, 671–678. Qiu, Z., Zhao, L., & Weatherley, L. (2010). Process intensification technologies in continu-
Hartman, R. L., & Jensen, K. F. (2009). Microchemical systems for continuous-flow synthe- ous biodiesel production. Chemical Engineering and Processing: Process Intensification,
sis. Lab on a Chip, 9, 2495–2507. 49, 323–330.
Hoang, D. A., Haringa, C., Portela, L. M., Kreutzer, M. T., Kleijn, C. R., & van Steijn, V. (2014). Rahimi, M., Aghel, B., Alitabar, M., Sepahvand, A., & Ghasempour, H. R. (2014). Optimiza-
Design and characterization of bubble-splitting distributor for scaled-out multiphase tion of biodiesel production from soybean oil in a micro-reactor. Energy Conversion
micro-reactors. Chemical Engineering Journal, 236, 545–554. and Management, 79, 599–605.
Jachuck, R., Pherwani, G., & Gorton, S. M. (2009). Green engineering: Continuous produc- Rahimi, M., Mohammadi, F., Basiri, M., Parsamoghadam, M. A., & Masahi, M. M. (2016).
tion of biodiesel using an alkaline catalyst in an intensified narrow channel reactor. Transesterification of soybean oil in four-way micromixers for biodiesel production
Journal of Environmental Monitoring: JEM, 11, 642–647. using a cosolvent. Journal of the Taiwan Institute of Chemical Engineers, 64, 203–210.
Jahnisch, K., Hessel, V., Lowe, H., & Baerns, M. (2004). Chemistry in microstructured reac- Rashid, W. N. W. A., Uemura, Y., Kusakabe, K., Osman, N. B., & Abdullah, B. (2014). Synthe-
tors. Angewandte Chemie, 43, 406–446. sis of biodiesel from palm oil in capillary millichannel reactor: Effect of temperature,
Jovanovic, G. (2006). http://oregonstate.edu/ua/ncs/archives/2006/feb/tiny-micro- methanol to oil molar ratio, and KOH concentration on FAME yield. Procedia
reactor-biodiesel-production-could-aid-farmers-nation. Chemistry, 9, 165–171.
Kaewchada, A., Pungchaicharn, S., & Jaree, A. (2016). Transesterification of palm oil in a Rathore, V., & Madras, G. (2007). Synthesis of biodiesel from edible and non-edible oils in
microtube reactor. Canadian Journal of Chemical Engineering, 94, 859–864. supercritical alcohols and enzymatic synthesis in supercritical carbon dioxide. Fuel,
Kalu, E. E., Chen, K. S., & Gedris, T. (2011). Continuous-flow biodiesel production using 86, 2650–2659.
slit-channel reactors. Bioresource Technology, 102, 4456–4461. Sahoo, H. R., Kralj, J. G., & Jensen, K. F. (2007). Multistep continuous-flow microchemical
Kashid, M., Gerlach, I., Goetz, S., Franzke, J., Acker, J., Platte, F., et al. (2005). Internal circu- synthesis involving multiple reactions and separations. Angewandte Chemie, 46,
lation within the liquid slugs of a liquid−liquid slug-flow capillary micro-reactor. 5704–5708.
Industrial & Engineering Chemistry Research, 44, 5003–5010. Saka, S., & Kusdiana, D. (2001). Biodiesel fuel from rapeseed oil as prepared in supercrit-
Kashid, M. N., Gupta, A., Renken, A., & Kiwi-Minsker, L. (2010). Numbering-up and mass ical methanol. Fuel, 80, 225–231.
transfer studies of liquid–liquid two-phase microstructured reactors. Chemical Santacesaria, E., Di Serio, M., Tesser, R., Tortorelli, M., Turco, R., & Russo, V. (2011). A sim-
Engineering Journal, 158, 233–240. ple device to test biodiesel process intensification. Chemical Engineering and
Kashid, M. N., & Kiwi-Minsker, L. (2009). Microstructured reactors for multiphase reac- Processing: Process Intensification, 50, 1085–1094.
tions: State of the art. Industrial and Engineering Chemistry Research, 48, 6465–6485. Santacesaria, E., Di Serio, M., Tesser, R., Turco, R., Tortorelli, M., & Russo, V. (2012). Biodie-
Khan, S. A., Günther, A., Schmidt, M. A., & Jensen, K. F. (2004). Microfluidic synthesis of sel process intensification in a very simple microchannel device. Chemical Engineering
colloidal silica. Langmuir, 20, 8604–8611. and Processing: Process Intensification, 52, 47–54.
Knothe, G., Sharp, C. A., & Ryan, T. W. (2006). Exhaust emissions of biodiesel, petrodiesel, Santacesaria, E., Vicente, G. M., Di Serio, M., & Tesser, R. (2012). Main technologies in
neat methyl esters, and alkanes in a new technology engine. Energy & Fuels, 20, biodiesel production: State of the art and future challenges. Catalysis Today, 195,
403–408. 2–13.
Kobayashi, J., Mori, Y., & Kobayashi, S. (2006). Multiphase organic synthesis in Santana, H. S., Tortola, D. S., Reis, É. M., Silva, J. L., & Taranto, O. P. (2016). Transesterification
microchannel reactors. Chemistry – An Asian Journal, 1, 22–35. reaction of sunflower oil and ethanol for biodiesel synthesis in microchannel reactor:
Koonin, S. E. (2006). Getting serious about biofuels. Science, 311, 435. Experimental and simulation studies. Chemical Engineering Journal, 302, 752–762.
Kralj, J. G., Sahoo, H. R., & Jensen, K. F. (2007). Integrated continuous microfluidic liquid– Santana, H. S., Tortola, D. S., Silva, J. L., & Taranto, O. P. (2017). Biodiesel synthesis in
liquid extraction. Lab on a Chip, 7, 256–263. micromixer with static elements. Energy Conversion and Management, 141, 28–39.
Kriel, F. H., Woollam, S., Gordon, R. J., Grant, R. A., & Priest, C. (2016). Numbering-up Y–Y Scheiff, F., Holbach, A., & Agar, D. W. (2013). Slug flow of ionic liquids in capillary
microfluidic chips for higher-throughput solvent extraction of platinum(IV) chloride. microcontactors: Fluid dynamic intensification for solvent extraction. Chemical
Microfluidics and Nanofluidics, 20, 138. Engineering and Technology, 36, 975–984.
Kurayama, F., Yoshikawa, T., Furusawa, T., Bahadur, N. M., Handa, H., Sato, M., et al. Schenk, R., Hessel, V., Hofmann, C., Kiss, J., Löwe, H., & Ziogas, A. (2004). Numbering-up of
(2013). Microcapsule with a heterogeneous catalyst for the methanolysis of rapeseed micro devices: A first liquid-flow splitting unit. Chemical Engineering Journal, 101,
oil. Bioresource Technology, 135, 652–658. 421–429.
Kurayama, F., Yoshikawa, T., Yamada, H., Furusawa, T., Sato, M., & Suzuki, N. (2010). Schenk, R., Hessel, V., Hofmann, C., Löwe, H., & Schönfeld, F. (2003). Novel liquid-flow
A new approach for biodiesel production using CaO-loaded microcapsules as a splitting unit specifically made for numbering-up of liquid/liquid chemical
solid base catalyst. Journal of the Society of Powder Technology, Japan, 47, microprocessing. Chemical Engineering and Technology, 26, 1271–1280.
594–599. Schürer, J., Bersch, D., Schlicker, S., Thiele, R., Wiborg, O., Ziogas, A., et al. (2016). Operation
Kusdiana, D., & Saka, S. (2001). Methyl esterification of free fatty acids of rapeseed oil as of a small-scale demonstration plant for biodiesel synthesis under supercritical con-
treated in supercritical methanol. Journal of Chemical Engineering of Japan, 34, ditions. Chemical Engineering and Technology, 39, 2151–2163.
383–387. Schürer, J., Thiele, R., Wiborg, O., Ziogas, A., & Kolb, G. (2014). Synthesis of biodiesel in mi-
Kusdiana, D., & Saka, S. (2004). Effects of water on biodiesel fuel production by supercrit- crostructures reactors under supercritical reaction conditions. Chemical Engineering
ical methanol treatment. Bioresource Technology, 91, 289–295. Transactions, 37, 541–546.
Leung, D., & Guo, Y. (2006). Transesterification of neat and used frying oil: Optimization Shaaban, W., El-Shazly, A., Elkady, M., & Ohshima, M. (2015). Investigation of factors af-
for biodiesel production. Fuel Processing Technology, 87, 883–890. fect biodiesel production in micro-reactor with T-mixer 88. (pp. 11–15), 11–15.
López-Guajardo, E., Ortiz-Nadal, E., Montesinos-Castellanos, A., & Nigam, K. D. (2017). Shah, S., Sharma, S., & Gupta, M. (2004). Biodiesel preparation by lipase-catalyzed
Process Intensification of Biodiesel Production Using a Tubular Micro-Reactor transesterification of Jatropha oil. Energy & Fuels, 18, 154–159.
(TMR): Experimental and Numerical Assessment. Chemical Engineering Shahid, E. M., & Jamal, Y. (2011). Production of biodiesel: A technical review. Renewable
Communications, 204, 467–475. and Sustainable Energy Reviews, 15, 4732–4745.
Lotero, E., Liu, Y., Lopez, D. E., Suwannakarn, K., Bruce, D. A., & Goodwin, J. G. (2005). Syn- Sharma, Y. C., Singh, B., & Upadhyay, S. N. (2008). Advancements in development and
thesis of biodiesel via acid catalysis. Industrial and Engineering Chemistry Research, 44, characterization of biodiesel: A review. Fuel, 87, 2355–2373.
5353–5363. Silva, C. d., Castilhos, F. d., Oliveira, J. V., & Filho, L. C. (2010). Continuous production of
Machsun, A. L., Gozan, M., Nasikin, M., Setyahadi, S., & Yoo, Y. J. (2011). Membrane micro- soybean biodiesel with compressed ethanol in a microtube reactor. Fuel Processing
reactor in biocatalytic transesterification of triolein for biodiesel production. Technology, 91, 1274–1281.
Biotechnology and Bioprocess Engineering, 15, 911–916. Silva, C. d, Vieitez, I., Jachmanián, I., Castilhos, F. d., Filho, L. C., & Oliveira, J. V. d. (2012).
Maddikeri, G. L., Pandit, A. B., & Gogate, P. R. (2012). Intensification approaches for biodie- Non-catalytic production of ethyl esters using supercritical ethanol in continuous mode.
sel synthesis from waste cooking oil: A review. Industrial and Engineering Chemistry Biodiesel-Feedstocks, Production and Applications: InTech.
Research, 51, 14610–14628. Silva, C., Colonelli, T. A. S., Silva, E. A., Cabral, V. F., Oliveira, J. V., & Cardozo-Filho, L. (2014).
Malsch, D., Kielpinski, M., Merthan, R., Albert, J., Mayer, G., Köhler, J., et al. (2008). μPIV- Continuous catalyst-free production of esters from Jatropha curcas L. oil under super-
analysis of Taylor flow in micro channels. Chemical Engineering Journal, 135, critical ethanol. Brazilian Journal of Chemical Engineering, 31, 727–735.
S166–S172. Silva, Cd, & Oliveira, J. V. (2014). Biodiesel production through non-catalytic supercritical
Martínez Arias, E. L., Fazzio Martins, P., Jardini Munhoz, A. L., Gutierrez-Rivera, L., & transesterification: Current state and perspectives. Brazilian Journal of Chemical Engi-
Maciel, Filho R. (2012). Continuous synthesis and in situ monitoring of biodiesel pro- neering, 31, 271–285.
duction in different microfluidic devices. Industrial and Engineering Chemistry Singh, K. K., Renjith, A. U., & Shenoy, K. T. (2015). Liquid–liquid extraction in
Research, 51, 10755–10767. microchannels and conventional stage-wise extractors: A comparative study.
Meher, L., Vidyasagar, D., & Naik, S. (2006). Technical aspects of biodiesel production by Chemical Engineering and Processing: Process Intensification, 98, 95–105.
transesterification—A review. Renewable and Sustainable Energy Reviews, 10, Slinn, M., & Kendall, K. (2009). Developing the reaction kinetics for a biodiesel reactor.
248–268. Bioresource Technology, 100, 2324–2327.
Mohammed-Dabo, I., Ahmad, M., Hamza, A., Muazu, K., & Aliyu, A. (2012). Cosolvent Sootchiewcharn, N., Attanatho, L., & Reubroycharoen, P. (2015). Biodiesel production
transesterification of Jatropha curcas seed oil. Journal of Petroleum Technology and from refined palm oil using supercritical ethyl acetate in a micro-reactor. Energy
Alternative Fuels, 3, 42–51. Procedia, 79, 697–703.
Naleini, N., Rahimi, M., & Heydari, R. (2015). Oleuropein extraction using microfluidic sys- Sun, J., Ju, J., Ji, L., Zhang, L., & Xu, N. (2008). Synthesis of biodiesel in capillary micro-
tem. Chemical Engineering and Processing: Process Intensification, 92, 1–6. reactors. Industrial & Engineering Chemistry Research, 47, 1398–1403.
Nguyen, N. -T., & Wu, Z. (2005). Micromixers—A review. Journal of Micromechanics and Sun, P. Y. (2010). Preparation and purification of biodiesel in microstructured reactors. (PhD
Microengineering, 15, R1–R16. dissertation) Nanjing University of Technology.
A. Tiwari et al. / Energy for Sustainable Development 43 (2018) 143–161 161

Sun, P., Sun, J., Yao, J., Zhang, L., & Xu, N. (2010). Continuous production of biodiesel from Xie, W., Peng, H., & Chen, L. (2006). Transesterification of soybean oil catalyzed by potas-
high acid value oils in microstructured reactor by acid-catalyzed reactions. Chemical sium loaded on alumina as a solid-base catalyst. Applied Catalysis A: General, 300,
Engineering Journal, 162, 364–370. 67–74.
Sun, P., Wang, B., Yao, J., Zhang, L., & Xu, N. (2009). Fast synthesis of biodiesel at high Xue, J., Grift, T. E., & Hansen, A. C. (2011). Effect of biodiesel on engine performances and
throughput in microstructured reactors. Industrial and Engineering Chemistry emissions. Renewable and Sustainable Energy Reviews, 15, 1098–1116.
Research, 49, 1259–1264. Yeh, S. I., Huang, Y. C., Cheng, C. H., Cheng, C. M., & Yang, J. T. (2016). Development of a
Tice, J. D., Song, H., Lyon, A. D., & Ismagilov, R. F. (2003). Formation of droplets and mixing millimetrically scaled biodiesel transesterification device that relies on droplet-
in multiphase microfluidics at low values of the Reynolds and the capillary numbers. based co-axial fluidics. Scientific Reports, 6, 29288.
Langmuir, 19, 9127–9133. Yue, J., Boichot, R., Luo, L., Gonthier, Y., Chen, G., & Yuan, Q. (2010). Flow distribution and
Trentin, C. M., Lima, A. P., Alkimim, I. P., da Silva, C., de Castilhos, F., Mazutti, M. A., et al. (2011). mass transfer in a parallel microchannel contactor integrated with constructal dis-
Continuous production of soybean biodiesel with compressed ethanol in a microtube re- tributors. AIChE Journal, 56, 298–317 (NA-NA).
actor using carbon dioxide as co-solvent. Fuel Processing Technology, 92, 952–958. Yue, J., Chen, G., Yuan, Q., Luo, L., & Gonthier, Y. (2007). Hydrodynamics and mass transfer
Tsaoulidis, D. A. (2015). Studies of intensified small-scale processes for liquid-liquid separa- characteristics in gas–liquid flow through a rectangular microchannel. Chemical
tions in spent nuclear fuel reprocessing. Engineering Science, 62, 2096–2108.
Van Gerpen, J. (2005a). The basics of diesel engines and diesel fuels. Urbana, IL: AOCS Press. Zhang, L., Jin, Q., Zhang, K., Huang, J., & Wang, X. (2012). The optimization of conversion of
Van Gerpen, J. (2005b). Biodiesel processing and production. Fuel Processing Technology, waste edible oil to fatty acid methyl esters in homogeneous media. Energy Sources,
86, 1097–1107. Part A: Recovery, Utilization, and Environmental Effects, 34, 711–719.
Vicente, G., Martınez, M., & Aracil, J. (2004). Integrated biodiesel production: A compari- Zhang, Y., Dube, M., McLean, D., & Kates, M. (2003). Biodiesel production from waste
son of different homogeneous catalysts systems. Bioresource Technology, 92, 297–305. cooking oil: 1. Process design and technological assessment. Bioresource Technology,
Wang, Y., Ou, S., Liu, P., & Zhang, Z. (2007). Preparation of biodiesel from waste cooking oil 89, 1–16.
via two-step catalyzed process. Energy Conversion and Management, 48, 184–188. Zhou, H., Lu, H., & Liang, B. (2006). Solubility of multicomponent systems in the biodiesel
Warabi, Y., Kusdiana, D., & Saka, S. (2004). Reactivity of triglycerides and fatty acids of production by transesterification of Jatropha curcas L. oil with methanol. Journal of
rapeseed oil in supercritical alcohols. Bioresource Technology, 91, 283–287. Chemical & Engineering Data, 51, 1130–1135.
Wen, Z., Yu, X., Tu, S. T., Yan, J., & Dahlquist, E. (2009). Intensification of biodiesel synthesis
using zigzag micro-channel reactors. Bioresource Technology, 100, 3054–3060.
Xie, T., Zhang, L., & Xu, N. (2012). Biodiesel synthesis in micro-reactors. Green Processing
and Synthesis, 1, 61–70.

You might also like