Next Article in Journal
The Nuclear Magnetic Flowmeter for Monitoring the Consumption and Composition of Oil and Its Complex Mixtures in Real-Time
Previous Article in Journal
Operation Algorithm for Protection Coordination Device in High-Voltage Customer with ESS for Demand Management
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Microwave-Assisted Pyrolysis of Biomass with and without Use of Catalyst in a Fluidised Bed Reactor: A Review

University of Lyon, CNRS, IRCELYON—UMR 5256, 2 Avenue Albert Einstein, CEDEX, F-69626 Villeurbanne, France
*
Author to whom correspondence should be addressed.
Energies 2022, 15(9), 3258; https://doi.org/10.3390/en15093258
Submission received: 14 March 2022 / Revised: 5 April 2022 / Accepted: 27 April 2022 / Published: 29 April 2022
(This article belongs to the Special Issue Developing Technologies for Fuels Production)

Abstract

:
Lignocellulosic biomass and waste, such as plastics, represent an abundant resource today, and they can be converted thermo-chemically into energy in a refinery. Existing research works on catalytic and non-catalytic pyrolysis performed in thermally-heated reactors have been reviewed in this text, along with those performed in microwave-heated ones. Thermally-heated reactors, albeit being the most commonly used, present various drawbacks such as superficial heating, high thermal inertia and slow response times. That is why microwave-assisted pyrolysis (MAP) appears to be a very promising technology, even if the process does present some technical drawbacks as well such as the formation of hot spots. The different types of catalysts used during the process and their impacts have also been examined in the text. More specifically, studies conducted in fluidised bed reactors (FBR) have been detailed and their advantages and drawbacks discussed. Finally, future prospects of MAP have been briefly presented.

1. Introduction

Since the start of the twenty-first century, rapid population and technological growth have put relentless pressure on already ever-dwindling fossil fuel resources. The peak in energy demand is not expected to diminish in the coming years. More people existing on the planet translates into more waste, and so, more biomass being produced. Hence, biomass as an energy source could be a viable replacement for fossil sources in the coming years. Various products such as biodiesel, bioethanol and biohydrogen have successfully been produced from biomass, supporting its potential as a reliable and clean energy source. Biofuels can be classified into three broad categories: first, second and third generations. Table 1 catalogues the different known biomass resources.
A plethora of conversion processes—physical, thermal, biochemical, microbial or chemical—have been used to transform biomass into fuels [2,3]. Among all the processes, thermochemical ones present a better choice so as to convert biomass feedstock more rapidly at a lesser cost [3]. Interesting thermochemical transformation processes include combustion or incineration, catalytic liquefaction, gasification and pyrolysis, which is the main focus of this review.

2. Pyrolysis and Catalytic Pyrolysis

Pyrolysis is a thermal conversion process that heats and decomposes a substance in an oxygen-inert atmosphere. This process is classified into three distinct technologies, and forms three main products, as illustrated in Figure 1 and Figure 2, respectively. The process has largely been examined in literature and can be optimised to maximise the production of any of its product constituents by altering process parameters such as the operational temperature and residence time, among others [4,5,6,7,8].
Different types of reactors have been used to conduct pyrolysis reactions over time [11]. The more common ones are: ablative, conical spouted bed, rotating cone, Auger and fluidised bed (bubbling and circulating) reactors [12,13,14]. Lam et al. [15] state that most studies pertaining to biomass pyrolysis at the laboratory scale are performed in batch reactors as the latter are simple to use, no high-pressure pump or compressor is needed, all types of biomass can be used and no complex separation means needs to be applied to recover the remaining bio-char. However, the drawbacks presented by this technology, extended reaction, heating and cooling times, product inconsistency and difficulty of large-scale production, have geared more recent research towards semi-continuous and continuous processes [16].
Now, it should be noted that the bio-oil produced from biomass pyrolysis has a high oxygen content, which translates to it also having high acidity, high viscosity, low calorific value and thermal instability [17]. The latter characteristics present a challenge for bio-oil to be used as a substitute for conventional fuels. Therefore, the main objective of most pyrolysis systems is to maximise the yield of bio-oil and improve the selectivity of the compounds within the bio-oil. Various methods have been studied in literature to enhance the properties of these bio-oils. One method commonly used in literature is catalytic fast pyrolysis (CFP). Different upgrading methods have also been researched in the literature; one of the most studied methods is catalytic hydrodeoxygenation (HDO).
Some studies have shown that catalytic HDO presents an increase in liquid hydrocarbons, gaseous olefin yields and a decrease in coke yield with respect to untreated bio-oil [18]. On the other hand, it does exhibit some drawbacks such as its need for rare metals and high-pressure hydrogen, and its rather expensive cost. Most research studies on catalytic HDO have been conducted on model compounds such as guaiacol, phenol or cresol; Ambursa et al. [19] and Wang et al. [20] present an exhaustive summary of these studies in their respective publications. Some of the most notable research works concerning catalytic HDO have been summarised in Table 2.
As for CFP, only the research performed on the upgrading of pyrolytic vapours, that is, ex situ CFP, have been considered. The most commonly used catalysts in literature are zeolite-based. Their efficiency in deoxygenating bio-oils and increasing their hydrocarbon content has been widely researched and proven, as shown in Table 3.
Nowadays, research is tending more and more towards a cleaner way of disposing of plastic waste, which has been increasing at an alarming rate in recent years. Pyrolysis, and hence, microwave pyrolysis, has emerged as a very promising technology to this end. Kaminsky [41] performed the chemical recycling of different types of plastics in an FBR. His results have been summarised in Table 4.
The authors found that PE and PP formed mostly waxy products around temperatures of 400–550 °C and gas and oil around 700 °C. By using pyrolysis gas as a fluidising agent, aromatics were produced. It was also observed that up to 98 wt. % of MMA could be recovered from the pyrolysis of pure PMMA wastes. As for styrene recovery from PS, it was limited to about 77 wt. % and the remainder was oligomers. PET was hydrolysed through this process with steam as a fluidising agent and PA, at low temperatures, produced aromatics and gas; soot formation was found to be low. Finally, PTFE was pyrolysed into high amounts of TFE and HEP, its monomers. Hence, it was concluded that the pyrolysis of plastics in a fluidised bed reactor was a real option for an industrial application.
Now, conventional pyrolysis systems making use of electrical heating have by far been the most-studied types of systems in the literature. However, these systems are energy-intensive, which compromises their economic viability on an industrial scale. Over recent years, research has been focusing on new, innovative ways to build energy-efficient systems for pyrolysis. One of the most promising routes found is microwave-assisted pyrolysis. It should also be noted that the addition of proper catalysts to a microwave pyrolysis process can dramatically influence the process’ product selectivity; gas and liquid yields, as well as heating performance, can be importantly promoted through this addition of catalysts [42]. The distinction between the effects of the two (microwave heating and catalyst use) will try to be made throughout the text.

3. Fundamentals on Microwaves

Microwaves are electromagnetic irradiation with a wavelength range of 0.01 to 1 m, and a frequency range of 0.3 to 300 GHz, correspondingly [43]. Most microwave reactors used for chemical syntheses, as well as all domestic microwave ovens, function at a frequency of 2.45 GHz, correlating to a 12.25 cm wavelength. This is conducted in order to prevent any interference with telecommunications and cellular phone frequencies [44]. Microwaves are typically defined as electromagnetic waves having two perpendicular components, electric and magnetic fields, as shown in Figure 3.
Microwave dielectrics are known as materials able to absorb microwave irradiation, and thus, microwave heating is called dielectric heating [46]. Dielectric heating generally functions by transforming electromagnetic energy into thermal energy; it is more of an energy conversion than direct heating. The microwaves penetrate materials and deposit energy; thus, the heat can be diffused throughout the volume of the material. In this case, the centre of the material is at a higher temperature than its surroundings [1]. On the other hand, during conventional heating, heat is transferred from the surface to the core of the material via conduction driven by temperature gradients [47]. This difference is illustrated in Figure 4.
Some advantages of microwave heating over conventional heating are presented in Table 5 [48].
Now, according to the dielectric properties of a material, it may be categorised in three possible ways [2,39]:
  • Insulators or microwave-transparent materials, through which microwaves may pass without any losses (e.g., quartz, Teflon and so on), commonly possess low dielectric loss factors and have very large penetration depths;
  • Conductors or reflectors, which microwaves cannot penetrate and are reflected, are generally materials with high conductance and low capacitance (e.g., metals) that have high dielectric loss factors and hence, near-zero penetration depth for microwaves, and;
  • Absorbers, where microwave irradiation may be absorbed most effectively (e.g., water, oils and so on), have dielectric loss factors in the middle of the conductivity range.
According to the literature [49,50], there are three ways of enhancing a chemical reaction using the microwave irradiation technology:
  • Thermal effects (the influence of a high reaction temperature rapidly attained when irradiating polar materials in a microwave field);
  • Specific effects (caused by the unique nature of the heating mechanism of microwave irradiation in a microwave field; this cannot be achieved using conventional heating) and;
  • Non-thermal, non-specific effects (chemical transformation accelerations that have not been attributed to either thermal or specific microwave effects).
Indeed, as a result of rapid heating within the material due to the use of microwaves, reaction times of many microwave-assisted organic reactions are shortened and allow higher yields and selectivity to be obtained. Another good reason to use microwave heating is that there is no direct contact between the energy source and the reactor walls as microwaves pass through the vessel walls, which are almost transparent to them, directly to interact with the reactants [50].

4. Microwave-Assisted Pyrolysis (MAP)

As mentioned previously, when compared to conventional heating, microwave heating presents various advantages, which influence significantly the yield and characteristics of the resulting products. Products from microwave pyrolysis obtained by using appropriate power levels may have high heating values; they may also yield higher gas and solid fractions, but lower liquid products than conventional pyrolysis. Indeed, through the use of microwave pyrolysis, almost half of the lignocellulosic biomass can be converted to gaseous products, mainly consisting of H2, CH4, CO and CO2 and this gas product will possess a higher energetic value because of its high H2 and CO yields [42].
Many researchers have devoted time and resources to microwave-assisted pyrolysis of biomass since recent years due to its advantages [2,6,7,51,52,53,54,55,56,57,58]. These studies have mostly focused on what the use of microwaves can bring to the pyrolysis process when compared to the use of conventional heating. This review will endeavour to group together significant research works performed in recent years concerning microwave-assisted catalytic pyrolysis while focusing mainly on the reactions carried out in fluidised bed reactors.
Setups used at the laboratory scale are typically fixed beds utilising continuous gas flow ones [59]. Microwave reactors may be differentiated based on applicator design, wavelength, flow and power source type [59]. Firstly, the applicator, also known as the cavity, is the structure where the dielectric material to be heated is placed. There are three types of applicators [59]:
  • Multimode applicators: the most commonly used types of applicators for domestic and laboratory purposes, where a large number of resonant modes for operational frequency are hosted.
  • Single-mode applicators: most widely used for heterogeneous gas-phase reactions; capacity to attain high temperatures at low powers. These applicators are able to deliver a highly concentrated energy field, leading to faster heating rates.
  • Travelling-wave applicators: microwaves travel in only one direction, ensuring no back-reflection of waves and hence, no non-uniform interference in the waveguide (channel through which microwave energy propagates from transmitter to receiver) and no standing wave formation.
Secondly, the type of waveguide used is important as each type can only support specific modes of wave propagation [59]. Rectangular ones are the most commonly used types owing to their compatibility with the operational modes needed and their ease of being fabricated [59]. Next to be considered are the generator types. Microwave reactors are typically powered by magnetron tubes. However, magnetrons function at a unique frequency and have short lifetimes (2000–10,000 h). New emerging technologies include solid-state amplifiers which can vary frequency during their operation. The advantage of changing frequency is to mitigate hot spot formation and random wave interactions [59]. Finally, a microwave system can be dimensioned and optimised for specified reaction flow conditions. Depending on various cases, fixed bed, batch or fluidised bed reactors may be used.
Lab-scale microwave setups include a microwave power generator, waveguides, a cavity withholding the reactor, an autotuner, sliding short, temperature and flow controllers, gas supply units and a cooling system [59]. Figure 5 illustrates an example of a common type of microwave setup used for the pyrolysis of biomass waste. The setup presented here was used by Mokhtar et al. [60].

4.1. Effect of MAP on Pyrolysis Product Distribution

Based on the literature, MAP is reported to be much better than conventional pyrolysis in terms of product distribution, yield and quality. Indeed, Ge et al. [55] found by comparing the effects of different heating sources (direct thermal, solar, infrared and microwave heating) on liquid yield obtained from pyrolysis under the same operational conditions that microwave heating was the most efficient heating method in order to achieve the highest bio-oil yield. Several studies have been performed on the effect of the intensity of power used for microwave heating on the products [61,62,63,64,65]. These have shown that the minimum power required to initiate the pyrolysis process is 300 W. However, powers less than 480 W yielded low bio-oil quantities. Increasing the microwave power used resulted in enhancing gas and bio-oil formation while diminishing char formation. It can also be noted that maximum gas yield is obtained at the maximum wattage used.
Domínguez et al. [65] studied the MAP and conventional pyrolysis of sewage sludge using graphite as a microwave absorber at temperatures ranging from 800 to 1000 °C. They used a multimode microwave cavity oven with a flow of 10 mL/min in a quartz reactor, maintaining a constant power input of 1000 W and microwave frequency of 2450 MHz. The temperature of the sample was monitored throughout the experiment through an infrared optical pyrometer. They found that a higher amount of bio-oil was obtained from MAP and these oils were found to contain virtually no polycyclic aromatic hydrocarbons (PAHs), which are undesirable since they are carcinogenic (present in significant amounts in oil obtained from conventional pyrolysis). Moreover, Fernández and Menéndez [66] examined the conventional pyrolysis and MAP of a series of biomass wastes: sewage sludges, coffee hulls and glycerol using the same reactor and conditions as Domínguez et al. [65]. They found higher quantities of syngas and less CO2 in the gaseous components from MAP as compared to those from conventional pyrolysis. They also observed that glycerol gave the highest concentration of syngas out of all the feedstock, but it also gave the lowest H2/CO ratio. Sewage sludges exhibited the opposite tendency while coffee hulls showed an intermediate behaviour.
Liu et al. [67] investigated the MAP and conventional pyrolysis of sewage sludge over a range of temperatures of 600 to 900 °C. The reactor used for conventional pyrolysis was a horizontal fixed-bed quartz tubular one, having a 60 mm internal diameter (i.d.) × 900 mm length. It was heated in a 2200 W electrical furnace. The power used for MAP was 1400 W, the reactor a 2.45 GHz multimode tubular quartz one with a 45 mm i.d. and 750 mm length. A flow of 300 mL/min of N2 was used to ensure an inert atmosphere. Despite the two setups not being identical, the authors found that when the temperature was increased, the product yields for conventional pyrolysis varied significantly, while those for microwave pyrolysis showed only a slight change. For instance, during conventional pyrolysis, H2 yield went from 1.26 mmol/g at 600 °C to 9.07 mmol/g at 900 °C whereas the yield during MAP changed from 1.84 mmol/g to 3.67 mmol/g at the same temperatures. This difference was attributed to the fact that more hydrogen atoms were converted to tar compounds instead of being released into the syngas under MAP, as observed by a higher H/C ratio under MAP.
Furthermore, microwave-induced pyrolysis of palm oil biomass using char as a microwave absorber was carried out by Salema and Ani [68]. The authors used a 1 kW modified domestic microwave system with a 2450 MHz frequency. The microwave cavity was modified to accommodate a fluidised bed quartz reactor (0.1 m i.d. × 0.15 m height). Their setup is illustrated in Figure 6. They used a microwave power of 450 W and an exposure time of 25 min for each experiment. Different ratios of biomass to microwave absorber (1:0.25, 1:0.5 and 1:1) and an N2 flow rate of 20 L/min were used in the experiment series. They found the maximum bio-oil yield and lowest char yield at a biomass to microwave absorber ratio of 1:0.5, showing that the quantity of microwave absorbers or catalysts used did affect the pyrolysis product yields. They also concluded that the use of an eco-friendly microwave absorber (bio-char, in this case) can help operate the microwave reactor at a lower power input, thus reducing the need for high power MAP.

4.2. MAP in Fluidised Bed Reactors

Besides the study led by Salema and Ani [68] mentioned earlier, other researchers have also tried out MAP experiments in FBRs. One such example is Adam et al. [69], who developed a methodology for a microwave fluidised bed system based on processing raw biomass, without adding any microwave absorbents (Figure 7). The authors determined the minimum fluidisation velocity (umf) of the biomass used, sycamore, through cold fluidisation tests in a 50 mm i.d. and 70 cm height acrylic column. They found umf values of 0.06–0.81 m/s corresponding to 45–2360 µm particle sizes, respectively.
The authors found that particle sizes <600 µm exhibited homogeneous fluidisation, with no channel formation, contrary to bigger particle sizes. It was also found that a minimum power density of 54 W/m3 was needed to reach 400 °C; a higher power density value would enable reaching the temperature more quickly. According to the simulations run by Adam et al. [66], the optimum dimensions and operating conditions of the microwave cavity would be a height of 30 cm, a width of 24 cm, waveguide position of 20.10 mm × 162.50 mm (Y and Z positions, respectively), an input power of 6 kW and a microwave frequency of 2450 MHz. However, the authors noted that a compromise would be needed on the input power when designing an actual experimental setup.

4.3. Catalytic MAP

It was found that through the use of catalysts during MAP, the yield and quality of bio-oil may be enhanced [70]. However, depending on the type of catalyst used, the production of bio-char or syngas may be favoured at the expense of that of bio-oil. Ge et al. [71] and Morgan et al. [57] have made an exhaustive list of relevant studies having been performed on different catalysts and process parameters used during MAP of biomass in recent years. The most significant and pertinent ones to the aim of this review have been detailed further down this text.
As with conventional pyrolysis, MAP also includes two processes for the use of catalysts: in situ and ex situ catalytic MAP [57]. The difference between the two is that during ex situ MAP, the biomass and catalyst are not in contact with one another; the volatiles generally pass through a packed-bed catalyst reactor after they leave the quartz reactor where pyrolysis takes place under microwaves.
The most common type of catalysts used in MAP, as with conventional catalytic pyrolysis, are zeolites such as HZSM-5, H-Ferrierite, H-Modernite and H-Y. These catalysts, as mentioned earlier, are very effective in enhancing the properties of bio-oil through cracking [72]. One noteworthy article written by Muley et al. [73] deals with the ex situ catalytic upgrading of pine sawdust pyrolysis vapours using a zeolite catalyst in a travelling wave microwave reactor. The authors then compared the results obtained with those from conventional pyrolysis. A catalyst:biomass ratio of 1:2 was used and it was noticed that when catalyst temperature increased, liquid yield slightly decreased. The latter amount was also found to be comparable to that obtained from conventional pyrolysis (25–35%). The use of catalysts resulted in higher amounts of aromatic and aliphatic hydrocarbons while the non-upgraded bio-oil mostly consisted of ketones, phenols, alcohols and aldehydes. The gas yield, on the other hand, increased as the catalyst bed temperature increased. Gas composition went from CO (the highest) to CO2 to CH4 to finally, C2-C5 gases (the lowest). No mention of H2 was made. As for coke, a high C/H ratio was observed for coke that was deposited on the catalyst surface, making it aromatic in nature. This ratio increased with temperature when MAP was used.
A selective summary of recent notable research works published on microwave pyrolysis using zeolites as a catalyst has been presented in Table 6. It should also be noted that the table also contains those performed on pyrolysis or co-pyrolysis of plastic wastes since, as mentioned earlier, more recent research studies have been focusing on this type of feedstock.
Other types of catalysts used in the catalytic MAP of biomass are metal oxides, most commonly NiO, CaO, CuO and MgO [57]. Wu et al. [72] performed the catalytic MAP of waste cooking oil under temperatures ranging from 350 to 550 °C, using CoO, NiO, ZrO2, SrO, CeO2 and CaO as catalysts, both in- and ex situ. They also studied the reaction using a mixture of different ratios of HZSM-5 and metal oxides. The reactor used was a catalytic MAP system with a frequency of 2450 MHz and a power input of 0–800 W. The microwave-absorber used was 300 g of 5 mm diameter SiC beads.
Their results showed that using higher pyrolytic temperatures increased monocyclic aromatic hydrocarbon formation despite diminishing overall bio-oil yield. However, increasing the temperature from 500 to 500 °C only resulted in a 1.16% increase in benzene-toluene-xylene (BTX) production. Using metal oxides, especially CaO, as a catalyst, promoted the formation of polycyclic aromatic hydrocarbons, which were produced in significantly less proportions when only HZSM-5 was used as a catalyst. The authors also found that an ex situ configuration suited co-catalysis of metal oxides and HZSM-5 more than an in situ one. The highest BTX yield found by the authors (702.20 mg/mL) was produced at 500 °C, using an HZSM-5:CaO ratio of 2:1 in an ex situ configuration. Other researchers have also used metal oxides during their catalytic MAP experimental works. Some significant findings have been summarised in Table 7.
Apart from zeolites and metal oxides, carbon-based catalysts have also been considered good candidates for catalytic MAP studies as they can act as both catalysts and microwave absorbers [58]. One important research work was led by Omoriyekomwan et al. [82]. The latter studied phenol production from palm kernel shell pyrolysis at 400, 500 and 600 °C using a 2000 W microwave generator. The authors used 10% activated carbon or 20% lignite char as a catalyst. It was found that the addition of the catalysts enabled pyrolysis temperature to increase as biomass alone is a poor microwave absorber. Bio-oil yield also improved with the presence of the catalysts. The highest concentration of total phenolics was 71.24%, with the highest concentration of phenol of 64.58% in bio-oil, was obtained at 500 °C using activated carbon. Other research works performed using carbon-based catalysts for MAP include those made by Zhang et al. [83] and Chen et al. [84]. All in all, significant results included that carbon-based catalysts gave rise to higher heating rates. When these catalysts are doped with metal precursors, syngas yield increased importantly, while bio-oil yield was mostly impacted positively by the use of activated carbon as catalyst.

5. Conclusions

Even though conventional and microwave pyrolysis technologies are difficult to put on par because of the presence of microwave irradiation, a comparison between their respective products is helpful to shed some light on their efficiency. Recent research works performed on microwave pyrolysis of different feedstocks with and without the use of catalysts were reviewed in this work. This technology presents very favourable odds of being successfully scaled up to an industrial scale, given its practicality and interesting results. Indeed, given the fact that MAP has proven to have a more rapid volumetric heating, have more precise control and be more energy saving, it can be established to be more interesting than the conventional means of performing pyrolysis. However, as of today, the most common type of reactor used for industrial pyrolysis applications is the fluidised bed reactor [85]. MAP being a fairly recent technology, most of the research was conducted on fixed bed reactors. Few studies have been performed in FBR and they have mostly shown that these reactors need a fair microwave power for efficient pyrolysis to take place. So as to find a valid compromise, the use of an eco-friendly microwave absorber such as bio-char may be warranted. However, more detailed studies such as studying the reaction kinetics and numerical simulation of the overall process need to be performed in this direction so as to prove this technology’s worth in being used in upcoming industrial applications.

Author Contributions

Investigation, C.M.; writing—original draft preparation, C.M.; writing—review and editing, N.G., D.L. and Y.S.; supervision, Y.S.; funding acquisition, Y.S. All authors have read and agreed to the published version of the manuscript.

Funding

This project has received funding from the European Union’s Horizon 2020 research and innovation programme “Waste2Road” under grant agreement N° 818120.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Motasemi, F.; Afzal, M.T. A review on the microwave-assisted pyrolysis technique. Renew. Sustain. Energy Rev. 2013, 28, 317–330. [Google Scholar] [CrossRef]
  2. Hemansi; Gupta, R.; Yadav, G.; Kumar, G.; Yadav, A.; Saini, J.K.; Kuhad, R.C. Second Generation Bioethanol Production: The State of Art. In Sustainable Approaches for Biofuels Production Technologies: From Current Status to Practical Implementation; Srivastava, N., Srivastava, M., Mishra, P.K., Upadhyay, S.N., Ramteke, P.W., Gupta, V.K., Eds.; Springer International Publishing: Cham, Switzerland, 2019; pp. 121–146. [Google Scholar] [CrossRef]
  3. Maj, G.; Najda, A.; Klimek, K.; Balant, S. Estimation of Energy and Emissions Properties of Waste from Various Species of Mint in the Herbal Products Industry. Energies 2020, 13, 55. [Google Scholar] [CrossRef] [Green Version]
  4. Bridgwater, A.V. Review of fast pyrolysis of biomass and product upgrading. Biomass Bioenergy 2012, 38, 68–94. [Google Scholar] [CrossRef]
  5. Mettler, M.S.; Vlachos, D.G.; Dauenhauer, P.J. Top ten fundamental challenges of biomass pyrolysis for biofuels. Energy Environ. Sci. 2012, 5, 7797–7809. [Google Scholar] [CrossRef]
  6. Oasmaa, A.; Czernik, S. Fuel Oil Quality of Biomass Pyrolysis OilsState of the Art for the End Users. Energy Fuels 1999, 13, 914–921. [Google Scholar] [CrossRef]
  7. Venderbosch, R.H. Fast Pyrolysis. In Thermochemical Processing of Biomass; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2019; pp. 175–206. [Google Scholar] [CrossRef]
  8. Onwudili, J.; Insura, N.; Williams, P. Composition of products from the pyrolysis of polyethylene and polystyrene in a closed batch reactor: Effects of temperature and residence time. J. Anal. Appl. Pyrolysis 2009, 86, 293–303. [Google Scholar] [CrossRef]
  9. Deglise, X. La gazéification thermochimique: Histoire et développement de la recherche. In Colloque ATEE/CIBE: Cogénération Biomasse Dans L’industrie Et Sur Les Réseaux de Chaleur; 2007; p. 39. Available online: https://cibe.fr/wp-content/uploads/2017/01/12_-_Historique_gazeification.pdf (accessed on 28 April 2022).
  10. Fagbemi, L.; Khezami, L.; Capart, R. Pyrolysis products from different biomasses: Application to the thermal cracking of tar. Appl. Energy 2001, 69, 293–306. [Google Scholar] [CrossRef]
  11. Rousset, P. From Biomass to Fuel, Power and Chemicals Brazilian Charcoal-Based Pig Iron. 2014. Available online: http://agritrop.cirad.fr/578836/ (accessed on 6 August 2018).
  12. Garcia-Nunez, J.A.; Pelaez-Samaniego, M.R.; Garcia-Perez, M.; Fonts, I.; Ábrego, J.; Westerhof, R.J.M. Historical Developments of Pyrolysis Reactors: A Review. Energy Fuels 2017, 31, 5751–5775. [Google Scholar] [CrossRef]
  13. Jiang, J.; Xu, J.; Song, Z. Review of the direct thermochemical conversion of lignocellulosic biomass for liquid fuels. Front. Agric. Sci. Eng. 2015, 2, 13–27. [Google Scholar] [CrossRef] [Green Version]
  14. Ronsse, F. Biochar Production. In Biochar; Bruckman, V.J., Apaydin Varol, E., Uzun, B.B., Liu, J., Eds.; Cambridge University Press: Cambridge, UK, 2016; pp. 199–226. [Google Scholar] [CrossRef]
  15. Lam, S.S.; Liew, R.K.; Jusoh, A.; Chong, C.T.; Ani, F.N.; Chase, H.A. Progress in waste oil to sustainable energy, with emphasis on pyrolysis techniques. Renew. Sustain. Energy Rev. 2016, 53, 741–753. [Google Scholar] [CrossRef]
  16. Qureshi, K.M.; Lup, A.N.K.; Khan, S.; Abnisa, F.; Daud, W.M.A.W. A technical review on semi-continuous and continuous pyrolysis process of biomass to bio-oil. J. Anal. Appl. Pyrolysis 2018, 131, 52–75. [Google Scholar] [CrossRef]
  17. Mohabeer, C.; Reyes, L.; Abdelouahed, L.; Marcotte, S.; Buvat, J.-C.; Tidahy, L.; Abi-Aad, E.; Taouk, B. Production of liquid bio-fuel from catalytic de-oxygenation: Pyrolysis of beech wood and flax shives. J. Fuel Chem. Technol. 2019, 47, 153–166. [Google Scholar] [CrossRef]
  18. Si, Z.; Zhang, X.; Wang, C.; Ma, L.; Dong, R. An Overview on Catalytic Hydrodeoxygenation of Pyrolysis Oil and Its Model Compounds. Catalysts 2017, 7, 169. [Google Scholar] [CrossRef] [Green Version]
  19. Ambursa, M.M.; Juan, J.C.; Yahaya, Y.; Taufiq-Yap, Y.; Lin, Y.-C.; Lee, H.V. A review on catalytic hydrodeoxygenation of lignin to transportation fuels by using nickel-based catalysts. Renew. Sustain. Energy Rev. 2021, 138, 110667. [Google Scholar] [CrossRef]
  20. Wang, X.; Arai, M.; Wu, Q.; Zhang, C.; Zhao, F. Hydrodeoxygenation of lignin-derived phenolics—A review on the active sites of supported metal catalysts. Green Chem. 2020, 22, 8140–8168. [Google Scholar] [CrossRef]
  21. Bui, V.N.; Laurenti, D.; Afanasiev, P.; Geantet, C. Hydrodeoxygenation of guaiacol with CoMo catalysts. Part I: Promoting effect of cobalt on HDO selectivity and activity. Appl. Catal. B Environ. 2011, 101, 239–245. [Google Scholar] [CrossRef]
  22. Bui, V.N.; Laurenti, D.; Delichère, P.; Geantet, C. Hydrodeoxygenation of guaiacol: Part II: Support effect for CoMoS catalysts on HDO activity and selectivity. Appl. Catal. B Environ. 2011, 101, 246–255. [Google Scholar] [CrossRef]
  23. Bykova, M.; Ermakov, D.; Kaichev, V.; Bulavchenko, O.; Saraev, A.; Lebedev, M.; Yakovlev, V. Ni-based sol–gel catalysts as promising systems for crude bio-oil upgrading: Guaiacol hydrodeoxygenation study. Appl. Catal. B Environ. 2012, 113-114, 296–307. [Google Scholar] [CrossRef]
  24. Oh, S.; Choi, H.S.; Choi, I.-G.; Choi, J.W. Evaluation of hydrodeoxygenation reactivity of pyrolysis bio-oil with various Ni-based catalysts for improvement of fuel properties. RSC Adv. 2017, 7, 15116–15126. [Google Scholar] [CrossRef] [Green Version]
  25. Coumans, A.; Hensen, E. A real support effect on the hydrodeoxygenation of methyl oleate by sulfided NiMo catalysts. Catal. Today 2017, 298, 181–189. [Google Scholar] [CrossRef]
  26. Wang, W.; Zhang, C.; Chen, G.; Zhang, R. Influence of CeO2 Addition to Ni–Cu/HZSM-5 Catalysts on Hydrodeoxygenation of Bio-Oil. Appl. Sci. 2019, 9, 1257. [Google Scholar] [CrossRef] [Green Version]
  27. Funkenbusch, L.T.; Mullins, M.E.; Salam, M.A.; Creaser, D.; Olsson, L. Catalytic hydrotreatment of pyrolysis oil phenolic compounds over Pt/Al2O3 and Pd/C. Fuel 2019, 243, 441–448. [Google Scholar] [CrossRef]
  28. Ly, H.V.; Kim, J.; Hwang, H.T.; Choi, J.H.; Woo, H.C.; Kim, S.-S. Catalytic Hydrodeoxygenation of Fast Pyrolysis Bio-Oil from Saccharina japonica Alga for Bio-Oil Upgrading. Catalysts 2019, 9, 1043. [Google Scholar] [CrossRef] [Green Version]
  29. Kavimonica, V.; Parasuraman, S.; Ravikrishnan, V. Kinetic Studies of Catalytic Upgradation of Biomass Model Compounds Using Analytical Py-gc/ms. Chem. Eng. Trans. 2020, 80, 1–6. [Google Scholar] [CrossRef]
  30. Tran, Q.K.; Han, S.; Ly, H.V.; Kim, S.-S.; Kim, J. Hydrodeoxygenation of a bio-oil model compound derived from woody biomass using spray-pyrolysis-derived spherical γ-Al2O3-SiO2 catalysts. J. Ind. Eng. Chem. 2020, 92, 243–251. [Google Scholar] [CrossRef]
  31. Williams, P.; Horne, P.A. The influence of catalyst type on the composition of upgraded biomass pyrolysis oils. J. Anal. Appl. Pyrolysis 1995, 31, 39–61. [Google Scholar] [CrossRef]
  32. Park, H.J.; Dong, J.I.; Jeon, J.K.; Yoo, K.S.; Yim, J.H.; Sohn, J.M.; Park, Y.K. Conversion of the Pyrolytic Vapor of Radiata Pine over Zeolites. J. Ind. Eng. Chem. 2007, 13, 182–189. [Google Scholar]
  33. Park, H.J.; Park, Y.K.; Kim, J.S.; Jeon, J.K.; Yoo, K.S.; Yim, J.H.; Sohn, J.M. Bio-oil upgrading over Ga modified zeolites in a bubbling fluidized bed reactor. In Studies in Surface Science and Catalysis; Rhee, H.-K., Nam, I.-S., Park, J.M., Eds.; Elsevier: Amsterdam, The Netherlands, 2006; Volume 159, pp. 553–556. [Google Scholar] [CrossRef]
  34. Park, H.J.; Heo, H.S.; Jeon, J.-K.; Kim, J.; Ryoo, R.; Jeong, K.-E.; Park, Y.-K. Highly valuable chemicals production from catalytic upgrading of radiata pine sawdust-derived pyrolytic vapors over mesoporous MFI zeolites. Appl. Catal. B Environ. 2010, 95, 365–373. [Google Scholar] [CrossRef]
  35. Aho, A.; Kumar, N.; Lashkul, A.; Eränen, K.; Ziolek, M.; Decyk, P.; Salmi, T.; Holmbom, B.; Hupa, M.; Murzin, D.Y. Catalytic upgrading of woody biomass derived pyrolysis vapours over iron modified zeolites in a dual-fluidized bed reactor. Fuel 2010, 89, 1992–2000. [Google Scholar] [CrossRef]
  36. Zhang, H.; Carlson, T.R.; Xiao, R.; Huber, G.W. Catalytic fast pyrolysis of wood and alcohol mixtures in a fluidized bed reactor. Green Chem. 2012, 14, 98–110. [Google Scholar] [CrossRef] [Green Version]
  37. Murata, K.; Liu, Y.; Inaba, M.; Takahara, I. Catalytic fast pyrolysis of jatropha wastes. J. Anal. Appl. Pyrolysis 2012, 94, 75–82. [Google Scholar] [CrossRef]
  38. Choi, S.J.; Park, S.H.; Jeon, J.-K.; Lee, I.G.; Ryu, C.; Suh, D.J.; Park, Y.-K. Catalytic conversion of particle board over microporous catalysts. Renew. Energy 2013, 54, 105–110. [Google Scholar] [CrossRef]
  39. Guda, V.K.; Toghiani, H. Catalytic upgrading of pinewood fast pyrolysis vapors using an integrated Auger–packed bed reactor system: Effects of acid catalysts on yields and distribution of pyrolysis products. J. For. Prod. Ind. 2015, 4, 33–43. [Google Scholar]
  40. Jia, L.Y.; Raad, M.; Hamieh, S.; Toufaily, J.; Hamieh, T.; Bettahar, M.M.; Mauviel, G.; Tarrighi, M.; Pinard, L.; Dufour, A. Catalytic fast pyrolysis of biomass: Superior selectivity of hierarchical zeolites to aromatics. Green Chem. 2017, 19, 5442–5459. [Google Scholar] [CrossRef]
  41. Kaminsky, W. Chemical recycling of plastics by fluidized bed pyrolysis. Fuel Commun. 2021, 8, 100023. [Google Scholar] [CrossRef]
  42. 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] [Green Version]
  43. Zlotorzynski, A. The Application of Microwave Radiation to Analytical and Environmental Chemistry. Crit. Rev. Anal. Chem. 1995, 25, 43–76. [Google Scholar] [CrossRef]
  44. van Loock, W. European Regulations, Safety Issues in RF and Microwave Power. In Advances in Microwave and Radio Frequency Processing; Springer: Berlin/Heidelberg, Germany, 2006; pp. 85–91. [Google Scholar] [CrossRef]
  45. CEM Corporation. Microwave Heating-Mechanism and Theory. 2020. Available online: https://cem.com/de/microwave-heating-mechanism-and-theory (accessed on 30 March 2022).
  46. Kappe, C.O.; Stadler, A.; Dallinger, D. Microwaves in Organic and Medicinal Chemistry; John Wiley & Sons: Hoboken, NJ, USA, 2012. [Google Scholar]
  47. Xie, Q. Fast Microwave-Assisted Thermochemical Conversion of Biomass for Biofuel Production. 2015. Available online: http://conservancy.umn.edu/handle/11299/177123 (accessed on 14 September 2021).
  48. Nomanbhay, S.; Salman, B.; Hussain, R.; Ong, M.Y. Microwave pyrolysis of lignocellulosic biomass––A contribution to power Africa. Energy Sustain. Soc. 2017, 7, 23. [Google Scholar] [CrossRef] [Green Version]
  49. Priecel, P.; Lopez-Sanchez, J.A. Advantages and Limitations of Microwave Reactors: From Chemical Synthesis to the Catalytic Valorization of Biobased Chemicals. ACS Sustain. Chem. Eng. 2018, 7, 3–21. [Google Scholar] [CrossRef] [Green Version]
  50. Quitain, A.T.; Sasaki, M.; Goto, M. Microwave-Based Pretreatment for Efficient Biomass-to-Biofuel Conversion. In Pretreatment Techniques for Biofuels and Biorefineries; Fang, Z., Ed.; Springer: Berlin/Heidelberg, Germany, 2013; pp. 117–130. [Google Scholar] [CrossRef]
  51. Lam, S.S.; Chase, H.A. A Review on Waste to Energy Processes Using Microwave Pyrolysis. Energies 2012, 5, 4209–4232. [Google Scholar] [CrossRef]
  52. Ding, K.; Liu, S.; Huang, Y.; Liu, S.; Zhou, N.; Peng, P.; Wang, Y.; Chen, P.; Ruan, R. Catalytic microwave-assisted pyrolysis of plastic waste over NiO and HY for gasoline-range hydrocarbons production. Energy Convers. Manag. 2019, 196, 1316–1325. [Google Scholar] [CrossRef]
  53. Ethaib, S.; Omar, R.; Kamal, S.M.M.; Biak, D.R.A.; Zubaidi, S.L. Microwave-Assisted Pyrolysis of Biomass Waste: A Mini Review. Processes 2020, 8, 1190. [Google Scholar] [CrossRef]
  54. Ge, S.; Shi, Y.; Xia, C.; Huang, Z.; Manzo, M.; Cai, L.; Ma, H.; Zhang, S.; Jiang, J.; Sonne, C.; et al. Progress in pyrolysis conversion of waste into value-added liquid pyro-oil, with focus on heating source and machine learning analysis. Energy Convers. Manag. 2021, 245, 114638. [Google Scholar] [CrossRef]
  55. Giorcelli, M.; Das, O.; Sas, G.; Försth, M.; Bartoli, M. A Review of Bio-Oil Production through Microwave-Assisted Pyrolysis. Processes 2021, 9, 561. [Google Scholar] [CrossRef]
  56. Morgan, H.M.; Bu, Q.; Liang, J.; Liu, Y.; Mao, H.; Shi, A.; Lei, H.; Ruan, R. A review of catalytic microwave pyrolysis of lignocellulosic biomass for value-added fuel and chemicals. Bioresour. Technol. 2017, 230, 112–121. [Google Scholar] [CrossRef]
  57. State, R.N.; Volceanov, A.; Muley, P.; Boldor, D. A review of catalysts used in microwave assisted pyrolysis and gasification. Bioresour. Technol. 2019, 277, 179–194. [Google Scholar] [CrossRef]
  58. Suresh, A.; Alagusundaram, A.; Kumar, P.S.; Vo, D.-V.N.; Christopher, F.C.; Balaji, B.; Viswanathan, V.; Sankar, S. Microwave pyrolysis of coal, biomass and plastic waste: A review. Environ. Chem. Lett. 2021, 19, 3609–3629. [Google Scholar] [CrossRef]
  59. Muley, P.D.; Wang, Y.; Hu, J.; Shekhawat, D. Microwave-assisted heterogeneous catalysis. Catalysis 2021, 33, 1–37. [Google Scholar] [CrossRef]
  60. Mokhtar, N.M.; Ethaib, S.; Omar, R. Effects of microwave absorbers on the products of microwave pyrolysis of oily sludge. J. Eng. Sci. Technol. 2018, 13, 3313–3330. [Google Scholar]
  61. Du, Z.; Li, Y.; Wang, X.; Wan, Y.; Chen, Q.; Wang, C.; Lin, X.; Liu, Y.; Chen, P.; Ruan, R. Microwave-assisted pyrolysis of microalgae for biofuel production. Bioresour. Technol. 2011, 102, 4890–4896. [Google Scholar] [CrossRef]
  62. Hu, Z.; Ma, X.; Chen, C. A study on experimental characteristic of microwave-assisted pyrolysis of microalgae. Bioresour. Technol. 2012, 107, 487–493. [Google Scholar] [CrossRef] [PubMed]
  63. Tian, Y.; Zuo, W.; Ren, Z.; Chen, D. Estimation of a novel method to produce bio-oil from sewage sludge by microwave pyrolysis with the consideration of efficiency and safety. Bioresour. Technol. 2011, 102, 2053–2061. [Google Scholar] [CrossRef] [PubMed]
  64. Yu, F.; Deng, S.; Chen, P.; Liu, Y.; Wan, Y.; Olson, A.; Kittelson, D.; Ruan, R. Physical and chemical properties of bio-oils from microwave pyrolysis of corn stover. Appl. Biochem. Biotechnol. 2007, 137–140, 957–970. [Google Scholar] [CrossRef]
  65. Domínguez, A.; Menéndez, J.A.; Inguanzo, M.; Bernad, P.L.; Pis, J.J. Gas chromatographic–mass spectrometric study of the oil fractions produced by microwave-assisted pyrolysis of different sewage sludges. J. Chromatogr. A 2003, 1012, 193–206. [Google Scholar] [CrossRef]
  66. Fernández, Y.; Menéndez, J. Influence of feed characteristics on the microwave-assisted pyrolysis used to produce syngas from biomass wastes. J. Anal. Appl. Pyrolysis 2011, 91, 316–322. [Google Scholar] [CrossRef]
  67. Liu, Y.; Chen, T.; Gao, B.; Meng, R.; Zhou, P.; Chen, G.; Zhan, Y.; Lu, W.; Wang, H. Comparison between hydrogen-rich biogas production from conventional pyrolysis and microwave pyrolysis of sewage sludge: Is microwave pyrolysis always better in the whole temperature range? Int. J. Hydrogen Energy 2021, 46, 23322–23333. [Google Scholar] [CrossRef]
  68. Salema, A.; Ani, F.N. Microwave induced pyrolysis of oil palm biomass. Bioresour. Technol. 2011, 102, 3388–3395. [Google Scholar] [CrossRef]
  69. Adam, M.; Beneroso, D.; Katrib, J.; Kingman, S.; Robinson, J.P. Microwave fluidized bed for biomass pyrolysis. Part I: Process design. Biofuels Bioprod. Biorefining 2017, 11, 601–612. [Google Scholar] [CrossRef]
  70. Ong, H.C.; Chen, W.-H.; Farooq, A.; Gan, Y.Y.; Lee, K.T.; Ashokkumar, V. Catalytic thermochemical conversion of biomass for biofuel production: A comprehensive review. Renew. Sustain. Energy Rev. 2019, 113, 109266. [Google Scholar] [CrossRef]
  71. Ge, S.; Yek, P.N.Y.; Cheng, Y.W.; Xia, C.; Mahari, W.A.W.; Liew, R.K.; Peng, W.; Yuan, T.-Q.; Tabatabaei, M.; Aghbashlo, M.; et al. Progress in microwave pyrolysis conversion of agricultural waste to value-added biofuels: A batch to continuous approach. Renew. Sustain. Energy Rev. 2021, 135, 110148. [Google Scholar] [CrossRef]
  72. Wu, Q.; Wang, Y.; Peng, Y.; Ke, L.; Yang, Q.; Jiang, L.; Dai, L.; Liu, Y.; Ruan, R.; Xia, D.; et al. Microwave-assisted pyrolysis of waste cooking oil for hydrocarbon bio-oil over metal oxides and HZSM-5 catalysts. Energy Conv. Manag. 2020, 220, 113124. [Google Scholar] [CrossRef]
  73. Muley, P.; Henkel, C.; Aguilar, G.; Klasson, K.; Boldor, D. Ex situ thermo-catalytic upgrading of biomass pyrolysis vapors using a traveling wave microwave reactor. Appl. Energy 2016, 183, 995–1004. [Google Scholar] [CrossRef] [Green Version]
  74. Zeng, Y.; Wang, Y.; Liu, Y.; Dai, L.; Wu, Q.; Xia, M.; Zhang, S.; Ke, L.; Zou, R.; Ruan, R. Microwave catalytic co-pyrolysis of waste cooking oil and low-density polyethylene to produce monocyclic aromatic hydrocarbons: Effect of different catalysts and pyrolysis parameters. Sci. Total Environ. 2022, 809, 152182. [Google Scholar] [CrossRef] [PubMed]
  75. Xue, Z.; Zhong, Z.; Zhang, B. Microwave-Assisted Catalytic Fast Pyrolysis of Biomass for Hydrocarbon Production with Physically Mixed MCM-41 and ZSM-5. Catalysts 2020, 10, 685. [Google Scholar] [CrossRef]
  76. Suriapparao, D.V.; Vinu, R.; Shukla, A.; Haldar, S. Effective deoxygenation for the production of liquid biofuels via microwave assisted co-pyrolysis of agro residues and waste plastics combined with catalytic upgradation. Bioresour. Technol. 2020, 302, 122775. [Google Scholar] [CrossRef]
  77. Huang, Y.-F.; Kuan, W.-H.; Chang, C.-C.; Tzou, Y.-M. Catalytic and atmospheric effects on microwave pyrolysis of corn stover. Bioresour. Technol. 2013, 131, 274–280. [Google Scholar] [CrossRef]
  78. Kuan, W.-H.; Huang, Y.-F.; Chang, C.-C.; Lo, S.-L. Catalytic pyrolysis of sugarcane bagasse by using microwave heating. Bioresour. Technol. 2013, 146, 324–329. [Google Scholar] [CrossRef]
  79. Yu, Y.; Yu, J.; Sun, B.; Yan, Z. Influence of catalyst types on the microwave-induced pyrolysis of sewage sludge. J. Anal. Appl. Pyrolysis 2014, 106, 86–91. [Google Scholar] [CrossRef]
  80. Ma, R.; Huang, X.; Zhou, Y.; Fang, L.; Sun, S.; Zhang, P.; Zhang, X.; Zhao, X. The effects of catalysts on the conversion of organic matter and bio-fuel production in the microwave pyrolysis of sludge at different temperatures. Bioresour. Technol. 2017, 238, 616–623. [Google Scholar] [CrossRef]
  81. Fan, L.; Chen, P.; Zhang, Y.; Liu, S.; Liu, Y.; Wang, Y.; Dai, L.; Ruan, R. Fast microwave-assisted catalytic co-pyrolysis of lignin and low-density polyethylene with HZSM-5 and MgO for improved bio-oil yield and quality. Bioresour. Technol. 2017, 225, 199–205. [Google Scholar] [CrossRef]
  82. Omoriyekomwan, J.E.; Tahmasebi, A.; Yu, J. Production of phenol-rich bio-oil during catalytic fixed-bed and microwave pyrolysis of palm kernel shell. Bioresour. Technol. 2016, 207, 188–196. [Google Scholar] [CrossRef] [PubMed]
  83. Zhang, S.; Dong, Q.; Zhang, L.; Xiong, Y. High quality syngas production from microwave pyrolysis of rice husk with char-supported metallic catalysts. Bioresour. Technol. 2015, 191, 17–23. [Google Scholar] [CrossRef] [PubMed]
  84. Chen, W.; Chen, Y.; Yin, J.; Li, Z. Bi-objective Mixed Optimal Planning for Distributed Energy Storage System of Active Distribution System. In Proceedings of the 2018 8th International Conference on Power and Energy Systems (ICPES), Colombo, Sri Lanka, 21–22 December 2018; pp. 34–39. [Google Scholar] [CrossRef]
  85. Guda, V.K.; Steele, P.H.; Penmetsa, V.K.; Li, Q. Chapter 7-Fast Pyrolysis of Biomass: Recent Advances in Fast Pyrolysis Technology. In Recent Advances in Thermo-Chemical Conversion of Biomass; Pandey, A., Bhaskar, T., Stöcker, M., Sukumaran, R.K., Eds.; Elsevier: Boston, MA, USA, 2015; pp. 177–211. [Google Scholar] [CrossRef]
Figure 1. Pyrolysis technologies classification (Adapted from [9]).
Figure 1. Pyrolysis technologies classification (Adapted from [9]).
Energies 15 03258 g001
Figure 2. Biomass pyrolysis products (Adapted from [10]).
Figure 2. Biomass pyrolysis products (Adapted from [10]).
Energies 15 03258 g002
Figure 3. Electric (E) and magnetic (H) field components in microwaves [45].
Figure 3. Electric (E) and magnetic (H) field components in microwaves [45].
Energies 15 03258 g003
Figure 4. Microwave and conventional heating systems (Adapted from [1]).
Figure 4. Microwave and conventional heating systems (Adapted from [1]).
Energies 15 03258 g004
Figure 5. Schematic of microwave reactor as used by Mokhtar et al. (Adapted from [60]).
Figure 5. Schematic of microwave reactor as used by Mokhtar et al. (Adapted from [60]).
Energies 15 03258 g005
Figure 6. MAP setup used by Salema and Ani (Adapted from [68]).
Figure 6. MAP setup used by Salema and Ani (Adapted from [68]).
Energies 15 03258 g006
Figure 7. MAP FBR used by Adam et al. (Adapted from [69]).
Figure 7. MAP FBR used by Adam et al. (Adapted from [69]).
Energies 15 03258 g007
Table 1. Biofuel resource classification (Adapted from [1]).
Table 1. Biofuel resource classification (Adapted from [1]).
First-Generation SourcesVirgin resourcesForest resourcesWood such as pine, woody and forest biomass such as willow, poplar, aspen
Oil cropsWheat, barley, corn, canola, rapeseed, palm oil, soybean, sugarcane, flax, oat, straw, pasture grasses
Second-Generation SourcesResiduesWood residuesBark, branches, leftover treetop, leaves from harvest, sawdust, shavings from pulp and sawmills
Agricultural residues and wasteResidual fraction from oil crop harvest, waste oil/fat
Livestock residuesLivestock excrement and carcass
Municipal solid waste (MSW)ResidentialCardboard and mixed paper, glass, various metals, electronics, plastics, tyres, organics
Non-residentialSewage sludge
Third-Generation SourcesAlgae-Macroalgae, microalgae
Table 2. Notable research works on catalytic HDO.
Table 2. Notable research works on catalytic HDO.
FeedReaction ConditionsResultsRef.
Catalyst UsedReactor UsedTemperature (°C), Pressure (bar), H2 PercentageWHSV *
(g feed/g cat. h), Contact Time (s) or F/C * Ratio
Bio-Oil Yield (wt. %)Products and Conclusions
GuaiacolMoS2Fixed bed300 °C, 40 bar,
100% H2
0.3–1 s-HDO transformation of guaiacol was significantly enhanced by promoting MoS2-based catalysts with Co. However, the use of Al2O3 as support gave rise to demethylation and production of methyl-substituted products, which render de-oxygenation more difficult.[21]
CoMoS
MoS2/Al2O3
CoMoS/Al2O3
GuaiacolMoS2/γ-Al2O3Fixed bed300 °C, 40 bar,
100% H2
0.3–1 s-Out of MoS2-based catalysts, the TiO2-supported one gave the most interesting improvements. As for CoMoS, ZrO2 proved to have the most significant catalytic activities. The CoMoS/ZrO2 system was very selective towards aromatic (C-O) hydrogenolysis as demethoxylation and direct de-oxygenation were both observed.[22]
MoS2/TiO2
MoS2/ZrO2
CoMo/γ-Al2O3
CoMo/TiO2
CoMo/ZrO2
GuaiacolNiCu/
CeO2-ZrO2
Batch280–360 °C, 170 bar, 100% H2337.5:1-NiCu/SiO2-ZrO2-La2O3 was the most efficient catalyst. Deoxygenation degree increased with temperature, while guaiacol conversion decreased due to catalyst coking at higher temperatures.[23]
NiCu/Al2O3
NiCu/SiO2
NiCu/SiO2-ZrO2-La2O3
Ni/SiO2
Ni-Cu-MgO
Yellow poplar
bio-oil
Ni/CAutoclave300 °C, 30 bar,
100% H2
25:163.8The mesoporous silica supports (SBA-15 and Al-SBA-15) showed greater catalytic activity than Ni/C. this could be correlated to a higher surface area and pore size, which caused uniform dispersion of NI particles.[24]
Ni/SBA-1576.0
Ni/Al-SBA-1574.8
Methyl oleate
(as model compound for green diesel)
NiMo/CFixed bed260 °C, 30 bar,
90% H2 and 10% H2S
6.5 g feed/g cat. h-NiMo/Al2O3 and NiMo/ASA showed the highest global HDO activity, however, they de-activated during the run—owing to Lewis acidity of Al surface active in methyl oleate hydrolysis. NiMo/SiO2 had similar overall activity than the two previous, but did not de-activate. NiMo/C did not de-activate either, and was more active than the others at the end of the run.[25]
NiMo/SiO2
NiMo/ASA
NiMo/Al2O3
Rice husk bio-oilNi-Cu/HZSM-5Autoclave270 °C, 20 bar,
100% H2
12:133.9Addition of CeO2 to Ni-Cu/HZSM-5 improved the performance of the catalyst. It caused an improvement in the Ni dispersion, redox ability and the Bronsted acidity ratio. It also decreased the particle size of the catalyst and diminished coke deposition on the catalyst surface, which causes catalyst deactivation. In total, 15% CeO2-Ni-Cu/HZSM-5 was the best in terms of catalyst efficiency.[26]
5% CeO2-Ni-Cu/
HZSM-5
37.4
15% CeO2-Ni-Cu/
HZSM-5
47.6
20 % CeO2-Ni-Cu/
HZSM-5
35.3
Anisole,
m-cresol and phenol
5% Pd/CParr batch250–350 °C, 50 bar,
100% H2
-Residence time of 4 h allowed the formation of high yields of deoxygenated compounds. Pt/Al2O3 promoted hydrogenation (ring saturation) and removal of the pendant groups, and a significant pathway shift was observed as temperature increased. Pd/C showed ring saturation followed by methanol abstraction.[27]
5% Pt/Al2O3
Saccharina Japonica bio-oilHZSM-5Autoclave350 °C, 3–15 bar,
100% H2
10:180.4Pressures of 3 to 15 bar were tested and 15 bar provided the highest bio-oil yield; all catalytic tests were hence performed at 15 bar. HHV* of HDO-upgraded bio-oils improved with use of Co/γ-Al2O3, but decreased with use of metal phosphide catalysts. The latter promoted decarboxylation while metal catalysts elevated demethylation. An increase in kerosene fraction of bio-oil was also observed with the use of catalytic HDO.[28]
Co/γ-Al2O377.6
Fe//γ-Al2O382.1
CoP//γ-Al2O382.7
Fe2P//γ-Al2O378.8
CoMoP//γ-Al2O368.0
Guaiacol and
furfural
Zeolite Y
(Si/Al: 5.1:1)
Pyroprobe (Py-GC/MS)500 °C, 1.01 bar,
100% H2
1:10-Conversion over catalysts with low Si/Al ratios gave rise to the production of aromatics (benzene, toluene, xylene and phenol). Highest benzene yield was 19.5 wt. % over zeolite Y (30:1). Trend of overall BTX yield was: Zeolite Y (30:1), 30.5% > Zeolite Y (5.1:1), 28.1% > Zeolite Y (80:1), 12.0% > Zeolite Y (60:1), 6.4%. It was also shown that zeolite acidity played an important role in the deoxygenation of guaiacol, but not in that of furfural.[29]
Zeolite Y
(Si/Al: 30:1)
Zeolite Y
(Si/Al: 60:1)
Zeolite Y
(Si/Al: 80:1)
Guaiacolγ-Al2O3-SiO2Fixed bed275 or 300 °C,
1.01 bar, 25% H2
6.50 g guaiacol/g cat. h-Catalyst with a 50:50 Al/Si ratio after calcination at 450 °C exhibited the highest guaiacol conversion (81.79%).[30]
* WHSV: Weight hourly space velocity. F/C ratio: Feed to catalyst ratio. HHV: Higher heating value.
Table 3. Notable research works on CFP using zeolite-based catalysts.
Table 3. Notable research works on CFP using zeolite-based catalysts.
FeedReaction ConditionsResultsRef.
Catalyst UsedReactor UsedTemperature (°C)WHSV
(g feed/g cat. h) or F/C Ratio
Bio-Oil Yield (wt. %)Product Quality, Oxygen Content and Conclusions
Waste wood shavingsHZSM-5Fluidised bed500–5501.05–1.14 g
biomass/g cat. h
28.8Good aromatics yield (15.9 wt. %)[31]
Na-ZSM-524.3Higher aromatics yield (21.3 wt. %)
Y-Zeolite28.6Average aromatics yield
(14.0 wt. %)
Activated
alumina
31.2Lowest aromatics yield (8.0 wt. %)
Radiata pine
sawdust
HZSM-5Bubbling
fluidised bed
475–6254 g biomass/g cat. h43.7Formation of mainly aromatic hydrocarbons[32,33]
H-Y8 g biomass/g cat. h45.7Almost no formation of aromatics
Ga/HZSM-54 g biomass/g cat. h51.3Higher selectivity for aromatics
(as compared to HZSM-5)
Radiata pine
sawdust
HZSM-5Fixed bed50010:146.6Reduction of non-phenolic oxygenates[34]
MMZHZSM-5 (mesoporous material from HZSM-5)50.6No change in phenols; reduction in other oxygenates
MFI zeolite45.9High aromatics yield; reduction in non-phenolic oxygenates and in phenols
1% Ga/MFI
5% Ga/MFI Similar to HZSM-5; less efficient than 1% Ga/MFI
Pine sawdustH-βDual
fluidised bed
4508.33:137.8β-zeolite was the most efficient for deoxygenation, followed by Y and ferrierite zeolites[35]
Fe-H-β37.0
H-Y39.7
Fe-H-Y34.3
H-Ferrierite43.8
Fe-H-Ferrierite44.5
Pine woodZSM-5Fluidised bed450–6000.35 g biomass/g cat. h-Increase in selectivity of aromatics[36]
Jatropha wastesHZSM-5Fixed bed5502:14.1–8.7Conversion of 76.7–91.6% found; production of mainly aromatics[37]
Particle boardHZSM-5Batch50010:142.5Reduction in oxygenates and increase in aromatics and phenolics[38]
Ga/HZSM-546.3Larger reduction in oxygenates and higher increase in aromatics (compared to HZSM-5); lower phenolic content
H-β44.6Similar to HZSM-5, but less efficient
Pine woodβ/Al2O3Auger45012 g biomass/g cat. h52.0Si/Al and β-zeolite contributed to increase of liquid product while HZSM-5 and H-Y led to formation of aromatics and higher gas percentage. HZSM-5 reduced oxygen content of 46.4 wt. % to 30 wt. %.[39]
Si/Al51.0
H-Y43.0
HZSM-541.0
Oak wood
cylinder particles
HZSM-5Micro
fluidised bed
5000.85:18.3–10.1High selectivity in monoaromatic compounds (4.4 wt. %); CO2:CO = 0.5[40]
Desilicated HZSM-5Higher selectivity in monoaromatic compounds (6.2 wt. %); same CO2:CO
HZSM-5Fixed bed-High selectivity in monoaromatics (4.1 wt. %), but less efficient than FBR *
Desilicated HZSM-5Higher selectivity than parent HZSM-5
(5.1 wt. %), but less efficient than FBR *
Beech wood (BW)HZSM-5Semi-
continuous
5004:150.0Fe-HZSM-5 was the most efficient in deoxygenation (33.82 to 17.50% for BW and 34.76 to 17.31% for FS). Zeolites were found to cause decomposition of carboxylic acids to form phenols, H2, H2O, CO2 and CO, mostly. Pt and CoMo-based catalysts did not show high efficiencies; these catalysts might be better suited for HDO applications.[17]
Fe-HZSM-558.8
H-Y49.0
Fe-H-Y61.0
Pt/Al2O360.0
CoMo/Al2O354.0
Flax shives (FS)HZSM-550.0
Fe-HZSM-558.8
H-Y41.0
Fe-H-Y51.0
Pt/Al2O352.0
CoMo/Al2O347.0
* FBR: Fluidised bed reactor.
Table 4. Results of pyrolysis of different plastics by Kaminsky [Adapted from 41].
Table 4. Results of pyrolysis of different plastics by Kaminsky [Adapted from 41].
FeedTemperature (°C)Fluidising GasProducts
Polyethylene (PE), Polypropylene (PP)400–550Nitrogen, steam and pyrolysis gas70–90% waxy products, oil,
2–9% gas
Polymethylmethacrylate (PMMA), Polystyrene (PS)400–500Pyrolysis gas70–95% monomers, 5–30% gas, oil
Polyethylene terephthalate (PET)400–550Steam45–60% monomers, 40–45% gas, oil
PE, PP700–800Nitrogen, steam70–80% gas (olefins), 20–30% oil
PE, PP, polyamide (PA)700–800Pyrolysis gas30–50% aromatics, 30–40% oil, gas, 1–10% soot
Table 5. Microwave-assisted heating compared to conventional heating.
Table 5. Microwave-assisted heating compared to conventional heating.
Microwave-Assisted HeatingConventional Heating
Conversion of energyTransfer of energy
Uniform heating at a molecular level from coreSuperficial heating through conduction, convection and radiation
Frequent hot spot formationRarer hot spot formation
RapidSlow
Higher electricity conversion efficiencyLow electricity conversion efficiency
SelectiveNon-selective
Dependent on material propertiesLess dependent
Precise heatingLess precise
Lower thermal inertia and faster responseHigher thermal inertia and slower response
Table 6. Summary of recent works on catalytic MAP using zeolites as catalyst.
Table 6. Summary of recent works on catalytic MAP using zeolites as catalyst.
FeedReaction ConditionsResultsRef.
Catalyst UsedReactor Used
(Power and Frequency)
Temperature (°C)WHSV
(g feed/g cat. h) or F/C Ratio
Bio-Oil Yield (wt. %)Products and Conclusions
Waste cooking oil (WCO) and low-density polyethylene (LDPE)HZSM-5Fixed bed (-)5502:1-WCO:LDPE used was 1:1. Reaction conditions used produced maximum yield of BTX and minimum yield of polyaromatic hydrocarbons. Efficiency of catalysts on formation increased as follows: SAPO-34 < H-β < H-Y < HZSM-5.[74]
H-β
H-Y
SAPO-34
Corn strawMCM-41 mixed with ZSM-5Fixed bed
(750 W and 2450 MHz)
450–6501:2-Carbon outputs of aromatics and olefins peaked at 550 °C. Addition of MCM-41 inhibited formation of coke on ZSM-5 surface.[75]
Rice strawHZSM-5Fixed bed
(3.2 kW and 2.45 GHz)
50044.4, 14.7, 11.1 g feed/g cat. h21.5Catalytic upgradation of pyrolytic vapours of biomass and polymer mixtures produced de-oxygenated bio-oils with properties similar to conventional fuel oil. WHSP of 11 yielded higher selectivity to unsaturated aliphatics and aromatic hydrocarbons.[76]
Bagasse22.4
PP74.2
PS92.3
LDPENiO2 stage fixed bed
(1.8 kW and 2.45 GHz)
450–60020:1, 15:1, 10:1, 5:132–57Optimum conditions were found to be 500 °C pyrolysis temperature, 450 °C catalytic temperature and LDPE:H-Y of 10:1. Results obtained at these conditions were 56.53 wt. % oil and 93.80% gasoline-range fraction. Addition of catalysts favoured formation of aromatics and inhibited that of aliphatics.[53]
H-Y
Table 7. Notable works on MAP using metal oxides as catalyst.
Table 7. Notable works on MAP using metal oxides as catalyst.
ConfigurationCatalystFeedstockMicrowave Power (W)Feedstock:Catalyst RatioPyrolytic Temperature (°C)FindingsRef.
In situNiO, CuO, CaO, MgOCorn stover50010:1–33:1450–520The authors found that pyrolysis under an N2 atmosphere was more efficient than one under a CO2 atmosphere due to the fact that CO2 possesses better heat absorbability and so, its use reduced heat for pyrolysis. The catalysts were found to increase maximum temperature and mass reduction ratio, but they also decreased the solids’ calorific values. Moreover, use of catalysts diminished PAHs formation, making the bio-oil less toxic.[77]
In situNiO, CuO, CaO, MgOSugarcane
bagasse
50010:1–33:1490–532Addition of catalysts resulted in an increase in mass reduction ratio and reaction rate, but in a decrease in maximum temperature. Use of NiO and CaO enhanced H2 production while use of NiO and CuO lessened it. Addition of CaO and MgO improved gas production while NiO and CuO favoured liquid production.[78]
In situCaO, CaCO3, NiO, Ni2O3, γ-Al2O3, TiO2Sewage sludge70010:1-Addition of catalysts was found to affect temperature evolution of sludge, product distribution and gas composition. The temperature rise rate was found to be highest with CaCO3, followed by, respectively, NiO, TiO2, Ni2O3 ≈ γ-Al2O3 and CaO, which caused virtually no temperature rise. Ni-based catalysts used favoured decomposition of organic matters in sludge and highly increased bio-oil and CO-rich syngas yields. CaO gave rise to a H2-rich syngas while γ-Al2O3 and TiO2 showed no impact on gas percentage or H2:CO ratio.[79]
In situCaO, Fe2O3Sludge-10:1500–900Fe2O3 was found to favour bio-oil production while CaO improved gas formation. The best quality of bio-oil was obtained at 800 °C while using CaO. It was also observed that CaO privileged H2 formation while Fe2O3 enhanced CH4 production.[80]
Ex situHZSM-5, MgOLow-density
polyethylene (LDPE)
7500, 1:2, 1:1 and 2:1450–600LDPE was found to be a good H2 donor to improve bio-oil properties; methoxy phenols were converted to phenols and alkylated phenols. HZSM-5 favoured aromatics production while MgO improved alkylation of phenols. Optimum parameters were found to be 500 °C and feedstock to catalyst ratio of 1:1.[81]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mohabeer, C.; Guilhaume, N.; Laurenti, D.; Schuurman, Y. Microwave-Assisted Pyrolysis of Biomass with and without Use of Catalyst in a Fluidised Bed Reactor: A Review. Energies 2022, 15, 3258. https://doi.org/10.3390/en15093258

AMA Style

Mohabeer C, Guilhaume N, Laurenti D, Schuurman Y. Microwave-Assisted Pyrolysis of Biomass with and without Use of Catalyst in a Fluidised Bed Reactor: A Review. Energies. 2022; 15(9):3258. https://doi.org/10.3390/en15093258

Chicago/Turabian Style

Mohabeer, Chetna, Nolven Guilhaume, Dorothée Laurenti, and Yves Schuurman. 2022. "Microwave-Assisted Pyrolysis of Biomass with and without Use of Catalyst in a Fluidised Bed Reactor: A Review" Energies 15, no. 9: 3258. https://doi.org/10.3390/en15093258

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