Next Article in Journal
The Bistable Behaviour of Pseudomonas putida KT2440 during PHA Depolymerization under Carbon Limitation
Previous Article in Journal
Three-Dimensional Culture Model of Skeletal Muscle Tissue with Atrophy Induced by Dexamethasone
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of Microwave-Assisted Reactions for Biodiesel Production

Institute of Sustainable Energy, Universiti Tenaga Nasional, 43000 Kajang, Selangor, Malaysia
*
Author to whom correspondence should be addressed.
Bioengineering 2017, 4(2), 57; https://doi.org/10.3390/bioengineering4020057
Submission received: 18 May 2017 / Revised: 9 June 2017 / Accepted: 12 June 2017 / Published: 15 June 2017

Abstract

:
The conversion of biomass into chemicals and biofuels is an active research area as trends move to replace fossil fuels with renewable resources due to society’s increased concern towards sustainability. In this context, microwave processing has emerged as a tool in organic synthesis and plays an important role in developing a more sustainable world. Integration of processing methods with microwave irradiation has resulted in a great reduction in the time required for many processes, while the reaction efficiencies have been increased markedly. Microwave processing produces a higher yield with a cleaner profile in comparison to other methods. The microwave processing is reported to be a better heating method than the conventional methods due to its unique thermal and non-thermal effects. This paper provides an insight into the theoretical aspects of microwave irradiation practices and highlights the importance of microwave processing. The potential of the microwave technology to accomplish superior outcomes over the conventional methods in biodiesel production is presented. A green process for biodiesel production using a non-catalytic method is still new and very costly because of the supercritical condition requirement. Hence, non-catalytic biodiesel conversion under ambient pressure using microwave technology must be developed, as the energy utilization for microwave-based biodiesel synthesis is reported to be lower and cost-effective.

1. Introduction

The microwave processing technology has revolutionized organic synthesis, and this technique has gained huge interest as a tool for synthesis and process development. In the last 30 years we have witnessed climate change—temperatures rising, weather becoming more erratic, and glaciers continuing to melt due to a higher level of greenhouse gas (GHG) emissions. Many integrated assessments have shown that increasing anthropogenic GHG emissions are largely caused by the use of fossil fuels, and hence there is a need for a range of renewable technology options to meet stringent global warming targets (e.g., keeping CO2 concentration below 440 ppm by 2050) [1]. The major component of GHGs is carbon dioxide, and there is a global concern about reducing carbon emissions. For thousands of years, the global carbon supply was essentially stable, as natural processes removed as much carbon as they released. Modern human activity—burning of fossil fuels, deforestation, and intensive agriculture—has led to huge quantities of carbon dioxide and other GHG emissions [2]. In order to rectify these problems, the dependency on fossil fuels should be scaled down. Switching to renewable energy will play a very significant role in combating climate change. Microwave technology can be used as an enabling tool to enhance reaction yield and selectivity for renewable energy processes. The need for sustainable development and green processing technologies will enhance interdisciplinary efforts and put microwave-based processes in the forefront.
Renewable energy is energy generated from natural resources—such as sunlight, wind, rain, tides, and geothermal heat—that are renewable (naturally replenished). Renewable energy technologies range from solar power, wind power, hydroelectricity/micro-hydro, biomass, and biofuels. This energy cannot be exhausted and is constantly renewed. In recent years, there has been a great interest in stimulating the use of bioenergy and biomaterials in order to reduce the GHG emissions associated with using fossil fuels. On the other hand, the biggest drawback of renewable energy is its relatively high cost compared to the conventional fossil-fuel energy sources. Additionally, with the current renewable energy technology, the amount of renewable electricity generated will not be as large as that produced by the conventional fossil fuels. Furthermore, the supply of renewable energy sources is not reliable, as most of them are based on the weather, which is unpredictable. Since 2011, the cost of renewable energy generation and energy storage equipment has fallen massively. Large reductions have occurred in the cost of lithium-ion batteries and related systems for energy storage. Cost reductions also come from economies of scale due to a large increase in production by China and other countries. Most importantly, advances in technology will keep prices falling after scale economies have been exhausted [3,4,5]. Any country with rich renewable energy resources and technological advancements will become a low-cost location for energy supplies in a low-carbon world economy. It will also make it an economically superior location for energy-intensive processing and manufacturing activity.
Recently, the rapid increase in chemical reaction rates by using microwave irradiation has attracted the attention of almost all fields of researchers, especially those in the chemistry area. Early researchers [6,7] reported that the microwave-assisted reaction rates increased, by a factor of 5, to 1000, when compared to the conventional reactions. The whole scientific and industrial community is interested with the application of microwave irradiation, not only due to its enhancement in the reaction rates, but also for its ability to revolutionize the chemical reactions to be performed with unexplainable results. In addition to these, there are several advantages of microwave processing [8,9,10], such as rapid heating, lower relative energy consumption, environmental friendliness, higher production yield, controllable processing, shorter processing time, and quality and properties improvement.
In this paper, some basic information on microwaves, such as their heating mechanisms and characteristics, are discussed in detail. Additionally, an overview of thermal and non-thermal effects of microwave heating is also presented. Furthermore, the application of microwave energy in various fields is also mentioned, particularly in biodiesel production, as biodiesel is a renewable fuel in liquid state, which provides advantages over the other renewable energies as most of the industrial areas and transportation use liquid fuel to operate. The purpose of this review paper is to provide a basic understanding regarding microwave energy and microwave-assisted biodiesel production for novice researchers in the biodiesel area.

2. Biodiesel Production

Biodiesel, or fatty acid alkyl ester (FFAE), is made up of methyl or ethyl esters produced from vegetable oil or animal fat. It is well-known to be an alternative sustainable fuel source to decrease the depletion rate of fossil fuels, due to its similar fuel properties to the diesel fuel. Biodiesel is a non-toxic and biodegradable fuel, which produces less harmful gases and thus reduces pollution to the environment.
Numerous kinds of feedstocks have been introduced to prepare biodiesel; for instance, refined oils [11], recycled waste cooking oils [12], and microalgae [13]. Commonly, these different types of feedstocks can be classified into three generations. The first-generation feedstock is the edible oils (vegetable oils), including soybean, palm tree, corn, coconut, rapeseed, sunflower oils, and others. Among these, the dominant feedstock is the soybean oil, while the palm tree is well-known due to its capability to produce the highest quantity of oils per cultivated area [14]. Nevertheless, vegetable oil conflict arose along with an upsurge of vegetable oil prices as its demand increased. Thus, the second-generation feedstocks that do not conflict with food interest have been presented: the oils of non-edible crops. Oils harvested from rubber seed, castor, jatropha, maize, and animal fat belong to this type. Some of these non-edible crops have been used in the production of commercial products such as cosmetics, and thus conflict arose again. Microalgae and other biodiesel feedstocks that do not conflict with the interests related to human consumption are considered to be third-generation feedstocks. Algae are predominant under this category due to their superior properties, and they bring numerous benefits. Algae have a higher oil content in comparison with other feedstocks such as sugar cane and corn. Similarly to a plant, algae consume carbon dioxide and release oxygen into the environment during their photosynthesis process. Thus they can be used to capture the carbon dioxide from the industrial flue gasses in order to reduce the air pollution. In addition to this, they have a strong adaptability to their living environment, even in polluted water, due to their unicellular or multicellular structure [15]. Furthermore, the fuels produced from algae are carbon-neutral. Different biodiesel feedstocks [16] are outlined in Table 1.
Biodiesel is environmentally friendly and thus it holds great promise in solving the environment-related problems, such as air pollution, and the greenhouse effect. Nevertheless, biodiesel production encounters one major challenge, which is the production cost in terms of its feedstock and oil-to-biodiesel conversion process. In order to reduce the cost of feedstocks, low-cost feedstocks such as palm oil and waste cooking oil have been used to produce biodiesel. The conversion process costs can also be reduced by conducting optimization and process enhancement.
Transesterification is a process in which the alcohol, usually methanol or ethanol, is used to break the triglycerides (oils or fats) into methyl or ethyl esters, the renewable biodiesels. This process can be conducted in either the presence or absence of an acid/alkali-based catalyst. Transesterification comprises three reversible reactions, Equations (1)–(3), that form diglyceride, monoglyceride, and glycerine consequently, and at the same time produce one ester molecule for each alcohol molecule consumed in each reaction step. Thus, the overall reaction, as shown in Equation (4), involves 1 mole of triglyceride and 3 moles of alcohol to form 3 moles of alkyl esters and 1 mole of glycerol as a by-product. Practically, however, an excess quantity of alcohol is added to complete the conversion process and to increase the alkyl ester (biodiesel) yields by shifting the reaction equilibrium towards the production of esters. Additionally, the excess alcohol allows the phase separation of biodiesel from the glycerol.
T r i g l y c e r i d e   ( o i l s   o r   f a t )   +   R O H     D i g l y c e r i d e   +   R C O O R
D i g l y c e r i d e   +   R O H     M o n o g l y c e r i d e   +   R C O O R
M o n o g l y c e r i d e   +   R O H     G l y c e r o l   +   R C O O R
T r i g l y c e r i d e   ( o i l s   o r   f a t )   +   3 R O H     G l y c e r o l   +   3 R C O O R
Generally, there are three types of heating methods to perform the transesterification process, which are conventional heating, supercritical heating, and microwave heating. Table 2 [14] compares and summarizes their characteristics and their respective pros and cons. Their details and mechanisms, however, will be discussed further in the later sections.

2.1. Catalytic Transesterification

As described above, transesterification can be performed with or without a catalyst. There are two major types of catalysts: the homogeneous type and the heterogeneous type. Moreover, they can be further categorized (Figure 1) into three groups: (a) acids, (b) bases, and (c) lipase enzymes. Catalysts are used to shorten the reaction time and to produce biodiesel of better quality. These catalysts are employed depending on the type and quality of the feedstock. The alkaline catalysts are usually preferable as they require a lower reaction temperature and provide a higher reaction rate and conversion efficiency when compared with the acid catalysts [17]. On the other hand, acid catalysts are insensitive to FFAs, compared to base catalysts. FFAs of the feedstock reduce the transesterification rate by reacting with the catalyst and forming unwanted soap instead of biodiesel. Hence, acid catalysts are superior for transesterification that involves vegetable oils with a higher percentage (more than 1%) of FFAs [17]. In addition, lipase enzymes are an alternative catalyst applied to accelerate the transesterification reaction. Lipase enzymes have outstanding advantages over the acid/alkali-based catalysts, as they reduce the complication of downstream processes such as product separation, purification, washing, and neutralization. They are also suitable for biodiesel feedstock with a high FFA content. Nevertheless, there are disadvantages to the lipase enzyme, such as that they are more expensive and require a longer reaction time [18].
Homogeneous-base catalysts such as sodium hydroxide (NaOH), potassium hydroxide (KOH), and sodium methoxide (CH3ONa) are the most common catalysts used for biodiesel production. Among these, CH3ONa is the most effective, as less is needed, in comparison to NaOH and KOH, in order to obtain the same catalytic effect. In addition, NaOH and KOH catalysts also form water, which leads to soap formation and slows down the reaction rate. However, CH3ONa is more expensive. Homogeneous-base catalytic transesterification involves the catalysts that can be dissolved in the alcohol. At the end of this reaction, these homogeneous-base catalysts will not be recovered for the reuse purpose. Additionally, problems such as a complicated separation downstream process and unwanted soap formation have encouraged researchers to discover the application of heterogenous catalysts in biodiesel production. Other than solving the problems mentioned above, heterogeneous catalysts have a higher adaptability to the cheaper feedstocks. Moreover, their downstream process is less complicated and the catalysts can be reused, hence resulting in a lower overall production cost.
All of the methods mentioned above are known as catalytic transesterifications, as catalysts or enzymes are applied to enhance the reaction rate. These catalytic processes are commonly assisted with conventional heating. The pros and cons of different types of catalysts used in transesterification for biodiesel production are summarized in Table 3 [19,20].

2.2. Non-Catalytic Transesterification with Supercritical Condition

In catalytic transesterification, the catalysts are not consumed, and so a complicated purification process is essential to separate them in order to collect the biodiesel. Additionally, this reaction is relatively time-consuming due to the phase separation of the oil-alcohol mixture. In order to avoid these problems, supercritical transesterification is introduced [21,22]. In the supercritical method, the temperature and pressure are maintained above the critical values of alcohol (e.g., 239 °C and 8.1 MPa, respectively) for methanol. This decreases the dielectric constant and solubility parameter of the alcohol, and thus the alcohol becomes more soluble with the oils. As a result, the transesterification reaction time is shortened significantly. Moreover, the complicated purification process is eliminated, as no catalysts are required for the supercritical method. Thus, the oil-biodiesel conversion using this method is also known as non-catalytic transesterification. Furthermore, the supercritical method has a higher adaptability with the low-cost feedstock (higher water content and FFAs) in comparison with the conventional catalytic transesterification [23]. The supercritical method is expensive, as a higher energy input is compulsory to operate at the critical temperature and pressure.

2.3. Microwave-Assisted Methods

Microwave-assisted methods have been developed as a promising way—that is rapid, energy-efficient, cost-saving, and environmentally friendly—to produce biodiesel. Numerous tests and optimization studies with different feedstocks have been conducted by researchers [24,25,26,27,28] in the recent years to maximize the advantages of microwaves for biodiesel production. Other than for transesterification reactions, microwave energy is also applied to assist the oil-extraction process in order to produce a greener and better quality of biodiesel [29].

2.3.1. Theory of Microwaves

Microwave dielectric heating was introduced as a new heating technique in the 1970s, and eventually became a powerful method to enhance chemical reactions in almost all areas of chemistry. Gedye et al. [6], the very first group of researchers that related microwave to organic chemistry, observed that the organic reactions could be carried out at a faster rate under microwave irradiation. Since this experiment, the application of microwave irradiation has been reviewed, and they observed that the microwave heating increased the reaction rate, and the formation of side products were reduced compared to the conventional heating method [30]. These acceleration, reactivity and selectivity changes observed during microwave-assisted reactions enabled energy saving and process intensification, and could be explained through the combination of thermal and non-thermal effects of microwaves. These effects will be explained in detail in the following section.
Electromagnetic waves are formed by the coupling of electric and magnetic fields in such a way that they are perpendicular to each other. Microwaves are electromagnetic waves with a frequency range between 0.3 and 300 GHz, with wavelengths ranging from 0.01 m to 1 m. Microwaves are located in-between infrared waves and radio waves on the electromagnetic spectrum. Electromagnetic waves with a higher frequency are associated with higher energy and shorter wavelengths. Figure 2 shows the electromagnetic wave spectrum.
Domestic applications such as microwave ovens are approved to be operated at a frequency of 2.45 GHz, while the industrial applications are between 0.915 GHz and 2.45 GHz. Most of the reported microwave chemistry experiments are carried out at 2450 MHz [32,33,34], as the maximum microwave energy is absorbed by liquid water near to this frequency. There are numerous studies that have been carried out on microwave applications in the organic synthesis sector; for instance, the application of microwaves in the synthesis of polymer [35] and metal oxides [7], solvent free reactions [36], homogeneous and heterogeneous catalytic reactions [37], and many other reactions involved in medicinal [38] and green chemistry [8].
The quantum energy of microwaves is given by Planck’s law:
E   =   h v
Within the frequency range of microwaves (0.3–300 GHz), respective energies associated are between 1.24 × 10−6 and 1.24 × 10−3 eV. These microwave energies are lower than the energy associated with Brownian motion. In other words, microwave energy is not sufficient to break chemical bonds and thus cannot induce chemical reactions. Table 4 gives the values of some bond energies and those of microwave radiation.
In conventional heating, heat is transferred through thermal conduction and convection, through the wall of the targeted object, to reach its internal materials. Thus, this heating method is also known as wall heating. In this case, the sample results in a non-uniform temperature and higher thermal gradients, as the heat is transferred through the other mediums before it reaches the internal sample. The heating effect is heterogeneous and it is also highly dependent on the materials’ properties, such as thermal conductivity, specific heat capacity, and density. Moreover, a large portion of energy [14] from the heat sources will be lost to the environment through conduction and convection. Hence, this is a relatively slow and unproductive way to transfer energy into the reacting system.
In contrast, microwave heating utilizes the ability of certain materials, usually solids or liquid, to transform electromagnetic energy into heat energy. The two components of microwaves (electric field and magnetic field) will interact differently with the material under different mechanisms. The electric and magnetic field components are responsible for the microwave dielectric heating and magnetic loss heating, respectively. Among these two heating effects, most of the studies reported the microwave dielectric heating effect [39], however, according to Cheng et al. [40,41], there are superior benefits of microwave magnetic-field heating for certain magnetic dielectric and conductive powder materials. Peng et al. [42] presented that the magnetic loss in the microwave heating of ferrites is up to approximately 4 times greater than the dielectric loss.
In the organic chemistry field, almost all of the microwave heating reported refers to dielectric heating. Essentially, there are two primary dielectric loss mechanisms for microwave dielectric heating: (a) dipolar polarization, and (b) ionic conduction, which lead to an instantaneous localized and rapid superheating of the reaction materials. Here, the molecules of the whole reaction mixture are coupled with the microwaves, resulting in a rapid increase in temperature. The heating of the reaction vessel is not involved in this process. Hence, microwave dielectric heating is not restricted by the thermal conductivity of the vessel, unlike the conventional heating method. However, this depends on the shape, dielectric properties and size of the reaction materials, and lastly, the nature of the microwave equipment used. For more details about the microwave heating mechanism, some excellent reviews can be found in [39] and [43]. Figure 3 illustrates the basic differences between conventional heating and microwave heating. The characteristics of microwave heating and conventional heating are also summarized in Table 5 to give an overall comparison of their features [44].

2.3.2. Thermal Effect of Microwaves

Several researchers conclude that the acceleration of a reaction is only due to the thermal effect, also known as the kinetic effect. This effect is caused by the rapid increase of temperature during the dipolar polarization and/or ionic conduction under microwave irradiation [45]. According to this group of researchers, it cannot be denied that the reaction time is dramatically reduced by using microwave heating when compared to conventional heating, but the kinetic mechanism of the same reactions remains identical. They explained that the acceleration of the reaction is due to the rapid, sudden and uncontrollable increasing of the temperature under microwave irradiation.
To explain the thermal effect of microwaves, the fundamental thermodynamic formula, which is the Arrhenius equation, is applied, as there is involvement of thermodynamics of reaction materials.
K   =   A × e Δ G R T
where K is the reaction rate, A is a collision frequency factor that indicates the molecular mobility based on the molecular vibration frequency during the reaction, ΔG represents the activation energy, R is the gas constant, and T is the temperature in kelvin. Based on Equation (6), there are two possible methods to enhance the chemical reaction rate, which are by increasing the collision frequency factor, A, and by decreasing the activation energy, ΔG.
Activation energy can also be expressed in term of enthalpy, H, and entropy, S, with this equation:
ΔG = ΔHTΔS
As shown in Equation (7), in order to decrease ΔG, ΔS (entropy) should be increased. This phenomenon can be attained in microwave-assisted reactions, as ΔS is higher due to the involvement of the interactions at the molecular level and also the rapid, random movement of the dipole. Additionally, activation energy can be reduced through the superheating condition achieved under microwave irradiation, as the result of the microwave energy dissipation over the whole reaction volume [46].
Some researchers [47,48] have claimed the existence of hot spots, which are certain particular areas with a higher energy and temperature, in the reaction mixture under microwave irradiation due to the inhomogeneity of the applied electromagnetic field. The non-homogenous distribution of microwave fields is created within the microwave ovens as a consequence of their reflections on the inner walls. Filter paper that is soaked with cobalt chloride solution, however, can be used to determine the location of a high-energy hotspot [43].
A hotspot usually acts as an active site to promote the catalytic activities and thus contribute to the improvement of the conversion rate and efficiency. Durka [47] concluded that for a given methanol conversion, a lower net heat input was applied to the reactor under microwave heating when compared to electric heating, due to the formation of hotspots. Moreover, Ren [49] has demonstrated this hotspot effect on the decomposition of hydrogen sulfide, H2S, over a molybdenum disulphide, MoS2, catalyst under microwave conditions. Generally, the H2S decomposition rate—or strictly speaking, the catalytic activity—is increased along with temperature. However, the results showed that the decomposition rate increased with the microwave power at the beginning of experiment, but that any further increment of the power supply after 100 W led to a decreasing of the decomposition rate due to the formation of hotspots where the temperatures were increased too rapidly until the active H and S may have gone through a new reacting pathway.
Microwave irradiation is also well-known due to its selective mode of heating. Generally, all of the interactions between materials and microwaves can be categorized into three groups, which are absorption, transmittance, and reflection. Naturally, in a reaction mixture under microwave irradiation, the microwave will be absorbed by the polar substances and heat is generated within them. At the same time, the nonpolar substances of the reaction mixture that do not absorb microwaves, however, are not heated. This selective mode of heating has been utilized and studied by numerous researchers in various fields [3,50,51,52], and can be grouped into these three categories: solvents, catalysts, and reagents [53]. Takuya et al. [54] illustrated the selective heating mode of microwave irradiation through their study on the size effect of water droplets in water/oil emulsion with sorbitan fatty acid monostearate surfactant. They observed that the water droplets’ temperature was 20 °C more than the whole reaction solution, and they concluded that this phenomenon occurred due to the selective microwave heating (Figure 4). Moreover, they found that in the water/oil emulsion, water droplets with a different size and concentration result in a different selective microwave heating rate. The application of selective microwave heating in order to enhance the extraction rate and yield was presented by Lee et al. [50]. A step-change of extraction yield was detected at the temperature for which the selective heating of the biomass, in this case the okra, occurred. Also, due to the principle of microwave selective heating, the extraction rate of okra with microwave heating was significantly higher, and at the same time produced a higher yield in comparison with the conventional method (Figure 5). Satoshi and his co-workers [55] emphasized that the dehydrogenation of organic hydride methylcyclohexane was at a faster rate and gave a higher yield as the result of the selective heating of the palladium/activated carbon (Pd/AC) catalyst under microwave irradiation. The Pd/AC catalyst was heated to 340 °C within 2 min, while the conventional heating took 35 min to reach a temperature at which the catalyst started to work efficiently.

2.3.3. Non-Thermal Effect of Microwaves

Some authors [56,57] claimed that the thermal effect alone is not sufficient to explain microwave results such as reaction rate enhancement and thus eventually they postulated the existence of non-thermal microwave effects. Even today, this issue remains a controversial topic among researchers [58,59]. It has been argued that microwave irradiation affects the pre-exponential factor or activation energy in the Arrhenius equation, A, due to the orientation effects of dipolar molecules in the presence of an electric field. In polar reaction mechanisms also, the polarity is increased under microwave irradiation, from the ground state to the transition state. This eventually causes the activation energy to be decreased and results in an enhancement of reactivity. This statement has been supported by the work of Xu et al. [60].
To prove the existence of microwave non-thermal effects, a model has been proposed by Zhou et al. [61] to study the reason for the reduction in the apparent activation energy, Ea′, under microwave irradiation, which results in an accelerating of chemical reactions. This model follows four hypotheses: (1) Matter is ideal and uniform. (2) A portion of microwave energy is absorbed by the molecules, and interacts with the molecular matter directly to reduce the Ea′, illustrating a vital microwave catalytic effect. (3) At the same time, another portion of microwave energy is released on the molecular matter by the thermal effect, causing the increasing of the reaction system’s temperature. (4) A dynamic balance between the absorbed and released microwave “energy molecular” on the molecular matter will be achieved under balanced circumstances.
Eventually, a formula that is used to estimate quantitatively the reduction of Ea′ due to the microwave energy was established successfully:
E a   =   E a 0     E a M W
where Ea0 represents the activation energy barrier of the reaction system without any microwave irradiation, and EaMW is the portion of microwave energy that is transformed to reduce the Ea′ directly.
EaMW can also be expressed as Equation (9), in terms of the microwave electromagnetic field (EMW), microwave energy power, (PMW), the total microwave energy absorbed by the matter (Et) and also a constant (k) that is determined by the microwave frequency and the dielectric properties of the reacting system. The details of the equation’s development can be found in [61].
E a M W   =   k E M W P M W 1   +   k E M W P M W   E t
In addition, there is some other evidence regarding the existence of a microwave non-thermal effect. For instance, Fukushima et al. [62] investigated the behavior of copper(II) oxide reduction under microwave irradiation, and the results suggested that the reduction energy was supplied by both thermal energy and microwave energy (magnetic, H, and electric, E, fields) as indicated in Figure 6. Strictly speaking, microwave energy could transform directly into reduction energy in the energy dissipation process by interacting directly with the valence electrons. Thus, they concluded that the reduction of copper(II) oxide involved thermal and non-thermal process.
Moreover, Mao et al. [63] reported that the synthesis of quaternary ammonium salt with the microwave-assisted process was far more effective due to its lower activation energy (41.44 kcal/mol) in comparison to the conventional thermal heating method (61.21 kcal/mol). This phenomenon was explained as the result of the non-thermal microwave effect. Furthermore, a sign of the non-thermal microwave effect was also observed by Nazari and Ghandi [57] in their study on oxidation of aromatic α-diketones. The activation energy and pre-exponential factor for microwave-assisted reactions were significantly changed when compared to the conventional oil bath method. This consequently resulted in larger rate constants. Since the experiments were conducted under similar conditions, other than in the presence of microwave irradiation, rapid heating (thermal effect) solely was not adequate to explain the outcomes of this study, and thus the existence of non-thermal effects was claimed.
Other than the reaction rate enhancement, there is further evidence of a non-thermal microwave effect, which is the alternation of electric conductivity in an aqueous solution. This effect was clearly detected in a sodium chloride, NaCl, aqueous solution under a microwave field with the improved experimental set-up by Tian and Huang [64]. They provided a direct indication of the presence of non-thermal microwave effects.
Liu et al. [65], however, proved the existence of non-thermal effects from the perspective of molecular dynamics. The results (Figure 7 and Figure 8) showed that the dynamics of the water structure did not alter monotonically with the variation of the microwave field strength, but a threshold effect was observed. Once the microwave field strength was greater than the threshold value, a significant distortion of the water structure was spotted. Moreover, there was an abnormal pattern of the microwave-electrolyte interaction due to the rapid increases of polarizability, and the microwave field changed the system structure.

2.3.4. Recent Study of Microwave-Assisted Biodiesel Production

Microwave-based biodiesel production (see Figure 9) usually starts with the mixing process of oil, alcohol, and a catalyst in the appropriate ratio. Then, the mixture is processed under microwave irradiation. Lastly, separation and purification are carried out to yield biodiesel. In microwave-assisted catalytic transesterification, methanol is mostly used due to its high capability to absorb microwaves, and it is also a high-polarity organic solvent. Hence, since solvents with a high dielectric loss factor are used, the transesterification process can be improved through dipolar polarization and ionic conduction under microwave irradiation, which has a higher dependency on the polarity and absorption capability of the reacting samples. The experimental studies of microwave-assisted transesterification with different feedstocks are summarized in Table 6.
Based on the study of Zahir et al. [74], it was proved that microwave irradiation has superior benefits over the conventional heating method in biodiesel preparation. Under microwave irradiation, the reaction time has reduced from 2 h to 6 min, and at the same time it produces a higher biodiesel conversion yield (4.1% higher than the conventional method). Moreover, the energy consumption for microwave heating (288 kJ) was reported to be lower than for the conventional heating (3150 kJ) by Patil and his co-worker [24] in their experiment on biodiesel production using sulfuric acid. Other than enhancing the reaction rate, producing higher yield, and consuming less energy, microwave-assisted transesterification produces less by-product, which results in a simplified downstream process and eventually reduces the time taken for product separation [14,24].
There are two major biodiesel standards that have been followed by people throughout the world, which are the American Standard Specification for Biodiesel (ASTM 6751), and the European Standard for Biodiesel (EN 14214). Microwave-based biodiesel properties were very well compared with these standards by several researchers, and are recorded in Table 7.
Furthermore, microwave energy can also be applied to enhance the supercritical reaction for the non-catalytic transesterification, especially for the feedstocks that contain water [14]. This is because water molecules consist of a dipole moment that is sensitive to the electric field. Under microwave irradiation, a dipole will try to align itself with the electric field applied by rotation, and as a result, local superheating is formed. Patil P. et al., demonstrated the direct conversion of dry [80] and wet [81] algal biomass (Nannochloropsis salina) into biodiesel with a non-catalytic transesterification method under microwave-mediated supercritical ethanol conditions (see Figure 10 and Figure 11). The results showed that the algal extraction had been improved dramatically. Moreover, this approach also increased the reaction efficiency and production yield, and at the same time, the extractive-transesterification time had been reduced. In addition, this technique resulted in highly purified and harmless products, and thus its separation and purification stages were less complicated. From there, the energy consumption and overall production cost of microwave-assisted reactions may be reduced in comparison to the conventional heating methods.

2.3.5. Future Outlook on Microwave Application in Biodiesel Production

As mentioned previously, transesterification can be performed with or without the presence of catalysts. The catalytic transesterification offers a high biodiesel conversion yield, however, it has several drawbacks, such as a complicated purification process, a low adaptability to cheaper feedstocks, and a large volume of waste water generated. To minimize these problems, non-catalytic transesterification has been proposed. Until now, non-catalytic biodiesel conversion methods have been limited to the use of supercritical conditions (i.e., temperatures above 350 °C and pressures above 15 MPa). In spite of the numerous advantages, the operational and equipment costs for producing biodiesel under supercritical conditions are huge obstacles. In this regard, non-catalytic biodiesel conversion with the conditions at temperatures around 250 °C and under ambient pressure (around 1 MPa) would be an alternative pathway to be explored. The non-catalytic transformation of lipids to biodiesel can be further justified as an economically viable process due to the high purity of glycerol that can be produced via this technology.
As described in this paper, numerous studies have been conducted to fully utilize the microwave energy for catalytic transesterification. Nevertheless, only very few studies focused on the implementation of microwaves into non-catalytic transesterification under supercritical conditions. However, both of the cases proved that microwave heating is a better heating method when compared to conventional heating, in terms of the reaction time, production yield, energy consumption, and so on. Hence, it is believed that the non-catalytic transesterification reactions at milder conditions can be achieved through the implementation of microwave energy. This technology has not been reported so far. Therefore, systematic and mechanistic investigation on the green process of microwave-assisted, catalyst-free transesterification under ambient conditions for biodiesel production is needed to be explored.

3. Conclusions

It is evident that the microwave-assisted reaction is extremely beneficial, in terms of both the reaction rate and energy consumption. Several reactions that do not occur under conventional heating can be achieved with high yields and better quality under microwave irradiation. These superior microwave results can be explained by the thermal and non-thermal effects of microwave irradiation, which cannot be reproduced easily by the conventional methods. The thermal effect refers to the outcome caused by the dielectric heating, which brings differences in the temperature regime, such as superheating, hotspot formation, and selective heating. The non-thermal effect, however, is defined as the specific radiation effect that is not caused by different temperature regimes, but is due to the non-thermal interactions between the substrates and microwave irradiation.
Besides the theoretical parts of microwave energy, this paper also highlights its great potential in biodiesel production. Numerous types of feedstocks have been introduced to prepare biodiesel: vegetable oil, non-edible oil, waste cooking oil, and algae. With the assistance of microwave technology, the advantages of microwaves have been brought over to biodiesel production. A rapid, energy-efficient and environmentally friendly way to produce biodiesel has been developed under microwave heating. The conversion efficiency and yield show that microwaves have the potential for large-scale biodiesel production, as they able to interact with various components. Furthermore, the non-catalytic transesterification at subcritical conditions (<300 °C, 1 bar) instead of supercritical conditions (>350 °C, 15 bar), with the assistance of microwave energy, should be developed in order to solve the cost problem of the conventional non-catalytic transesterification due to its higher energy input to achieve the supercritical condition.
It is undeniable that most of the results collected so far are from the laboratory-scale experiments, and so the reproducibility of similar outcomes at the industry level is still doubtful. Nevertheless, it seems very plausible that the biodiesels used in the future may involve microwaves at some points in their production processes. Certainly, more in-depth research is necessary for the scaling-up purpose in terms of the process design and optimization, the reaction kinetics and thermodynamics, and the protocols of biodiesel analysis.

Acknowledgments

The authors wish to acknowledge the financial support provided by the Ministry of Science, Technology & Innovation (Project code: 03-02-03-SF0287), and Malaysian Ministry of Higher Education (My Master Scholarship). Additionally, we would like to thank Universiti Tenaga Nasional for the research facilities.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gerssen-Gondelach, S.J.; Saygin, D.; Wicke, B.; Patel, M.K.; Faaij, A.P.C. Competing uses of biomass: Assessment and comparison of the performance of bio-based heat, power, fuels and materials. Renew. Sustain. Energy Rev. 2014, 40, 964–998. [Google Scholar] [CrossRef]
  2. Harvey, M.; Pilgrim, S. The new competition for land: Food, energy, and climate change. Food Policy 2011, 36, S40–S51. [Google Scholar] [CrossRef]
  3. Hong, I.K.; Jeon, H.; Kim, H.; Lee, S.B. Preparation of waste cooking oil based biodiesel using microwave irradiation energy. J. Ind. Eng. Chem. 2016, 42, 107–112. [Google Scholar] [CrossRef]
  4. Aly, M.M.; Abdel-Akher, M.; Said, S.M.; Senjyu, T. A developed control strategy for mitigating wind power generation transients using superconducting magnetic energy storage with reactive power support. Int. J. Electr. Power Energy Syst. 2016, 83, 485–494. [Google Scholar] [CrossRef]
  5. Holtsmark, N.; Agheb, E.; Molinas, M.; Høidalen, H.K. High frequency wind energy conversion system for offshore DC collection grid—Part II: Efficiency improvements. Sustain. Energy Grids Netw. 2016, 5, 177–185. [Google Scholar] [CrossRef]
  6. Gedye, R.; Smith, F.; Westaway, K.; Ali, H.; Baldisera, L.; Laberge, L.; Rousell, J. The use of microwave ovens for rapid organic synthesis. Tetrahedron Lett. 1986, 27, 279–282. [Google Scholar] [CrossRef]
  7. Mirzaei, A.; Neri, G. Microwave-assisted synthesis of metal oxide nanostructures for gas sensing application: A review. Sens. Actuators B Chem. 2016, 237, 749–775. [Google Scholar] [CrossRef]
  8. Kusuma, H.S.; Mahfud, M. Preliminary study: Kinetics of oil extraction from basil (Ocimum basilicum) by microwave-assisted hydrodistillation and solvent-free microwave extraction. S. Afr. J. Chem. Eng. 2016, 21, 49–53. [Google Scholar] [CrossRef]
  9. Mubarak, N.M.; Sahu, J.N.; Abdullah, E.C.; Jayakumar, N.S. Plam oil empty fruit bunch based magnetic biochar composite comparison for synthesis by microwave-assisted and conventional heating. J. Anal. Appl. Pyrolysis 2016, 120, 521–528. [Google Scholar] [CrossRef]
  10. Ghasali, E.; Yazdani-rad, R.; Asadian, K.; Ebadzadeh, T. Production of Al-SiC-TiC hybrid composites using pure and 1056 aluminum powders prepared through microwave and conventional heating methods. J. Alloys Compd. 2016, 690, 512–518. [Google Scholar] [CrossRef]
  11. Salamatinia, B.; Mootabadi, H.; Hashemizadeh, I.; Abdullah, A.Z. Intensification of biodiesel production from vegetable oils using ultrasonic-assisted process: Optimization and kinetic. Chem. Eng. Process. Process Intensif. 2013, 73, 135–143. [Google Scholar] [CrossRef]
  12. Farooq, M.; Ramli, A.; Subbarao, D. Biodiesel production from waste cooking oil using bifunctional heterogeneous solid catalysts. J. Clean. Prod. 2013, 59, 131–140. [Google Scholar] [CrossRef]
  13. Im, H.; Lee, H.; Park, M.S.; Yang, J.-W.; Lee, J.W. Concurrent extraction and reaction for the production of biodiesel from wet microalgae. Bioresour. Technol. 2014, 152, 534–537. [Google Scholar] [CrossRef] [PubMed]
  14. Gude, V.; Patil, P.; Martinez-Guerra, E.; Deng, S.; Nirmalakhandan, N. Microwave energy potential for biodiesel production. Sustain. Chem. Process. 2013, 1, 5. [Google Scholar] [CrossRef]
  15. Richmond, A.; Hu, Q. Handbook of Microalgal Culture: Applied Phycology and Biotechnology, 2nd ed.; John Wiley & Sons Ltd.: Chichester, UK, 2013. [Google Scholar]
  16. Singh, S.P.; Singh, D. Biodiesel production through the use of different sources and characterization of oils and their esters as the substitute of diesel: A review. Renew. Sustain. Energy Rev. 2010, 14, 200–216. [Google Scholar] [CrossRef]
  17. Motasemi, F.; Ani, F.N. A review on microwave-assisted production of biodiesel. Renew. Sustain. Energy Rev. 2012, 16, 4719–4733. [Google Scholar] [CrossRef]
  18. Leung, D.Y.C.; Wu, X.; Leung, M.K.H. A review on biodiesel production using catalyzed transesterification. Appl. Energy 2010, 87, 1083–1095. [Google Scholar] [CrossRef]
  19. Saifuddin, N.; Samiuddin, A.; Kumaran, P. A Review on Processing Technology for Biodiesel Production. Trends Appl. Sci. Res. 2015, 10, 1–37. [Google Scholar] [CrossRef]
  20. Lam, M.K.; Lee, K.T.; Mohamed, A.R. Homogeneous, heterogeneous and enzymatic catalysis for transesterification of high free fatty acid oil (waste cooking oil) to biodiesel: A review. Biotechnol. Adv. 2010, 28, 500–518. [Google Scholar] [CrossRef] [PubMed]
  21. Cheng, J.; Yu, T.; Li, T.; Zhou, J.; Cen, K. Using wet microalgae for direct biodiesel production via microwave irradiation. Bioresour. Technol. 2013, 131, 531–535. [Google Scholar] [CrossRef] [PubMed]
  22. García-Martínez, N.; Andreo-Martínez, P.; Quesada-Medina, J.; de los Ríos, A.P.; Chica, A.; Beneito-Ruiz, R.; Carratalá-Abril, J. Optimization of non-catalytic transesterification of tobacco (Nicotiana tabacum) seed oil using supercritical methanol to biodiesel production. Energy Convers. Manag. 2016, 131, 99–108. [Google Scholar] [CrossRef]
  23. Ding, J.; He, B.; Li, J. Biodiesel production from acidified oils via supercritical methanol. Energies 2011, 4, 2212–2223. [Google Scholar] [CrossRef]
  24. Patil, P.D.; Gude, V.G.; Reddy, H.K.; Muppaneni, T.; Deng, S. Biodiesel Production from Waste Cooking Oil Using Sulfuric Acid and Microwave Irradiation Processes. J. Environ. Prot. 2012, 3, 107–113. [Google Scholar] [CrossRef]
  25. Cancela, A.; Maceiras, R.; Urrejola, S.; Sanchez, A. Microwave-assisted transesterification of macroalgae. Energies 2012, 5, 862–871. [Google Scholar] [CrossRef]
  26. Hincapié, G.M.; Valange, S.; Barrault, J.; Moreno, J.A.; López, D.P. Effect of microwave-assisted system on transesterification of castor oil with ethanol. Univ. Sci. 2014, 19, 193–200. [Google Scholar] [CrossRef]
  27. Azcan, N.; Yilmaz, O. Microwave assisted transesterification of waste frying oil and concentrate methyl ester content of biodiesel by molecular distillation. Fuel 2013, 104, 614–619. [Google Scholar] [CrossRef]
  28. Li, Y.; Ye, B.; Shen, J.; Tian, Z.; Wang, L.; Zhu, L.; Ma, T.; Yang, D.; Qiu, F. Optimization of biodiesel production process from soybean oil using the sodium potassium tartrate doped zirconia catalyst under Microwave Chemical Reactor. Bioresour. Technol. 2013, 137, 220–225. [Google Scholar] [CrossRef] [PubMed]
  29. Gimbun, J.; Ali, S.; Kanwal, C.C.S.C.; Shah, L.A.; Ghazali, N.H.M.; Cheng, C.K.; Nurdin, S. Biodiesel Production from Rubber Seed Oil using Activated Cement Clinker as Catalyst. Procedia Eng. 2013, 53, 13–19. [Google Scholar] [CrossRef]
  30. De Souza, R.O.M.A. Theoretical Aspects of Microwave Irradiation Practices; Springer: Dordrecht, The Netherlands, 2015; pp. 3–16. [Google Scholar]
  31. Abadal, G.; Alda, J.; Agust, J. Electromagnetic Radiation Energy Harvesting—The Rectenna Based Approach. In ICT-Energy-Concepts Towards Zero-Power Information and Communication Technology; InTech: Rijeka, Croatia, 2014. [Google Scholar]
  32. Huang, Y.-F.; Chiueh, P.-T.; Lo, S.-L. A review on microwave pyrolysis of lignocellulosic biomass. Sustain. Environ. Res. 2016, 26, 103–109. [Google Scholar] [CrossRef]
  33. Chen, H.-Y.; Cui, Z.-W. A Microwave-Sensitive Solid Acid Catalyst Prepared from Sweet Potato via a Simple Method. Catalysts 2016, 6, 211. [Google Scholar] [CrossRef]
  34. Arshanitsa, A.; Akishin, Y.; Zile, E.; Dizhbite, T.; Solodovnik, V.; Telysheva, G. Microwave treatment combined with conventional heating of plant biomass pellets in a rotated reactor as a high rate process for solid biofuel manufacture. Renew. Energy 2016, 91, 386–396. [Google Scholar] [CrossRef]
  35. Huang, S.; Xu, H.; Wang, M.; Zhong, S.; Zeng, C. Rapid microwave synthesis and photoluminescence properties of rare earth-based coordination polymer core-shell particles. Opt. Mater. 2016, 62, 538–542. [Google Scholar] [CrossRef]
  36. Valizadeh, H.; Dinparast, L.; Noorshargh, S.; Heravi, M.M. Microwave assisted synthesis of hydroxychromenes using imidazole-functionalized silica nanoparticles as a catalyst under solvent-free conditions. Comptes Rendus Chim. 2016, 19, 395–402. [Google Scholar] [CrossRef]
  37. Ye, W.; Gao, Y.; Ding, H.; Liu, M.; Liu, S.; Han, X.; Qi, J. Kinetics of transesterification of palm oil under conventional heating and microwave irradiation, using CaO as heterogeneous catalyst. Fuel 2016, 180, 574–579. [Google Scholar] [CrossRef]
  38. Zhang, D.; Zhang, Y.; Zhao, T.; Li, J.; Hou, Y.; Gu, Q. A rapid and efficient solvent-free microwave-assisted synthesis of pyrazolone derivatives containing substituted isoxazole ring. Tetrahedron 2016, 72, 2979–2987. [Google Scholar] [CrossRef]
  39. Sun, J.; Wang, W.; Yue, Q. Review on microwave-matter interaction fundamentals and efficient microwave-associated heating strategies. Materials 2016, 9, 231. [Google Scholar] [CrossRef]
  40. Cheng, J.; Roy, R.; Agrawal, D. Radically different effects on materials by separated microwave electric and magnetic fields. Mater. Res. Innov. 2002, 5, 170–177. [Google Scholar] [CrossRef]
  41. Cheng, J.; Roy, R.; Agrawal, D. Experimental proof of major role of magnetic field losses in microwave heating of metal and metallic composites. J. Mater. Sci. Lett. 2001, 20, 1561–1563. [Google Scholar] [CrossRef]
  42. Peng, Z.; Hwang, J.-Y.; Andriese, M. Magnetic Loss in Microwave Heating. Appl. Phys. Express 2012, 5, 27304. [Google Scholar] [CrossRef]
  43. Surat, M.A.; Jauhari, S.; Desak, K.R. A brief review: Microwave assisted organic reaction. Appl. Sci. Res. 2012, 4, 645–661. [Google Scholar]
  44. De la Hoz, A.; Loupy, A. Microwaves in Organic Synthesis; Wiley-VCH: Weinheim, Germany, 2012. [Google Scholar]
  45. Chen, C.-C.; Reddy, P.M.; Devi, C.S.; Chang, P.-C.; Ho, Y.-P. Study of microwave effects on the lipase-catalyzed hydrolysis. Enzyme Microb. Technol. 2016, 82, 164–172. [Google Scholar] [CrossRef] [PubMed]
  46. Baghurst, D.R.; Mingos, D.M.P. Superheating effects associated with microwave dielectric heating. J. Chem. Soc. Chem. Commun. 1992, 20, 674–677. [Google Scholar] [CrossRef]
  47. Durka, T.; Stefanidis, G.D.; Van Gerven, T.; Stankiewicz, A.I. Microwave-activated methanol steam reforming for hydrogen production. Int. J. Hydrogen Energy 2011, 36, 12843–12852. [Google Scholar] [CrossRef]
  48. Varisli, D.; Korkusuz, C.; Dogu, T. Microwave-assisted ammonia decomposition reaction over iron incorporated mesoporous carbon catalysts. Appl. Catal. B Environ. 2017, 201, 370–380. [Google Scholar] [CrossRef]
  49. Ren, L. Directly Catalytic Decomposition of H2S to Sulfur and Hydrogen under Microwave Conditions. Adv. Mater. Res. 2010, 129, 317–321. [Google Scholar] [CrossRef]
  50. Lee, C.S.; Binner, E.; Winkworth-Smith, C.; John, R.; Gomes, R.; Robinson, J. Enhancing natural product extraction and mass transfer using selective microwave heating. Chem. Eng. Sci. 2016, 149, 97–103. [Google Scholar] [CrossRef]
  51. Nagao, D.; Fukushima, J.; Hayashi, Y.; Takizawa, H. Synthesis of homologous compounds Fe2O3(ZnO)m (m = 6, 8, 34) by various selective microwave heating conditions. Ceram. Int. 2015, 41, 14021–14028. [Google Scholar] [CrossRef]
  52. Wang, Y.; Stang, J.; Yu, M.; Tsvetkov, M.; Wu, C.-C.; Qin, X.; Chung, E.; Moghaddam, M.; Wu, W. Microwave Selective Heating Enhancement for Cancer Hyperthermia Therapy Based on Lithographically Defined Micro/Nanoparticles. Adv. Mater. Technol. 2016, 1, 1600038. [Google Scholar] [CrossRef]
  53. De La Hoz, A.; Díaz-Ortiz, A.; Moreno, A. Review On Non-thermal Effects of Microwave Irradiation in Organic Synthesis. J. Microw Power Electr. Energy 2007, 41, 44–46. [Google Scholar] [CrossRef]
  54. Sumi, T.; Horikoshi, S. Microwave selective heating for size effect of water droplet in W/O emulsion with sorbitan fatty acid monostearate surfactant. Radiat. Phys. Chem. 2015, 114, 31–37. [Google Scholar] [CrossRef]
  55. Horikoshi, S.; Kamata, M.; Sumi, T.; Serpone, N. Selective heating of Pd/AC catalyst in heterogeneous systems for the microwave-assisted continuous hydrogen evolution from organic hydrides: Temperature distribution in the fixed-bed reactor. Int. J. Hydrogen Energy 2016, 41, 12029–12037. [Google Scholar] [CrossRef]
  56. Rasuli, R.; Mokarian, Z.; Karimi, R.; Shabanzadeh, H.; Abedini, Y. Wettability modification of graphene oxide by removal of carboxyl functional groups using non-thermal effects of microwave. Thin Solid Films 2015, 589, 364–368. [Google Scholar] [CrossRef]
  57. Nazari, S.; Ghandi, K. Solvent and microwave effects on oxidation of aromatic α-diketones. J. Ind. Eng. Chem. 2015, 21, 198–205. [Google Scholar] [CrossRef]
  58. Kappe, C.O.; Pieber, B.; Dallinger, D. Microwave Effects in Organic Synthesis: Myth or Reality? Angew. Chem. Int. Ed. 2013, 52, 1088–1094. [Google Scholar] [CrossRef] [PubMed]
  59. Bhattacharjee, M.K.; Delsol, J.K. Does microwave sterilization of growth media involve any non-thermal effect? J. Microbiol. Methods 2014, 96, 70–72. [Google Scholar] [CrossRef] [PubMed]
  60. Xu, W.; Zhou, J.; Su, Z.; Ou, Y.; You, Z. Microwave catalytic effect: A new exact reason for microwave-driven heterogeneous gas-phase catalytic reactions. Catal. Sci. Technol. 2016, 6, 698–702. [Google Scholar] [CrossRef]
  61. Zhou, J.; Xu, W.; You, Z.; Wang, Z.; Luo, Y.; Gao, L.; Yin, C.; Peng, R.; Lan, L. A new type of power energy for accelerating chemical reactions: The nature of a microwave-driving force for accelerating chemical reactions. Sci. Rep. 2016, 6, 25149. [Google Scholar] [CrossRef] [PubMed]
  62. Fukushima, J.; Kashimura, K.; Takayama, S.; Sato, M.; Sano, S.; Hayashi, Y.; Takizawa, H. In-situ kinetic study on non-thermal reduction reaction of CuO during microwave heating. Mater. Lett. 2013, 91, 252–254. [Google Scholar] [CrossRef]
  63. Mao, T.Y.; Zheng, C.; Yu, J. Effects of Microwave Irradiation on Activation Energy and Reaction Rate of Synthesis of Dodecylmethyldihydroxyl Ammonium Bromide. Adv. Mater. Res. 2011, 396, 1044–1047. [Google Scholar] [CrossRef]
  64. Tian, W.-Y.; Huang, K.-M. An improved experimental set-up based on ridged waveguide for microwave non-thermal effects. J. Instrum. 2011, 6, T02001. [Google Scholar] [CrossRef]
  65. Liu, J.; Jia, G. Non-thermal effects of microwave in sodium chloride aqueous solution: Insights from molecular dynamics simulations. J. Mol. Liq. 2017, 227, 31–36. [Google Scholar] [CrossRef]
  66. Suryanto, A.; Suprapto; Mahfud. The Production of Biofuels from Coconut Oil Using Microwave. Mod. Appl. Sci. 2015, 9, 93–98. [Google Scholar] [CrossRef]
  67. Barekati-Goudarzi, M.; Boldor, D.; Nde, D.B. In-situ transesterification of seeds of invasive Chinese tallow trees (Triadica sebifera L.) in a microwave batch system (GREEN3) using hexane as co-solvent: Biodiesel production and process optimization. Bioresour. Technol. 2016, 201, 97–104. [Google Scholar] [CrossRef] [PubMed]
  68. Choedkiatsakul, I.; Ngaosuwan, K.; Assabumrungrat, S.; Mantegna, S.; Cravotto, G. Biodiesel production in a novel continuous flow microwave reactor. Renew. Energy 2015, 83, 25–29. [Google Scholar] [CrossRef]
  69. Marwan; Indarti, E. Hydrated calcined Cyrtopleura costata seashells as an effective solid catalyst for microwave-assisted preparation of palm oil biodiesel. Energy Convers. Manag. 2016, 117, 319–325. [Google Scholar]
  70. Encinar, J.M.; González, J.F.; Martínez, G.; Sánchez, N.; Pardal, A. Soybean oil transesterification by the use of a microwave flow system. Fuel 2012, 95, 386–393. [Google Scholar] [CrossRef]
  71. Kumar, R.; Ravi Kumar, G.; Chandrashekar, N. Microwave assisted alkali-catalyzed transesterification of Pongamia pinnata seed oil for biodiesel production. Bioresour. Technol. 2011, 102, 6617–6620. [Google Scholar] [CrossRef] [PubMed]
  72. Wahidin, S.; Idris, A.; Shaleh, S.R.M. Rapid biodiesel production using wet microalgae via microwave irradiation. Energy Convers. Manag. 2014, 84, 227–233. [Google Scholar] [CrossRef]
  73. Kanitkar, A.V. Parameterization of Microwave Assisted Oil Extraction and Its Transesterification to Biodiesel; Louisiana State University, Biological & Agricultural Engineering Department: Baton Rouge, LA, USA, 2010. [Google Scholar]
  74. Duz, M.Z.; Saydut, A.; Ozturk, G. Alkali catalyzed transesterification of safflower seed oil assisted by microwave irradiation. Fuel Process. Technol. 2011, 92, 308–313. [Google Scholar] [CrossRef]
  75. Jin, L.; Zhang, Y.; Dombrowski, J.P.; Chen, C.-H.; Pravatas, A.; Xu, L.; Perkins, C.; Suib, S.L. ZnO/La2O2CO3 layered composite: A new heterogeneous catalyst for the efficient ultra-fast microwave biofuel production. Appl. Catal. B Environ. 2011, 103, 200–205. [Google Scholar] [CrossRef]
  76. Hussain, A.; Ali, S.; Ahmed, I.; Gimbun, J.; Albeirutty, M.H.; Rehan, Z.A. Microwave Reinforced Transesterification of Rubber Seed Oil Using Waste Cement Clinker Catalyst. Curr. Nanosci. 2016, 12, 576–585. [Google Scholar] [CrossRef]
  77. Zhang, Y.; Xia, X.; Duan, M.; Han, Y.; Liu, J.; Luo, M.; Zhao, C.; Zu, Y.; Fu, Y. Green deep eutectic solvent assisted enzymatic preparation of biodiesel from yellow horn seed oil with microwave irradiation. J. Mol. Catal. B Enzym. 2016, 123, 35–40. [Google Scholar] [CrossRef]
  78. Kapilan, N. Production of Biodiesel from Vegetable Oil Using Microware Irradiation. Acta Polytech. 2012, 52, 46–50. [Google Scholar]
  79. Venkatesh Kamath, H.; Regupathi, I.; Saidutta, M.B. Optimization of two step karanja biodiesel synthesis under microwave irradiation. Fuel Process. Technol. 2011, 92, 100–105. [Google Scholar] [CrossRef]
  80. Patil, P.D.; Reddy, H.; Muppaneni, T.; Ponnusamy, S.; Cooke, P.; Schuab, T.; Deng, S. Microwave-mediated non-catalytic transesterification of algal biomass under supercritical ethanol conditions. J. Supercrit. Fluids 2013, 79, 67–72. [Google Scholar] [CrossRef]
  81. Patil, P.D.; Reddy, H.; Muppaneni, T.; Schaub, T.; Holguin, F.O.; Cooke, P.; Lammers, P.; Nirmalakhandan, N.; Li, Y.; Lu, X.; et al. In situ ethyl ester production from wet algal biomass under microwave-mediated supercritical ethanol conditions. Bioresour. Technol. 2013, 139, 308–315. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Biodiesel production classification.
Figure 1. Biodiesel production classification.
Bioengineering 04 00057 g001
Figure 2. Electromagnetic spectrum [31].
Figure 2. Electromagnetic spectrum [31].
Bioengineering 04 00057 g002
Figure 3. Heating mechanisms for: (a) conventional heating, and (b) microwave heating [43].
Figure 3. Heating mechanisms for: (a) conventional heating, and (b) microwave heating [43].
Bioengineering 04 00057 g003
Figure 4. Selective heating of water/oil emulsion (water droplets are represented by the orange bubbles) [54].
Figure 4. Selective heating of water/oil emulsion (water droplets are represented by the orange bubbles) [54].
Bioengineering 04 00057 g004
Figure 5. Effect of extraction time on extract yield (DB) at 70 °C [50].
Figure 5. Effect of extraction time on extract yield (DB) at 70 °C [50].
Bioengineering 04 00057 g005
Figure 6. Reaction coordinates of CuO reduction under conventional heating, E-field microwave heating, and H-field microwave heating.
Figure 6. Reaction coordinates of CuO reduction under conventional heating, E-field microwave heating, and H-field microwave heating.
Bioengineering 04 00057 g006
Figure 7. Schematic illustration of the tetrahedral structure of water before and after microwave irradiation.
Figure 7. Schematic illustration of the tetrahedral structure of water before and after microwave irradiation.
Bioengineering 04 00057 g007
Figure 8. The field dependence of the tetrahedral order parameter and lifetime of hydrogen bonds.
Figure 8. The field dependence of the tetrahedral order parameter and lifetime of hydrogen bonds.
Bioengineering 04 00057 g008
Figure 9. Microwave-assisted catalytic biodiesel production.
Figure 9. Microwave-assisted catalytic biodiesel production.
Bioengineering 04 00057 g009
Figure 10. The microwave-mediated transesterification of algal biomass under supercritical ethanol conditions to yield ethyl ester.
Figure 10. The microwave-mediated transesterification of algal biomass under supercritical ethanol conditions to yield ethyl ester.
Bioengineering 04 00057 g010
Figure 11. Experimental protocol for the microwave-mediated supercritical ethanol transesterification process for algal biomass.
Figure 11. Experimental protocol for the microwave-mediated supercritical ethanol transesterification process for algal biomass.
Bioengineering 04 00057 g011
Table 1. Different feedstocks for biodiesel production.
Table 1. Different feedstocks for biodiesel production.
Vegetable/Edible OilsNon-Edible OilsAnimal FatsOther Feedstocks
SoybeansAlmondLardBacteria
RapeseedAbutilon muticumTallowAlgae
CanolaAndirobaPoultry FatFungi
SafflowerBabassuFish oilMicroalgae
BarleyBrassica carinata Tarpenes
CoconutB. napus Latexes
CopraCamelina Cooking oil (Yellow Grease)
Cotton seedCumaru Microalgae (Chlorellavulgaris)
GroundnutCynara cardunculus
OatJatropha curcas
RiceJatropha nana
SorghumJojoba oil
WheatPongamiaglabra
Winter rapeseed oilLaurel
Lesquerellafendleri
Mahua
Piqui
Palm
Karang
Tobacco seed
Rubber plant
Rice bran
Sesame
Salmon oil
Table 2. Comparison between three types of heating methods for biodiesel production.
Table 2. Comparison between three types of heating methods for biodiesel production.
Characteristic/ParameterConventional HeatingSupercritical HeatingMicrowave Heating
Reaction time1 to 2 h (Long)Less than 1 h (Short)0.05 to 0.1 h (Very short)
Reaction temperature40 °C to 100 °C250 °C to 400 °C40 °C to 100 °C
Reaction pressureAtmospheric35 MPa to 60 MPa (High)Atmospheric *
Catalyst requiredYesNoYes/No
Heat lossesHighModerateLow
Energy formElectrical energy to thermal energyElectrical energy to thermal energyElectrical energy applied through microwave
Process efficiencyLowModerateHigh
Catalyst removalYesNoYes
Soap removalYesNoYes
AdvantagesSimple to operate, uses low energy sourceShort reaction time, easy product separationShort reaction time, cleaner products, energy efficient
LimitationsHigh energy requirement, saponified productsHigh capital costs, pressure vessel safetyMay not be efficient with feedstock containing solids
* Reactions at high temperature and pressure with no catalysts are possible.
Table 3. Advantages and disadvantages of different types of catalysts in biodiesel production.
Table 3. Advantages and disadvantages of different types of catalysts in biodiesel production.
Type of CatalystAdvantagesDisadvantages
Homogeneous base catalyst
  • Very fast reaction rate: around 4000 times faster than acid-catalyzed transesterification.
  • Reaction can be carried out at mild reaction conditions (less energy intensive).
  • Catalysts such as NaOH and KOH are relatively cheap and widely available.
  • Sensitive to free fatty acid (FFA) content in the oil.
  • Unwanted soap formation if the FFA percentage in the feedstock is more than 2 wt %.
  • Decrease in the biodiesel yield due to soap formation.
  • The downstream purification process raises problems such as producing a large amount of wastewater.
Heterogeneous base catalyst
  • Relatively fast reaction rate in comparison to acid-catalyzed transesterification.
  • Reaction can be carried out at mild reaction conditions (less energy intensive).
  • Easy separation of catalyst after the reaction.
  • High possibility to reuse and regenerate the catalyst.
  • Poisoning of catalyst when exposed to the surrounding air.
  • Sensitive to FFA content in the oil due to its basicity property.
  • Unwanted soap formation if the FFA percentage in the feedstock is more than 2 wt %.
  • Decrease in the biodiesel yield due to soap formation.
  • Leaching of catalyst active sites may lead to product contamination.
Homogeneous acid catalyst
  • Insensitive to FFAs and water content in the feedstock.
  • Preferable method for cheaper feedstock.
  • Simultaneous esterification and transesterification processes.
  • Reaction can be carried out at mild reaction conditions (less energy intensive).
  • Very slow reaction rate.
  • Corrosive catalyst such as H2SO4 leads to corrosion on reactor and pipelines.
  • Separation of catalyst from product is problematic.
Heterogeneous acid catalyst
  • Insensitive to FFA and water content in the feedstock.
  • Preferable method for cheaper feedstock.
  • Simultaneous esterification and transesterification processes.
  • Easy separation of catalyst from the product.
  • High possibility to reuse and regenerate the catalyst.
  • Complicated catalyst synthesis process results in higher overall production cost.
  • Normally, a high reaction temperature, high alcohol-to-oil molar ratio, and long reaction time are required.
  • Energy intensive.
  • Leaching of catalyst active sites may lead to product contamination.
Enzymes
  • FFAs are converted to biodiesel.
  • Not affected by the water content in the feedstock.
  • High biodiesel yield, usually around 90%.
  • Simple glycerol recovery and produces high grade glycerol.
  • Easy catalyst recovery and the reusability proved but not sufficiently studied.
  • Low reaction temperature at 20–50 °C.
  • Low environmental impact as no need for wastewater treatment.
  • Low to moderately-high reaction rate, depending on the parameters.
  • Relatively high cost of catalysts if enzymes cannot be recovered and reused.
  • Possible enzyme inhibition by alcohols.
Table 4. Energy of different bonds and microwaves.
Table 4. Energy of different bonds and microwaves.
Bonding and MicrowaveEnergy (eV)
Brownian motion at 37 °C2.7 × 10−3
Biological compound13.6
Covalent bond (e.g., OH)5
Hydrogen bond2
Van der Waals intermolecular interactions<2
Microwave at 0.3 GHz1.24 × 10−6
Microwave at 2.45 GHz1 × 10−5
Microwave at 300 GHz1.24 × 10−3
Table 5. Comparison between microwave heating and conventional heating.
Table 5. Comparison between microwave heating and conventional heating.
Microwave HeatingConventional Heating
Energetic couplingConduction/convection
Coupling at molecular levelSuperficial/wall heating
RapidSlow
VolumetricSuperficial
SelectiveNon-selective
Dependent on material’s propertiesLess dependent
Table 6. Experimental studies of microwave-assisted transesterification with different feedstocks.
Table 6. Experimental studies of microwave-assisted transesterification with different feedstocks.
FeedstockFeedstock SolventReaction Time (min)Reaction Temperature (°C)CatalystsMicrowave Power (W)Biodiesel Yield (%)Reference
Coconut oil1:6 Methanol5 1% NaOH10097.76[66]
Waste frying oil1:6 Methanol2601% NaOCH3 98.87[27]
1:12 Methanol6 2% BaO80096[24]
1:9 Methanol6 2% KOH80092[24]
Chinese tallow tree1:3 Methanol (w/v) (co-solvent: hexane)2058.11.74% NaOH 96.62[67]
Palm oil1:12 Methanol1.75701% NaOH40099.4[68]
1:5 Methanol10 0.5% Ca(OH)2 from seashells 96[69]
1:9 Methanol60655% CaO15089.9[37]
Soybean oil1:12 Methanol2701% KOH20099[70]
1:14 Methanol306510% C4H4O6KNa doped ZrO2 catalyst 94.75[28]
Castor oil1:10 Ethanol10601.5% KOH 80.1[26]
Pongamia pinnata1:6 Methanol5600.5% NaOH 96[71]
1:6 Methanol5601% KOH 96[71]
Wet microalgae1 g : 4 mL Methanol (co-solvents: chloroform, sulfuric acid)3060 40011% of dry-mass[21]
Microalgae oil1:1 Methanol-hexane (v/v)1065NaOH 86.41[72]
Rice bran oil1:5 Methanol20600.15% NaOH 98[73]
Safflower oil1:10 Methanol6601% NaOH30098.4[74]
Canola1:1 Methanol (w/w)51001% ZnO/La2O2CO3 95[75]
Rubber seed oil1:5 Methanol60606% Cement clinker catalyst 96.8[76]
Yellow horn seed oil1:6 Methanol120508% immobilized Novozym 435 (in green deep eutectic solvent)40095[77]
Table 7. Microwave-based biodiesel properties.
Table 7. Microwave-based biodiesel properties.
PropertyUnitsASTM D6751 EN 14214 Microwave HeatingRegular Diesel [78]
Waste Cooking Oil [24]Chinese Tallow Tree [67]Karanja [79]Soybean [70]Jatropha [78]
Densitykg/m3 at 15 °C870–890860–900870–880880887877889846
Viscositymm2/s at 40 °C1.9–63.5–52.25–3.102.024.34.224.212.28
Pour point°C −4 to −12 6−15
Flash point°C>93>101 14517313268
Cetane index >47>5155.45–56.1062.7356.350.9
Copper strip corrosion index No. 3 maxClass 1 No. 2 No. 1No. 1
Iodine valueg I2/100 g <120 83115.3
Heating valueMJ/kg 32.5–36.145.08–45.24 37.9542.71
Saponification valuemg KOH/g 195181.3
Acid valueMg KOH/g<0.5<0.5 0.340.4050.140.420.34
Water content%<0.05<0.05 6.00% 2.00%0.01290.0102

Share and Cite

MDPI and ACS Style

Nomanbhay, S.; Ong, M.Y. A Review of Microwave-Assisted Reactions for Biodiesel Production. Bioengineering 2017, 4, 57. https://doi.org/10.3390/bioengineering4020057

AMA Style

Nomanbhay S, Ong MY. A Review of Microwave-Assisted Reactions for Biodiesel Production. Bioengineering. 2017; 4(2):57. https://doi.org/10.3390/bioengineering4020057

Chicago/Turabian Style

Nomanbhay, Saifuddin, and Mei Yin Ong. 2017. "A Review of Microwave-Assisted Reactions for Biodiesel Production" Bioengineering 4, no. 2: 57. https://doi.org/10.3390/bioengineering4020057

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop